Physical Geology, First University of Saskatchewan Edition

Physical Geology, First University of Saskatchewan Edition

Adapted from Physical Geology written by Steven Earle for the BCcampus Open Textbook Project

Karla Panchuk

Contents

1

Download PDF Files

Welcome!

image

Click to download this textbook as a PDF file. (File size: 100 MB)

Individual Chapter Downloads

Use the following links to download PDF files of individual chapters.

Chapter 1. Introduction to Geology (1st U of S Ed.) Updated 10-01-2019  Get PDF or Read online

Chapter 2. The Origin of Earth and the Solar System (1st U of S Ed.) Updated 10-01-2019  Get PDF or Read online

Chapter 3. Earth’s Interior (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 4. Plate Tectonics (1st U of S Ed.) Updated 10-01-2019 Get PDF or  Read online

Chapter 5. Minerals (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 6. The Rock Cycle (1st U of S Ed.) Updated 10-01-2019  Get PDF or Read online

Chapter 7. Igneous Rocks (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 8. Weathering, Sediment, and Soil (1st U of S Ed.) Updated 10-01-2019 Get PDF or  Read online

Chapter 9. Sedimentary Rocks (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 10. Metamorphism and Metamorphic Rocks (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 11. Volcanism (1st U of S Ed.) Updated 10-01-2019  Get PDF or Read online

Chapter 12. Earthquakes  (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 13. Geological Structures and Mountain Building (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 14. Streams and Floods (1st U of S Ed.) Updated 17-1-2019 Get PDF or Read online

Chapter 15. Mass Wasting (1st U of S Ed.) Updated 17-1-2019 Get PDF or  Read online

Chapter 16. Earth-System Change (1st U of S Ed.) Updated 10-01-2019. Get PDF or Read online

Chapter 17. Glaciation (1st U of S Ed.) Updated 10-01-2019 Get PDF or Read online

Chapter 18. Geological Resources (2nd Adapted Ed.) Updated 1-5-2017 Get PDF or Read online

Chapter 19. Measuring Geological Time (1st U of S Ed.). Updated 10-01-2019 Get PDF or Read online

Links to Chapters in Physical Geology by Steven Earle (2015):

Chapter 14. Groundwater

Chapter 18. Geology of the Oceans

Chapter 21. Geological History of Western Canada

 

Cover image. View of Howe Sound a fjord on the south coast of B.C., Canada. This photo was taken from the South Summit of Stawamous Chief, a granodiorite pluton that is part of the coast mountains of B.C. Source: Joyce M. McBeth (2018) CC BY 4.0.

3

Acknowledgments

An open textbook for physical geology is something I had been considering ever since taking the Introduction to Learning Technologies course at the Gwenna Moss Centre for Teaching and Learning at the University of Saskatchewan. Adapting an open textbook is a far less daunting task than starting from scratch so I was excited to hear of the textbook Physical Geology by Steven Earle, written for the BCcampus Open Textbook project. Steven’s original edition was a comprehensive and solid foundation on which to build this adapted work.  Thanks to Amanda Coolidge of BCcampus for saving me an enormous amount of time by explaining how to modify the text and sending me the exported files from Steven’s version of the textbook.

Many thanks go to Heather Ross and Nancy Turner at the Gwenna Moss Centre for their support and encouragement on this project and for discussions with them about open textbooks. The University of Saskatchewan Open Educational Resources Fund provided funding to support my work on this project. In-kind work and assistance on the project to match my time for this funding were provided by Joyce McBeth and Tim Prokopiuk of the Department of Geological Sciences.

This book has benefited from the work of numerous contributors at the University of Saskatchewan who have assisted with editing the document and providing new images to include in this edition. Tim Prokopiuk contributed edits and selected rock samples for me to photograph from the department’s collection. Joyce McBeth provided numerous edits to this edition and adapted Chapters 14, 15, and 17. Lyndsay Hauber provided assistance with updates to image attributions for the chapter on plate tectonics. Donna Beneteau and Doug Milne of the College of Engineering, and Zoli Hajnal of Geological Sciences gave me a tour of the Geological Engineering Rock Mechanics Facility, and helped me to photograph their experiments.

Image Sources

This project would not be possible without the generosity of many individuals and organizations who shared their work with a Creative Commons license or under other open licensing terms. The following is a list of valuable image resources, as much as it is an acknowledgement of contributions:

Roger Weller has made available thousands of his high-quality rock and mineral photographs through his website hosted by Cochise College, and granted permission for their non-commercial educational use. His photos have been used extensively throughout this project. Roger’s usage stipulation has led to thoughtful discussions about what the appropriate way is to license derivative materials that make use of non Creative-Commons content. We have concluded that the best way to ensure that his wishes are respected is to license materials I make with his photographs as CC BY-NC-SA. This permits free sharing and remixing, but stipulates no commercial use, and that all derivative works must be shared with a non-commercial license.

James St. John is a geologist and paleontologist who has contributed (at the time of this writing) more than 59,000 high-quality geology-related photographs to the photo-sharing website Flickr. His photographs cover a wide range of rocks and minerals, and rarely has there been an image that I needed but couldn’t find in his work. His Flickr account is remarkable for the abundance and quality of photographs, but also because he includes detailed descriptions of his images, making it possible for me to verify that an image is what I think it is, and gather useful background information. He has shared his images with a CC BY license, which I appreciate greatly because it allows me to combine them with content having more restrictive licenses.

The U. S. Geological Survey has contributed innumerable images to the public domain. The Hawaiian Volcano Observatory in particular is my go-to source for both the latest in volcano photos, and for fascinating historical images. Data and images from the USGS Earthquake Hazards Program Latest Earthquakes map have been invaluable.

I have used NASA images for views of Earth as much as I have for views of space and other planets. It is truly remarkable that in spite of the vast resources and expertise needed to acquire these photographs, they are free to view, use, and learn from.

Among the many teaching resources offered by IRIS (Incorporated Research Institutions for Seismology) are beautifully designed images for explaining earthquakes and seismology.

When all other sources failed, the odds were good that Robert Lavinsky (www.iRocks.com), Mike Norton, or Michael Rygel had contributed exactly the right photograph to Wikimedia Commons.

4

Preface to the First University of Saskatchewan Edition

Karla Panchuk

The First University of Saskatchewan Edition of Physical Geology is the product of several years’ work iteratively adapting Steven Earle’s original Physical Geology textbook. Edits since the spring of 2017 were supported financially through the University of Saskatchewan’s Open Educational Resources Fund.

Key aspects of this latest version include:

Thus far, this textbook (including previous adapted versions we’ve prepared) has been used by nearly a thousand students at the University of Saskatchewan, saving them tens of thousands of dollars in textbook costs. If you are considering adopting this version of this textbook in your courses or adapting it, please get in touch. We’d love to talk to you about what we’ve done so far and what we are planning for the next edition.

Karla Panchuk

January 2019

5

Preface to the Original Edition

This book was born out of a 2014 meeting of earth science educators representing most of the universities and colleges in British Columbia, and nurtured by a widely shared frustration that many students are not thriving in our courses because textbooks have become too expensive for them to buy. But the real inspiration comes from a fascination for the spectacular geology of western Canada and the many decades that I have spent exploring this region along with colleagues, students, family, and friends. My goal has been to provide an accessible and comprehensive guide to the important topics of geology, richly illustrated with examples from western Canada. Although this text is intended to complement a typical first-year course in physical geology, its contents could be applied to numerous other related courses.

As a teacher for many years, and as someone who is constantly striving to discover new things, I am well aware of that people learn in myriad ways, and that for most, simply reading the contents of a book is not one of the most effective ones. For that reason, this book includes numerous embedded exercises and activities that are designed to encourage readers to engage with the concepts presented, and to make meaning of the material under consideration. It is strongly recommended that you try the exercises as you progress through each chapter. You should also find it useful, whether or not assigned by your instructor, to complete the questions at the end of each chapter.

Over many years of teaching earth science I have received a lot of feedback from students. What gives me the most pleasure is to hear that someone, having completed my course, now sees Earth with new eyes, and has discovered both the thrill and the value of an enhanced understanding of how our planet works. I sincerely hope that this textbook will help you see Earth in a new way.

Steven Earle, Gabriola Island, 2015

I

Chapter 1. Introduction to Geology

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Badlands in southern Saskatchewan. Erosion has exposed layers of rock going back more than 65 million years.
Figure 1.1 Badlands in southern Saskatchewan. Erosion has exposed layers of rock going back more than 65 million years. Source: Karla Panchuk (2017) CC BY-SA 4.0. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

 

1

1.1 What Is Geology? 

Geologists study Earth — its interior and its exterior surface, the rocks and other materials around us, and the processes that formed those materials. They study the changes that have occurred over the vast time-span of Earth’s history, and changes that might take place in the near future.

Geology is a science, meaning that geological questions are investigated with deductive reasoning and scientific methodology. Geology is arguably the most interdisciplinary of all of the sciences because geologists must understand and apply other sciences, including physics, chemistry, biology, mathematics, astronomy, and more.

An aspect of geology that is unlike most of the other sciences is the role played by time — deep time — billions of years of it. When geologists study the evidence around them, they are often observing the results of are observing the results of events that took place thousands, millions, and even billions of years in the past, and which may still be ongoing. Many geological processes happen at incredibly slow rates — millimetres per year to centimetres per year — but because of the amount of time available, tiny changes can result in expansive oceans forming, or entire mountain ranges being worn away.

Geology on a Grand Scale in the Canadian Rocky Mountains

The peak on the right of the photographs in Figure 11.2 is Rearguard Mountain, which is a few kilometres northeast of Mount Robson. Mount Robson is the tallest peak in the Canadian Rockies, at 3,954 m. The large glacier in the middle of the photo is the Robson Glacier. The river flowing from Robson Glacier drains into Berg Lake in the bottom right.

Many geological features are shown here. The rocks that these mountains are made of formed in ocean water over 500 million years ago. A few hundred million years later, the rocks were pushed east for tens to hundreds of kilometres, and thousands of meters upward in a great collision between Earth’s tectonic plates.

Over the past two million years this area, like most of the rest of Canada, has been repeatedly covered by glaciers that scoured away rocks to form the valley to the left of Rearguard Mountain. The Robson Glacier itself is now only a fraction of its  size during the Little Ice Age of the 15th to 18th centuries. And, like almost all other glaciers on Earth, it is now receding even more rapidly because of climate change. Figure 11.2 (right) taken around 1908 by the Canadian geologist and artist Arthur Philemon Coleman, gives an indication of how much the glacier has receded in the last hundred years.

Rearguard Mountain and Robson Glacier in Mount Robson Provincial Park, BC. Left: Robson Glacier today, retreating up the valley. Right: Robson Glacier circa 1908 is much larger..
Figure 11.2 Rearguard Mountain and Robson Glacier in Mount Robson Provincial Park, BC. Left: Robson Glacier today, retreating up the valley. Right: Robson Glacier circa 1908. Sources: Left- Karla Panchuk (2017) CC BY-SA 4.0 with photo by Steven Earle (2015) CC BY 4.0 view source. Right: A.P. Coleman (c. 1908) Public Domain. Click the image for more attributions.

Geology is about understanding the evolution of Earth through time. It is about discovering resources such as metals and energy, and minimizing the environmental implications of our use of resources. It is about learning to mitigate the hazards of earthquakes, volcanic eruptions, and slope failures. All of these aspects of geology, and many more, are covered in this textbook.

References

Victoria University Library (2009) A. P. Coleman Exhibition. Retrieved 25 August 2017. Visit the website

2

1.2 Why Study Earth?

Why?  Because Earth is our home — our only home for the foreseeable future — and in order to ensure that it continues to be a great place to live, we need to understand how it works. Another answer is that some of us can’t help but study it because it’s fascinating. But there is more to it than that.

The Importance of Geological Studies for Minimizing Risks to the Public

Figure 1.3 shows a slope failure that took place in January 2005 in the Riverside Drive area of North Vancouver. The steep bank beneath the house shown gave way, and a slurry of mud and sand flowed down. It destroyed another house below, and killed one person. The slope failure happened after a heavy rainfall, which is a common occurrence in southwestern B.C. in the winter.

Aftermath of a deadly debris flow in the Riverside Drive area of North Vancouver in January, 2005. Source: The Province (2005), used with permission.
Figure 1.3 Aftermath of a deadly debris flow in the Riverside Drive area of North Vancouver in January, 2005. Source: The Province (2005), used with permission.

A geological report written in 1980 warned the District of North Vancouver that the area was prone to slope failure, and that steps should be taken to minimize the risk to residents. Unfortunately, not enough was done in the intervening 25 years to prevent a tragedy.

 

 

3

1.3 What Do Geologists Do?

Geologists do a lot of different things.  Many of the jobs are the things you would expect.  Geologists work in the resource industry, including mineral exploration and mining, and exploring for and extracting sources of energy. They do hazard assessment and mitigation (e.g., assessment of risks from slope failures, earthquakes, and volcanic eruptions).  They study the nature of the subsurface for construction projects such as highways, tunnels, and bridges. They use information about the subsurface for water supply planning, development, and management; and to decide how best to contain contaminants from waste.

Geologists also do the research that makes practical applications of geology possible.  Some geologists spend their summers trekking through the wilderness to make maps of the rocks in a particular location, and collect clues about the geological processes that occurred there.  Some geologists work in laboratories analyzing the chemical and physical properties of rocks to understand how the rocks will behave when forces act on them, or when water flows through them.  Some geologists specialize in inventing ways to use complex instruments to make these measurements.  Geologists study fossils to understand ancient animals and environments, and go to extreme environments to understand how life might have originated on Earth.  Some geologists help NASA understand the data they receive from objects in space.

Geological work can be done indoors in offices and labs, but some people are attracted to geology because they like to be outdoors.  Many geological opportunities involve fieldwork in places that are as amazing to see as they are interesting to study. Sometimes these are locations where few people have ever set foot, and where few ever will again.

Geologists at work on the island of Spitsbergen, part of the Svalbard archipelago. The islands are located in the Arctic Ocean north of Norway.
Figure 1.4 Geologists at work on the island of Spitsbergen, part of the Svalbard archipelago. The islands are located in the Arctic Ocean north of Norway. Source: Gus MacLeod (2007) CC BY-NC-ND 2.0 view source

4

1.4 We Study Earth Using the Scientific Method

There is no single method of inquiry that is specifically the scientific method.  Furthermore, scientific inquiry is not necessarily different from serious research in other disciplines. The key features of serious inquiry are the following:

An Example of the Scientific Method at Work

Consider a field trip to the stream shown in Figure 1.5. Notice that the rocks in and along the stream are rounded off rather than having sharp edges. We might hypothesize that the rocks were rounded because as the stream carried them, they crashed into each other and pieces broke off.

Figure 1.5 Hypothesizing about the origin of round rocks in a stream. Source: Steven Earle (2015) CC BY 4.0 view source

If the hypothesis is correct, then the further we go downstream, the rounder and smaller the rocks should be. Going upstream we should find that the rocks are more angular and larger. If we were patient we could also test the hypothesis by marking specific rocks and then checking back to see if those rocks have become smaller and more rounded as they moved downstream.

If the predictions turn out to be correct, we must still be careful about how much certainty to attach to our hypothesis.  Although our hypothesis might seem to us to be the only reasonable explanation, someone could argue that we have the mechanism wrong, and the rocks weren’t rounded by bumping into each other. If our experiment didn’t specifically check for the mechanism (e.g., by looking to see if chips fall off the rocks and the rocks are made smoother) then we would have to acknowledge the possibility.  We needn’t abandon the hypothesis as a useful tool for making predictions, but it is necessary to be open to the possibility that other things might be going on. If someone demonstrates conclusively that our hypothesis is wrong, then we have to discard the hypothesis and come up with a better one.

A good hypothesis is testable.  Someone might argue that an extraterrestrial organization creates rounded rocks and places them in streams when nobody is looking. There is no practical way to test this hypothesis to confirm it, and there is no way to prove it false. Even if we never see aliens at work, we still can’t say they haven’t been, because according to the hypothesis they only work when people aren’t looking. Compare this to our original hypothesis which allows us to make testable predictions such as rocks getting smaller and rounder downstream. Our original hypothesis gives us a way to see how realistic it is, whereas the alien hypothesis gives us no way to know if it makes sense or not.

Theories and Laws

Two other terms appear in discussions of the scientific method: theory and law. A theory starts out as a hypothesis, but over a long period of time and a great many tests, it has never come up short. That doesn’t mean it never will, but the odds of that are very unlikely given our present (and conceivable future) state of knowledge.  You may have heard someone dismiss an idea by saying it is “just a theory,” but they are using the term incorrectly if they mean to say it’s a wild and unproven guess.

A law is a description of a phenomenon rather than an explanation of it.  For example, you could do thousands of tests by dropping an object with known mass and measuring its acceleration and the force with which it hits the ground.  Again and again your results will yield the formula force = mass x acceleration.  However, that doesn’t mean you know what is responsible for the force accelerating it toward the ground.  Yes, we say that gravity is pulling it toward the Earth’s surface, but why? A law is true regardless of why a phenomenon happens as long as it describes the outcome of that phenomenon.

5

1.5 Three Big Ideas: Geological Time, Uniformitarianism, and Plate Tectonics

In geology there are three big ideas that are fundamental to the way we think about how Earth works.  The ideas are like the sound track to a movie- sometimes we might not even notice them, but at the same time they affect our perception of what is happening.  In the rest of this book these ideas may be mentioned explicitly in some cases, but in other cases it will be helpful for you to realize that they are relevant, even if they are not being discussed by name.

Geological Time (Deep Time)

Earth is approximately 4.57 billion years old (4,570,000,000 years), which is a long time for geological events to unfold and changes to happen. The changes themselves might be tiny. For example, over a year, a chemical reaction might eat away a few layers of atoms at the surface of a rock. But over time the changes accumulate and have a great impact. Over hundreds of millions of years the chemical reaction could cause a mountain range to crumble into grains of sand, and be swept away by rivers.

For geologists who study very, very slow processes, 10 million years might be a short time, and 1 million years might be trivial.  For these geologists, intervals of 1 million years aren’t even useful to consider, because the changes over that time are too small to see in the rocks that accumulated.

As you read through this book, keep in mind that the well of geologic time is indeed deep, and “ancient” is defined in a whole new way.

Expressing Geological Time in Numbers

Special notation is used for geological time because, as you might imagine, writing all those zeroes can become tiresome.  Table 1.1 shows common abbreviations you will see throughout this book.

Table 1.1 Abbreviations Used to Describe Geological Time
Abbreviation Meaning Example
Ga giga annum
 or billions of years
Earth is 4.57 Ga old.
Ma mega annum
or millions of years
Earth is 4,570 Ma old.
ka kilo annum or thousands of years The last glacial cycle ended 11,700 years ago, or 11.7 ka.

Expressing Geological Time Using the Geological Time Scale

The geological time scale (Figure 1.6) is a way of breaking down geological time according to important events in Earth’s history.  Time is divided into eons, eras, periods, and epochs, and these intervals are referred to by names rather than by years.  Giving intervals of geologic time names rather than using numbers makes sense because we won’t always know the age in years (the absolute age) of a rock or fossil, but we can place it in context based on our knowledge of the geological record.  We can describe its relative age by saying that it is older than or younger than another rock or fossil.

Geologic Society of America Geologic Time Scale, 2012
Figure 1.6 Geologic Society of America Geologic Time Scale, 2012. Source: Walker, J.D., Geissman, J.W., Bowring, S.A., and Babcock, L.E., compilers (2012) Geologic Time Scale v. 4.0: Geological Society of America, doi: 10.1130/2012.CTS004R3C. Download PDF

The tricky thing about the geologic time scale is that the boundaries are always changing.  As our knowledge of the absolute age of an event improves with new discoveries, it might be necessary to nudge a boundary earlier or later.  Sometimes the original reason for defining a boundary no longer holds, but we agree to use it anyway.  For example, the Phanerozoic Eon (the last 542 million years) is named for the time during which visible (phaneros) life (zoi) is present in the geological record, and its start was meant to mark the first appearance of these organisms. In fact, we now have evidence that large organisms — those that leave fossils visible to the naked eye — have existed longer than that, first appearing by 600 Ma at the latest.

An Early Definition of the Proterozoic

Notice that in Figure 1.6 the Proterozoic Eon precedes the Phanerozoic Eon. This was not always the case. Figure 1.7 shows an excerpt from a periodical published in 1879, in which the Proterozoic is defined as covering the Cambrian through Silurian. The author refers to “the most extreme adherents of the Murchisonian party in geology,” a reference to the contentious assertion by Scottish geologist Roderick Murchison (1792-1871) that the Silurian Period should encompass the Cambrian and Ordovician periods as well.

Classification of the Lower Paleozoic Rocks. The systems at present assigned to the Paleozoic age fall into two main groups- an older group, including the Cambrian, Ordovician, and Silurian systems, and a younger group, including the Devonian, Carboniferous, and Permian. The period duringwhich the former were deposited may be deonimated the Lower Paleozoic or Proterozoic Age; that in which the latter were laid down may be called the Upper Paleozoic or Deuterozoic. Broadly speaking, the Proterozoic rocks include all the sedimentary formations to which the name Silurian has at any time been applied by the most extreme adherents of the Murchisonian party in geology.
Figure 1.7 An excerpt from the periodical The Annals and Magazine of Natural History (1879) in which the name “Proterozoic” is assigned to the Cambrian, Ordovician, and Silurian periods instead of to the time preceding the Cambrian. Source: Karla Panchuk (2017) CC BY 4.0 Read the book

A Way To Think About Geological Time

A useful mechanism for understanding geological time is to scale it down into one year. The origin of the solar system and Earth at 4.57 Ga would be represented by January 1, and the present year would be represented by the last tiny fraction of a second on New Year’s Eve. At this scale, each day of the year represents 12.5 million years; each hour represents about 500,000 years; each minute represents 8,694 years; and each second represents 145 years. Some significant events in Earth’s history, as expressed on this time scale, are summarized in Table 1.2.

Table 1.2  Some Important Dates Expressed As If All of Geological Time Were Condensed Into One Year
Event Approximate Date Calendar Equivalent
Formation of oceans and continents 4.5 – 4.4 Ga first week of January
Evolution of the first primitive life forms 3.8 Ga end of February
Formation of Saskatchewan’s oldest rocks 3.4 Ga end of March
Evolution of the first multi-celled animals 600 Ma beginning of November
Animals first crawled onto land 360 Ma end of November
Vancouver Island reached North America and the Rocky Mountains were formed 90 Ma December 16
Extinction of the non-avian dinosaurs 65 Ma December 18
Beginning of the Pleistocene ice age 2 Ma 10:10 p.m., December 31
Oldest radiocarbon date from people living in Canada (British Columbia) 13.8 ka 11:58 p.m., December 31
Earliest evidence of human activity in Saskatchewan 11.5 ka 48 seconds before midnight, December 31
The last of the glacial ice retreats from Saskatchewan 6 ka 41 seconds before midnight, December 31
Hudson’s Bay Company establishes a permanent settlement at Cumberland House in northern Saskatchewan 243 years ago 2 seconds before midnight, December 31
Source: Karla Panchuk (2017) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view original

Uniformitarianism

Uniformitarianism is the notion that the geological processes occurring on Earth today are the same ones that occurred in the past.  This is an important idea because it means that observations we make today about geological processes can be used to interpret and understand the rock record.  While this idea might not seem remarkable today, it was ground breaking and even controversial for its time.  Many people who heard about it for the first time thought about the age of the Earth in thousands of years, but uniformitarianism required them to think on timescales almost too vast to comprehend.  For some, this implied questioning their most deeply held religious beliefs.

The Scottish geologist James Hutton initially presented the idea in 1785Read James Hutton's abstract at http://bit.ly/1j6tIAN. Note that the typeface prints an "s" like an "f.".  Charles Lyell, also a Scottish geologist, paraphrased this idea as “the present is the key to the past” in his book Principles of Geology.The 7th edition of Charles Lyell's Principles of Geology (1847) can be found at http://bit.ly/1l3T6Zh  This is how it is often described today.

To be clear, “the present is the key to the past” can be viewed as an oversimplification. Not all geological processes occurring today occurred at all times in the geological past.  For example, some important chemical reactions that happened on Earth’s surface today require abundant oxygen in the atmosphere, and could not have occurred prior to Earth developing an oxygen-rich atmosphere.  Conversely, there was a time in Earth’s history when continents as we know them hadn’t yet developed. Some events, such as devastating impacts by objects from space, have never been witnessed on the same scale by humans. We must be cognizant of the fact that conditions were different at different times in Earth’s history, and take that into account when interpreting the rock record.

Despite the different past conditions on Earth as a whole, there still exist environments today where some of these conditions are present. These environments are like little samples of what Earth used to be like.  This means we can still use present conditions to inform us about the past, but we have to think carefully about ways that such environments today differ from the ancient environments that no longer exist.

Plate Tectonics

It is only within the last 50 years or so that we have been able to answer questions like, “How did that mountain range get there?” and “Why do earthquakes happen where they do?”  The theory of plate tectonics– the idea that Earth’s surface is broken into large moving fragments, called plates– profoundly changed our perspective on how the Earth works.  Figure 1.8 shows Earth’s 15 largest tectonic plates, along with arrows indicating the plates’ direction of motion, and how fast they go.  (Longer arrows mean faster motion.)  There are many more plates on Earth that are too small to show conveniently in Figure 1.8. A more detailed map of Earth’s tectonic plates can be found at here.

Figure 1.8 Earth’s fifteen largest tectonic plates. Black arrows show the direction of plate motions. The length of the arrow indicates velocity. Red arrows show how plates move relative to each other. Source: Steven Earle (2015) CC BY 4.0. view source Modified after U. S. Geological Survey (1996) Public Domain view original

Prior to plate tectonics, we made observations but could only guess at mechanisms.  It was like watching the hands on a clock and trying to guess what moves them.  After plate tectonics it was like being able to open the clock and not only watch the gears turn, but realize for the first time that there are such things as gears. Plate tectonics not only explains why things have happened, but also allows us to predict what might happen in the future.

Plate tectonics is covered in more detail later, however the key point is that Earth’s outer layer consists of rigid plates that are constantly interacting with each other as they move around the Earth.  The boundaries of plates move away from each other in some places, collide in others, and sometimes just slide past each other (illustrated by the red arrows in Figure 1.8). The plates can move because they are floating on a layer of weak rock that deforms as the plates travel, much the same way the filling in a peanut butter and jelly sandwich allows you to slide the top layer of bread across the bottom layer.

Whether the plates move away from each other, collide, or just slide past each other determines things like the locations of mountain belts and volcanoes, where earthquakes happen, and the shapes and sizes of oceans and continents.

References

Cottrell, M. (2006) History of Saskatchewan. Retrieved 26 August 2017. Visit the website

6

Chapter 1 Summary

The topics covered in this chapter can be summarized as follows:

1.1 What is Geology?

Geology is the study of Earth. It is an integrated science that involves the application of many of the other sciences. Geologists must take into account the fact that the geological features we see today may have formed thousands, millions, or even billions of years ago, and over very long time spans.

1.2 Why Study Earth?

Geologists study Earth out of curiosity and for other, more practical reasons, including understanding the evolution of life on Earth; searching for resources; understanding risks from geological events such as earthquakes, volcanoes, and slope failures; and documenting past environmental and climate changes so that we can understand how human activities are affecting Earth.

1.3 What Do Geologists Do?

Geologists work in the resource industry, and in efforts to protect the environment. Geologists work to minimize the risks from geological hazards (e.g., earthquakes), and to help the public understand those risks. Geologists investigate Earth materials in the field, in and in the lab.

1.4 We Study Earth Using the Scientific Method

Scientific inquiry requires a careful process of making a hypothesis and then testing it. If a hypothesis doesn’t pass the test, it’s time for a new one. A theory is a hypothesis that has been tested repeatedly and never failed a test. A law is a description of a natural process.

1.5 Three Big Ideas: Geological Time, Uniformitarianism, and Plate Tectonics

Geological time: Earth is approximately 4,570,000,000 years old; that is, 4.57 billion years or 4.57 Ga or 4,570 Ma. It’s such a huge amount of time that even extremely slow geological processes can have an enormous impact.

Uniformitarianism: Processes that occur today also occurred in the geologic past.  We can use our observations of the present to understand the processes that shaped the Earth throughout its history.

Plate tectonics: Earth’s surface is broken into plates that move and interact with each other.  The interactions between these plates are key for understanding the mechanisms behind geologic processes.

Review Questions

  1. How does the element of time make geology different from the other sciences, such as chemistry and physics?
  2. List three ways in which geologists can contribute to society.
  3. The following dates are written with the abbreviations Ga, Ma, and ka. Express the dates in years. (For example, 2.3 Ma = 2,300,000 years)
    1. 2.75 ka
    2. 0.93 Ga
    3. 4.2 Ma
    4. 0.2 ka.
  4. Dinosaurs first appear in the geological record in rocks from about 215 Ma and then most became extinct at 65 Ma. What percentage of geological time does this represent?
  5. If sediments typically accumulate at a rate of 1 mm/year, what thickness of sediment could accumulate over a period of 30 million years?
  6. Does uniformitarianism mean that conditions on Earth are uniform, and never change?
  7. Summarize the main idea behind plate tectonics.

7

Answers to Chapter 1 Review Questions

  1. Geology requires that we consider vast amounts of time, and think about the effects that accumulate over thousands, millions, or even billions of years.
  2. There are many ways that geologists contribute. Geologists provide information to reduce the risk of harm from hazards such as earthquakes, volcanoes, and slope failures; they play a critical role in the discovery of important resources; they contribute to our understanding of life and its evolution through paleontological studies; and they play a leading role in the investigation of climate change, past and present and its implications.
  3. Ages in years
    1. 2.75 ka = 2,750 years
    2. 0.93 Ga = 930,000,000 years
    3. 14.2 Ma = 14,200,000 years
    4. 0.2 ka = 200 years.
  4. 215 – 65 = 150 Ma. Since the age of the Earth is 4570 Ma, this represents 150/4,570 = 0.033 or 3.3% of geological time.
  5. At 1 mm/y 30,000,000 mm of sediment would accumulate over that 30 million years. This is equivalent to 30,000 m or 30 km. Few sequences of sedimentary rock are even close to that thickness because most sediments accumulate at much lower rates, more like 0.1 mm/y. Also, over time the sediments are compressed.
  6. No. Uniformitarianism means that we can use the processes we observe today to help us understand what happened in the past.
  7. Plate tectonics is the idea that Earth’s outer layer is broken into rigid plates. The plates move around and interact with each other along their margins.

II

Chapter 2. The Origin of Earth and the Solar System

By Karla Panchuk

Figure 2.1 Earthrise, October 12, 2015. The Lunar Reconnaissance Orbiter Camera captured images of the lunar surface with Earth in the background. Source: NASA Lunar Reconnaissance Orbiter Science Team (2015) Public Domain. view source

 

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

The story of how Earth came to be is a fascinating contradiction. On the hand, many things had to go just right for Earth to turn out the way it did, and for life to develop. On the other hand, the formation of planets similar to Earth is an entirely predictable consequence of the physical and chemical processes taking place around stars. In fact, it has happened more than once.

This chapter starts Earth’s story from the beginning — the very beginning — to explain why, for billions of years, generations of stars had to be born, then die explosive deaths before Earth could exist. How stars form and burn, and affect the objects around them are fundamental to Earth’s story, as is the rough neighbourhood in which Earth spent its early years.

8

2.1 Starting with a Big Bang

According to the big bang theory, the universe blinked violently into existence 13.8 billion years ago. The big bang is often described as an explosion, but imagining it as an enormous fireball isn’t accurate. The big bang started with a sudden expansion of energy and space from a single point. The kind of Hollywood explosion that might come to mind involves expansion of matter and energy within space, but during the big bang, energy, space, and matter were created. In Figure 2.2 the pointed base of the universe “vessel” represents the big bang. Time advances moving up the diagram. The vessel gets wider as time progresses, reflecting the expansion of the universe.

Figure 2.2 The big bang. The universe began 13.8 billion years ago as a rapid expansion of space, energy, and matter. It continues to expand. Left: Timeline of the universe. The point at the base of the “vessel” represents the moment of the big bang. The vessel gets wider as time progresses, representing the expansion of the universe. Right: Mollwiede projection of the cosmic microwave background, a “fog” from when the universe was still very dense. Temperature variations correspond to clumping of matter in the early universe. Source: Karla Panchuk (2018) CC BY 4.0 modified after Ryan Kaldari (2006) Public Domain view source, derivative of NASA/WMAP Science Team (2006) Public Domain view source. CMB map by NASA/WMAP Science Team (2006) Public Domain view source. Click the image for data sources.

You might wonder how a universe can be created out of nothing. Creating a universe out of nothing is mostly beyond the scope of this chapter, but there is a way to think about it. The particles that make up the universe have opposites that cancel each other out, similar to the way that we can add the numbers 1 and -1 to get zero (also known as “nothing”). As far as the math goes, having zero is exactly the same as having a 1 and a -1. It is also exactly the same as having a 2 and a -2, a 3 and a -3, two -1s and a 2, and so on. In other words, nothing is really the potential for something if you divide it into its opposite parts.

Composition of the Universe

In Figure 2.2, the “contents” of the vessel change as time progresses. A few minutes after the big bang, the universe was still too hot and dense to be anything but a sizzle of particles smaller than atoms. But as it expanded, it also cooled. Eventually particles that collided were able to stick together to form atoms, rather than being smashed apart again when other particles crashed into them. Those collisions produced hydrogen and helium, the most common elements in the universe.

For a long time after the big bang, clouds of hydrogen and helium atoms drifted about a dark universe. The “dark ages” (bottom of Figure 2.2) were a time when the ingredients for stars existed, but stars themselves did not. It took approximately 500 million years for enough hydrogen atoms to clump together in clouds to allow the first stars to form and begin to shine.

Looking Back to the Early Stages of the Big Bang

The notion of seeing the past is often used metaphorically when we talk about ancient events, but in this case it is meant literally. In our everyday experience, when we watch an event take place, we perceive that we are watching it as it unfolds in real time. In fact, this isn’t true. To see the event, light from that event must travel to our eyes. Light travels very rapidly, but it does not travel instantly. If we were watching a digital clock 1 m away from us change from 11:59 a.m. to 12:00 p.m., we would actually see it turn to 12:00 p.m. three billionths of a second after it happened.

This isn’t enough of a delay to cause us to be late for an appointment, but the universe is a very big place, and the “digital clock” in question is often much, much farther away. In fact, the universe is so big that it is convenient to describe distances in terms of light years, or the distance light travels in one year. What this means is that light from distant objects takes so long to get to us that we see those objects as they were at some considerable time in the past. For example, the star Proxima Centauri is 4.24 light years from the sun. If you viewed Proxima Centauri from Earth on January 1, 2018, you would actually see it as it appeared in early October 2013.

We now have tools that are powerful enough to look deep into space and see the arrival of light from early in the universe’s history. Astronomers can detect light from approximately 380,000 years after the big bang is thought to have occurred. Physicists tell us that if the big bang happened, then particles within the universe would still be very close together at this time. They would be so close that light wouldn’t be able to travel far without bumping into another particle and getting scattered in another direction. The effect would be to fill the sky with glowing fog, the “afterglow” from the formation of the universe.

In fact, this is exactly what we see when we look at light from 380,000 years after the big bang. The fog is referred to as the cosmic microwave background (or CMB), and it has been carefully mapped throughout the sky. In Figure 2.2, the colourful patch at the base of the diagram represents the fog that is measured today as the CMB. Figure 2.2 (right) is a CMB map of the universe in Mollweide projection. This is a projection that is used to show Earth’s geography on a flat surface. In this case, the map of the CMB represents a sphere surrounding Earth rather than what’s beneath our feet.

Colour variations in the CMB map represent temperature variations. These variations translate to differences in the density at which matter was distributed in the early universe. The red patches are the highest density regions and the blue patches are the lowest density. Higher density regions represent the eventual beginnings of stars and planets. The CMB map in Figure 2.2 has been likened to a baby picture of the universe.

The Universe is Still Expanding

The expansion that started with the big bang never stopped. It continues today, and we can see it happen by observing that large clusters of billions of stars, called galaxies, are moving away from us. (An exception is the Andromeda galaxy with which we are on a collision course.) The astronomer Edwin Hubble came to this conclusion when he observed that the light from other galaxies was red-shifted. The red shift is a consequence of the Doppler effect. This refers to how we see waves when the object that is creating the waves is moving toward us or away from us.

Before looking at the Doppler effect as it pertains to light, it can be useful to see how it works on something more tangible. The duckling swimming in Figure 2.3 is generating waves as it moves through the water. It is generating waves that move forward as well as back, but notice that the ripples ahead of the duckling are closer to each other than the ripples behind the duckling. The distance from one ripple to the next is called the wavelength. The wavelength is shorter in the direction that the duckling is moving, and longer as the duckling moves away.

Figure 2.3 A duckling illustrates the Doppler effect in water. The ripples made in the direction the duckling is moving (blue lines) are closer together than the ripples behind the duckling (red lines). Source: Karla Panchuk (2015) CC BY 4.0. Photo by M. Harkin (2013) CC BY 2.0 view source

When waves are in air as sound waves rather than in water as ripples, the different wavelengths manifest as sounds with different pitches — the short wavelengths have a higher pitch, and the long wavelengths have a lower pitch. This is why an observer will hear a change in the pitch of a car’s engine as the car races past.

For light waves, wavelength translates to colour. In the spectrum of light that we can see, shorter wavelengths are on the blue end of the spectrum, and longer wavelengths are on the red end of the spectrum. In Figure 2.4, the longer or shorter wavelengths of the water ripples at the top of the diagram reflect the longer or shorter wavelengths of light in the visible spectra below. Does this mean that galaxies look red because they are moving away from us? No, but the colour we see is shifted toward the red end of the spectrum and longer wavelengths.

Figure 2.4 Red shift in light from the supercluster BAS11 compared to the sun’s light. Black lines represent wavelengths absorbed by atoms (mostly hydrogen and helium). For BAS11 the black lines are shifted toward the red end of the spectrum compared to the sun. Source: Karla Panchuk (2018) CC BY 4.0, spectra by Georg Wiora (2011) CC BY-SA 2.5 view source

Notice that the sun’s spectrum in the upper part of Figure 2.4 has black lines in it. The black lines are there because some colours are missing from the sun’s light that reaches Earth. Different elements absorb light of specific wavelengths, and many of the black lines in Figure 2.4 represent colours that are absorbed by hydrogen and helium within the sun. This means the black lines are like a bar code that can tell us what a star is made of.

The lower spectrum in Figure 2.4 is the light coming from BAS11, an enormous cluster of approximately 10,000 galaxies located 1 billion light years away. The black lines represent the same elements as in the sun’s spectrum, but they are shifted to the right toward the red end of the spectrum, because BAS11 is moving away from us as the universe continues to expand. To summarize, because almost all of the galaxies we can see have light that is red-shifted, it means they are all moving away from us. In fact, the farther away they are, the faster they are going. This is evidence that the universe is still expanding.

References

European Space Agency (2015). Planck reveals first stars were born late. Visit website

Knop, R. (2010). The History of the Universe. Visit blog

Lawrence, C. R. (2015, March). Planck 2015 Results. Paper presented to the Astrophysics Subcommittee, NASA HQ. View slides

9

2.2 Forming Planets from the Remnants of Exploded Stars

Only four elements account for 95% of Earth’s mass: oxygen (O), magnesium (Mg), silicon (Si), and iron (Fe). Most of the remaining 5% comes from aluminum (Al), calcium (Ca), nickel (Ni), hydrogen (H), and sulphur (S). We know that the big bang made hydrogen, but where did the rest of the elements come from?

The answer is that the other elements were made by stars. Sometimes stars are said to “burn” their fuel, but burning is not what is going on within stars. The burning that happens when wood in a campfire is turned to ash and smoke is a chemical reaction — heat causes the atoms that were in the wood and in the surrounding atmosphere to exchange partners. Atoms group in different ways, but the atoms themselves do not change. What stars do is change the atoms. The heat and pressure within stars cause smaller atoms to smash together and fuse into new, larger atoms. For example, when hydrogen atoms smash together and fuse, helium is formed. Large amounts of energy are released when atoms fuse within stars, and this is what causes stars to shine. Stars can form large quantities of elements as heavy as iron during their normal burning process. Side reactions can form heavier elements in small amounts.

It takes larger stars to make elements as heavy as iron in large quantities. Our sun is an average star. After it uses up its hydrogen fuel to make helium, and some of that helium is fused to make small amounts of other elements, it will be at the end of its life. It will stop making new elements and will cool down and bloat until its middle reaches the orbit of Mars. In contrast, large stars end their lives in spectacular fashion. They explode as supernovae, casting off newly formed atoms into space, and triggering side reactions to make even more heavy atoms. It took many generations of stars creating heavier elements and casting them into space before heavier elements were abundant enough for planets like Earth to form.

Until recently, astronomers have only been able to see stars that already contain heavier elements in small amounts, but not the first-generation stars that started out before any of the heavier elements were produced. That changed in 2015 when it was announced that a distant galaxy called CR7 had been found that contained stars made only of hydrogen and helium. The galaxy is so far away that it shows us a view of the universe from approximately 800 million years after the big bang. Since then, more galaxies like CR7 have been discovered.

References

Pilipenko, S. V. (2013). Paper-and-pencil cosmological calculator. arXiv:1303.5961v1 [astro-ph.CO]

Royal Astronomical Society (2016, June 28). CR7 is not alone—A team of super bright galaxies in the early universe. Phys.org Visit website

Sobral, D., Matthee, J.,  Darvish, B., Schaerer, D., Mobasher, B., Röttgering, H., Santos, S., & Hemmati, S. (2015). Evidence for PopIII-like stellar populations in the most luminous Lyman-α emitters at the epoch of re-ionisation: spectroscopic confirmation. The Astrophysical Journal 808(2) doi: 10.1088/0004-637x/808/2/139.

 

10

2.3 How to Build a Solar System

A solar system consists of a collection of objects orbiting one or more central stars. All solar systems start out the same way. They begin in a cloud of gas and dust called a nebula. Nebulae are some of the most beautiful objects that have been photographed in space. They have vibrant colours from the gases and dust they contain, and brilliant twinkling from the many stars that have formed within them (Figure 2.5). The gas consists largely of hydrogen and helium, and the dust consists of tiny mineral grains, ice crystals, and organic particles.

Figure 2.5 The Pillars of Creation within the Eagle Nebula viewed in visible light (left) and near infrared light (right). Near infrared light captures heat from stars and allows us to view stars that would otherwise be hidden by dust. This is why the picture on the right appears to have more stars than the picture on the left. Source: NASA, ESA, and the Hubble Heritage Team (STScI/AURA) (2015) Public Domain. view source

Step 1: Collapse a Nebula

A solar system begins to form when a small patch within a nebula (small by the standards of the universe, that is) begins to collapse upon itself. Exactly how this starts isn’t clear, although it might be triggered by the violent behaviour of nearby stars as they progress through their life cycles. Energy and matter released by these stars might compress the gas and dust in nearby neighbourhoods within the nebula.

Once it is triggered, the collapse of gas and dust within that patch continues for two reasons. One of those reasons is that gravitational force pulls gas molecules and dust particles together. But early in the process, those particles are very small, so the gravitational force between them isn’t strong. So how do they come together? The answer is that dust first accumulates in loose clumps for the same reason dust bunnies form under the bed: static electricity. Given the role of dust bunnies in the early history of the solar system, one might speculate that an accumulation of dust bunnies poses a substantial risk to one’s home (Figure 2.6). In practice, however, this is rarely the case.

Figure 2.6 Public service announcement. If you don’t think housekeeping is important, then you don’t understand the gravity of the situation. Source: Karla Panchuk (2018) CC BY 4.0. Planets modified after NASA/JPL (2008) Public Domain. view source

Step 2: Make a Disk with a Star at Its Centre

As the small patch within a nebula condenses, a star begins to form from material drawn into the centre of the patch, and the remaining dust and gas settle into a protoplanetary disk that rotates around the star. The disk is where planets will eventually form. Figure 2.7 (upper left) is an artist’s impression of a protoplanetary disk, and Figure 2.7 (upper right) is an actual protoplanetary disk surrounding the star HL Tauri. Notice the dark rings within the HL Tauri protoplanetary disk. These are gaps formed by the collection of dust and debris by incipient planets, called protoplanets, as they orbit the star. There is an analogy for this in our own solar system, because the dark rings are akin to the gaps in the rings of Saturn (Figure 2.7, lower left), where moons can be found (Figure 2.7, lower right).

Figure 2.7 Protoplanetary disks and Saturn’s rings. Upper left: Artist’s impression of a protoplanetary disk containing gas and dust, surrounding a new star. Upper right: A photograph of the protoplanetary disk surrounding HL Tauri. The dark rings within the disk are thought to be gaps where newly forming planets are sweeping up dust and gas. Lower left: A photograph of Saturn showing similar gaps within its rings. The bright spot at the bottom is an aurora, similar to the northern lights on Earth. Lower right: a close-up view of a gap in Saturn’s rings showing a moon as a white dot. Source: Upper left- NASA/JPL-Caltech (2008) Public Domain view source; Upper right- ALMA (ESO/NAOJ/NRAO) (2014) CC BY 4.0 view source; Lower left- NASA, ESA, J. Clarke (Boston University), and Z. Levay (STScI) (2005) Public Domain view source; Lower right- NASA/JPL/Space Science Institute (2005) Public Domain view source

Step 3: Build Some Planets

In general, planets can be classified into three categories based on what they are made of (Figure 2.8). Terrestrial planets are those planets like Earth, Mercury, Venus, and Mars that have a core of metal surrounded by rock. Jovian planets (also called gas giants) are those planets like Jupiter and Saturn that consist predominantly of hydrogen and helium. Ice giants are planets such as Uranus and Neptune that consist largely of water ice, methane (CH4) ice, and ammonia (NH3) ice, and have rocky cores. Often, the ice giant planets Uranus and Neptune are grouped with Jupiter and Saturn as gas giants; however, Uranus and Neptune are very different from Jupiter and Saturn.

Figure 2.8 Three types of planets. Jovian (or gas giant) planets such as Jupiter consist mostly of hydrogen and helium. They are the largest of the three types. Ice giant planets such as Uranus are the next largest. They contain water, ammonia, and methane ice, and have rocky cores. Terrestrial planets such as Earth are the smallest, and they have metal cores covered by rocky mantles. Source: Karla Panchuk (2015) CC BY 4.0. Click the image for more attributions.

These three types of planets are not mixed together randomly within our solar system. Instead they occur in a systematic way, with terrestrial planets closest to the sun, followed by the Jovian planets and then the ice giants (Figure 2.9). Smaller solar system objects follow this arrangement as well. The asteroid belt contains bodies of rock and metal. Bodies ranging from metres to hundreds of metres in diameter are classified as asteroids, and smaller bodies are referred to as meteoroids. In contrast, the Kuiper belt (Kuiper rhymes with piper), and the Oort cloud (Oort rhymes with sort), which are at the outer edge of the solar system, contain bodies composed of large amounts of ice in addition to rocky fragments and dust.

Figure 2.9 Our solar system. Top: The solar system shown with distances to scale. Distances are in astronomical units (AU), where 1 AU is the average distance from Earth to the sun. The edge of the Kuiper belt extends to 50 AU (7.5 billion km), but this distance is minuscule compared to the size of the solar system as a whole, which extends to the edge of the Oort cloud, thought to be 15 trillion km away. Bottom: Solar system with the sun and planets to scale. The gas giants are the largest planets, followed by the ice giants, and then the terrestrial planets. Note that the planets in this diagram likely do not reflect the entire population of planets in our solar system because evidence suggests that large planets are present beyond the Kuiper belt. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0, Milky Way photo by R@pp (2017) CC BY-NC-SA 2.0 view source, planet photographs courtesy of NASA. Click the image for planet photo sources and attributions.

Part of the reason for this arrangement is the frost line (also referred to as the snow line). The frost line marks the division between the inner part of the protoplanetary disk closer to the sun, where it was too hot to permit anything but silicate minerals and metal to crystalize, and the outer part of the disk farther from the sun, where it was cool enough to allow ice to form. As a result, the objects that formed in the inner part of the protoplanetary disk consist largely of rock and metal, while the objects that formed in the outer part consist largely of gas and ice. The young sun blasted the solar system with raging solar winds (winds made up of energetic particles), which helped to drive lighter molecules toward the outer part of the protoplanetary disk.

Rules of the Accretion Game

The objects in our solar system formed by accretion. Early in this process, particles collected in fluffy clumps because of static electricity. As the clumps grew larger, gravity became more important and collected clumps into solid masses, and solid masses into larger and larger bodies. If you were one of these bodies in the early solar system, and participating in the “accretion game” with the goal of becoming a planet, you would have to follow some key rules:

You would also have to watch out for some dangers:

Winners and Losers

The outcome of the game is evident in Figure 2.9. Today eight official winners are recognized, with Jupiter taking the grand prize, followed closely by Saturn. Both planets have trophy cases with more than 60 moons each, and each has a moon that is larger than Mercury. Prior to 2006, Pluto was also counted a winner, but in 2006 a controversial decision revoked Pluto’s planet status. The reason was a newly formalized definition of a planet, which stated that an object can only be considered a planet if it is massive enough to have swept its orbit clean of other bodies. Pluto is situated within the icy clutter of the Kuiper belt, so it does not fit this definition.

Pluto’s supporters have argued that Pluto should have been grandfathered in, given that the definition came after Pluto was declared a planet, but to no avail. Pluto has not given up, and on July 13, 2015, it launched an emotional plea with the help of the NASA’s New Horizons probe. New Horizons sent back images of Pluto’s heart (Figure 2.10). On closer inspection, Pluto’s heart was discovered to be broken.

Figure 2.10 Photographs of Pluto. Left: The heart-shaped region called Tombaugh Regio is outlined. This region is named after Pluto’s discoverer Clyde Tombaugh. Right: False-colour images show compositional variations in Tombaugh Regio. Source: Karla Panchuk (2015) CC BY 4.0. Left photo- NASA/APL/SwRI (2015) Public Domain view source, Right photo- NASA/APL/SwRI (2015) Public Domain. view source.

 The Accretion Game and the Solar System Today

The rules and dangers of the planet-forming game help to explain many features of our solar system today.

11

2.4 Earth's First 2 Billion Years

If you were to get into a time machine and visit Earth shortly after it formed (around 4.5 billion years ago), you would probably regret it. Large patches of Earth’s surface would still be molten, which would make landing your time machine very dangerous indeed. If you happened to have one of the newer time-machine models with hovering capabilities and heat shields, you would still face the inconvenience of having nothing to breathe but a tenuous wisp of hydrogen and helium gas, and depending on how much volcanic activity was going on, volcanic gases such as water vapour and carbon dioxide. Some ammonia and methane might be thrown in just to make it interesting, but there would be no oxygen. Assuming you had the foresight to purchase the artificial atmosphere upgrade for your time machine, it would all be for naught if you materialized just in time to see an asteroid, or worse yet another planet, bearing down on your position. The moral of the story is that early Earth was a nasty place, and a time machine purchase is not something to take lightly.

Why was early Earth so nasty?

Earth Was Hot

Earth’s heat comes from the decay of radioactive elements within Earth, as well as from processes associated with Earth’s formation. Formation processes contributed heat in the following ways:

Heating had an important consequence for Earth’s structure. As Earth grew, it collected a mixture of rocky silicate mineral grains as well as iron and nickel. These materials were scattered throughout Earth. That changed when Earth began to heat up: it got so hot that the metals melted and trickled down through the rocky silicate material toward Earth’s centre, becoming Earth’s core. The silicate material became Earth’s crust and mantle. In other words, Earth unmixed itself. The separation of silicate minerals and metals into a rocky outer layer and a metallic core, respectively, is called differentiation. Friction from metal melts moving through Earth caused it to heat up even more.

Earth’s high temperature early in its history also means that early tectonic processes were accelerated compared to today, and Earth’s surface was more geologically active.

Earth Was Bombarded by Objects from Space

Although Earth had swept up a substantial amount of the material in its orbit as it was accreting, unrest within the solar system caused by changes in the orbits of Saturn and Jupiter was still sending many large objects on cataclysmic collision courses with Earth. The energy from these collisions repeatedly melted and even vaporized minerals in the crust, and blasted gases out of Earth’s atmosphere. Very old scars from these collisions are still detectable, although we have to look carefully to see them. For example, the oldest impact site discovered is the 3 billion year old Maniitsoq “crater” in west Greenland, although there is no crater to see. What is visible are rocks that were 20 km to 25 km below Earth’s surface at the time of the impact, but which nevertheless display evidence of deformation that could only be produced by intense, sudden shock.

The evidence of the very worst collision that Earth experienced is not subtle at all. In fact, you have probably looked directly at it hundreds of times already, perhaps without realizing what it is. That collision was with a planet named Theia, which was approximately the size of Mars. Not long after Earth formed, Theia struck Earth (Figure 2.11). When Theia slammed into Earth, Theia’s metal core merged with Earth’s core, and debris from the outer silicate layers was cast into space, forming a ring of rubble around Earth. The material within the ring coalesced into a new body in orbit around Earth, giving us our moon. Remarkably, the debris may have coalesced in 10 years or fewer! This scenario for the formation of the moon is called the giant impact hypothesis.

Figure 2.11 Artist’s impression of a collision between planets. A similar collision between Earth and the planet Theia might have given us our moon. Source: NASA/ JPL-Caltech (2009) Public Domain. view source.

Today’s Atmosphere Took a Long Time to Develop

Earth’s first experiment with having an atmosphere did not succeed. It started out with a thin veil of hydrogen and helium gases that came with the material it accreted. But hydrogen and helium are very light gases, and they bled off into space.

Earth’s second experiment with having an atmosphere went much better. Volcanic eruptions built up the atmosphere by releasing gases. The most common volcanic gases are water vapour and carbon dioxide (CO2), but volcanoes release a wide variety of gases. Other important contributions include sulphur dioxide (SO2), carbon monoxide (CO), hydrogen sulphide (H2S), hydrogen gas, and methane (CH4). Meteorites and comets also brought substantial amounts of water and nitrogen to Earth. It is not clear what the exact composition of the atmosphere was after Earth’s second experiment, but carbon dioxide, water vapour, and nitrogen were likely the three most abundant components.

One thing we can say for sure about Earth’s second experiment is that there was effectively no free oxygen (O2, the form of oxygen that we breathe) in the atmosphere. We know this in part because prior to 2 billion years ago, there were no rocks stained red from oxidized iron minerals. Iron minerals were present, but not in oxidized form. At that time, O2 was produced in the atmosphere when the sun’s ultraviolet rays split water molecules apart. However, chemical reactions removed the oxygen as quickly as it was produced.

It wasn’t until well into Earth’s third experiment — life — that the atmosphere became oxygenated. Photosynthetic organisms used the abundant CO2 in the atmosphere to manufacture their food, and released O2 as a by-product. At first all of the oxygen was consumed by chemical reactions as before, but eventually the organisms released so much O2 that it overwhelmed the chemical reactions. Oxygen began to accumulate in the atmosphere, although present levels of 21% oxygen didn’t occur until about 350 million years ago. Today the part of our atmosphere that isn’t oxygen consists largely of nitrogen (78%).

The oxygen-rich atmosphere on our planet is life’s signature. If geologic processes were the only ones controlling our atmosphere, it would consist mostly of carbon dioxide, like the atmosphere of Venus. It is an interesting notion (or a disconcerting one, depending on your point of view) that for the last 2 billion years the light reflected from our planet has been beaming a bar code out to the universe, similar to the ones in Figure 2.4, except ours says “oxygen.” For 2 billion years, our planet has been sending out a signal that could cause an observer from another world to say, “That’s odd… I wonder what’s going on over there.”

12

2.5 Are There Other Earths?

As of January 2018, 6,355 possible exoplanets– extrasolar planets, or planets outside of our solar system- have been detected by preliminary tests. Further tests have confirmed that 3,726 of those candidates are indeed planets. If “other Earths” are defined as planets where we could walk out of a spaceship with no equipment other than a picnic basket, and enjoy a pleasant afternoon on a grassy slope near a stream, then it remains to be seen whether any of these planets fit the description. On the other hand, if “other Earths” refers to rocky worlds approximately Earth’s size, and orbiting within their star’s habitable zone (the zone in which liquid water, and potentially life, can exist), then there is cautious optimism that we have found at least 53 such worlds.

Part of the uncertainty about the 53 possible Earth-like worlds is related to their composition. We don’t yet know their composition; however, it is tempting to conclude that they are rocky because they are similar in size to Earth. Remember the rules of the accretion game: you can only begin to collect gas once you are a certain size, and how much matter you collect depends on how far away from the sun you are. Given how large our gas giant and ice giant planets are compared to Earth, and how far away they are from the sun, we would expect that a planet similar in size to Earth, and a similar distance from its star, should be rocky.

But it isn’t quite as simple as that. We are finding that the rules to the accretion game can result in planetary systems very different from our own. For example, in the planetary systems we have observed, it is common to have planets larger than Earth orbiting closer to their star than Mercury does to the sun. Planets as large as Jupiter are rare, and where large planets do exist, they are much closer to their star than Jupiter is to the sun. To summarize, we need to be cautious about drawing conclusions from our own solar system, just in case we are basing those conclusions on something truly unusual.

On the other hand, the seemingly unique features of our solar system would make planetary systems like ours difficult to spot. One of the ways exoplanets are detected is by measuring the brightness of stars, and looking for tiny variations in brightness that could be caused by a planet passing between the star it orbits and the instrument observing the star. Small planets are harder to detect because they block less of a star’s light than larger planets. Larger planets farther from a star, like our gas giant planets, are difficult to spot because they don’t go past the star as frequently. For example, Jupiter goes around the sun once every 12 years. If someone were observing our solar system, they might have to watch for 12 years to see Jupiter go past the sun once. For Saturn, they might have to watch for 30 years.

If Habitable Zone Planets Are Terrestrial, Could We Live There?

The operational definition of “other Earths” involving a terrestrial composition, a size constraint of one to two times that of Earth, and location within a star’s habitable zone, does not preclude worlds incapable of supporting life as we know it. By those criteria, Venus is an “other Earth,” albeit right on the edge of the habitable zone for our sun. Venus is much too hot for us, with a constant surface temperature of 465°C (lead melts at 327°C). Its atmosphere is almost entirely carbon dioxide, and the atmospheric pressure at its surface is 92 times higher than on Earth. Any liquid water on its surface boiled off long ago. Yet the characteristics that make Venus a terrible picnic destination aren’t entirely things we could predict from its distance from the sun. They depend in part on the geochemical evolution of Venus- at one time Venus might have been a lot more like a youthful Earth. These are the kinds of things we won’t know about until we can look carefully at the atmospheres and compositions of habitable-zone exoplanets.

Keep Up-To-Date on the Exoplanet Count

Look up the latest count of potential and confirmed exoplanets in the Extrasolar Planets Catalog.

Look up the latest number of potentially habitable exoplanets in the Habitable Exoplanets Catalog.

13

Chapter 2 Summary

The topics covered in this chapter can be summarized as follows:

2.1 Starting With a Big Bang

The universe began 13.8 billion years ago when energy, matter, and space expanded from a single point. Evidence for the big bang is the cosmic “afterglow” from when the universe was still very dense, and red-shifted light from distant galaxies, which tell us the universe is still expanding.

2.2 Forming Planets from the Remnants of Exploding Stars

The big bang produced hydrogen, helium, but heavier elements come from nuclear fusion reactions in stars. Large stars make elements such as silicon, iron, and magnesium, which are important in forming terrestrial planets. Large stars explode as supernovae and scatter the elements into space.

2.3 How to Build a Solar System

Solar systems begin with the collapse of a cloud of gas and dust. Material drawn to the centre forms a star, and the remainder forms a disk around the star. Material within the disk clumps together to form planets. In our solar system, rocky planets are closer to the sun, and ice and gas giants are farther away. This is because temperatures near the sun were too high for ice to form, but silicate minerals and metals could solidify.

2.4 Earth’s First 2 Billion Years

Early Earth was heated by radioactive decay, collisions with bodies from space, and gravitational compression. Heating caused molten metal to sink to Earth’s centre and form a core, and silicate minerals to form the mantle and crust. A collision with a planet the size of Mars knocked debris into orbit around Earth, and the debris coalesced into the moon. Earth’s atmosphere is the result of volcanic degassing, contributions by comets and meteorites, and photosynthesis.

2.5 Are There Other Earths?

The search for exoplanets has identified 53 planets that are similar in size to Earth and within the habitable zone of their stars. These are thought to be rocky worlds like Earth, but the compositions of these planets are not known for certain.

Review Questions

1. How can astronomers view events that happened in the universe’s distant past?

2. In this image of three spectra, one is from the sun, and the other two are from galaxies. One of the galaxies is the Andromeda galaxy. Which spectrum is from Andromeda?

Spectra for the sun and two galaxies. [KP]
Spectra for the sun and two galaxies. Source: Karla Panchuk (2015) CC BY 4.0.

3. Astronomers looking for some of the earliest stars in the universe were surprised to find a planetary system called HIP 11952, which existed 12.8 billion years ago. This was very early in the universe’s history, when stars still consisted largely of hydrogen and helium. Do you think there were terrestrial planets in this system? Why or why not?

4. Summarize the trends in size and composition of objects in the solar system.

5. What is the frost line, and what does it help to explain?

6.  Why is Pluto not considered a planet?

7. What is differentiation?

8. The exoplanet Kepler-452b is within the habitable zone of its star. In our solar system, planets a similar distance from the Sun are terrestrial planets. Why can we not say for certain that Kepler-452b’s distance from its star means it is a terrestrial planet?

9. Of the planetary systems discovered thus far, none are exactly like our solar system. Does this mean our solar system is unique in the universe?

14

Answers to Chapter 2 Review Questions

1. To see an event, light from that event must reach our eyes. Light travels very quickly (about 300,000,000 m/s), but the universe is very, very large. Depending on how far away the event was, it could take billions of years for light to travel from the event to our eyes so we can see it. Astronomers take advantage of this fact to view the universe’s past.

2. B is the spectrum from the Andromeda galaxy. We know that one spectrum represents the sun, which is not moving toward or away from us. (Our orbit is not perfectly circular, but the small eccentricity is not a factor in this comparison.) We know that the Andromeda galaxy is on a collision course with us, so it is the exception to the rule that galaxies are moving away from us, and their light is red-shifted. That means the spectrum B which is shifted furthest to the left (blue-shifted) is Andromeda, and spectrum A which is furthest to the right (red-shifted) is a galaxy moving away from us. That means C is the sun.

Spectra for the sun and two galaxies. [KP]
Spectra for the sun and two galaxies. Source: Karla Panchuk (2015) CC BY 4.0

3. The planetary system consisted of two Jupiter-sized gas giant planets. Gas giant planets contain large amounts of hydrogen, and hydrogen was plentiful in the early universe. In contrast, terrestrial planets have heavier elements, especially silica, iron, magnesium, and nickel, that had yet to be manufactured by stars. Those elements were not present in sufficient abundance to form terrestrial planets until much later.

4. Closest to the sun we find the small, rocky, terrestrial planets with metal cores. Further out are the gas giant planets, which are the largest in the solar system. They consist mostly of hydrogen, and have cores of rock and ice. Beyond the gas giant planets are the ice giant planets, which are next largest. They have a mantle of ice (not just water ice but ammonia and methane ice), and a rocky core. Smaller objects in the solar system include rocky bodies within the asteroid belt between Mars and Jupiter, and bodies of ice and dust in the Kuiper belt and Oort cloud beyond Neptune.

5. The frost line marks the distance from the sun beyond which temperatures were cool enough to allow ice to form. This helps to explain why the terrestrial planets are closer to the sun, and the Jovian and ice giant planets farther away. Mineral grains could solidify and begin to accrete closer to the sun, forming terrestrial planets, because they have higher melting points. In contrast, water vapour, methane, and ammonia had to be farther from the sun before they could freeze and begin to accrete.

6. Planets are defined as having cleared their orbits of debris. Pluto is located within the Kuiper belt, so it shares its orbit with other objects. There are two other criteria in the definition of a planet: planets in our solar system must orbit the sun, and they must have a spherical shape. Pluto satisfies both these criteria, but sadly the people deciding whether or not Pluto should be a planet are not amenable to a “best two out of three” compromise.

7. Differentiation is the separation of materials within a planet such that dense materials sink to the core. In Earth’s case, the denser materials are iron and nickel.

8. The fact that we have terrestrial planets close to the sun makes sense in terms of the frost line, but it does not seem to be a hard-and-fast rule in other planetary systems. Therefore, we can’t conclude from Kepler-452b’s position alone that it is a terrestrial planet.

9. The rules of the accretion game mean that there are many complex interactions, so even a small difference in the starting conditions or in how the game goes in the beginning could have major implications in the end. For that reason, we shouldn’t expect to find a planetary system that matches ours in every minute detail. However, just because we haven’t found a similar planetary system does not mean one does not exist. Our planet-finding methods are biased toward discovering large planets orbiting close to their stars, whereas our solar system has small planets close to the sun and larger ones farther away. That doesn’t mean our methods won’t eventually turn up a system like ours, just that they are more likely to turn up systems that are different.

 

III

Chapter 3. Earth's Interior

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 3.1 The red rocks of the Tablelands in Gros Morne National Park are a sample of Earth’s mantle. Top: The red rocks of the Tablelands are on the right, and stand in contrast with the green surroundings. Bottom: A closer view of Tablelands terrain, showing rocks weathered red, and a near absence of plant life.  Source: Top photograph- Leos Kral (2008) CC BY-NC-SA 2.0 view source; Bottom photograph: Tara Joyce (2013) CC BY-SA 2.0 view source. Click the image for more attributions.

 

Learning Objectives

After reading this chapter and completing the review questions at the end, you should be able to:

 

The barren red rocks of the Tablelands stand in stark contrast to their lush green surroundings in Gros Morne National Park (Figure 3.1, top). If the Tablelands appear out of place, it’s because they are. The Tablelands are one of few places on Earth where you can walk directly on the rocks of Earth’s mantle, thanks to an accident of plate tectonics that happened hundreds of millions of years ago. The red colour of the Tablelands rocks comes from iron-bearing minerals reacting with oxygen. Unaltered, the rocks are dark green (Figure 3.2). The rocks lack vegetation because the chemical composition of the rocks does not provide adequate nutrients for plants.

Figure 3.2 Tablelands mantle rock with reddish weathering rind, and dark green fresh surface. Scale in cm. Source: Karla Panchuk (2017) CC BY 4.0

Locations like the Tablelands are one way we can learn about Earth’s interior. Meteorites derived from smashed differentiated bodies (asteroids that separated into mantle and core) are another. Asteroids that formed at a similar distance from the sun as Earth had a mineral composition akin to Earth’s. When these objects were shattered in giant collisions, the result was stony meteorites from fragmented mantle rock, and iron meteorites from fragmented core. Some fragments sampled the result of violent encounters that mixed the two (Figure 3.3).

Figure 3.3 Cut and polished slab of a stony-iron meteorite called a pallasite, thought to have formed in a collision that smashed mantle rocks against the metal core of an asteroid early in the solar system’s history. Green and brown crystals are the mineral olivine. The metal between the olivine crystals is an iron-nickel mineral. Source: Muséum de Toulouse (2012) CC BY-NC 2.0 view source

We also get information about the structure of Earth’s interior by analyzing the speeds and paths of earthquake vibrations, called seismic waves.

We need to know something about the inside of our planet— what it’s made of, and what happens within it— in order to understand how Earth works, especially the mechanisms of plate tectonics. It is fortunate that there are many ways for geologists gather information about Earth’s interior, because one thing they can’t do is go down and look at it.

15

3.1 Earth's Layers: Crust, Mantle, and Core

Earth consists of three main layers: the crust, the mantle, and the core (Figure 3.4).  The core accounts for almost half of Earth’s radius, but it amounts to only 16.1% of Earth’s volume.  Most of Earth’s volume (82.5%) is its mantle, and only a small fraction (1.4%) is its crust.

Figure 3.4 Earth’s interior. Right- crust, mantle, and outer and inner core to scale.  Left- Cutaway showing continental and ocean crust, and upper mantle layers. The lithosphere is the crust plus the uppermost layer of the mantle. Source: Karla Panchuk (2018) CC BY 4.0. Earth photo by NASA (n.d.) Public Domain view source

Crust

The Earth’s outermost layer, its crust, is rocky and rigid. There are two kinds of crust: continental crust, and ocean crust. Continental crust is thicker, and predominantly felsic in composition, meaning that it contains minerals that are richer in silica. The composition is important because it makes continental crust less dense than ocean crust.

Ocean crust is thinner, and predominantly mafic in composition.  Mafic rocks contain minerals with less silica, but more iron and magnesium. Mafic rocks (and therefore ocean crust) are denser than the felsic rocks of continental crust.

The crust floats on the mantle.  Continental crust floats higher in the mantle than ocean crust because of the lower density of continental crust.  An important consequence of the difference in density is that if tectonic plates happen to bring ocean crust and continental crust into collision, the plate with ocean crust will be forced down into the mantle beneath the plate with continental crust.

Mantle

The mantle is almost entirely solid rock, but it is in constant motion, flowing very slowly. It is ultramafic in composition, meaning it has even more iron and magnesium than mafic rocks, and even less silica.  Although the mantle has a similar chemical composition throughout, it has layers with different mineral compositions and different physical properties.  It can have different mineral compositions and still be the same in chemical composition because the increasing pressure deeper in the mantle causes mineral structures to be reconfigured.

Rocks higher in the mantle are typically composed of peridotite, a rock dominated by the minerals olivine and pyroxene. The Tablelands rock in Figure 3.2 is a type of peridotite. Lower in the mantle, extreme pressures transform minerals and create rocks like eclogite (Figure 3.5), which contains garnets.

Figure 3.5 Eclogite from the Swiss-Italian Alps. Reddish brown spots are garnets. Source: James St. John (2014) CC BY 2.0 view source

Lithosphere

The lithosphere can’t be classified neatly as either crust or mantle because it consists of both.  It is formed from the crust as well as the uppermost layer of the mantle which is stuck to the underside of the crust.  Tectonic plates are fragments of lithosphere.

Asthenosphere

Beneath the lithosphere is the asthenosphere.Tiny amounts of melted rock dispersed through the otherwise solid asthenosphere make the asthenosphere weak compared to the lithosphere. The weakness of the asthenosphere is important for plate tectonics because it deforms as fragments of lithosphere move around upon and through it. Without a weak asthenosphere, plates would be locked in place, unable to move as they do now.

D”

The D” (dee double prime) layer is a mysterious layer beginning approximately 200 km above the boundary between the core and mantle.  (This boundary is referred to as the core-mantle boundary.)  We know it exists because of how seismic waves change speed as they move through it, but it isn’t clear why it’s different from the rest of the mantle.  One idea is that it is minerals are undergoing another transition in this region because of pressure and temperature conditions, similar to the transition between the upper and lower mantle. Other ideas are that small pools of melt are present, or that the differences in seismic properties are due to subducted slabs of lithosphere resting on the core-mantle boundary.

Core

The core is primarily composed of iron, with lesser amounts of nickel. Lighter elements such as sulfur, oxygen, or silicon may also be present. The core is extremely hot (~3500° to more than 6000°C). But despite the fact that the boundary between the inner and outer core is approximately as hot as the surface of the sun, only the outer core is liquid. The inner core is solid because the pressure at that depth is so high that it keeps the core from melting.

 

16

3.2 Imaging Earth's Interior

Seismology is the study of vibrations within Earth. These vibrations are caused by events such as earthquakes, extraterrestrial impacts, explosions, storm waves hitting the shore, and tides. Seismology is applied to the detection and study of earthquakes, but seismic waves also provide important information about Earth’s interior.

Seismic waves travel through different materials at different speeds, and we can apply knowledge of how they interact with different materials to understand Earth’s layers and internal structures. Similar to the way that ultrasound is used to image the human body, we can measure how long it takes for seismic waves to travel from their source to a recording station.

Another feature of seismic waves is that some, called P-waves, can travel rapidly though both liquids and solids, but others, called S-waves, can only travel though solids, and are slower than P-waves. Observing where P-waves travel, and S-waves do not, allows us to identify regions within Earth that are melted.

Seismic Wave Paths

Seismic waves travel in all directions from their source, but it is convenient to imagine the path traced by one point on the wave front, and represent that path as a seismic ray (arrows, Figure 3.6).

Figure 3.6 Seismic waves and seismic rays. The paths of seismic waves can be represented as rays. Seismic ray paths are bent when they enter a rock layer with a different seismic velocity. Source: Karla Panchuk (2018) CC BY 4.0

When seismic waves encounter a different rock layer, some might bounce off the layer, or reflect, as in the bottom layer of Figure 3.6. But some waves will travel through the layer. If the wave travels at a different speed in the new layer, its path will be bent, or refracted, as it crosses into the new layer. If the wave can travel faster in the new layer, it will be bent slightly toward the contact between the two layers. In Figure 3.6, the ray can travel progressively faster in each layer as it goes down through the layers, and it is bent slightly upward each time it crosses into the next layer. The reverse happens if the wave slows down. On the right side of the diagram, the wave is moving upward through slower and slower layers. It is bent away from the faster layer each time, causing it to take a more direct path to the surface.

Seismic velocities are higher in more rigid layers, so broadly speaking, they get faster deeper within Earth, because higher pressures make layers more rigid. They tend to take curved paths through the Earth because refraction bends their path until they are reflected and directed upward again, as in Figure 3.6.

Discoveries with Seismic Waves

The Moho: Where Crust Meets Mantle

One of the first discoveries about Earth’s interior made through seismology was in the early 1900s by Croatian seismologist Andrija Mohorovičić (pronounced Moho-ro-vi-chich).  He noticed that sometimes, seismic waves arrived at seismic stations (measuring locations) farther from an earthquake before they arrived at closer ones.  He reasoned that the waves that traveled farther were faster because they bent down and traveled faster through different rocks (those of the mantle) before being bent upward back into the crust (Figure 3.7).

Figure 3.7 Depiction of seismic waves emanating from an earthquake (red star). Some waves travel through the crust to the seismic station (at ~6 km/s), while others go down into the mantle (where they travel at ~8 km/s) and are bent upward toward the surface, reaching the station before the ones that travelled only through the crust. Source: Steven Earle (2016) CC BY 4.0 view source

The boundary between the crust and the mantle is now known as the Mohorovičić discontinuity (or Moho). Its depth is between 60 – 80 km beneath major mountain ranges, 30 – 50 km beneath most of the continental crust, and 5 – 10 km beneath ocean crust.

The Core-Mantle Boundary

Arguments for a liquid outer core were supported by a distinctive signature in the global distribution of seismic waves from earthquakes. When an earthquake occurs, there is a zone on the opposite side of Earth where S-waves are not measured. This S-wave shadow zone begins 103° on either side of the earthquake, for a total angular distance of 154° (Figure 3.8, left). There is also a P-wave shadow zone on either side of the earthquake, from 103° to 150° (Figure 3.8, right).

Figure 3.8 Patterns of seismic wave propagation through Earth’s mantle and core. S-waves do not travel through the liquid outer core, so they leave a shadow on Earth’s far side. P-waves do travel through the core, but because the waves that enter the core are refracted, there are also P-wave shadow zones. Source: Steven Earle (2016) CC BY 4.0 view source

The S-wave shadow zone occurs because S-waves cannot travel through the liquid outer core. The P-wave shadow zone occurs because seismic velocities are much lower in the liquid outer core than in the overlying mantle, and the P-waves are refracted in a way that leaves a gap. Not only do the shadow zones tell us that the outer core is liquid, the size of the shadow zones allows us to calculate the size of the core, and the location of the core-mantle boundary.

Seismic Portrait of Earth’s Layers

The change seismic wave velocity with depth in Earth (Figure 3.9) has been determined over the past several decades by analyzing seismic signals from large earthquakes all around the world. Earth’s layers are detectable as changes in velocity with depth. The asthenosphere is visible as a low velocity zone within the upper mantle (Figure 3.9, left). There is an abrupt increase in P-wave velocity at 420 km, showing the depth at which minerals transform into structures that are more stable at higher pressures and temperatures.

Figure 3.9 P-wave and S-wave velocity variations with depth from the crust through the upper mantle (left) and from the crust through to the core (right). Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source left/ right

The boundary between the upper and lower mantle is visible at 660 km as a sudden change from rapidly increasing P- and S-wave velocities to slow or no change in P-wave and S-wave velocities (Figure 3.9, right). The core-mantle boundary (CMB in Figure 3.9) is apparent as a sudden drop in P-wave velocities, where seismic waves move from solid mantle to liquid outer core. The boundary between the outer core and inner core is marked by a sudden increase in P-wave velocity after 5000 km, where seismic waves move from a liquid back into a solid again.

Seismic Images of Plate Tectonic Structures

Using data from many seismometers and hundreds of earthquakes, it is possible to create images from the seismic properties of the mantle. This technique is known as seismic tomography. Tomography can be used to map out slabs of lithosphere that are entering the mantle, or have disappeared within it. Those slabs are cooler, and therefore more rigid than surrounding mantle rocks, so seismic waves travel through them faster. In Figure 3.10, higher-than-average seismic velocities in cool slabs are indicated in dark blue.

Figure 3.10 P-waves and S-waves used to map out the location of the Cocos slab of lithosphere. The slab appears in dark blue, indicating higher than average seismic wave velocities. Left- Tomograms showing seismic wave anomalies for a 1290 km surface. Right- Cross-sections along the transect marked X-Y on the globe. Source: Karla Panchuk (2018) CC BY 4.0, modified after van der Meer et al. (2018) CC BY 4.0 view source

Thanks to the tomograms, we can see that the Cocos plate, which is colliding with Central America, is part of a much larger slab of lithosphere that has already settled onto the mantle. Tomograms representing a surface at 1290 km depth (Figure 3.10, left) show that at that level, the Cocos slab is beneath the Caribbean Sea. The tomograms on the right show a vertical view along the line X-Y marked on the globe. The vertical tomograms show us that the Cocos slab extends all the way down to the core-mantle boundary.

Visit the Underworld

What is the Atlas of the Underworld?

The Atlas of the Underworld is a catalog of more than 90 slabs of lithosphere that have been imaged within the mantle using seismic tomography. The Atlas includes tomographic images, locator maps, and geological histories for each slab. The catalog can be searched online at http://www.atlas-of-the-underworld.org/ or viewed in the original publication by van der Meer et al. (2018). The Atlas of the Underworld is an open-access resource.  Visit the Atlas of the Underworld

The HADES Underworld Explorer

Create your own tomographic cross-sections for locations anywhere in the world by using this intuitive drag-and-drop tool. Visit the HADES Underworld Explorer

References

van der Meer, D.G., van Hinsbergen, D.J.J., and Spakman, W., (2018). Atlas of the Underworld: slab remnants in the mantle, their sinking history, and a new outlook on lower mantle viscosity. Tectonophysics 723, p. 309-448. https://doi.org/10.1016/j.tecto.2017.10.004

17

3.3 Earth's Interior Heat

Earth Gets Hotter the Deeper You Go

Earth’s temperature increases with depth, but not at a uniform rate (Figure 3.11). Earth’s geothermal gradient is 15° to 30°C/km within the crust.  It then drops off dramatically through the mantle, increases more quickly at the base of the mantle, and then increases slowly through the core. The temperature is approximately 1000°C at the base of the crust, around 3500°C at the base of the mantle, and approximately 6,000°C at Earth’s centre.

Figure 3.11 Geothermal gradient (change in temperature with depth). Left- Geothermal gradient in the crust and upper mantle. The geothermal gradient remains below the melting temperature of rock, except in the asthenosphere. There, temperatures are high enough to melt some of the minerals. Right- Geothermal gradient throughout Earth. Rapid changes occur in the uppermost mantle, and at the core-mantle boundary. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source left/ right

The temperature gradient within the lithosphere varies depending on the tectonic setting. Gradients are lowest in the central parts of continents, higher where plates collide, and higher still at boundaries where plates are moving away from each other.

In spite of high temperatures within Earth, mantle rocks are almost entirely solid. High pressures keep them from melting. The red dashed line in Figure 3.11 (right) shows the minimum temperature at which dry mantle rocks will melt. Rocks at temperatures to the left of the line will remain solid. In rocks at temperatures to the right of the line, some minerals will begin to melt. Notice that the red dashed line goes further to the right for greater depths, and therefore greater pressures. Now compare the geothermal gradient with the red dashed line. The geothermal gradient is to the left of the red line, except in the asthenosphere, where small amounts of melt are present.

Convection Helps to Move Heat Within Earth

The fact that the temperature gradient is much lower in the main part of the mantle than in the lithosphere has been interpreted as evidence of convection in the mantle. When the mantle convects, heat is transferred through the mantle by physically moving hot rocks. Mantle convection is the result of heat transfer from the core to the base of the lower mantle. As with a pot of soup on a hot stove (Figure 3.12), the material near the heat source (the soup at the bottom of the pot) becomes hot and expands, making it less dense than the material above. Buoyancy causes it to rise, and cooler material flows in from the sides. Of course, convection in the soup pot is much faster than convection in the mantle. Mantle convection occurs at rates of centimetres per year.

Figure 3.12 Convection in a pot of soup on a hot stove (left). As long as heat is being transferred from below, the liquid will convect. If the heat is turned off (right), the liquid remains hot for a while, but convection will cease. Source: Steven Earle (2015) CC-BY 4.0 view source

Convection carries heat to the surface of the mantle much faster than heating by conduction. Conduction is heat transfer by collisions between molecules, and is how heat is transferred from the stove to the soup pot. A convecting mantle is an essential feature of plate tectonics, because the higher rate of heat transfer is necessary to keep the asthenosphere weak. Earth’s mantle will stop convecting once the core has cooled to the point where there is not enough heat transfer to overcome the strength of the rock. This has already happened on smaller planets like Mercury and Mars, as well as on Earth’s moon. When mantle convection stops, the end of plate tectonics will follow.

Models of Mantle Convection

In the soup pot example, convection moves hot soup from the bottom of the pot to the top. Some geologists think that Earth’s convection works the same way— hot rock from the base of the mantle moves all the way to the top of the mantle before cooling and sinking back down again. This view is referred to as whole-mantle convection (Figure 3.8, left). Other geologists think that the upper and lower mantle are too different to convect as one. They point to slabs of lithosphere that are sinking back into the mantle, some of which seem to perch on the boundary between the upper and lower mantle, rather than sinking straight through. They also note chemical differences in magma originating in different parts of the mantle— differences that are not consistent with the entire mantle being well stirred. They argue that double-layered convection is a better fit with the observations (Figure 3.13, right).  Still others argue that there may be some locations where convection goes from the bottom of the mantle to the top, and some where it doesn’t (Figure 3.13, middle).

Figure 3.13 Models of mantle convection. Left- whole mantle convection. Rocks rise from the core-mantle boundary to the top of the mantle, then sink to the bottom again. Right- Two-layer convection, in which upper and lower mantle convect at different rates. Middle- Convection paths vary depending on the circumstances. Source: Karla Panchuk (2018) CC BY 4.0

Why Is Earth Hot Inside?

The heat of Earth’s interior comes from a variety of sources. These include the heat contained in the objects that accreted to form Earth, and the heat produced when they collided. As Earth grew larger, the increased pressure on Earth’s interior caused it to compress and heat up.  Heat also came from friction when melted material was redistributed within Earth, forming the core and mantle.

A major source of Earth’s heat is radioactivity, the energy released when the unstable atoms decay. The radioactive isotopes uranium-235 (235U), uranium-238 (238U), potassium-40 (40K), and thorium-232 (232Th) in Earth’s mantle are the primary source. Radioactive decay produced more heat early in Earth’s history than it does today, because fewer atoms of those isotopes are left today (Figure 3.14). Heat contributed by radioactivity is now roughly a quarter what it was when Earth formed.

Figure 3.14 Production of heat within the Earth over time by radioactive decay of uranium, thorium, and potassium. Heat production has decreased over time as the abundance of radioactive atoms has decreased. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Arevalo et al. (2009)

References

Arevalo, R., McDonough, W., & Luong, M. (2009). The K/U ratio of Earth: Insights into mantle composition, structure and thermal evolution. Earth and Planetary Science Letters, 278(3-4), 361-369. https://doi.org/10.1016/j.epsl.2008.12.023

18

3.4 Earth's Magnetic Field

Earth’s liquid iron core convects because it is heated from beneath by the inner core. Because iron is a metal and conducts electricity (even when molten), its motion generates a magnetic field.

Earth’s magnetic field is defined by north and south poles representing lines of magnetic force flowing into Earth in the northern hemisphere and out of Earth in the southern hemisphere (Figure 3.15). Because of the shape of the field lines, the magnetic force is oriented at different angles to the surface in different locations. The tilt, or inclination of magnetic field lines is represented by the tilt of compass needles in Figure 3.15. At the north and south poles, the force is vertical. The force is horizontal at the equator. Everywhere in between, the magnetic force is at an intermediate angle to the surface.

Figure 3.15 Earth’s magnetic field depicted as the field of a bar magnet coinciding with the core. The south pole of the magnet points to Earth’s magnetic north pole. The red and white compass needles represent the orientation of the magnetic field at various locations on Earth’s surface. Source: Karla Panchuk (2018) CC BY-SA 4.0, modified after Steven Earle (2015) CC BY-SA 4.0 view source, and T. Stein (2008) CC BY-SA 3.0 view source

 

Exercise: Magnetic Inclination

Regular compasses point only to the north magnetic pole, but if you had a magnetic dip meter (or a smartphone with the appropriate app), you could also measure the angle of the magnetic field at your location in the up-and-down sense. However, you don’t need a dip meter or app to do this exercise!

Using Figure 3.15 as a guide, describe the general location on Earth where the vertical angles would be as follows:

  1. Straight down
  2. Down at a steep angle
  3. Up at a steep angle
  4. Parallel to flat ground

Earth’s magnetic field is generated within the outer core by the convective movement of liquid iron, but although convection is continuous, the magnetic field is not stable. Periodically, the magnetic field decays and then becomes re-established. When it does re-establish, the polarity may have reversed (i.e., your compass would point south rather than north). Over the past 250 Ma, there have been hundreds of magnetic field reversals, and their timing has been anything but regular. The shortest ones that geologists have been able to identify lasted only a few thousand years, and the longest one was more than 30 million years, during the Cretaceous Period (Figure 3.16).

Figure 3.16 Magnetic field reversal chronology for the past 170 Ma. Black stripes mark times when the magnetic field was oriented the same as today. Source: Steven Earle (2015) CC BY 4.0 view source, modified after AnomieX (2010) Public Domain view source

Changes in Earth’s magnetic field have been studied using mathematical models that simulate convection in the outer core (Figure 3.17).  Reversals happened spontaneously when the model was run to simulate a period of several hundred thousand years. Spontaneous reversals can happen because convection does not occur in an orderly way, in spite of what the bar magnet analogy may suggest. Many small-scale variations occur in convection patterns within the inner core, and Earth’s magnetic field over all is the sum of those variations. Magnetic reversals do not happen as frequently as they might, if not for the solid inner core. Magnetic field changes take much longer within the inner core, so reversals in the outer core do not always coincide with reversals in the inner core. Both are required in order for Earth’s magnetic field to flip.

Figure 3.17 Earth’s magnetic field between reversals (left) and during a reversal (right). The lines represent magnetic field lines: blue where the field points toward Earth’s centre and yellow where it points away. The rotation axis of Earth is vertical, and the outline of the core is shown as a dashed white circle. Source: NASA (2007) Public Domain view source

References

British Geological Survey, Natural Environment Research Council (n.d.). Reversals: Magnetic Flip. Visit website

Glatzmaier, G. A. (n.d.) The Geodynamo. Visit website

19

3.5 Isostasy

Lithospheric Plates Float on the Mantle

The mantle is able to convect because it can deform by flowing over very long timescales. This means that tectonic plates are floating on the mantle, like a raft floating in the water, rather than resting on the mantle like a raft sitting on the ground. How high the lithosphere floats will depend on the balance between gravity pulling the lithosphere down, and the force of buoyancy as the mantle resists the downward motion of the lithosphere. Isostasy is the state in which the force of gravity pulling the plate toward Earth’s centre is balanced by the resistance of the mantle to letting the plate sink.

To see how isostasy works, consider the rafts in Figure 3.18. The raft on the right is sitting on solid concrete. The raft will remain at the same elevation whether there are two people on it, or four, because the concrete is too strong to deform. In contrast, isostasy is in play for the rafts on the left, which are floating in a swimming pool full of peanut butter. With only one person on board, the raft floats high in the peanut butter, but with three people, it sinks dangerously low. Peanut butter, rather than water, is used in this example because the viscosity of peanut butter (its stiffness or resistance to flowing) more closely represents the relationship between the tectonic plates and the mantle. Although peanut butter has a similar density to water, it’s higher viscosity means that if a person is added to a raft, it will take longer for the raft to settle lower into the peanut butter that it would take the raft to sink into water.

Figure 3.18 Illustration of isostatic relationships between rafts and peanut butter (left), and a non-isostatic relationship between a raft and solid ground (right). Source: Steven Earle (2015) CC BY 4.0 view source

The relationship of Earth’s crust to the mantle is similar to the relationship of the rafts to the peanut butter. The raft with one person on it floats comfortably high. Even with three people on it the raft is less dense than the peanut butter, so it floats, but it floats uncomfortably low for those three people. The crust, with an average density of around 2.6 g/cm3, is less dense than the mantle (average density of ~3.4 g/cm3 near the surface, but more at depth), and so it is floating on the mantle. When weight is added to the crust through the process of mountain building, the crust slowly sinks deeper into the mantle, and the mantle material that was there is pushed aside (Figure 3.19, left). When erosion removes material from the mountains over tens of millions of years, decreasing the weight, the crust rebounds and the mantle rock flows back (Figure 3.19, right).

Figure 3.19 Isostatic relationship between the crust and the mantle. Mountain building adds mass to the crust, and the thickened crust sinks down into the mantle (left). As the mountain chain is eroded, the crust rebounds (right). Green arrows represent slow mantle flow. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source

Isostasy and Glacial Rebound

The crust and mantle respond in a similar way to glaciation. Thick accumulations of glacial ice add weight to the crust, and the crust subsides, pushing the mantle out of the way. The Greenland ice sheet, at over 2,500 m thick, has depressed the crust below sea level (Figure 3.20a). When the ice eventually melts, the crust and mantle will slowly rebound (Figure 3.20b), but full rebound will likely take more than 10,000 years (3.20c).

Figure 3.20 Cross-section through the crust in the northern part of Greenland. a) Up to 2,500 m of ice depresses the crust downward (red arrows) and below sea level. b) After complete melting. Isostatic rebound would be slower than the rate of melting, leaving central Greenland at or below sea level for thousands of years. c) Complete rebound after ~10,000 years raises central Greenland above sea level again. Source: Steven Earle (2015) CC BY 4.0 view source a/ b/ c

Large parts of Canada are still rebounding as a result of the loss of glacial ice over the past 12,000 years, as are other parts of the world (Figure 3.21). The highest rate of uplift is in a large area west of Hudson Bay, where the Laurentide Ice Sheet was the thickest, at over 3,000 m. Ice finally left this region around 8,000 years ago, and the crust is currently rebounding at nearly 2 cm/year. Strong isostatic rebound is also occurring in northern Europe where the Fenno-Scandian Ice Sheet was thickest, and in the eastern part of Antarctica, which also experienced significant ice loss during the Holocene.

Glacial rebound in one location means subsidence in surrounding areas (Figure 3.21, yellow through red regions). Regions surrounding the former Laurentide and Fenno-Scandian Ice Sheets that were lifted up when mantle rock was forced aside and beneath them are now subsiding as the mantle rock flows back.

Figure 3.21 Current rates of post-glacial isostatic uplift (green, blue, and purple shades) and subsidence (yellow and orange). Subsidence is taking place where the mantle is slowly flowing back toward areas that are experiencing post-glacial uplift. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Erik Ivins, JPL (2010) Public Domain view source

How Can the Mantle Be Both Solid and Plastic?

You might be wondering how it is possible that Earth’s mantle is solid, rigid rock, and yet it convects and flows like a very viscous liquid. The explanation is that the mantle behaves as a non-Newtonian fluid, meaning that it responds differently to stresses depending on how quickly the stress is applied.

A good example of non-Newtonian behaviour is the deformation of Silly Putty, which can bounce when it is compressed rapidly when dropped, and will break if you pull on it sharply. But will deform in a liquid manner if stress is applied slowly. The force of gravity applied over a period of hours can cause it to deform like a liquid, dripping through a hole in a glass tabletop (Figure 3.22). Similarly, the mantle will flow when placed under the slow but steady stress of a growing (or melting) ice sheet.

Figure 3.22 Silly Putty exhibiting plastic behavior when acted upon by gravity over several hours. Source: Erik Skiff (2006) CC BY-SA 2006 view source

20

Chapter 3 Summary

The topics covered in this chapter can be summarized as follows:

3.1 Earth’s Layers

Earth is divided into a rocky crust and mantle, and a core consisting largely of iron. The crust and the uppermost mantle form the lithosphere, which is broken into tectonic plates.  The next layer, the asthenosphere, allows the plates to move because it deforms by flowing.

3.2 Imaging Earth’s Interior

Seismic waves that travel through Earth are either P-waves or S-waves. P-waves are faster than S-waves, and can pass through fluids. Earth’s layers can be identified by looking at changes in the velocity of seismic waves. Seismic wave shadow zones contributed to knowledge of the depth of the core-mantle boundary, and the knowledge that the outer core is liquid.  Plate tectonic structures within Earth can also be mapped using the seismic waves generated by earthquakes.

3.3 Earth’s Interior Heat

Earth’s temperature increases with depth (to around 6000°C at the centre), but the rate of increase is not the same everywhere. In the lithosphere, thickness and plate tectonic setting are are factors. Deeper within the mantle, convection currents are more important.

3.4 Earth’s Magnetic Field

Earth’s magnetic field is generated by convection of the liquid outer core. The magnetic field is similar to that of a bar magnet, and has force directions that vary with latitude. The polarity of the field is not constant, meaning that the positions of the north and south magnetic poles have flipped from “normal” (as it is now) to reversed and back many times in Earth’s history.

3.5 Isostasy

The plastic nature of the mantle, which allows for mantle convection, also determines the nature of the relationship between the crust and the mantle. The crust floats on the mantle in an isostatic relationship. Where the crust becomes thicker and heavier because of mountain building, it pushes farther down into the mantle. Oceanic crust, being denser than continental crust, floats lower on the mantle than continental crust.

Questions for Review

  1. What parts of Earth are most closely represented by typical stony meteorites and typical iron meteorites?
  2. Draw a simple diagram of Earth’s layers, and label the approximate locations of the following boundaries: crust/mantle, mantle/core, outer core/inner core.
  3. How do P-waves and S-waves differ?
  4. Why does P-wave velocity decrease dramatically at the core-mantle boundary?
  5. Why do both P-waves and S-waves gradually bend as they move through the mantle?
  6. What is the evidence for mantle convection, and what causes mantle convection?
  7. How is Earth’s magnetic field generated?
  8. When were the last two reversals of Earth’s magnetic field?
  9. What property of the mantle is essential for the isostatic relationship between the crust and the mantle?
  10. How would you expect the depth to the crust-mantle boundary in the area of the Rocky Mountains to differ from that in central Saskatchewan?
  11. British Columbia is still experiencing weak post-glacial isostatic uplift, especially in the interior, but also along the coast (see Figure 3.21). Meanwhile, offshore areas are experiencing weak isostatic subsidence. Why?

21

Answers to Chapter 3 Review Questions

1. Stony meteorites are similar in composition to Earth’s mantle, while iron meteorites are similar to the core.

2. Compare your answer to Figure 3.4.

3. P-waves can pass through a liquid, and travel approximately twice as fast as S-waves (which cannot pass through a liquid).

4. P-wave velocity decreases at the core-mantle boundary because the outer core is liquid.

5. The mantle gets increasingly dense and strong with depth because of the increasing pressure. This difference affects both P-wave and S-wave velocities, and they are refracted toward the lower density mantle material (meaning they are bent out toward Earth’s surface).

6. The key evidence for mantle convection is that the rate of temperature increase with depth within the mantle is less than expected. This can only be explained by a mantle that is mixing by convection. The mechanism for convection is the transfer of heat from the core to the mantle, causing the to mantle flow.

7. Earth’s magnetic field is generated within the liquid outer core because liquid metal is convecting.

8. The last two reversals of Earth’s magnetic field were at the beginning of the present Brunhes normal chron (0.78 Ma), and at the end of the Jaramillo normal subchron (0.90 Ma). 

9. The isostatic relationship between the crust and the mantle is dependent on the fact that over very long timescales, the mantle deforms by flowing.

10. In the area of the Rocky Mountains the crust is thickened and pushed down into the mantle. In Saskatchewan the crust is thinner and does not extend as far into the mantle.

11. During the Pleistocene glaciation, British Columbia was pushed down by glacial ice. Mantle rock flowed slowly out from under the weighted-down crust and toward the ocean floor. Now that the land area is rebounding, that mantle rock is flowing back and the offshore areas are subsiding.

IV

Chapter 4. Plate Tectonics

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 4.1 Iceland is known for its volcanoes, which are present because Iceland is located on the Mid-Atlantic Ridge, where the Atlantic Ocean is spreading apart and new crust is forming. In fact, Iceland exists because that volcanic activity has built up the island from the ocean floor. Iceland is cut by rift zones (white lines on the map at left) where the island is splitting apart along with the rest of the Atlantic Ocean. Rift zones are marked by belts of young volcanic rocks (dark green). You can stand on a rift zone if you visit Thingvellir National Park (right). Rifting has produced a valley where the crust has settled downward. The margins of the North American and Eurasian tectonic plates are visible as ridges on either side of the valley. The photographer was standing on a ridge on the North American side. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photo: Ruth Hartnup (2005) CC BY 2.0 view source. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

Plate tectonics is the model or theory that we use to understand how our planet works: it explains the origins of continents and oceans, the origins of folded rocks and mountain ranges, the presence of different kinds of rocks, the causes and locations of earthquakes and volcanoes, and changes in the positions of continents over time. So… everything!

The theory of plate tectonics was proposed to the geological community more than 100 years ago, so it may come as a surprise that an idea underpinning the study of Earth today did not become an accepted part of geology until the 1960s. It took many decades for this theory to become accepted for two main reasons. First, it was a radically different perspective on how Earth worked, and geologists were reluctant to entertain an idea that seemed preposterous in the context of the science of the day. The evidence and understanding of Earth that would have supported plate tectonic theory simply didn’t exist until the mid-twentieth century. Second, their opinion was affected by their view of the main proponent, Alfred Wegener. Wegener was not trained as a geologist, so he lacked credibility in the eyes of the geological community. Alfred Wegener was also German, whereas the geological establishment was centred in Britain and the United States- and Britain and the United States were at war with Germany in the first part of the 20th century. In summary, plate tectonics was an idea too far ahead of its time, and delivered by the wrong messenger.

References

Thordarson, T., and Larsen, G. (2007) Volcanism in Iceland in historical time: Volcano types, eruption styles and eruptive history. Journal of Geodynamics 43, 118–152. Full text

22

4.1 Alfred Wegener's Arguments for Plate Tectonics

Alfred Wegener (1880-1930; Figure 4.2) earned a PhD in astronomy at the University of Berlin in 1904, but had a keen interest in geophysics and meteorology, and focused on meteorology for much of his academic career.

Alfred Wegener during a 1912-1913 expedition to Greenland. [Source: Alfred Wegener Institute (Public domain)]
Figure 4.2 Alfred Wegener during a 1912-1913 expedition to Greenland. Source: Alfred Wegener Institute (2008) Public Domain view source

In 1911 Wegener happened upon a scientific publication that described matching Permian-aged terrestrial fossils in various parts of South America, Africa, India, Antarctica, and Australia.  He concluded that because these organisms could not have crossed the oceans to get from one continent to the next, the continents must have been joined in the past, permitting the animals to move from one to the other (Figure 4.3).  Wegener envisioned a supercontinent made up of all the present day continents, and named it Pangea (meaning “all land”). He described the motion of the continents reconfiguring themselves as continental drift.

Figure 4.2 The distribution of several Permian terrestrial fossils that are present in various parts of continents that are now separated by oceans. During the Permian, the supercontinent Pangea included the supercontinent Gondwana, shown here, along with North America and Eurasia.
Figure 4.3 The distribution of several Permian terrestrial fossils that are present in various parts of continents now separated by oceans. During the Permian, the supercontinent Pangea included the supercontinent Gondwana, shown here, along with North America and Eurasia. Source: J.M. Watson, USGS (1999) Public Domain view source

Wegener pursued his idea with determination, combing libraries, consulting with colleagues, and making observations in an effort to find evidence in support of it. He relied heavily on matching geological patterns across oceans, such as sedimentary strata in South America matching those in Africa, North American coalfields matching those in Europe, the mountains of Atlantic Canada matching those of northern Britain—both in structure and rock type—and comparisons of rocks in the Canadian Arctic with those of Greenland (Figure 4.4).

Figure 4.4 Diagram from Alfred Wegener’s book Die Entstehung der Kontinente und Ozeane comparing rock types on Canadian Arctic Islands and Greenland. Source: Karla Panchuk (2018) CC BY 4.0. Click the image for more attributions.

Wegener also called upon evidence for the Carboniferous and Permian (~300 Ma) Karoo Glaciation from South America, Africa, India, Antarctica, and Australia (Figure 4.5). He argued that this could only have happened if these continents were once all connected as a single supercontinent. He also cited evidence (based on his own astronomical observations) that showed that the continents were moving with respect to each other, and determined a separation rate between Greenland and Scandinavia of 11 m per year, although he admitted that the measurements were not accurate. (The separation rate is actually about 2.5 cm per year.)

Figure 4.5 Carboniferous and Permian Karoo Glaciation in the southern hemisphere. Paleogeographic reconstruction for 306 million years ago. Source: Cropped from C. R. Scotese, PALEOMAP Project (www.scotese.com) view source. Click the image for terms of use.

Wegener first published his ideas in 1912 in a short book called Die Entstehung der Kontinente (The Origin of Continents), and then in 1915 in Die Entstehung der Kontinente und Ozeane (The Origin of Continents and Oceans). He revised this book several times up to 1929. It was translated into French, English, Spanish, and Russian in 1924.

The main criticism of Wegener’s idea was that he could not explain how continents could move. Remember that, as far as anyone was concerned, Earth’s crust was continuous, not broken into plates. Thus, any mechanism Wegener could think of would have to fit with that model of Earth’s structure. Geologists at the time were aware that continents were made of different rocks than the ocean crust, and that the material making up the continents was less dense, so Wegener proposed that the continents were like icebergs floating on the heavier ocean crust.  He suggested that the continents were moved by the effect of Earth’s rotation pushing objects toward the equator, and by the lunar and solar tidal forces, which tend to push objects toward the west. However, it was quickly shown that these forces were far too weak to move continents, and without any reasonable mechanism to make it work, Wegener’s theory was quickly dismissed by most geologists of the day.

Alfred Wegener died in Greenland in 1930 while carrying out studies related to glaciation and climate. At the time of his death, his ideas were tentatively accepted by a small minority of geologists, and firmly rejected by most. But within a few decades that was all to change.

Resources

On The Shoulders of Giants: Alfred Wegener

References

Wegener, A. (1920). Die Entstehung der Kontinente und Ozeane. Braunschweig, Germany: Friedr. Vieweg & Sohn. Full text at Project Gutenberg

23

4.2 Global Geological Models of the Early 20th Century

The untimely death of Alfred Wegener did not solve any problems for those who opposed his ideas, because they still had some inconvenient geological truths to deal with. One of those was explaining the distribution of terrestrial species across five continents that are currently separated by hundreds or thousands of kilometres of ocean water, and another was explaining the origin of extensive fold-belt mountains, such as the Appalachians, the Alps, the Himalayas, and the Canadian Rockies.

Before we continue, it is important to know what was generally believed about global geology before plate tectonics. At the beginning of the 20th century, geologists had a good understanding of how most rocks were formed and understood their relative ages through interpretation of fossils, but there was considerable controversy regarding the origin of mountain chains, especially fold-belt mountains. At the end of the 19th century, one of the prevailing views on the origin of mountains was the theory of contractionism — the idea that since Earth is slowly cooling, it must also be shrinking. In this scenario, mountain ranges had formed like the wrinkles on a dried-up apple. Oceans formed above parts of former continents that had settled downward and become submerged.

While this hypothesis helped to address the dilemma of the terrestrial fossils by explaining how continents once connected could now be separated by oceans, it came with its own set of problems.  One problem was that Earth wasn’t cooling fast enough to create the necessary amount of shrinking.  Another problem was the principle of isostasy (already understood for several decades; see Section 3.5 for a review of isostasy), which wouldn’t allow blocks of continental crust to sink in the way necessary for oceans to form in accordance with contractionist theory.

Another widely held view was permanentism, the idea that the continents and oceans have always been generally the same as they are today. This view incorporated a mechanism for creation of mountain chains known as the geosyncline theory. A geosyncline is a thick (potentially 1000s of metres) deposit of sediments and sedimentary rocks, typically situated along the edge of a continent, and derived from continental weathering (Figure 4.6).

image
Figure 4.6 The development of a geosyncline along a continental margin. (Note that a geosyncline is not related to a syncline, which is a downward fold in rocks.) Source: Steven Earle (2015) CC BY 4.0 view source

The idea that geosynclines developed into fold-belt mountains originated in the middle of the 19th century. It was first proposed by James Hall and later elaborated upon by Dwight Dana, both of whom worked extensively in the Appalachian Mountains of the eastern United States. The process of converting a geosyncline into a mountain belt was believed to involve compression by forces pushing from either side, causing sedimentary layers within the geosyncline to fold up. In 1937, Philip Kuenen published a paper of experiments with layers of paraffin wax to test how this might work. He was able to cause layers within a geosyncline to fold up as the geosyncline deepened and became more tightly folded during the experiment (Figure 4.7).

Figure 4.7 Simulation of mountain building within a geosyncline using layers of wax. Left- A sequence of photographs showing deformation in the wax layers as pistons apply increasing amounts of compression from the side. Right- Close-up view of slices through the wax layers at the end of the experiment, showing that stiffer white layers of wax folded in a way that resembled the folds in mountain belts. Source: Karla Panchuk (2018) CC BY 4.0. Photographs from Kuenen (1937) Public Domain view source.

The problem with the geosynclinal hypothesis for mountain building is that the lateral forces required to cause the compression were never adequately explained. Kuenen compressed the wax layers in his experiment by using pistons that pushed horizontally from either side, but described that mechanism as unrealistic. He explained that the pistons were just to get the process started within the experiment, and that in nature the main force was likely that of gravity pulling the geosyncline downward, drawing the sides together as it folded. When the sides were drawn together, this provided the compression to fold the sediments within the geosyncline. He couldn’t specify how the process got started in nature, but suggested that there could be a variety of reasons for an irregularity in the crust to respond to forces in such a way that would trigger downward sagging and folding.

Proponents of the geosyncline theory of mountain formation—and there were many well into the 1960s—also had the problem of explaining the intercontinental terrestrial fossil matchups. The explanation offered was that land bridges had once linked the continents, permitting animals and plants to migrate back and forth. One proponent of this idea was the American naturalist Ernest Ingersoll. Referring to evidence of past climate changes, Ingersoll contributed the following to the 1918-20 edition of the Encyclopedia Americana:

The most interesting feature of these changes, however, is that by which, now and again, the Old World was connected with the New by necks or spaces of land, known as “land-bridges”; especially as these permitted an interchange of plants and animals, giving to us many new ones from the other side of the ocean, including, finally, man himself.

A problem with the land-bridge hypothesis is that there is no evidence of land bridges that could account for the fossil distribution patterns. The world’s oceans are approximately 4 km deep on average, so the underwater slopes leading up to a land bridge would have to have been at least 10s of km wide in most places, and many times that in others. Even if flooded, a land bridge of that size would still be visible in the shape of ocean-floor terrain.  Isostasy would not permit such a land bridge to sink down without leaving a trace.

We do know the locations of some past land bridges, but they were very different from the ones that would be required for this hypothesis. They are bridges such as the flooded Bering Strait land bridge, which is beneath only 30 to 50 m of water, and was exposed when sea levels were much lower because of water being locked in polar ice caps during the last major glaciation event.  The narrowest point of the Bering Strait is 82 km wide. The shortest distance between South America and Africa is more than 2800 km.

 

Exercise: Fitting the Continents Together

The main continents around the Atlantic Ocean are shown in Figure 4.8 with the shapes that they might have had during the Mesozoic, including the extents of their continental shelves. Cut these shapes out and see how well you can fit them together in the positions that these areas occupied within Pangea. Refer to a map of Pangea to help you make the fit. Note that the fit of the continents is even better than this, as distortions are introduced when rendering Earth onto a flat map. A better fit could be accomplished if you were to do this exercise upon the surface of a globe.

Figure 4.8 Mesozoic continent shapes. Source: Steven Earle (2015) CC BY 4.0 view source

References

Ingersoll, E. (1919). Land-Bridges Across the Oceans. In The Encyclopedia Americana (Vol. XVI, pp. 692-694). New York, NY: Encyclopedia Americana Corporation. Full text

Kuenen, P. H. (1937) The negative isostatic anomalies in the East Indies (with Experiments). Leidse Geologische Mededelingen 8(2), 169-214. Full text

 

24

4.3 Geological Renaissance of the Mid-20th Century

Two key areas of research ultimately led to the acceptance of continental drift, and the formulation of plate tectonic theory.  One was the study of paleomagnetism, the record of Earth’s magnetic field through time.  The other was exploration of the ocean floor.

Paleomagnetism (Remnant Magnetism)

Figure 4.6 Rock layers recording remnant magnetism. The red arrows represent the direction of the vertical component of Earth's magnetic field. The oldest rock has a magnetic dip characteristic of the southern hemisphere, but over time the dip changes, indicating that the rocks moved toward magnetic north. [SE]
Figure 4.9 Rock layers recording remnant magnetism. The red arrows represent the direction of the vertical component of Earth’s magnetic field. The oldest rock has a magnetic dip characteristic of the southern hemisphere, but over time the dip changes, indicating that the rocks moved toward magnetic north. Source: Steven Earle (2015) CC BY 4.0 view source

When rocks form, some of the minerals that make them up can become aligned with the Earth’s magnetic field, just like a compass needle pointing to north.  This happens to the mineral magnetite (Fe3O4) when it crystallizes from magma.  Once the rock cools the crystals are locked in place.  This means that if the rock moves, the crystals can’t realign themselves, and they retain a remnant magnetism. This would be like jamming your compass needle so that if you turned away from north, the needle would turn with you rather than continuing to point north.

Rocks like basalt, which cool from a high temperature and commonly have relatively high levels of magnetite, are particularly susceptible to being magnetized in this way.  However, even sediments and sedimentary rocks can take on remnant magnetism as long as they have small amounts of magnetic minerals, because the magnetic grains can gradually become lined up with Earth’s magnetic field as the sediments are deposited.

By studying both the horizontal and vertical components of the remnant magnetism, one can tell not only the direction to magnetic north at the time of the rock’s formation, but also the latitude where the rock formed relative to magnetic north.  Remember that the vertical component of the magnetic field points more sharply downward the closer it is to the magnetic north pole.  Figure 4.9 shows the vertical component of remnant magnetism in a sequence of rocks.  Notice that the arrow starts out at 500 Ma pointing slightly upward.  This means that the rocks were in the southern hemisphere.  As the rocks get younger, the arrow tilts toward horizontal, and then points downward.  This indicates that the rocks were getting progressively closer to the north magnetic pole.

Apparent Polar Wandering Paths

In the early 1950s, a group of geologists from Cambridge University, including Keith Runcorn, Ted Irving,Ted Irving later set up a paleomagnetic lab at the Geological Survey of Canada in Sidney BC, and did important work on understanding the geology of western North America. and several others, started looking at the remnant magnetism of Phanerozoic British and European volcanic rocks, and collecting paleomagnetic data. Using an analysis similar to that in Figure 4.9, they noticed that rocks of different ages sampled from the same general area showed very different magnetic pole positions (the green line in Figure 4.10). They assumed this meant that Earth’s magnetic pole had moved around significantly over time along polar wandering paths, rather than staying close to the geographic north pole as it does today.  At the time, geophysical models suggested that the magnetic poles did not need to be aligned with the rotational poles, so this wasn’t an unreasonable conclusion, given what was known.

image
Figure 4.10 Apparent polar-wandering paths (APWP) for Eurasia and North America. The view is from the geographic north pole (black dot) looking down. Dots along each path show the location of magnetic north as determined from paleomagnetic data. Left- Data from Eurasia and North America agree on the location of magnetic north today (time 0), but not at any time in the past. Right- Once continent motion has been accounted for, there is agreement in data from Eurasia and North America on the location of magnetic north over the past 500 million years. Source: Steven Earle (2015) CC BY 4.0 view source

Runcorn and colleagues extended their paleomagnetic studies to North America, and began to realize that their initial conclusion had a problem.  Notice that on the right of Figure 4.10 the polar wandering path for North America (in red) does not match the path for Eurasia (in green).  For example, data from North America suggest that 200 Ma ago, magnetic north was somewhere in China, whereas data from Europe said it was in the Pacific Ocean.  There could only have been one magnetic north pole position at 200 Ma, therefore the only way to explain the discrepancy was if Europe and North America moved along different paths during this time while the pole stayed in more or less the same location.

The polar wandering paths were not actually records of the pole moving, they just looked that way, so the paths are now referred to as apparent polar wandering paths (APWP).  Subsequent paleomagnetic work showed that unique apparent polar wandering paths can be derived from rocks in South America, Africa, India, and Australia. In 1956, Runcorn became a proponent of continental drift.  There was simply no other way to explain the data.

This paleomagnetic work of the 1950s was the first new evidence in favour of continental drift, and it led a number of geologists to start thinking that the idea might have some merit. Nevertheless, for a majority of geologists, this type of evidence was not sufficiently convincing to get them to change their views.

Ocean Basin Geology and Geography

During the 20th century, our knowledge and understanding of the ocean basins and their geology increased dramatically. Before 1900 we knew virtually nothing about the bathymetry (the hills and valleys of the ocean floor) and geology of the oceans. By the end of the 1960s, we had detailed maps of the topography of the ocean floors, a clear picture of the geology of ocean floor sediments and the solid rocks beneath them, and almost as much information about the geophysical nature of ocean rocks as of continental rocks.

Acoustic Depth Sounding

Up until the 1920s, ocean depths were measured using weighted lines dropped overboard. In deep water this is a painfully slow process and the number of soundings in the deep oceans was probably fewer than 1,000. That is roughly one depth sounding for every 350,000 square kilometres of the ocean. To put that in perspective, it would be like trying to describe the topography of British Columbia with elevation data for only a half a dozen points!

The voyage of the Challenger in 1872 and the laying of trans-Atlantic cables had shown that there were mountains beneath the seas, but most geologists and oceanographers still believed that the oceans were essentially vast basins with flat bottoms, filled with thousands of metres of sediments.

Following development of acoustic depth sounders (Figure 4.11) in the 1920s, the number of depth readings increased by many orders of magnitude, and by the 1930s there was no doubt that major mountain chains ran through all of the world’s oceans. During and after World War II, there was a well-organized campaign to study the oceans, and by 1959, sufficient bathymetric data had been collected to produce detailed maps of all the oceans (Figure 4.12).

image
Figure 4.11 A ship-borne acoustic depth sounder. The instrument emits sound (black arcs) that reflects off the sea floor and returns to the surface (white arcs). The time interval between emitting the sound and detecting it on receivers on the ship is proportional to the water depth. Source: Steven Earle (2015) CC BY 4.0 view source
Figure 4.9 Ocean floor bathymetry (and continental topography). Inset (a): the mid-Atlantic ridge, (b): the Newfoundland continental shelf, (c): the Nazca trench adjacent to South America, and (d): the Hawaiian Island chain. [SE after NOAA, http://bit.ly/1OtRMc0]
Figure 4.12 Ocean floor bathymetry (and continental topography). Inset (a): the mid-Atlantic ridge, (b): the Newfoundland continental shelf, (c): the Nazca trench adjacent to South America, and (d): the Hawaiian Island chain. Source: Steven Earle (2015) CC BY 4.0 view source; Basemap after NOAA (2006) Public Domain view source

The important physical features of the ocean floor are:

Seismic Reflection Sounding

Seismic reflection sounding involves transmitting high-energy sound bursts and then measuring the echoes with a series of receivers called geophones towed behind a ship. The technique is related to acoustic sounding as described above, however, much more energy is transmitted and the sophistication of the data processing is much greater. As the technique evolved, and the amount of energy was increased, it became possible to see through the sea-floor sediments and map the bedrock topography and crustal thickness. This allowed sediment thicknesses to be mapped (Figure 4.13).

Figure 4.10 Map of global sediment thickness. [Source: NOAA, http://1.usa.gov/1Ywxxz6]
Figure 4.13 Map of global sediment thickness. Source: NOAA (2003) Public Domain view source

It was soon discovered that although the sediments were up to several 1000s of m thick near the continents, they were relatively thin — or even non-existent — along ocean ridges (Figure 4.14). The seismic studies also showed that the crust is relatively thin under the oceans (5 km to 6 km) compared to the continents (30 km to 60 km) and geologically very consistent, composed almost entirely of basalt.

image
Figure 4.14 Topographic section at an ocean ridge based on reflection seismic data. Sediments are not thick enough to be detectable near the ridge, but get thicker on either side. The diagram represents approximately 50 km width, and has a 10x vertical exaggeration. Source: Steven Earle (2015) CC BY 4.0 view source

Heat Flow Rates

In the early 1950s, Edward Bullard—who spent time at the University of Toronto but is mostly associated with Cambridge University—developed a probe for measuring the flow of heat from the ocean floor. Bullard and colleagues found higher than average heat-flow rates along the ridges, and lower than average rates in trenches. These data were interpreted as evidence of mantle convection, with areas of high heat flow corresponding to upward convection of hot mantle material, and areas of low heat flow corresponding to downward convection.

Earthquake Belts

With developments of networks of seismographic stations in the 1950s, it became possible to plot the locations and depths of both major and minor earthquakes with great accuracy. A remarkable correspondence was observed between earthquake locations and both the mid-ocean ridges and the deep ocean trenches. In 1954 Gutenberg and Richter showed that the ocean-ridge earthquakes were all relatively shallow, and confirmed what had first been shown by Benioff in the 1930s, that earthquakes in the vicinity of ocean trenches were both shallow and deep, but that the deeper ones were situated progressively farther inland from the trenches (Figure 4.15).

image
Figure 4.15 Aleutian Island subduction zone earthquakes. Left- Map view with earthquakes marked as dots. Red dots are the shallowest earthquakes and blue are the deepest. Quakes get deeper further inland from the trench. Right- Cross-section through a-b. Coloured dots show the depth of earthquakes. Colours correspond to dots in the left figure. Earthquake depth is related to the position of the Pacific plate as it travels beneath the North American plate. Source: Steven Earle (2015) CC BY 4.0 view source

Magnetic Stripes on the Sea Floor

In the 1950s, scientists from the Scripps Oceanographic Institute in California persuaded the United States Coast Guard to include magnetometer readings on one of their expeditions to study ocean floor topography. The first comprehensive magnetic data set was compiled in 1958 for an area off the coasts of BC and Washington State. This survey revealed a bewildering pattern of low and high magnetic field intensity in sea-floor rocks (Figure 4.16). When the data were first plotted on a map in 1961, nobody understood them — not even the scientists who collected them. Many 1000s of km of magnetic surveys were conducted over the next several years.

Figure 4.16 Pattern of sea-floor magnetic field intensity off the west coast of British Columbia and Washington. Black regions have higher than average magnetic field instensity, and white regions have lower than average intensity. Source: Steven Earle (2015) CC BY-SA 4.0, modified after U. S. Geological Survey (n.d.) Public Domain view source (adapted from Raff and Mason, 1961).

The wealth of new data from the oceans began to significantly influence geological thinking in the 1960s. In 1960, Harold Hess from Princeton University, advanced a hypothesis with many of the elements that we now accept as plate tectonics. He maintained some uncertainty about his proposal however, and in order to deflect criticism from mainstream geologists, he labelled it “geopoetry.” In fact, until 1962, Hess didn’t even put his ideas in writing — except internally to the U.S. Navy (which funded his research) — but presented them mostly in lectures and seminars.

Hess proposed that new sea floor was generated from mantle material at the ocean ridges, and that old sea floor was dragged down at the ocean trenches and re-incorporated into the mantle. He suggested that the process was driven by mantle convection currents, rising at the ridges and descending at the trenches (Figure 4.17). He also suggested that the less-dense continental crust did not descend with oceanic crust into trenches, but that colliding landmasses were thrust up to form mountains.

Hess’s hypotheses formed the basis for our ideas on sea-floor spreading and continental drift, but did not go so far as to claim that the crust is made up of separate plates. The Hess model was not roundly criticized, but also not widely accepted, partly because evidence was still lacking.

image
Figure 4.17 A representation of Harold Hess’s model for sea-floor spreading and subduction. Source: Steven Earle (2015) CC BY 4.0 view source

Collection of magnetic data from the oceans continued in the early 1960s, but the striped patterns remained unexplained. Some assumed that, as with continental crust, the stripes were related to compositional variations in rock, such as variations in the amount of magnetite. The first real understanding of the significance of the striped anomalies was the interpretation by Fred Vine, a Cambridge graduate student. Vine was examining magnetic data from the Indian Ocean and, like others before him, noted that the magnetic patterns were symmetrical on either side of the ridge.

At the same time, other researchers led by groups in California and New Zealand were studying the phenomenon of reversals in Earth’s magnetic field. They were trying to determine when such reversals had taken place over the past several million years by analyzing the magnetic characteristics of hundreds of samples from basaltic flows. As discussed in Chapter 3, Earth’s magnetic field periodically weakens, then becomes virtually non-existent before becoming re-established with the reverse polarity. During periods of reversed polarity, a compass would point south instead of north.

The time scale of magnetic reversals is irregular. The present “normal” event, known as the Bruhnes magnetic chron, has persisted for about 780,000 years. It was preceded by a 190,000-year reversed event; a 50,000-year normal event known as Jaramillo; and then a 700,000-year reversed event (see Figure 3.16).

In a paper published in September 1963, Vine and his PhD supervisor Drummond Matthews proposed that the patterns associated with ridges were related to the magnetic reversals, and that oceanic crust created from cooling basalt during a normal event would have polarity aligned with the present magnetic field, and would produce a positive anomaly (a black stripe on the sea-floor magnetic map). Oceanic crust created during a reversed event would have polarity opposite Earth’s present field and thus produce a negative magnetic anomaly (a white stripe). The same idea had been put forward a few months earlier by Lawrence Morley, of the Geological Survey of Canada. However, Morley’s papers submitted earlier in 1963 to Nature and The Journal of Geophysical Research were rejected. The idea is sometimes referred to as the Vine-Matthews-Morley (VMM) hypothesis.

Vine, Matthews, and Morley were the first to show this type of correspondence between the relative widths of the stripes and the durations of the magnetic reversals. The VMM hypothesis was confirmed within a few years when magnetic data were compiled from spreading ridges around the world. It was shown that the same general magnetic patterns were present straddling each ridge, although the widths of the anomalies varied according to the spreading rates characteristic of the different ridges. It was also shown that the patterns corresponded with the known timeline of Earth’s magnetic field reversals.

In 1963, J. Tuzo Wilson of the University of Toronto proposed the idea of a mantle plume or hot spot — a place where hot mantle material rises in a stationary and semi-permanent plume, and affects the overlying crust. He based this hypothesis partly on the distribution of the Hawai’ian and Emperor Seamount island chains in the Pacific Ocean (Figure 4.18). The volcanic rock making up these islands becomes progressively younger toward the southeast, culminating in the island of Hawai’i itself, which consists of rock that is almost all younger than 1 Ma.

Ages of the Hawaiian Islands and the Emperor Seamounts in relation to the location of the Hawaiian mantle plume. Hawaii: 0 Ma; Necker: 10.3 Ma; Midway: 27.7 Ma; Koko: 48.1 Ma; Suiko: 64.7 Ma
Figure 4.18 The ages of the Hawai’ian Islands and the Emperor Seamounts in relation to the location of the Hawai’ian mantle plume. Source: Steven Earle (2015) CC BY 4.0 view source; Base map from the National Geophysical Data Centre/USGS (2005) Public Domain view source

Wilson suggested that a stationary plume of hot upwelling mantle material is the source of the Hawaiian volcanism, and that the ocean crust of the Pacific Plate is moving toward the northwest over this hot spot. Near the Midway Islands, the chain makes a pronounced change in direction, from northwest-southeast for the Hawai’ian Islands, to nearly north-south for the Emperor Seamounts. This change has been ascribed to a change in direction of the Pacific Plate moving over the stationary mantle plume. An alternative hypothesis is that rather than the Pacific Plate having undergone a sudden change in motion, the plume itself has moved at least 2,000 km south over the period between 81 and 45 Ma (Tarduno et al., 2003).

There is evidence of many such mantle plumes around the world (Figure 4.19). Most are within ocean basins, including places like Hawai’i, Iceland, and the Galapagos Islands, but some are under continents. One example is the Yellowstone hot spot in the west-central United States, and another is the one responsible for the Anahim Volcanic Belt in central British Columbia. It is evident that mantle plumes are very long-lived phenomena, lasting for at least tens of millions of years, and possibly for hundreds of millions of years in some cases.

image
Figure 4.19 Mantle plume locations.  Selected Mantle plumes: 1: Azores, 3: Bowie, 5: Cobb, 8: Eifel, 10: Galapagos, 12: Hawai’i, 14: Iceland, 17: Cameroon, 18: Canary, 19: Cape Verde, 35: Samoa, 38: Tahiti, 42: Tristan, 44: Yellowstone, 45: Anahim. Source: Ingo Wölbern (2007) Public Domain view source

Oceanic spreading ridges appear to be curved features on Earth’s surface, but the ridges are in fact composed of a series of straight-line segments, offset at intervals by faults perpendicular to the ridge (Figure 4.20). In a paper published in 1965, Tuzo Wilson termed these features transform faults.

Figure 4.20 Part of the Mid-Atlantic ridge near the equator. Transform faults (red lines) are in between the ridge segments (double white lines), where the yellow arrows (indicating relative plate movement) point in opposite directions. Solid white lines are fracture zones. Source: Steven Earle (2015) CC BY 4.0 view source

In the same 1965 paper, Wilson introduced the idea that the crust can be divided into a series of rigid plates, and is thus responsible for the term plate tectonics.

Exercise: Paper Transform Fault Model

J. Tuzo Wilson used a paper model similar to the one in Figure 4.21 to explain transform faults to his colleagues. To use this model, print Figure 4.21, cut around the outside, and then slice along the line A-B (the fracture zone) with a sharp knife. Fold down the top half where shown, and then pinch together in the middle. Do the same with the bottom half. When you’re done, you should have two folds of paper extending downward as in Figure 4.22.

Figure 4.21 Transform fault model. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Stewart (1990).

 

Figure 4.22 Use of the transform fault model. Source: Steven Earle (2015) CC BY 4.0 view source

Find someone else to pinch those folds with two fingers just below each ridge crest, and then gently pull apart where shown. As you do, the oceanic crust will emerge from the middle, and you will see that the parts of the fracture zone between the ridge crests will be moving in opposite directions (this is the transform fault) while the parts of the fracture zone outside of the ridge crests will be moving in the same direction. You will also see that the oceanic crust is being magnetized as it forms at the ridge. The magnetic patterns represent the last 2.5 Ma of geological time.

There are other versions of this model available at https://web.viu.ca/earle/transform-model/. For more information see Earle (2004).

References

Earle, S. (2004). A simple paper model of a transform fault at a spreading ridge. Journal of Geoscience Education 52, 391-392.

Raff, A., & Mason, R. (1961) Magnetic survey off the west coast of North America, 40˚ N to 52˚ N latitude. Geological Society of America Bulletin 72, 267-270.

Stewart, J. A. (1990). Drifting continents and colliding paradigms. Bloomington IN: Indiana University Press.

Tarduno, J. A., Duncan, R. A., Scholl, D. W., Cottrell, R. D., Steinberger, B., Thordarson, T., Kerr, B. C., Neal, C. R., Frey, F. A., Torii, M., and Carvallo, C. (2003). The Emperor Seamounts: Southward Motion of the Hawaiian Hotspot Plume in Earth’s Mantle. Science 301(5636), 1064–1069. DOI: 10.1126/science.1086442

Wilson, J. T. (1965). A new class of faults and their bearing on continental drift. Nature 207, 343-347.

 

25

4.4 Plates, Plate Motions, and Plate-Boundary Processes

The ideas of continental drift and sea-floor spreading became widely accepted by 1965, and more geologists started thinking in these terms. By the end of 1967, Earth’s surface had been mapped into a series of plates (Figure 4.23). The major plates are Eurasian, Pacific, Indian, Australian, North American, South American, African, and Antarctic plates. There are also numerous small plates (e.g., Juan de Fuca, Nazca, Scotia, Philippine, Caribbean), and many very small plates or sub-plates. The Juan de Fuca Plate is actually three separate plates (Gorda, Juan de Fuca, and Explorer), all moving in the same general direction but at slightly different rates.

Figure 4.18 A detailed map of Earth's tectonic plates. [Source: NASA, http://bit.ly/1PZHRMZ]
Figure 4.23 A detailed map of Earth’s tectonic plates. Click on the map to enlarge. Source: Paul Lowman and Jacob Yates, NASA Goddard Space Flight Center (2002) Public Domain view source

Plate motions can be tracked using Global Positioning System (GPS) data from different locations on Earth’s surface. Rates of motions of the major plates range from less than 1 cm/y to more than 10 cm/y. The Pacific Plate is the fastest, moving at more than 10 cm/y in some areas, followed by the Australian and Nazca Plates. The North American Plate is one of the slowest, averaging ~1 cm/y in the south up to almost 4 cm/y in the north.

Plates move as rigid bodies, so it may seem surprising that the North American Plate can be moving at different rates in different places. The explanation is that plates rotate as they move; the North American Plate, for example, rotates counter-clockwise, while the Eurasian Plate rotates clockwise.

Boundaries between the plates are of three types: divergent (moving apart), convergent (moving together), and transform (moving side by side). The plates are made up of crust and lithospheric mantle (Figure 4.24). Even though the plates are in constant motion, and move in different directions, there is never a significant amount of space between them. Plates move along the lithosphere-asthenosphere boundary, because the asthenosphere is relatively weak. It deforms as the plates move, rather than locking them in place.

image
Figure 4.24 The crust and upper mantle. Tectonic plates consist of lithosphere, which includes the crust and the lithospheric (rigid) part of the mantle. Source: Steven Earle (2015) CC BY 4.0 view source

At spreading centres, the lithospheric mantle is relatively thin. The upward convective motion of hot mantle material generates temperatures that are too high for the existence of a significant thickness of rigid lithosphere at the same time that the plates are falling away from each other (Figure 4.17).

The fact that plates include both crustal material and lithospheric mantle material makes it possible for a single plate to be include both oceanic and continental crust. The North American Plate includes most of North America, plus half of the northern Atlantic Ocean. Similarly the South American Plate extends across the western part of the southern Atlantic Ocean, while the European and African plates each include part of the eastern Atlantic Ocean. The Pacific Plate is almost entirely oceanic, but it does include the part of California west of the San Andreas Fault.

Divergent Boundaries

Divergent boundaries are spreading boundaries, where new oceanic crust is created from magma derived from partial melting of the mantle. The partial melting happens when hot mantle rock is moved from deep within Earth where pressures are too high for it to be liquid, to shallower depths where the pressure is much lower (Figure 4.25, bottom left).

The triangular zone of partial melting near the ridge crest is approximately 60 km thick and the proportion of magma is about 10% of the rock volume, thus producing crust that is about 6 km thick once the melt escapes from the rock in which it formed, and ascends. Most divergent boundaries are located in the oceans, and the crustal material created at a spreading boundary is always oceanic in character; in other words, it is mafic igneous rock (basalt or gabbro, with minerals rich in iron and magnesium). Spreading rates vary considerably, from 1 cm/y to 3 cm/y in the Atlantic, to between 6 cm/y and 10 cm/y in the Pacific. Some of the processes taking place in this setting include (Figure 4.25, top):

image
Figure 4.25 Divergent boundary. Lower left- General processes taking place along divergent boundaries. Top- Expanded view of the white box showing divergent boundary processes and materials. Bottom right- Pillow basalts from the ocean floor of Hawai’i. Source: Lower left- Steven Earle (2015) CC BY 4.0 view source; Top- Steven Earle (2015) CC BY 4.0 view source modified after Sinton and Detrick (1992); Lower right- NOAA (1988) Public Domain view source

Spreading is thought to start with lithosphere being warped upward into a dome by buoyant material from an underlying mantle plume or series of mantle plumes. The buoyancy of the mantle plume causes the dome to fracture in a radial pattern, with three arms spaced at approximately 120° (Figure 4.26).

rift formation
Figure 4.26 Depiction of the process of dome and three-part rift formation (left) and of continental rifting between the African and South American parts of Pangea at around 200 Ma (right) Source: Steven Earle (2015) CC BY 4.0 view source

When a series of mantle plumes exists beneath a large continent, the resulting rifts may align and lead to the formation of a rift valley, such as the present-day Great Rift Valley in eastern Africa. This type of valley may eventually develop into a linear sea (such as the present-day Red Sea), and finally into an ocean (such as the Atlantic). It is likely that as many as 20 mantle plumes, many of which still exist, were responsible for the initiation of the rifting of Pangea along what is now the mid-Atlantic ridge (see the Atlantic Ocean mantle plume locations in Figure 4.19).

Convergent Boundaries

Convergent boundaries, where two plates are moving toward each other, are of three types, depending on whether ocean or continental crust is present on either side of the boundary. The types are ocean-ocean, ocean-continent, and continent-continent.

Ocean-Ocean Convergent Boundaries

At an ocean-ocean convergent boundary, a plate margin consisting of oceanic crust and lithospheric mantle is subducted, or travels beneath, the margin of the plate with which it is colliding (Figure 4.27). Often it is the older and colder plate that is denser and subducts beneath the younger and hotter plate. Ocean trenches commonly form along these boundaries.

image
Figure 4.27 Configuration and processes of an ocean-ocean convergent boundary Source: Steven Earle (2015) CC BY 4.0 view source

As the subducting crust is heated and the pressure increases, water is released from within the subducting material. This water comes primarily from alteration of the minerals pyroxene and olivine to serpentine near the spreading ridge shortly after the rock’s formation. The water mixes with the overlying mantle, which lowers the melting point of mantle rocks, causing magma to form. This process is called flux melting or fluid-induced melting.

The newly produced magma, which is lighter than the surrounding mantle rocks, rises through the mantle and sometimes through the overlying oceanic crust to the ocean floor where it creates a chain of volcanic islands known as an island arc. A mature island arc develops into a chain of relatively large islands (such as Japan or Indonesia) as more and more volcanic material is extruded and sedimentary rocks accumulate around the islands. The largest earthquakes occur near the surface where the subducting plate is still cold and strong.

Examples of ocean-ocean convergent zones are subduction of the Pacific Plate south of Alaska (Aleutian Islands) and west of the Philippines, subduction of the Indian Plate south of Indonesia, and subduction of the Atlantic Plate beneath the Caribbean Plate.

Ocean-Continent Convergent Boundaries

At an ocean-continent convergent boundary, the oceanic plate is subducted beneath the continental plate in the same manner as at an ocean-ocean boundary. Rocks and sediment on the continental slope are thrust up into an accretionary wedge, and compression leads to faults forming within the continental plate (Figure 4.28). The mafic magma produced adjacent to the subduction zone rises to the base of the continental crust and leads to partial melting of the crustal rock. The resulting magma ascends through the crust, producing a mountain chain with many volcanoes.

Figure 4.28 Configuration and processes of an ocean-continent convergent boundary Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

Examples of ocean-continent convergent boundaries are subduction of the Nazca Plate under South America (which has created the Andes Range) and subduction of the Juan de Fuca Plate under North America (creating the mountains Garibaldi, Baker, St. Helens, Rainier, Hood, and Shasta, collectively known as the Cascade Range).

Continent-Continent Convergent Boundary

A continent-continent collision occurs when a continent or large island that has been moved along with subducting oceanic crust collides with another continent (Figure 4.29). The colliding continental material will not be subducted because it is not dense enough, but the root of the oceanic plate will eventually break off and sink into the mantle. There is tremendous deformation of the pre-existing continental rocks, and creation of mountains from that rock, as well as from any sediments that had accumulated along the shores of both continental masses, and commonly also from some ocean crust and upper mantle material.

image
Figure 4.29 Configuration and processes of a continent-continent convergent boundary Source: Steven Earle (2015) CC BY 4.0 view source

Examples of continent-continent convergent boundaries are the collision of the India Plate with the Eurasian Plate, creating the Himalaya Mountains, and the collision of the African Plate with the Eurasian Plate, creating the series of ranges extending from the Alps in Europe to the Zagros Mountains in Iran.

When a subduction zone is jammed shut by a continent-continent collision, plate tectonic stresses that are still present can sometimes cause a new subduction zone to develop outboard of the colliding plate.

Transform Boundaries

Transform boundaries exist where one plate slides past another without producing or destroying crust, except in the special case where the transform boundary has bends and jogs. There will be collisions and divergence on a small scale as the jogs crash into the bends, or open up small windows to deeper crust.

Most transform faults connect segments of mid-ocean ridges and are thus ocean-ocean plate boundaries (Figure 4.20). Some transform faults connect continental parts of plates. An example is the San Andreas Fault, which connects the southern end of the Juan de Fuca Ridge with the northern end of the East Pacific Rise (a ridge) in the Gulf of California (Figures 4.30 and 4.31). The part of California west of the San Andreas Fault and all of Baja California are on the Pacific Plate. But transform faults do not just connect divergent boundaries; the Queen Charlotte Fault connects the north end of the Juan de Fuca Ridge, starting at the north end of Vancouver Island, to the Aleutian subduction zone.

image
Figure 4.30 The San Andreas Fault extends from the north end of the East Pacific Rise in the Gulf of California to the southern end of the Juan de Fuca Ridge. All of the red lines on this map are transform faults. Source: Steven Earle (2015) CC BY 4.0 view source
image
Figure 4.31 The San Andreas Fault at Parkfield in central California. The person with the orange shirt is standing on the Pacific Plate and the person at the far side of the bridge is on the North American Plate. The bridge is designed to slide on its foundation. Source: Steven Earle (2015) CC BY 4.0 view source

 

Exercise: A Different Type of Transform Fault

Figure 4.32 shows the Juan de Fuca (JDF) and Explorer plates off the coast of Vancouver Island. The JDF Plate is moving toward the North American Plate at ~ 4.5 cm/y. We think that the Explorer Plate is also moving east, but we don’t know the rate, and there is evidence that it is slower than the JDF Plate.

The boundary between the two plates is the Nootka Fault, which is the location of frequent small-to-medium earthquakes (up to magnitude ~5), as depicted by the red stars. Explain why the Nootka Fault is a transform fault, and show the relative sense of motion along the fault with two small arrows.

Figure 4.32 Juan de Fuca and Explorer plates separated by the Nootka Fault (marked with red stars). Source: Steven Earle (2015) CC BY 4.0 view source

Plate Tectonics and Supercontinent Cycles

The present continents were once all part of a supercontinent that Alfred Wegener named Pangea (all land). More recent studies of continental matchups and the magnetic ages of ocean-floor rocks have enabled us to reconstruct the history of the break-up of Pangea.

Pangea began to rift apart along a line between Africa and Asia and between North America and South America at around 200 Ma (Figure 4.33). During the same period the Atlantic Ocean began to open up between northern Africa and North America, and India broke away from Antarctica. Between 200 and 150 Ma, rifting started between South America and Africa and between North America and Europe, and India moved north toward Asia. By 80 Ma, Africa had separated from South America, and most of Europe had separated from North America. By 50 Ma, Australia had separated from Antarctica, and shortly after that, India collided with Asia.

 

Figure 4.33 Sequence of paleogeographic reconstructions showing the breakup of Pangea. Source: Karla Panchuk (2017) CC BY-NC-SA 4.0. Maps from C. R. Scotese, PALEOMAP  Project (www.scotese.com). Click the image for map sources and terms of use.

Within the past few million years, rifting has occurred in the Gulf of Aden and the Red Sea, and also within the Gulf of California. Incipient rifting has begun along the Great Rift Valley of eastern Africa, extending from Ethiopia and Djibouti on the Gulf of Aden (Red Sea) all the way south to Malawi.

Pangea was not the first supercontinent. It was preceded by Pannotia (600 to 540 Ma), Rodinia (1,100 to 750 Ma), and by others before that. In fact, in 1966, Tuzo Wilson proposed that supercontinents are part of an on-going cycle, which we now refer to as a Wilson cycle. In a Wilson cycle, continents break up, and fragments drift apart only to collide again and make a new continent.

At present we are in the stages of a Wilson cycle where fragments are drifting and changing their configuration. North and South America, Europe, and Africa are moving with their respective portions of the Atlantic Ocean. The eastern margins of North and South America and the western margins of Europe and Africa are called passive margins because there is no subduction taking place along them. Because the oceanic crust formed by spreading along the mid-Atlantic ridge is not currently being subducted (except in the Caribbean), the Atlantic Ocean is slowly getting bigger, and the Pacific Ocean is getting smaller.

This situation may not continue for too much longer, however. As the Atlantic Ocean floor gets weighed down around its margins by great thickness of continental sediments, it will be pushed farther and farther into the mantle, and eventually the oceanic lithosphere may break away from the continental lithosphere and begin to subduct (Figure 4.34).

image067
Figure 4.34 Development of a subduction zone at a passive margin. Times A, B, and C are separated by tens of millions of years. Once the oceanic crust breaks off and starts to subduct, the continental crust (North America in this case) may no longer be pushed to the west and could start to move east because the rate of spreading in the Pacific basin is faster than along the Mid-Atlantic Ridge. Source: Steven Earle (2015) CC BY 4.0 view source

A subduction zone will develop, and the oceanic plate will begin to descend under the continent. Once this happens, the continents will no longer continue to move apart because the spreading at the mid-Atlantic ridge will be taken up by subduction. If spreading along the mid-Atlantic ridge continues to be slower than spreading within the Pacific Ocean, the Atlantic Ocean will start to close up, and eventually (in a 100 million years or more) North and South America will collide again with Europe and Africa. If this continues without changing for another few hundred million years, we will be back to where we started, with one supercontinent (Figure 4.35).

Figure 4.35 Sequence of reconstructions showing the possible future configuration of land masses on Earth at 50, 150, and 250 million years from now. Movements culminate in the formation of a new supercontinent called Pangea Ultima. Source: Karla Panchuk (2017) CC BY-NC-SA 4.0. Maps from C. R. Scotese, PALEOMAP  Project (www.scotese.com). Click the image for map sources and terms of use.

There is strong evidence around the margins of the Atlantic Ocean that this process has taken place before. There are roots of ancient mountain belts along the eastern margin of North America, the western margin of Europe, and the north-western margin of Africa, which show that these landmasses once collided with each other to form a mountain chain. The mountain chain might have been as big as the Himalayas.

The apparent line of collision runs between Norway and Sweden, between Scotland and England, through Ireland, through Newfoundland and the Maritimes, through the north-eastern and eastern states, and across the northern end of Florida. When rifting of Pangea started at approximately 200 Ma, the fissuring was along a different line from the line of the earlier collision. This is why some of the mountain chains formed during the earlier collision can be traced from Europe to North America and from Europe to Africa.

It is probably no coincidence that the Atlantic Ocean rift may have occurred in approximately the same place during two separate events several hundred million years apart. The series of hot spots that has been identified in the Atlantic Ocean may also have existed for several hundred million years, and thus may have contributed to rifting in roughly the same place on at least two separate occasions (Figure 4.36).

Wilson cycle
Figure 4.36 A scenario for the Wilson cycle. The cycle starts with continental rifting above a series of mantle plumes (red dots, A). The continents separate (B), and then re-converge some time later, forming a fold-belt mountain chain. Eventually rifting is repeated, possibly because of the same set of mantle plumes (D), but this time the rift is in a different place. Source: Steven Earle (2015) CC BY 4.0 view source

References

Sinton, J. M., and Detrick, R. S. (1992). Mid-Ocean Ridge Magma Chambers. Journal of Geophysical Research 97(B1), 197-216.

26

4.5 Mechanisms for Plate Motion

Mantle convection is often said to be critical to plate tectonics. While this is almost certainly so, there is still debate about the actual forces that make the plates move. One side of the argument holds that the plates are only moved by the traction caused by mantle convection, and that friction between the asthenosphere and lithosphere pulls the lithosphere along as the mantle convects. The other side holds that traction plays only a minor role and that ridge-push and slab-pull are more important (Figure 4.37).

Ridge-push refers to gravity causing lithosphere to slide downhill away from the elevated mid-ocean ridges. Slab-pull refers to the weight of subducting slabs dragging the rest of the plate down into the mantle.

image
Figure 4.37 Models for plate motion mechanisms. Source: Steven Earle (2015) CC BY 4.0 view source

Kearey and Vine (1996) have listed some compelling arguments in favour of the ridge-push/slab-pull model:

References

Kearey, P., & Vine, F. (1996). Global Tectonics (2nd E.). Oxford: Blackwell Science Ltd.

27

Chapter 4 Summary

The topics covered in this chapter can be summarized as follows:

4.1 Alfred Wegener’s Arguments for Plate Tectonics

The evidence for continental drift in the early 20th century included the matching of continental shapes on either side of the Atlantic, and the geological and fossil matchups between continents that are now thousands of kilometres apart.

4.2 Global Geological Models of the Early 20th Century

The established theories of global geology were permanentism and contractionism, but neither of these theories was able to explain some of the evidence that supported the idea of continental drift.

4.3 Geological Renaissance of the Mid-20th Century

Giant strides were made in understanding Earth during the middle decades of the 20th century, including discovering magnetic evidence of continental drift, mapping the topography of the ocean floor, describing the depth relationships of earthquakes along ocean trenches, measuring heat flow differences in various parts of the ocean floor, and mapping magnetic reversals on the sea floor. By the mid-1960s, the fundamentals of the theory of plate tectonics were in place.

4.4 Plates, Plate Motions, and Plate-Boundary Processes

Earth’s lithosphere is made up of over 20 plates that are moving in different directions at rates of between 1 cm/y to greater than 10 cm/y. The three types of plate boundaries are divergent (plates moving apart and new crust forming), convergent (plates moving together and one possibly being subducted), and transform (plates moving side by side). Divergent boundaries form where existing plates are rifted apart, and it is hypothesized that this is caused by a series of mantle plumes. Subduction zones can form where accumulation of sediment at a passive margin leads to separation of oceanic and continental lithosphere. Supercontinents form and break up through these processes.

4.5 Mechanisms for Plate Motion

It is widely believed that ridge-push and slab-pull are the main mechanisms for plate motion, as opposed to traction by mantle convection. Mantle convection is a key factor for producing the conditions necessary for ridge-push and slab-pull.

Review Questions

  1. List some of the evidence used by Wegener to support his idea of moving continents.
  2. What was the primary weakness in Wegener’s continental drift hypothesis?
  3. How were mountains thought to be formed (a) by contractionists and (b) by permanentists?
  4. How were the trans-Atlantic paleontological matchups explained in the late 19th century?
  5. How did we learn about the topography of the sea floor in the early part of the 20th century?
  6. What evidence from paleomagnetic studies provided support for continental drift?
  7. Which parts of the oceans are the deepest?
  8. Why is there less sediment along ocean ridges than on other parts of the sea floor?
  9. How were the oceanic heat-flow data related to mantle convection?
  10. Describe the spatial distribution of earthquakes at ocean ridges and ocean trenches.
  11. In the model for ocean basins developed by Harold Hess, what happens at oceanic ridges and what happens at oceanic trenches?
  12. What aspect of plate tectonics was not included in the Hess model?
  13. What is a mantle plume and what is its expected lifespan?
  14. Describe the nature of movement at an ocean ridge transform fault (a) between the ridge segments, and (b) outside of ridge segments.
  15. How is it possible for a plate to include both oceanic and continental crust?
  16. What is the likely relationship between mantle plumes and the development of a continental rift?
  17. Why does subduction not take place at a continent-continent convergence zone?
  18. Where are Earth’s most recent sites of continental rifting and creation of new ocean floor?
  19. What geological situation might eventually lead to the generation of a subduction zone at a passive ocean-continent boundary such as the eastern coast of North America?

28

Answers to Chapter 4 Review Questions

  1. The evidence used by Wegener to support his idea of moving continents included matching continental shapes and geological features on either side of the Atlantic; common terrestrial fossils in South America, Africa, Australia, and India; and data on the rate of separation between Greenland and Europe.
  2. The primary weakness of Wegener’s theory was that he had no realistic mechanism for making continents move.
  3. Contractionists believed that mountains formed because the crust wrinkled into mountains as Earth cooled and contracted. Permanentists believed that mountains formed by the geosynclinal process.
  4. In the late 19th century the trans-Atlantic paleontological matchups were explained by assuming that there must have been land bridges between the continents at some time in the past.
  5. Prior to 1920, ocean depths were measured by dropping a weighted line over the side of ship. Echo sounding techniques were developed at around that time and greatly facilitated the measurement of ocean depths.
  6. Paleomagnetic studies showed that old rocks on the continents indicated different locations for magnetic north than the position of magnetic north today. They also showed that the difference in pole position from data on different continents increased progressively for older and older rocks. This implied that either Earth had more than one magnetic pole moving around, or that the continents had moved.
  7. Trenches associated with subduction zones are the deepest parts of the oceans.
  8. The ocean ridge areas are the youngest parts of the sea floor and thus there hasn’t been time for much sediment to accumulate.
  9. It was (and still is) thought that high heat flow exists where mantle convection cells are moving hot rock from the lower mantle toward the surface, and that low heat flow exists where there is downward movement of mantle rock.
  10. Earthquakes are consistently shallow and relatively small at ocean ridges. At ocean trenches earthquakes become increasingly deep in the direction that the subducting plate is moving.
  11. In the Hess model new crust was formed at ocean ridges. Crust was recycled back into the mantle at the trenches.
  12. Hess’s model did not include the concept of tectonic plates.
  13. A mantle plume is a column of hot rock (not magma) that ascends toward the surface from the lower mantle. Mantle plumes last tens of millions of years to hundreds of millions of years.
  14. (a) Between the ridge segments there is movement in opposite directions along a transform fault. (b) Outside of the ridge segments the two plates are moving in the same direction and likely at about the same rate. These regions are known as fracture zones.
  15. Tectonic plates are made up of crust and the lithospheric (rigid) part of the underlying mantle. The mantle part ensures that the very different oceanic and continental crust sections of a plate can act as one unit.
  16. A mantle plume beneath a continent can cause the crust to form a dome that might eventually split open. Several mantle plumes along a line within a continent could lead to rifting.
  17. Subduction does not take place at a continent-continent convergent zone because neither plate is dense enough to sink into the mantle.
  18. Continental rifting is taking place along the East Africa Rift, and sea floor has recently been created in the Red Sea and also in the Gulf of California.
  19. The accumulation of sediment at a passive ocean-continent boundary will lead to the depression of the lithosphere and could eventually result in the separation of the oceanic and continental parts of the plate and the beginning of subduction.

V

Chapter 5. Minerals

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 5.1 Giant crystals of gypsum in the Naica Mine in Mexico. The crystals formed in volcanically heated water, and became accessible when the cave was drained as part of mining activities. The cave was very hot, making it fatal for visitors to enter without cooling equipment and respirators. When mining activities ceased, caverns were allowed to flood again. Source: Karla Panchuk CC BY-NC-SA 4.0. Photograph- Paul Williams (2009) CC BY-NC 2.0 view source. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

 

What Is a Mineral?

Minerals are all around us: the graphite in your pencil, the salt on your table, the plaster on your walls, and the trace amounts of gold in your computer. Minerals can be found in a wide variety of consumer products such as paper, medicine, processed foods, and cosmetics. And of course, everything made of metal is also derived from minerals.

A mineral is a naturally occurring solid made of specific elements, and arranged in a particular repeating three-dimensional structure.

“Naturally occurring” means that minerals can be formed from substances and under conditions found in nature. Substances that can only be made by humans—classified as anthropogenic materials—do not count as minerals, nor do substances produced by natural processes acting upon anthropogenic materials.

In the context of the definition of minerals, “solid” means solid at 25ºC. There are some exceptions to this rule, made for substances defined as minerals before 1959, prior to strict procedures being established for determining what is or isn’t a mineral. One example is ice, which is only solid at or below 0 °C. Another is mercury, which is solid below -39 ºC. Mercury that is present in rocks at temperatures above -39 ºC appears as silvery blobs of liquid (Figure 5.2).

Figure 5.2 Droplets of native mercury (pure mercury, Hg), also called quicksilver, amid waxy red crystals of cinnabar (HgS). Cinnabar is a mercury ore mineral. Source: Parent Géry (2012) CC BY-SA 3.0 view source

“Specific elements” means that minerals have a specific chemical formula or composition. The mineral pyrite, for example, is FeS2 (two atoms of sulphur for each atom of iron), and any significant departure from that formula would make it a different mineral. Some minerals can have variable compositions within a specific range. The mineral olivine, for example, has a formula written as (Fe,Mg)2SiO4, because the composition of olivine can range all the way from Fe2SiO4 to Mg2SiO4,  and have any proportion of iron and magnesium in between. This type of substitution is known as solid solution.

Most important of all, the atoms within a mineral are arranged in a specific repeating three-dimensional structure or lattice. This regular structure means that all minerals are crystals. The mineral halite, which we use as table salt, has a relatively simple crystal lattice (Figure 5.3). Atoms of sodium (Na, purple) alternate with atoms of chlorine (Cl, green). The chemical bonds holding the Na and Cl atoms together are all at 90º to each other. Even tiny crystals, like the ones in your salt shaker, have lattices that extend in three dimensions for thousands of repetitions. Halite will always have this structure, and will always have the formula NaCl.

Figure 5.3 Halite crystal lattice. Halite is the mineral in table salt. Source: Steven Earle (2015) CC BY 4.0

Some mineral-like materials do not have a regular internal atomic arrangement. Opal (Figure 5.4) is one example. In many respects it fits the definition of a mineral: it has a specific chemical composition (SiO2·nH2O, where n means that there can be varying amounts of water in the structure), forms naturally through geological processes, and is solid at 25 ºC. However, the structure of opal consists of closely packed spheres (Figure 5.4, right) rather than a lattice like halite. Substances like opal, which are mineral-like, but which do not have a crystalline structure, are called mineraloids.

Figure 5.4 Opal is mineral-like, but does not have a crystalline structure. Instead, it is made up of layers of closely packed spheres (right). Source: Left- James St. John (2016) CC BY 2.0 view source; Right- Mineralogy Division, Geological and Planetary Sciences, Caltech (n.d.) CC BY-NC view source/ view context
Note: Element symbols such as Na and Cl are used extensively in this book. In Appendix A you can find a list of the symbols, the names of the elements common in minerals, and a copy of the periodic table of elements.

References

Nickel, E. H. (1995). The Definition of a Mineral. The Canadian Mineralogist 33, 698-690. Read paper

Williams, P. (2010, July 28). Deadliest place on Earth? Surviving Cueva de los Cristales – The Giant Crystal Cave. Visit website

29

5.1 Atoms

Protons Are What Make Elements Distinct

All matter, including mineral crystals, is made up of atoms.  All atoms are made up of three main particles: protons, neutrons, and electrons. Protons have a positive charge, neutrons have no charge, and electrons have a negative charge. Protons and neutrons have approximately the same mass, but electrons have a mass that is 10,000 times smaller.

The element hydrogen (H) has the simplest atoms. Most hydrogen atoms have just one proton and one electron. The proton forms the nucleus (the centre of the atom), while the electron orbits around it (Figure 5.5, left). All other elements have more than one proton in their nucleus. Protons repel each other because they are positively charged, but it is possible to have more than one proton in a nucleus because neutrons hold them together. The next most complex atom, helium (He) has two protons and two neutrons in its nucleus, in its most common form. Some atoms of the same element can have different numbers of neutrons. For example, forms of hydrogen exist with one and two neutrons, and a tiny fraction of He atoms have only one neutron. Forms of an element with different numbers of neutrons are called isotopes.

Figure 5.5 Atomic structure of hydrogen and helium showing protons (p+), neutrons (n), and electrons (e-). Source: Bruce Blaus (2014) CC BY 3.0 view source

The number of protons in an atom determines what element it will be, so the number of protons is called the atomic number of that element. The total number of protons and neutrons in the nucleus is the mass number. The mass number distinguishes between isotopes of an element. Isotopes of an element are denoted by putting the mass number as a subscript in front of the symbol for that element. For example, the isotopes of hydrogen are 1H (1 proton), 2H (1 proton + 1 neutron), and 3H (1 proton + 2 neutrons).

For most of the 16 lightest elements (up to oxygen) the number of neutrons is equal to the number of protons. For most of the remaining elements, there are more neutrons than protons. This is because the more protons that are concentrated in a small space, the more neutrons are needed to keep the nucleus together.  The most common isotope of uranium (U), for example, is 238U. It has 92 protons, but requires 146 neutrons to keep them together.  The neutrons are only partly successful.  Uranium is radioactive, meaning that its nucleus will eventually split apart and release energy. What remains of the nucleus has fewer protons, so after decay the atom is a different element.

Electrons Are What Control How Atoms Interact

Electrons orbiting around the nucleus of an atom are arranged in shells (also called energy levels). The first shell can hold only two electrons (as in H and He in Figure 5.5), but the next shell holds up to eight electrons. An atom can have many shells of electrons, but there are never more than 8 outermost electrons interacting with surrounding atoms.

The outermost electrons determine how atoms can be bonded together. Elements that have a full outer shell (e.g., neon, Figure 5.6 right)  are inert because they do not react with other elements to form compounds. These are the noble gases (including helium, argon, krypton, and radon, in addition to neon) in the far-right column of the periodic table. For elements that do not have a full outer shell (e.g., lithium, Figure 5.6 left), the outermost electrons can interact with the outermost electrons of nearby atoms to create chemical bonds.

Figure 5.6 The number of electrons in an atom’s outermost shell (or energy level) determine whether it will bond to other atoms, and how it will bond. Right- Neon has a completely filled outer shell with 8 electrons. It does not bond with other atoms. Left- Lithium has only one electron in its outer shell. It bonds with other atoms. Source: Bruce Blaus (2014) CC BY 3.0 view source

The electron shell configurations for 29 of the first 36 elements are listed in Table 5.1. Note that some of the shells in the table below have more than 8 electrons.  This is because they contain subshells.  For example, the third shell can hold up to 18 electrons because it contains one subshell that can hold 2 electrons, and two subshells that can hold 8 electrons each.

Table 5.1 Electron shell configurations of some of the elements up to krypton. Inert elements (those with filled outer shells) are shaded.
 Number of Electrons in Each Shell
Element Symbol Atomic Number First Second Third Fourth
Hydrogen H 1 1
Helium He 2 2
Lithium Li 3 2 1
Beryllium Be 4 2 2
Boron B 5 2 3
Carbon C 6 2 4
Nitrogen N 7 2 5
Oxygen O 8 2 6
Fluorine F 9 2 7
Neon Ne 10 2 8
Sodium Na 11 2 8 1
Magnesium Mg 12 2 8 2
Aluminum Al 13 2 8 3
Silicon Si 14 2 8 4
Phosphorus P 15 2 8 5
Sulphur S 16 2 8 6
Chlorine Cl 17 2 8 7
Argon Ar 18 2 8 8
Potassium K 19 2 8 8 1
Calcium Ca 20 2 8 8 2
Scandium Sc 21 2 8 9 2
Titanium Ti 22 2 8 10 2
Vanadium V 23 2 8 11 2
Chromium Cr 24 2 8 13 1
Manganese Mn 25 2 8 13 2
Iron Fe 26 2 8 14 2
. . . . . . .
Selenium Se 34 2 8 18 6
Bromine Br 35 2 8 18 7
Krypton Kr 36 2 8 18 8

 

30

5.2 Bonding and Lattices

Atoms seek to have a full outer shell. For hydrogen and helium, a full outer shell means two electons. For other elements, it means 8 electrons. Filling the outer shell is accomplished by transferring or sharing electrons with other atoms in chemical bonds.  The type of chemical bond is important for the study of minerals because the type of bond will determine many of a mineral’s physical and chemical properties.

Ionic Bonds

Consider the example of halite again, which is made up of sodium (Na) and chlorine (Cl).  Na has 11 electrons: two in the first shell, eight in the second, and one in the third (Figure 5.7, top). Na readily gives up the third shell electron so it can have the second shell with 8 electrons as its outermost shell.  When it loses the electron, the total charge from the electrons is -10, but the total charge from the protons is +11, so it is left with a +1 charge over all.

Figure 5.7 Electron configuration of sodium and chlorine atoms (top). Sodium gives up an electron to become a cation (bottom left) and chlorine accepts an electron to become an anion (bottom right). Source: Steven Earle (2015) CC BY 4.0 view source

Chlorine has 17 electrons: two in the first shell, eight in the second, and seven in the third. Cl readily accepts an eighth electron to fill its third shell, and therefore becomes negatively charged because it has a total charge of -18 from electrons, and a total charge of +17 from protons.

In changing their number of electrons, these atoms become ions — the sodium loses an electron to become a positive ion or cation,You can remember that a cation is positive by remembering that a cat has paws (paws sounds like "pos" in "positive"). You could also think of the "t" in "cation" as a plus sign. and the chlorine gains an electron to become a negative ion or anion (Figure 5.7, bottom). Because negative and positive charges attract, sodium and chlorine ions stick together, creating an ionic bond. In an ionic bond, electrons can be thought of as having transferred from one atom to another.

 

Exercise: Cation or Anion?

A number of elements are listed below along with their atomic numbers (the number of protons, and therefore also the number of electrons in the atom). Assuming that the first electron shell can hold two electrons and subsequent electron shells can hold eight electrons, sketch the electron configurations for these elements, as in the example for fluorine (Figure 5.8). If you fill a shell and have electrons left over, draw another shell around the atom. Predict whether the element is likely to form a cation or an anion, and what charge it would have (e.g., +1, +2, –1).

Figure 5.8 How to draw the electron configuration for fluorine, with an atomic number of 9. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source
1. Lithium (3) 5. Beryllium (4)
2. Magnesium (12) 6. Oxygen (8)
3. Argon (18) 7. Sodium (11)
4. Chlorine (17)

 

Covalent Bonds

An element like chlorine can also form bonds without forming ions. For example, two chlorine atoms can each complete their outer shells by sharing electrons.  Chlorine gas (Cl2, Figure 5.9) is formed when two chlorine atoms form a covalent bond.

Figure 5.9 A covalent bond between two chlorine atoms. The electrons are black in the left atom, and blue in the right atom. Two electrons are shared (one black and one blue) so that each atom appears to have a full outer shell. Source: Steven Earle (2015) CC BY 4.0 view source

Carbon is another atom that participates in covalent bonding.  An uncharged carbon atom has six protons and six electrons. Two of the electrons are in the inner shell and four are in the outer shell (Figure 5.10, left). Carbon would need to gain or lose four electrons to have a filled outer shell, and this would create too great a charge imbalance. Instead, carbon atoms share electrons to create covalent bonds (Figure 5.10, right).

Figure 5.10 The electron configuration of carbon (left) and the sharing of electrons in covalent C bonding (right). The electrons shown in blue are shared between adjacent C atoms. Source: Steven Earle (2015) CC BY 4.0 view source

In the mineral diamond (Figure 5.11, left), the carbon atoms are linked together in a three-dimensional framework, where one carbon atom is bonded to four other carbon atoms, and every bond is a very strong covalent bond.

Figure 5.11 Covalently bonded structures. Left: Diamond with three-dimensional structure of covalently bonded carbon. Right: Graphite with covalently bonded sheets of carbon. Sheets are held together by weaker van der Waals forces. Source: Karla Panchuk (2018) CC BY 4.0, modified after Materialscientist (2009) CC BY-SA 3.0 view source

Other Types of Bonds

Most minerals are characterized by ionic bonds, covalent bonds, or a combination of the two, but there are other types of bonds that are important in minerals. Consider the mineral graphite (Figure 5.11, right): the carbon atoms are linked together in sheets or layers in which each carbon atom is covalently bonded to three others. Graphite-based compounds are strong because of the covalent bonding between carbon atoms within each layer, which is why they are used in high-end sports equipment such as ultralight racing bicycles. Graphite itself is soft, however, because the layers themselves are held together by relatively weak Van der Waals forces

Van der Waals forces, like hydrogen bonds, work because molecules can be electrostatically neutral, but still have an end that is slightly more positive and an end that is slightly more negative. In water molecules (Figure 5.12, left), the bent shape puts the hydrogen atoms on one side of the molecule, and the oxygen atom, with more electrons, on the other. The charge is distributed asymmetrically across the water molecule. Contrast this with the straight carbon dioxide (Figure 5.12, right) molecule. The slightly more negative oxygen atoms on the ends are distributed symmetrically on either side of the carbon atom.  

Figure 5.12 Hydrogen bonding. Water molecules (left) are polar molecules (their charge is distributed asymmetrically). Slightly negative parts of the molecule are attracted to slightly positive parts of other water molecules. Carbon dioxide (right) is a non-polar molecule. The slightly negative oxygen atoms are distributed symmetrically on either side of the carbon atom. Source: Karla Panchuk (2018) CC BY-SA 4.0. Modified after Querter (2011) CC BY-SA 3.0 view source and Jynto (2011) CC0 1.0 view source

Metallic bonding occurs in metallic elements because they have outer electrons that are relatively loosely held. (The metals are highlighted on the periodic table in Appendix 1.) When bonds between such atoms are formed, the dissociated electrons can move freely from one atom to another. This feature accounts for two very important properties of metals: their electrical conductivity and their malleability (they can be deformed and shaped).

Figure 5.13 Metallic bonding. Dissociated electrons (grey dots) move between metal atoms. Source: Karla Panchuk (2018) CC BY-SA 4.0. Nucleus by Fornax (2010) CC BY-SA 3.0 view source

31

5.3 Mineral Groups

Minerals are organized according to the anion or anion group (a group of atoms with a net negative charge, e.g., SO42–) they contain, because the anion or anion group has the biggest effect on the properties of the mineral.  Silicates, with the anion group SiO44-, are by far the most abundant group in the crust and mantle. (They will be discussed in Section 5.4). The different mineral groups along with some examples of minerals in each group are summarized below.

Oxide Minerals: O2- Anion

Oxide minerals (Figure 5.14) have oxygen (O2–) as their anion.  They don’t include anion groups with other elements, such as  the carbonate (CO32–), sulphate (SO42–), and silicate (SiO44–) anion groups. The iron oxides hematite and magnetite are two examples that are important ores of iron. Corundum is an abrasive, but can also be a gemstone in its ruby and sapphire varieties. If the oxygen is also combined with hydrogen to form the hydroxyl anion (OH), the mineral is known as a hydroxide. Some important hydroxides are limonite and bauxite, which are ores of iron and aluminum, respectively.

Oxide minerals shown are hematite (Fe2O3), magnetite (Fe3O4), corundum (Al2O3), limonite (2Fe2O3-3H2O), bauxite (Al2O3-2H2O)
Figure 5.14 Oxide minerals include metal ore minerals, industrial minerals, and gemstones. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

Sulphide Minerals: S2- Anion

Sulphide minerals (Figure 5.15) include galena, sphalerite, chalcopyrite, and molybdenite, which are the most important ores of lead, zinc, copper, and molybdenum, respectively. Some other sulphide minerals are pyrite, bornite, stibnite, and arsenopyrite. Sulphide minerals tend to have a metallic sheen.

Sulphide minerals include galena (PbS), sphalerite (ZnS), chalcopyrite (CuFeS2), molybdenite (MoS2), pyrite (FeS2), bornite (Cu5FeS4), stibnite (Sb2S3), and arsenopyrite (FeAsS).
Figure 5.15 Sulphide minerals often have a metallic lustre and include metal ores. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

Sulphate Minerals: SO42- Anion Group

Many sulphate minerals form when sulphate-bearing water evaporates. A deposit of sulphate minerals may indicate that a lake or sea has dried up at that location.  Sulphates with calcium include anhydrite, and gypsum (Figure 5.16). Sulphates with barium and strontium are barite and celestite, respectively. In all of these minerals, the cation has a +2 charge, which balances the –2 charge on the sulphate ion.

Figure 5.16 Sulphate minerals. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Click the image for photo sources.

Halide Minerals: Anions from the Halogen Group

The anions in halides are the halogen elements including chlorine, fluorine, and bromine. Examples of halide minerals are cryolite, fluorite, and halite (Figure 5.17).  Halide minerals are made of ionic bonds. Like the sulphates, some halides also form when mineral-rich water evaporates.

Halides include halite (NaCl), cryolite (Na3AlF6), and fluorite (CaF2).
Figure 5.17 Halide minerals. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Click the image for photo sources.

Carbonate Minerals: CO32- Anion Group

The carbonate anion group combines with +2 cations to form minerals such as calcite, magnesite, dolomite, and siderite (Figure 5.18). The copper minerals malachite and azurite are also carbonates.  The carbonate mineral calcite is the main component of rocks formed in ancient seas by organisms such as corals and algae.

Carbonate minerals include calcite (CaCO3), magnesite (MgCO3), dolomite ((Ca,Mg)CO3), and siderite (FeCO3). Malachite and azurite are hydrated copper carbonates.
Figure 5.18 Carbonate minerals. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photos by Rob Lavinsky, iRocks.com, CC BY-SA 3.0. Click the image for photo sources.

Phosphate Minerals: PO43- Anion

The apatite group of phosphate minerals (Figure 5.19, left) includes hydroxyapatite, which makes up the enamel of your teeth. Turquoise is also a phosphate mineral (Figure 5.19, right).

Figure 5.19 Phosphate minerals. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

Silicates (SiO44)

The silicate minerals include the elements silicon and oxygen in varying proportions . These are discussed at length in Section 5.4.

Native Element Minerals

These are minerals made of a single element, such as gold, copper, silver, or sulphur (Figure 5.20).

Figure 5.20 Native element minerals are made up of a single element. Source: Karla Panchuk (2018) CC BY-SA 4.0. Click the image for photo sources.

Exercise: Mineral Groups

Minerals are grouped according to the anion part of the mineral formula, and mineral formulas are always written with the anion part last. For example, for pyrite (FeS2), Fe2+ is the cation and S is the anion. This helps us to know that it’s a sulphide, but it is not always that obvious. Hematite (Fe2O3) is an oxide; that’s easy, but anhydrite (CaSO4) is a sulphate because SO42– is the anion, not O. Similarly, calcite (CaCO3) is a carbonate, and olivine (Mg2SiO4) is a silicate. Minerals with only one element (such as S) are native minerals, while those with an anion from the halogen column of the periodic table (Cl, F, Br, etc.) are halides. Provide group names for the following minerals:

Mineral Formula Group
sphalerite ZnS
magnetite Fe3O4
pyroxene MgSiO3
anglesite PbSO4
sylvite KCl
silver Ag
fluorite CaF2
ilmenite FeTiO3
siderite FeCO3
feldspar KAlSi3O8
sulphur S
xenotime YPO4

 

32

5.4 Silicate Minerals

Silicon and oxygen bond covalently to create a silicate tetrahedron (SiO44-), which is a four-sided pyramid shape with oxygen at each corner and silicon in the middle (Figure 5.21). This structure is the building block of many important minerals in the crust and mantle. Silicon has a charge of +4, and oxygen has a charge of -2, so the total charge of the silicate anion is -4.

Figure 5.21 The silica tetrahedron is the building block of all silicate minerals. Source: Karla Panchuk (2018) CC BY-SA 4.0. Modified after Helgi (2013) CC BY-SA 3.0 view source

In silicate minerals, these tetrahedra are arranged and linked together in a variety of ways, from single units to chains, rings, and more complex frameworks.  In the rest of this section we will discuss the structures of the most common silicate minerals in Earth’s crust and mantle.

Exercise: Make a Tetrahedron

Download this PDF file with the tetrahedron pattern below. Cut around the outside of the shape (solid lines and dotted lines), and then fold along the solid lines to form a tetrahedron.

If you have glue or tape, secure the tabs to the tetrahedron to hold it together. If you don’t have glue or tape, make a slice along the thin grey line and insert the pointed tab into the slit.

If you’re feeling ambitious, make several tetrahedra and and use toothpicks through the corners to make the configurations discussed below.

Figure 5.22 Pattern for a tetrahedron. Source: Steven Earle (2015) CC BY 4.0 view source

Isolated Tetrahedra

The simplest silicate structure, that of the mineral olivine (Figure 5.23), is composed of isolated tetrahedra bonded to iron and/or magnesium ions (Figure 5.23 left). In olivine, the –4 charge of each silica tetrahedron is balanced by two iron or magnesium cations, each with a charge of +2. Olivine can be pure Mg2SiO4 or Fe2SiO4, or a combination of the two, written as (Mg,Fe)2SiO4. Magnesium and iron can substitute for each other because they both have a charge of +2, and they are similar in size. Magnesium cations have a radius of 0.73 Å, and iron cations have a radius of 0.62 Å Å stands for Ångstrom, a unit commonly used to express atomic-scale dimensions. One angstrom is 10–10 m or 0.0000000001 m..

Figure 5.23 Olivine is a silicate mineral made of isolated silica tetrahedra bonded to Fe and Mg ions (left). Olivine crystals (centre) can often be found in the volcanic igneous rock called basalt (right). Source: Karla Panchuk (2018) CC BY-SA 4.0. Left- modified after Steven Earle (2015) CC BY 4.0 view source. Click the image for photo sources.

Although the iron and magnesium ions are similar in size, allowing them to substitute for each other in some silicate minerals, the common ions in silicate minerals have a wide range of sizes (Figure 5.24).  Ionic radii are critical to the composition of silicate minerals, because the structure of the silicate mineral will determine the size of spaces available.

Figure 5.24 The ionic radii in angstroms of some of the common ions in silicate minerals. Radii shown to scale. Notice that iron appears twice with two different radii. This is because iron can exist as a +2 ion (if it loses two electrons when it becomes an ion) or a +3 ion (if it loses three). Fe2+ is known as ferrous iron. Fe3+ is known as ferric iron. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view source

Chain Silicates

Pyroxene (Figure 5.25 bottom left) is an example of a single-chain silicate.  The structure of chain silicates is shown in Figure 5.25 (top). In pyroxene, silica tetrahedra form a chain because one oxygen from each tetrahedron is shared with the adjacent tetrahedron. This means there are fewer oxygens in the structure. This can be expressed as an oxygen-to-silicon ratio (O:Si). The O:Si is lower than in olivine (3:1 instead of 4:1), and the net charge per silicon atom is less (–2 instead of –4), because fewer cations are necessary to balance that charge.

Figure 5.25 Chain silicate minerals. Top: Arrangement of silica tetrahedra in single and double chains. Bottom left: Pyroxene crystals (dark crystals) of the variety aegirine (acmite). Bottom right: Amphibole crystal (dark) of the variety hornblende. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Top left- modified after Steven Earle (2015) CC BY 4.0. Top right- modified after Klein & Hurlbut (1993). Photos by R. Weller/ Cochise College. Click the image for sources.

Pyroxene compositions have the silica tetrahedra represented as SiO3 (e.g., MgSiO3, FeSiO3, and CaSiO3The variation in composition can also be written as (Mg,Fe,Ca)SiO3, where the elements in the brackets can be present in any proportion..)  In other words, pyroxene has one cation for each silica tetrahedron (e.g., MgSiO3) while olivine has two (e.g., Mg2SiO4).  The structure of pyroxene is more “permissive” than that of olivine, meaning cations with a wider range of ionic radii can fit into it. That’s why pyroxenes can have calcium cations (radius 1.00 Å) substitute for iron (0.63 Å) and magnesium (0.72 Å) .

In amphibole (Figure 5.25 bottom right), the silica tetrahedra are linked in a double chain that has an oxygen-to-silicon ratio lower than that of pyroxene, and hence still fewer cations are necessary to balance the charge. Amphibole is even more permissive than pyroxene and its compositions can be very complex, as shown by the formula for the hornblende group of amphibole minerals in Figure 5.25 (bottom right).

Exercise: Oxygen to Silicon Ratio

Figure 5.26 shows single chain and double chain structures. Count the number of tetrahedra versus the number of oxygen ions (yellow spheres) for each. Each tetrahedron has one silicon atom.
  1. Confirm for yourself that the ratio of silicon to oxygen in the single chain is 1:3.
  2. What is the O:Si for the double chain?
Figure 5.26 Single and double chains of tetrahedra. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 single chain/ double chain

 

Sheet Silicates

In mica structures the silica tetrahedra are arranged in continuous sheets (Figure 5.27), where each tetrahedron shares three oxygen anions with adjacent tetrahedra. Because even more oxygens are shared between adjacent tetrahedra, fewer charge-balancing cations are needed for sheet silicate minerals. Bonding between sheets is relatively weak, and this accounts for the tendency of mica minerals to split apart in sheets (Figure 5.27 bottom right). Two common micas in silicate rocks are biotite (Figure 5.27 bottom left), which contains iron and/or magnesium, making it a dark mineral; and muscovite (Figure 5.27 right), which contains aluminum and potassium, and is light in colour. All of the sheet silicate minerals have water in their structure, in the form of the hydroxyl (OH-) anion.

Figure 5.27 Micas are sheet silicates and split easily into thin layers along planes parallel to the sheets. Biotite mica (lower left) is has Fe and Mg cations. Muscovite mica (lower right) has Al and K instead. The muscovite mica shows how thin layers can split away in a sheet silicate. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Top left- modified after Steven Earle (2015) CC BY 4.0. Top right- modified after Klein & Hurlbut (1993). Photos by R. Weller/ Cochise College. Click the image for sources.

Some sheet silicates typically occur in clay-sized fragments (i.e., less than 0.004 mm). These include the clay minerals kaolinite, illite, and smectite, which are important components of rocks and especially of soils.

Framework Silicates

In framework silicates, tetrahedra are connected to each other in three-dimensional structures rather than in two-dimensional chains and sheets.

Feldspar

Feldspars are a group of very abundant framework silicates in Earth’s crust. They include alumina tetrahedra as well as silicate tetrahedra. In alumina tetrahedra, there is an aluminum cation at the centre instead of a silicon cation.

Feldspars are classified using a ternary (3-fold) system with three end-members (“pure” feldspars). This system is illustrated with a triangular diagram that has each end-member at one corner (Figure 5.28). The distance along a side of the diagram represents the relative abundance of the composition of each end-member.

 

Figure 5.28 Ternary diagram showing the feldspar group of framework silicate minerals. Alkali feldspars are those with compositions ranging between albite (with a Na cation) and orthoclase and its polymorphs (with a K cation. Plagioclase feldspars are those with compositions ranging between albite and anorthite (with a Ca cation). Source: Karla Panchuk (2018) CC BY-SA 4.0. Ternary diagram modified after Klein & Hurlbut (1993). Click the image for photo sources and a ternary diagram without mineral images.

One end-member is potassium feldspar (also referred to as K-feldspar), which has the composition KAlSi3O8. Depending on the temperature and rate of cooling, K-feldspar can occur as one of three polymorphs: orthoclase, sanidine, or microcline.  Another end member is albite, which has sodium instead of potassium (formula NaAlSi3O8). As is the case for iron and magnesium in olivine, there is a continuous range of compositions (referred to as a solid-solution series) between albite and orthoclase. Feldspars in this series are referred to as alkali feldspars. Potassium cations are much larger than sodium cations (1.37 Å versus 0.99 Å, respectively), so high temperatures are required to form alkali feldspars with intermediate compositions.

The third end-member is anorthite and it has calcium instead of potassium or sodium (formula CaAl2Si3O8). Feldspars in the solid-solution series between albite and anorthite are called plagioclase feldspars. Calcium and sodium cations are nearly the same size (1.00 Å and 0.99 Å, respectively), so from that perspective it makes sense that they substitute readily for each other, and that any intermediate compositions between CaAl2Si3O8 and NaAlSi3O8 can exist. However, calcium and sodium ions don’t have the same charge (Ca2+ versus Na+), making it surprising that they substitute so easily. The difference in charge is accommodated by substituting some Al3+ for Si4+.  Albite has one Al and three Si in its formula, while anorthite is has two Al and two Si.  Plagioclase feldspars of intermediate composition also have intermediate proportions of Al and Si.

Quartz

Quartz (SiO2; Figure 5.29) contains only silica tetrahedra. In quartz, each silica tetrahedron is bonded to four other tetrahedra (with an oxygen shared at every corner of each tetrahedron), making a three-dimensional framework.  As a result, the ratio of silicon to oxygen is 1:2. Because the one silicon cation has a +4 charge and the two oxygen anions each have a –2 charge, the charge is balanced. There is no need to add cations to balance the charge. The hardness of quartz and the fact that it breaks irregularly (notice the bottom of the crystal in Figure 5.29 right) and not along smooth planes result from the strong covalent/ionic bonds characteristic of the silica tetrahedron.

Figure 5.29 Quartz is another silicate mineral with a three-dimensional framework of silica tetrahedra. Sometimes quartz occurs as well-developed crystals (left), but it also occurs in common rocks such as granite (right). In addition to quartz, the granite contains potassium feldspar, albite, and amphibole. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

References

Klein, C. & Hurlbut, C. S., Jr. (1993). Manual of Mineralogy (after J. D. Dana). New York, NY: John Wiley & Sons, Inc.

33

5.5 How Minerals Form

The following criteria are required for mineral crystals to grow:

Physical and chemical conditions include factors such as temperature, pressure, amount of oxygen available, pH, and the presence of water. The presence of water makes it easier for ions to move to where there are needed, and can lead to the formation of larger crystals over shorter time periods, as with the gypsum crystals at the beginning of this chapter. Time is one of the most important factors because it takes time for atoms to line themselves up into an orderly structure. If time is limited, the mineral grains may remain very small.

Most of the minerals that make up the rocks in the crust and mantle formed through the cooling of molten rock, known as magma. At the high temperatures that exist deep within Earth, some geological materials are liquid. As magma rises up through the crust, either by volcanic eruption or by more gradual processes, it cools and minerals crystallize. When cooling is rapid and many crystals form at once, only small mineral grains will form before the rock becomes solid. The resulting rock will be fine-grained (i.e., crystals less than 1 mm). When cooling is slow, or when few crystals are growing at a time, relatively large crystals will develop.

Minerals can also form in several other ways:

34

5.6 Mineral Properties

Minerals are universal. A crystal of hematite on Mars will have the same properties as one on Earth, and the same as one on a planet orbiting another star. That’s good news for geology students who are planning interplanetary travel, because they can use the same properties to identify minerals anywhere. That doesn’t mean that it’s easy, however. Identification of minerals takes practice. Some of the mineral properties that are useful for identification are colour, streak, lustre, hardness, habit, cleavage or fracture, and density.

Colour

Some minerals have distinctive colours that useful as diagnostic criteria. The mineral sulphur (Figure 5.30 left) is always a characteristic bright yellow. For other minerals, colour might vary. Hematite is an example of a mineral for which colour is not necessarily diagnostic. In some forms hematite is a deep dull red (a fairly unique colour), but in others it is a metallic silvery black (5.30, right).

Figure 5.30 Colour is a useful diagnostic property for sulphur (left) and for some types of hematite (right) because the yellow and dark red colours are unique to those minerals. In contrast, silvery metallic forms of hematite are similar in appearance to many other minerals. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

For other minerals, the problem is that a single mineral can have a wide range of colours. The colour variations can be the result of varying proportions of trace elements within the mineral, or structural defects within the crystal lattice. In the case of quartz (Figure 5.31), milky quartz gets its white colour from millions of tiny fluid-filled cavities. Smoky quartz gets its grey colour from structural damage caused by natural radiation. Amethyst and citrine get their colours from trace amounts of iron, and rose quartz gets its pink hue from manganese.

Figure 5.31 The many colours of quartz.Quartz can be colourless, milky, a greyish smoky colour, purple, yellow, and pink. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

Streak

The colour of a mineral is what you see when light reflects off the surface of the sample. One reason that colour can be so variable is that the surface textureis variable. A way to get around this problem is to grind a small amount of the sample to a powder and observe the colour of the powder. This colour is the mineral’s streak. The mineral can be powdered by scraping the sample across a piece of unglazed porcelain called a streak plate (Figure 5.32). In Figure 5.32, two samples of hematite have been scraped across the streak plate. Even though one sample is metallic and the other is deep red, both have a similar reddish-brown streak.

Figure 5.32 Hematite leaves a distinctive reddish-brown streak whether the sample is metallic or deep red. Source: Karla Panchuk (2015) CC BY 4.0

 

Streak is an especially helpful property when minerals look similar. In Figure 5.33 all of the minerals are dark in colour, with varying degrees of metallic sheen. The streaks of these minerals are much more distinctive.

Figure 5.33 Similar dark-grey minerals with varying degrees of metallic sheen leave different colours of streaks. The minerals are from upper left clockwise: hematite, magnetite, sphalerite, and galena. Source: Karla Panchuk (2015) CC BY 4.0

Lustre

Lustre is the way light reflects off the surface of a mineral, and the degree to which it penetrates into the interior. The key distinction is between metallic and non-metallic lustre. Light does not pass through metals, and that is the main reason they look metallic (e.g., the hematite on left of Figure 5.32). Even a thin sheet of metal — such as aluminum foil — will be not permit light to pass through it. Many non-metallic minerals may look as if light will not pass through them, but if you look closely at a thin slice of the mineral you will see that the mineral is translucent or transparent.

If a non-metallic mineral has a shiny, reflective surface, it is said to have a glassy lustre.  The quartz crystals in Figure 5.31 are examples of minerals with glassy lustre. If the mineral surface is dull and non-reflective, it has an earthy lustre (like the hematite on the right of Figure 5.32). Other types of non-metallic lustres are silky, pearly, and resinous. Lustre is a good diagnostic property because most minerals will always appear either metallic or non-metallic, although as Figure 5.31 shows, there are exceptions.

Hardness

One of the most important diagnostic properties of a mineral is its hardness. In practical terms, hardness determines whether or not a mineral can be scratched by a particular material.

In 1812 German mineralogist Friedrich Mohs came up with a list of 10 minerals representing a wide range of hardness, and numbered them 1 through 10 in order of increasing hardness (Figure 5.34, horizontal axis). While each mineral on the list is harder than the one before it, the measured hardness (vertical axis) is not linear. Notice that apatite is about three times harder than fluorite, and diamond is three times harder than corundum.

Figure 5.34 Minerals and reference materials in the Mohs scale of hardness. The measured hardness values are Vickers Hardness numbers. Source: Steven Earle (2015) CC BY 4.0 view source

Some commonly available reference materials are also shown on this diagram, including a typical fingernailNote that artificial fingernails may be much harder than natural fingernails. Some materials used for artificial nails are harder than quartz. (2.5), a piece of copper wire (3.5), a knife blade or piece of window glass (5.5), a hardened steel file (6.5), and a porcelain streak plate (7). These are tools that a geologist can use to measure the hardness of unknown minerals: if you have a mineral that you can’t scratch with your fingernail, but you can scratch with a copper wire, then its hardness is between 2.5 and 3.5. The minerals themselves can be used to test other minerals.

Crystal Habit

When minerals form within rocks, there is a possibility that they will form in distinctive crystal shapes if they are not crowded out by other pre-existing minerals. Every mineral has one or more distinctive crystal habits determined by their atomic structure, although it is not that common in ordinary rocks for the shapes to be obvious.

Quartz, for instance, will form six-sided prisms with pointed ends (Figure 5.35 left), but this typically happens only when it crystallizes from a hot water solution within a cavity in an existing rock. Pyrite can form cubic crystals (Figure 5.35 centre), but can also form crystals with 12 faces, known as dodecahedra. The mineral garnet also forms many-sided crystals with an over-all rounded shape (Figure 5.35 right).

Figure 5.35 Hexagonal prisms of quartz (left), intergrown cubic crystals of pyrite (centre), and 24-sided crystals of garnet (right). Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources.

Some of the terms that are used to describe habit include bladed, botryoidal (grape-like), dendritic (branched), drusy (an encrustation of crystals), equant (similar size in all dimensions), fibrous, platy, prismatic (long and thin), and stubby.

Cleavage and Fracture

Cleavage and fracture describe how a mineral breaks. These characteristics are the most important diagnostic features of many minerals, and often the most difficult to understand and identify. Cleavage is what we see when a mineral breaks along a plane or planes, while fracture is an irregular break. Some minerals tend to cleave along planes at various fixed orientations. Some, like quartz, do not cleave at all, only fracture. Minerals that have cleavage can also fracture along surfaces that are not parallel to their cleavage planes.

The way minerals break is determined by the arrangement of atoms within them, and more specifically by the orientation of weaknesses within their crystal lattice. Graphite and mica break off in parallel sheets (Figure 5.36).

Figure 5.36 One direction of cleavage (basal cleavage). Left: Schematic of basal cleavage. Right: Muscovite showing basal cleavage. The white dashed line marks the edge of the cleavage plane. Source: Karla Panchuk (2018) CC BY-SA 4.0. Cleavage diagram modified after M.C. Rygel (2010) CC BY-SA 3.0 view source

Other minerals have two directions of cleavage, classified as two directions at 90° (Figure 5.37 top) and two directions not at 90° (Figure 5.37 bottom). While the diagrams of planes on the left of Figure 5.37 make this difference clear, it may be less obvious in practice. The minerals in Figure 5.37 both have two planes of cleavage that are very close to 90°.  The white dashed lines mark the edges of the planes, as with Figure 5.36.  See if you can find the planes repeated in the images.  The images are close-up views of the minerals, only a few cm across. Sometimes you must look very carefully to find cleavage planes.

Figure 5.37 Two directions of cleavage. Top: Two directions at 90° in pyroxene. Bottom: two directions not at 90° in plagioclase feldspar. Edges of cleavage planes marked with dashed lines. Source: Karla Panchuk (2018) CC BY-SA 4.0. Cleavage diagrams modified after M.C. Rygel (2010) CC BY-SA 3.0 view source

Some minerals have many directions of cleavage.  Figure 5.38 shows  minerals with three directions of cleavage.  Halite (Figure 5.38 top) has three directions at 90° and calcite (Figure 5.38 bottom) has three directions not at 90°.

Figure 5.38 Three directions of cleavage. Top: Three directions at 90° in halite. Bottom: Three directions not at 90° in calcite. Source: Karla Panchuk (2018) CC BY-SA 4.0. Cleavage diagrams modified after M.C. Rygel (2010) CC BY-SA 3.0 view source

There are a few common difficulties that students encounter when learning to recognize and describe cleavage.  One is that it might be necessary to look very closely at a sample to see mineral cleavage.  The key features in Figure 5.37 are only cm or mm in scale.  If crystals are very small, it may not be possible to see cleavage at all. Another issue is that sometimes cleavage is present, but it is poor, meaning the cleavage surface isn’t perfectly flat. Finally it can be difficult to know whether a flat surface on a crystal is a cleavage plane, a crystal face, or simply a surface that happens to be flat. Cleavage planes tend to repeat themselves at different depths throughout the mineral, so if you are unsure whether the surface you are looking at is a cleavage plane, try rotating the mineral in bright light. If cleavage is present, you will generally find that, for a given cleavage direction, all of the cleavage surfaces will glint in the light simultaneously. Crystal faces will also glint in light, but they do not repeat themselves at depth throughout the mineral. The best way to overcome all of these problems is to look at lots of examples.  It’s worth it to be able to identify cleavage and fracture, because cleavage is a reliable diagnostic property for most minerals.

Density

Density is a measure of the mass of a mineral per unit volume, and it is a useful diagnostic tool in some cases. Most common minerals, such as quartz, feldspar, calcite, amphibole, and mica, are of average density (2.6 to 3.0 g/cm3), and it would be difficult to tell them apart on the basis of their density. On the other hand, many of the metallic minerals, such as pyrite, hematite, and magnetite, have densities over 5 g/cm3. If you picked up a sample of one of these minerals, it would feel much heavier compared to a similarly sized sample of a mineral with average density. A limitation of using density as a diagnostic tool is that one cannot assess it in minerals that are a small part of a rock with other minerals in it.

Other properties

Several other properties are useful for identification of some minerals. Some of these are:

35

Chapter 5 Summary

The topics covered in this chapter can be summarized as follows:

5.1 Atoms

An atom is made up of protons and neutrons in the nucleus, and electrons arranged in energy shells around the nucleus. The first shell holds two electrons, and outer shells hold more. Atoms strive to have eight electrons in their outermost shell (or two for H and He). Atoms gain, lose, or share electrons to achieve this. In so doing they become either positively charged cations (if they lose electrons) or negatively charged anions (if they gain them).

5.2 Bonding and Lattices

The main types of bonding in minerals are ionic bonding (electrons transferred) and covalent bonding (electrons shared). Some minerals have metallic bonding or weak Van der Waals forces. Minerals form in three-dimensional lattices. The configuration of the lattices and the type of bonding within help determine mineral properties.

5.3 Mineral Groups

Minerals are grouped according to the anion part of their formula. Some common types are: oxides, sulphides, sulphates, halides, carbonates, phosphates, silicates, and native minerals.

5.4 Silicate Minerals

Silicate minerals are the most common minerals in Earth’s crust and mantle. They all have silica tetrahedra (four oxygens surrounding a single silicon atom) arranged in different structures (chains, sheets, etc).

5.5 How Minerals Form

Most minerals in the crust form from the cooling and crystallization of magma. Some form from hot water solutions, during metamorphism or weathering, or through organic processes. More rarely, minerals precipitate directly from a gas, such as at a volcanic vent.

5.6 Mineral Properties

Some of the important properties for mineral identification include hardness, cleavage/fracture, density, lustre, colour, and streak colour.

Review Questions

1. What is the electrical charge of a proton? A neutron? An electron? What are their relative masses?

2. Explain how the need for an atom’s outer shell to be filled with electrons contributes to bonding.

3. Why are helium and neon non-reactive?

4. What is the difference in the role of electrons in an ionic bond compared to a covalent bond?

5. How do cations differ from anions?

6. What chemical feature is used in the classification of minerals into groups?

7. Name the mineral group for the following minerals:

calcite biotite pyrite
gypsum galena orthoclase
hematite graphite magnetite
quartz fluorite olivine

8. What is the net charge on an unbonded silica tetrahedron?

9. What allows magnesium to substitute freely for iron in olivine?

10. How are the silica tetrahedra structured differently in pyroxene and amphibole?

11. Why is biotite called a ferromagnesian mineral, while muscovite is not?

12. What are the names and compositions of the two end-members of the plagioclase series?

13. Why does quartz have no additional cations (other than Si+4)?

14. Why is colour not necessarily a useful guide to mineral identification?

15. You have an unknown mineral that can scratch glass but cannot scratch a porcelain streak plate. What is its approximate hardness?

36

Answers to Chapter 5 Review Questions

1. Charges: proton: +1, neutron: 0, electron: -1, Masses: proton: 1, neutron: 1, electron: almost 0.

2. The element’s atomic number will determine the extent to which its outer layers are populated with electrons. If the outer shell is not quite full, the atom may gain electrons to fill them and become an anion (negative charge). If the outer shell has only a few electrons, it may lose them and become a cation (positive charge). Cations and anions attract each other to form molecules with ionic bonding.

3. Helium and neon (and the other noble gases) have complete outer shells and therefore no tendency to form ionic bonds.

4. Electrons are transferred from one atom to another to form an ionic bond. Electrons are shared between atoms to form a covalent bond.

5. An anion has a negative charge and a cation has a positive charge.

6. Minerals are classified into groups based on their anion or anion group.

7. Name the mineral group for the following minerals:

calcite CaCO3   carbonate biotite silicate pyrite FeS2 sulphide
gypsum CaSO4 sulphate galena PbS sulphide orthoclase KAlSi3O8 silicate
hematite Fe2O3 oxide graphite C native magnetite Fe3O4 oxide
quartz SiO2 silicate fluorite CaF2 halide olivine MgSiO4 silicate

8. An unbonded silica tetrahedron has one Si ion (+4 charge) and 4 oxygens (-2 charge each) so the overall charge is 4 – 8 = -4 for SiO4-4


9. Magnesium can substitute freely for iron in olivine and several other minerals because they have similar charges (+2) and similar ionic radii.

10. Pyroxene is made up of single chains of tetrahedra while amphibole is made up of double chains.

11. The two end-members of the plagioclase series are Albite (NaAlSi3O8) and Anorthite (CaAl2Si2O8)

12. In quartz each silica tetrahedron is bonded to four other tetrahedra, and because oxygens are shared at each bond the overall ratio is silicon (+4) to two oxygens (2 x -2 = -4), which is balanced.

13. Some minerals have distinctive colours, but many have a wide range of colours due to differing impurities.

14. Glass has a Mohs hardness of about 5.5 while porcelain is close to 6.5. The mineral is between these two, so it must be close to 6.

VI

Chapter 6. The Rock Cycle

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 6.1 A petrified beach near Rock Springs, Wisconsin, U. S. A. The wrinkled face of this vertical cliff displays ripples from an ancient beach. Flowing water moved sand grains to form ripples, and over time the sand was transformed into a solid sedimentary rock. The petrified beach was buried deeper and deeper, and the higher pressures and temperatures caused the sand grains to lose their individual boundaries and merge together. Thus, the sedimentary rock was transformed into a different type of rock, called a metamorphic rock. Source: Karla Panchuk (2017) CC BY-SA 4.0. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the Review Questions at the end, you should be able to:

 

 

37

6.1 What Is A Rock?

A rock is a solid mass of geological materials. Geological materials include individual mineral crystals, inorganic non-mineral solids like glass, pieces broken from other rocks, and even fossils. The geological materials in rocks may be inorganic, but they can also include organic materials such as the partially decomposed plant matter preserved in coal. A rock can be composed of only one type of geological material or mineral, but many are composed of several types. Figure 6.2 shows a rock made of three different kinds of minerals.

Rocks are grouped into three main categories based on how they form. Igneous rocks form when melted rock cools and solidifies. Sedimentary rocks form when fragments of other rocks are buried, compressed, and cemented together; or when minerals precipitate from solution, either directly or with the help of an organism. Metamorphic rocks form when heat and pressure alter a pre-existing rock. Although temperatures can be very high, metamorphism does not involve melting of the rock.

This close-up view of the igneous rock pegmatite shows black biotite crystals, colourless quartz crystals, and pink potassium feldspar crystals. Crystals are mm to cm in scale.
Figure 6.2 This close-up view of the igneous rock pegmatite shows black biotite crystals, colourless quartz crystals, and pink potassium feldspar crystals. Crystals are mm to cm in scale. Source: R. Weller/ Cochise College (2011) Permission for non-commercial educational use. (labels added) view source

38

6.2 The Rock Cycle

The rock components of the crust are slowly but constantly being changed from one form to another. The processes involved are summarized in the rock cycle (Figure 6.3). The rock cycle is driven by two forces:

  1. Earth’s internal heat, which causes material to move around in the core and mantle, driving plate tectonics.
  2. The hydrological cycle– movement of water, ice, and air at the surface. The hydrological cycle is powered by the sun.
Figure 6.3 The rock cycle describes processes that form the three types of rock: igneous, sedimentary, and metamorphic. These same processes can turn one type of rock into another. Source: Karla Panchuk (2017) CC BY-SA 4.0. Click the image for more attributions.

The rock cycle is still active on Earth because our core is hot enough to keep the mantle moving, the atmosphere is relatively thick, and there is liquid water. On some other planets or their satellites (e.g., Mercury), the rock cycle is virtually dead because the core is no longer hot enough to drive mantle convection, and there is no atmosphere or liquid water.

We can start anywhere we like to describe the rock cycle, but it’s convenient to start with magma. Magma is melted rock located within the Earth.  Rock can melt at between about 800 °C and 1300 °C, depending on the minerals in the rock, and the pressure the rock is under.  If it cools slowly within the Earth (over centuries to millions of years), magma forms intrusive igneous rocks.  If magma erupts onto the surface, we refer to it as lava.  Lava cools rapidly on Earth’s surface (within seconds to years) and forms extrusive igneous rocks (Figure 6.4).Remember the difference between intrusive and extrusive igneous rocks by recalling that INtrusive rocks form withIN the Earth, and EXtrusive rocks form when lava EXits the Earth's crust.

Figure 6.4 Lava flowing from Kīlauea Volcano, Hawai`i. Source: J. D. Griggs, U. S. Geological Survey (1985) Public Domain view source

Mountain building lifts rocks upward where they are acted upon by weathering. Weathering includes chemical processes that break rocks apart, as well as physical processes. Figure 6.5 shows the result of rocks in mountains being broken apart when water gets into cracks, freezes, and forces the cracks wider. Uplift through mountain building is how rocks once buried deep within Earth can be exposed at Earth’s surface.

Figure 6.5 Mountains being broken apart by the wedging action of ice near La Madaleta Glacier, Spain. Source: Luis Paquito (2006) CC BY-SA 2.0 view source

The weathering products — mostly small rock and mineral fragments — are eroded, transported, and then deposited as sediments. Transportation and deposition occur through the action of glaciers, streams, waves, wind, and other agents. Figure 6.6 shows transportation of fine-grained sediment particles by wind during the Great Depression in the 1930s.

Figure 6.6 Wind transports sediment in a dust storm near Okotoks, Alberta, Canada in July of 1933. Source: Glenbow Museum Archives, File Number NA-2199-1 (1933) Public Domain view source

Sediments are deposited in stream channels, lakes, deserts, and the ocean. Some depositional settings result in characteristic sedimentary structures, such as the ripples that formed when flowing water moved sand along the bottom of the South Saskatchewan River (Figure 6.7).

Figure 6.7 Sand ripples along the South Saskatchewan River, near Saskatoon SK (dog for scale). Source: Karla Panchuk (2008) CC BY-SA 4.0 view source

Unless they are re-eroded and moved along, sediments will eventually be buried by more sediments. At depths of hundreds of metres or more, the sediments become compressed, forcing particles closer together. Mineral crystals grow around and between the particles, binding them together (cementing them). The hardened cemented sediments are sedimentary rock. Figure 6.8 shows an example of an ancient sedimentary rock in which ripple structures are preserved, and visible in cross-section as wavy lines.

Figure 6.8 Ripples preserved in 1.2 Ga old sandstone. Notice the wavy lines above the coin. This is a side view of the ripples. Source: Anne Burgess (2008) CC BY-SA 2.0 view source

Rocks that are buried very deeply within the crust can reach pressures and temperatures much higher than those at which sedimentary rocks form. Existing rocks that are heated up and squeezed under those extreme conditions are transformed into metamorphic rocks (Figure 6.9). The transformation to a metamorphic rock can happen through physical changes, such as when the minerals making up an existing rock re-form into larger crystals of the same mineral. It can also happen through chemical changes, when minerals within the rock react to form new minerals.

Figure 6.9 Limestone, a sedimentary rock formed in marine waters, has been altered by metamorphism into this marble visible on Quadra Island, BC. Source: Steven Earle (2015) CC BY 4.0 view source 

39

Chapter 6 Summary

The topics covered in this chapter can be summarized as follows:

6.1 What Is a Rock?

A rock is a solid mass of geological materials. Geological materials include individual mineral crystals, inorganic non-mineral solids like glass, pieces broken from other rocks, and even fossils.

6.2 The Rock Cycle

There are three main types of rock. Igneous rocks form when melted rock cools and solidifies. Sedimentary rock forms from fragments of other rocks, or when crystals precipitate from solution. Metamorphic rocks form when existing rocks are altered by heat, pressure, and/or chemical reactions. The rock cycle summarizes the processes that contribute to transformation of rock from one type to another. The rock cycle is driven by Earth’s internal heat, and by processes happening at the surface that are driven by solar energy.

Review Questions

  1. What processes must take place to transform rocks into sediment?
  2. What processes normally take place in the transformation of sediments to sedimentary rock?
  3. What are the processes that lead to the formation of a metamorphic rock?

40

Answers to Chapter 6 Review Questions

  1. The rock must be exposed at surface. This means uplift and removal of overlying rocks and sediments is required. Once exposed, chemical and/or physical weathering can reduce the rock to smaller loose fragments (sediments). The sediments can be eroded and then transported by a variety of mechanisms.
  2. Sediments are buried beneath other sediments, where pressure compacts the sediments, forcing grains closer together. Mineral cement forms around the grains, binding them to each other and into solid rock.
  3. Rock is buried deeply in the crust and exposed to very high temperatures and pressures. Under those conditions, a new type of rock is formed when minerals undergo physical changes and chemical reactions.

VII

Chapter 7. Igneous Rocks

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 7.1 Lava lake of Mount Nyiragongo, a volcano in the Democratic Republic of Congo. Igneous rocks form when melted rock freezes. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo by Baron Reznik (2015) CC BY-NC-SA 2.0 view source. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

 

 

41

7.1 Magma and How It Forms

Igneous rocks form when melted rock cools. Melted rock originates within Earth as magma.  Magma compositions vary, but will have eight main elements in different proportions. The most abundant elements are oxygen and silicon, followed by aluminum, iron, calcium, sodium, magnesium, and potassium. These eight elements are also the most abundant in Earth’s crust (Figure 7.2).  All magmas have varying proportions of lighter elements such as hydrogen, carbon, and sulphur. Lighter elements are converted into gases like water vapour, carbon dioxide, hydrogen sulphide, and sulphur dioxide as the magma cools.

Figure 7.2 Average composition of Earth’s crust by mass. Source: Steven Earle (2016) CC BY 4.0 view source

Magma composition depends on the composition of the rocks that melted to form the magma, and on the conditions under which the melting happened. Most igneous rock in Earth’s crust comes from magmas that formed through partial melting of existing rock, either in the upper mantle or the crust. During partial melting, only some of the minerals within a rock melt. This happens because different minerals have different melting temperatures. The melt is less dense than the surrounding rock, and will percolate upward without the source rock having melted completely. The result is magma with a different composition than the original rock. Partial melting produces melt that has more silica than the original rock, because minerals higher in silica have lower melting points.

To see how partial melting works, consider the mix of materials in Figure 7.3a. It contains white blocks of candle wax, black plastic pipe, green beach glass, and pieces of aluminum wire. When the mixture is heated to 50 °C in a warm oven, the wax melts into a clear liquid (Figure 7.3b), but the other materials remain solid. This is partial melting.

Figure 7.3 An experiment to illustrate partial melting. (a) The original components are white candle wax, black plastic pipe, green beach glass, and aluminum wire. (b) After heating to 50˚C for 30 minutes only the wax has melted. (c) After heating to 120˚C for 60 minutes much of the plastic has melted and the two liquids have mixed. (d) The liquid has been poured off and allowed to cool, making a solid with a different overall composition from the original mixture. Source: Steven Earle (2015) CC BY 4.0 view source

When the mixture is heated to 120 °C, the plastic melts and mixes with the wax, but the aluminum and glass still remain solid (Figure 7.3c). This is still considered partial melting because solid materials remain. When the plastic and wax mixture is poured into a separate container and allowed to cool, the resulting solid has a very different composition from the original mixture (Figure 7.3d). The plastic and wax are analogous to more silica-rich minerals with relatively lower melting points than other minerals in the same rock.

Of course, partial melting in the real world isn’t as simple as the example in Figure 7.3. Many rocks are much more complex than the four-component system used here. Some mineral components of rocks may have similar melting temperatures, and begin to melt at the same time. The melting temperature of a mineral may change in the presence of other minerals. Also, when rocks melt, the process can take millions of years, unlike the 90 minutes required to melt the pipe and wax in the experiment in Figure 7.3.

Why Rocks Melt

The magma that is produced by partial melting is less dense than the surrounding rock. Magma from partial melting of mantle rocks rises upward through the mantle, and may pool at the base of the crust, or rise through the crust. Moving magma carries heat with it, and some of that heat is transferred to surrounding rocks. If the melting temperature of a rock is less than the temperature of the magma, the rock will begin to melt, and the composition of the magma may change to reflect a mixture of sources. But adding heat is not the only way to trigger melting.

Decompression Melting

Earth’s mantle is almost entirely solid rock, in spite of temperatures that would cause rock at Earth’s surface to melt. Mantle rock remains solid at those temperatures because the rock is under high pressure. This means that melting can be triggered without adding heat if the rock is already hot enough, and the pressure is reduced (Figure 7.4, left, white dashed boxes). Melting triggered by a reduction in pressure is called decompression melting.

Figure 7.4 Melting triggers. Left- Decompression melting occurs when rock rises or the overlying crust thins. Right- Flux-induced melting occurs when volatile compounds such as water are added. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Steven Earle (2016) CC BY 4.0 view source

Pressure is reduced when mantle rocks move upward due to convection, or rise as a plume within the mantle. Pressure is also reduced where the crust thins, such as along rift zones.

Flux-induced Melting

When a substance such as water is added to hot rocks, the melting points of the minerals within those rocks decreases. If a rock is already close to its melting point, the effect of adding water can be enough to trigger partial melting. The added water is a flux, and this type of melting is called flux-induced melting. In Figure 7.4 (right), the rock (represented by the dashed box) is not hot enough to be right of the line where dry mantle rocks melt, but it is to the right of the line where wet mantle rocks melt.

Flux-induced partial melting of rock happens in subduction zones. Minerals are transformed by chemical reactions under high pressures and temperatures, and a by-product of those transformations is water. Relatively little water is required to trigger partial melting. In laboratory studies of the conditions of partial melting in the Japanese volcanic arc, rocks with only 0.2% of their weight consisting of water melted by up to 25%.

Cooling Magma Becomes More Viscous

Viscosity refers to the ease with which a substance flows. A substance with low viscosity is runnier than a substance with high viscosity. At temperatures over 1300°C, most magma is entirely liquid because there is too much energy for the atoms to bond together. As magma loses heat to the surrounding rocks and its temperature drops, things start to change. Silicon and oxygen combine to form silica tetrahedra.  With further cooling, the tetrahedra start to link together into chains, or polymerize. These silica chains make the magma more viscous. Magma viscosity has important implications for the characteristics of volcanic eruptions.

Exercise: Making Magma Viscous

This is a quick and easy experiment that you can do at home to help you understand the properties of magma. It will only take about 15 minutes, and all you need is half a cup of water and a few tablespoons of flour.

If you’ve ever made gravy, white sauce, or roux, you’ll know how this works.

Place 1/2 cup (125 mL) of water in a saucepan over medium heat. Add 2 teaspoons (10 mL) of white flour and stir while continuing to heat the mixture until boiling. The white flour represents silica. The mixture should thicken like gravy because the gluten in the flour becomes polymerized into chains during this process.

Now add more “silica” to see how this changes the viscosity of your magma: take another 4 teaspoons (20 mL) of flour and mix it thoroughly with 4 teaspoons (20 mL) of water in a cup. Add that mixture to the rest of the water and flour in the saucepan. Stir while bringing it back up to nearly boiling temperature, and then allow it to cool. This mixture should slowly become much thicker (Figure 7.5) because there is more gluten, and more chains have formed.

Figure 7.5 Thick mixture of flour and water. Source: Steven Earle (2016) CC BY 4.0 view source

References

Kushiro, I. (2007). Origins of magmas in subduction zones: a review of experimental studies. Proceedings of the Japan Academy, Series B, Physical and Biological Sciences 83(1), 1-15. Read the paper

 

42

7.2 Crystallization of Magma

The minerals that make up igneous rocks crystallize (solidify, freeze) at a range of different temperatures. This explains why cooling magma can have some crystals within it and yet remain predominantly liquid. The sequence in which minerals crystallize from a magma as it cools is known as Bowen’s reaction series (Figure 7.6).

Figure 7.6 Bowen’s reaction series describe the sequence in which minerals form as magma cools. Source: Steven Earle (2016) CC BY 4.0 view source

How Did We Get Bowen’s Reaction Series?

Understanding how the reaction series was derived is key to understanding what it means.

Figure 7.7 Norman Bowen in his laboratory. Source: University of Chicago Photographic Archive, apf1-00841, Special Collections Research Center, University of Chicago Library. Click the image for source and terms of use.

Norman Levi Bowen (Figure 7.7) was born in Kingston Ontario. He studied geology at Queen’s University and then at Massachusetts Institute of Technology. In 1912 he joined the Carnegie Institution in Washington, D.C., where he carried out ground-breaking experiments into how magma cools.

Working mostly with mafic magmas (magmas rich in iron and magnesium), he determined the order of crystallization of minerals as the temperature drops. First, he melted the rock completely in a specially made kiln. Then he allowed it to cool slowly to a specific temperature before quenching (cooling it quickly) so that no new minerals could form. The rocks that formed were studied under the microscope and analyzed chemically. This was done over and over, each time allowing the magma to cool to a lower temperature before quenching.

The result of these experiments was the reaction series which, even a century later, is still an important basis for our understanding of igneous rocks.

Discontinuous and Continuous Series

Bowen’s reaction series (Figure 7.6) has two pathways for minerals to form as magma cools: on the left is the discontinuous series. This refers to the fact that one mineral is transformed into a different mineral through chemical reactions. On the right is the continuous series, where plagioclase feldspar goes from being rich in calcium to being rich in sodium.

Continuous Series

At about the point where pyroxene begins to crystallize, plagioclase feldspar also begins to crystallize. At that temperature, the plagioclase is calcium-rich (toward the anorthite end-member). As the temperature drops, and providing that there is sodium left in the magma, the plagioclase that forms is a more sodium-rich variety (toward the albite end-member). The series is continuous because the mineral is always plagioclase feldspar, but the series involves a transition from calcium-rich to sodium-rich.

When cooling happens relatively quickly, instead of getting crystals which are of uniform composition, individual plagioclase crystals can be zoned from calcium-rich in the centre to more sodium-rich around the outside (Figure 7.8). This occurs because calcium-rich early-forming plagioclase crystals become coated with progressively more sodium-rich plagioclase as the magma cools.

Figure 7.8 Plagioclase crystal exhibiting compositional zones. Source: Akademia Górniczo-Hutnicza w Krakowie Otwartych Zasobów Edukacyjnych (n.d.) CC BY-NC-SA view source/ view context

Discontinuous Series

Olivine begins to form at just below 1300°C, but as the temperature drops, olivine becomes unstable. The early-forming olivine crystals react with silica in the remaining liquid and are converted into pyroxene, something like this:

Mg2SiO4 + SiO2 goes to 2MgSiO3

As long as there is silica remaining and the rate of cooling is slow, this process continues down the discontinuous branch: olivine reacts to form pyroxene, and the pyroxene reacts to form amphibole. Under the right conditions amphibole will form to biotite. Finally, if the magma is quite silica-rich to begin with, there will still be some left at around 750 °C to 800 °C, and from this last magma, potassium feldspar, quartz, and maybe muscovite mica will form.

Notice that the sequence of minerals that form goes from isolated tetrahedra (olivine) toward increasingly complex arrangements of silica tetrahedra. Pyroxene consists of single chains, amphibole has double chains, mica has sheets of tetrahedra, and potassium feldspar and quartz at the bottom of the series have tetrahedra connected to each other in three dimensions.

If the magma cools enough, the first minerals to form will be completely used up in later chemical reactions.  This is why igneous rocks do not normally have both olivine (at the top of the series) and quartz (at the bottom). Exceptions can occur when rocks that crystallized early in the series come into contact with magmas representing compositions later in the series, such as with the dark green olivine-rich xenoliths included within the quartz- and feldspar-rich rock in Figure 7.9. The dark line around the xenoliths is amphibole, which formed as the olivine reacted with the melt. In some of the smaller xenoliths within this boulder, the olivine has been completely transformed into amphibole.

Figure 7.9 Boulder with olivine-rich xenoliths surrounded by silica-rich rock. Black rims on the xenoliths are where the olivine has reacted with the silica-rich melt, forming amphibole. Right- Enlarged view of the amphibole reaction rim. Source: Karla Panchuk (2018) CC BY 4.0

Magma Composition: Mafic, Intermediate, and Felsic

The composition of the original magma determines how far the reaction process can continue before all of the magma is used up. In other words, it determines which minerals will form. Magma compositions are reported in terms of the fraction of mass of oxides (e.g., Al2O3 rather than just Al; Figure 7.10). On average, mafic"Mafic" combines the words MAgnesium and FerrIC (containing iron). magma (Figure 7.10, left) is approximately half SiO2 by mass, and more than 25% iron, magnesium, and calcium oxides by mass. Average felsic"Felsic" combines the words FELdspar and SIliCa. magmas (Figure 7.10, right) are closer to 75% SiO2 by mass, and have approximately 5% iron, magnesium, and calcium oxides. Sodium and potassium oxides account for approximately 10% of felsic magmas by mass, but only 5% of mafic magmas. Magmas that fall between mafic and felsic magmas have an intermediate composition (Figure 7.10, centre).

Figure 7.10 Chemical compositions of typical mafic, intermediate, and felsic magmas. Source: Karla Panchuk (2018) CC BY 4.0 modified after Steven Earle (2016) CC BY 4.0 view source

 

Exercise: Mafic, Intermediate, or Felsic?

The proportions of the main chemical components of felsic, intermediate, and mafic magmas are listed in the table below. (The values are similar to those shown in Figure 7.10).

Oxide Felsic Magma Intermediate Magma Mafic Magma
SiO2 65-75% 55-65% 45-55%
Al2O3 12-16% 14-18% 14-18%
FeO 2-4% 4-8% 8-12%
CaO 1-4% 4-7% 7-11%
MgO 0-3% 2-6% 5-9%
Na2O 2-6% 3-7% 1-3%
K2O 3-5% 2-4% 0.5-3%

Chemical composition by mass for four rock samples are shown in the following table. Compare these with those in the table above to determine whether each of these samples is felsic, intermediate, or mafic.

SiO2 Al2O3 FeO CaO MgO Na2O K2O Type?
55% 17% 5% 6% 3% 4% 3%
74% 14% 3% 3% 0.5% 5% 4%
47% 14% 8% 10% 8% 1% 2%
65% 14% 4% 5% 4% 3% 3%

 

What Determines the Composition of Magma?

Why Is There No Ultramafic Magma Anymore?

Refer back to Bowen’s reaction series in Figure 7.6. Notice that on the far right-hand side of the diagram under “Rock Types,” mafic, intermediate, and felsic magma compositions are listed.  At the very top of the list is ultramafic. Ultramafic rocks have higher MgO than mafic rocks, and even less SiO2.
The vast majority of silicate rocks in Earth’s lithosphere are ultramafic rocks, because the mantle is composed of ultramafic rock. However, ultramafic magma is not encountered in modern volcanic environments, and ultramafic rocks are relatively rare at Earth’s surface. The reason is that although Earth was once hot enough to have ultramafic magma, it is no longer hot enough to melt ultramafic rocks. Ultramafic volcanic rocks—called komatiites—do exist, but with two notable exceptions, the youngest of these is 2 billion years old.The komatiites of the Song Da zone in northwestern Vietnam are 270 million years old, and those on Gorgona Island, Columbia are 89 million years old. Exactly how they formed is still a bit of a mystery. See Table 1 of arXiv:physics/0512118v2 [physics.geo-ph] for a compilation of komatiite ages with references.

Partial Melting Makes Magma That Is Richer in Silica

In partial melting, some components of a mixture melt before others do.  In the case of mafic magma, it is produced when ultramafic rocks undergo partial melting. In general, silicate minerals with more silica will melt before those with less silica.  This means the partial melt will have more silica than the rock as a whole.

Fractional Crystallization Also Makes Magma Richer In Silica

A number of processes that take place within a magma chamber can affect the types of rocks that form once magma cools and crystallizes. If the magma has a low viscosity— which is likely if the magma is mafic—the crystals that form early, such as olivine (Figure 7.11a), may slowly settle toward the bottom of the magma chamber (Figure 7.11b). This process is called fractional crystallization.

Figure 7.11 Formation of a zoned magma chamber. a- Olivine crystals form. b- Olivine crystals settle to the base of the magma chamber, leaving the upper part of the chamber richer in silica. c- Olivine crystals remelt, making magma at the base of the chamber more mafic. Source: Steven Earle (2015) CC BY 4.0 view source

The formation of olivine removes iron- and magnesium-rich components, leaving the overall composition of the magma near the top of the magma chamber more felsic. The crystals that settle might either form an olivine-rich layer near the bottom of the magma chamber. Or, because the lower part of the magma chamber is likely to be hotter than the upper part, the crystals might remelt. If remelting happens, crystal settling will make the magma at the bottom of the chamber more mafic than it was to begin with (Figure 7.11c).

Magma Composition Also Changes When Other Rocks Are Melted And Mixed In

Magma chambers aren’t isolated from their surroundings.  If the rock in which the magma chamber is located (called the country rock) is more felsic than the magma, the country rock may also melt, adding to the magma already in the magma chamber (Figure 7.12).  Sometimes magma carries fragments of unmelted rock, called xenoliths, within it.  Melting of xenoliths can also alter the composition of magma, as can re-melting of crystals that have settled out of the magma.

Figure 7.12 The composition of magma in a magma chamber is affected by fractional crystallization within the magma chamber, but it can also be affected by partial melting of the rock surrounding the magma chamber, melting of xenoliths within the magma, or re-melting of crystals that have settled to the bottom of the magma chamber. Source: Steven Earle (2015) CC BY 4.0 view source

43

7.3 Classification of Igneous Rocks

Classification By Mineral Abundance

Igneous rocks can be divided into four categories based on their chemical composition: felsic, intermediate, mafic, and ultramafic. The diagram of Bowen’s reaction series (Figure 7.6) shows that differences in chemical composition correspond to differences in the types of minerals within an igneous rock.  Igneous rocks are given names based on the proportion of different minerals they contain.  Figure 7.13 is a diagram with the minerals from Bowen’s reaction series, and is used to decide which name to give an igneous rock.

Figure 7.13 Classification diagram for igneous rocks. Igneous rocks are classified according to the relative abundances of minerals they contain. A given rock is represented by a vertical line in the diagram. In the mafic field, the arrows represent a rock containing 48% pyroxene and 52% plagioclase feldspar. The name an igneous rock gets depends not only on composition, but on whether it is intrusive or extrusive. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0, modified after Steven Earle (2015) CC BY 4.0 view source and others, with photos by R. Weller/Cochise College. Click the image for links to photos and notes on image construction. High-resolution version.

To see how Figure 7.13 works, first notice the scale in percent along the vertical axis.  The interval between each tick mark represents 10% of the minerals within a rock by volume.  An igneous rock can be represented as a vertical line drawn through the diagram, and the vertical scale used to break down the proportion of each mineral it contains.  For example, the arrows in the mafic field of the diagram represent a rock containing 48% pyroxene and 52% plagioclase feldspar. An igneous rock at the boundary between the mafic and ultramafic fields (marked with a vertical dashed line) would have approximately 20% olivine, 50% pyroxene, and 30% Ca-rich plagioclase feldspar by volume.

Classification By Grain Size

The name an igneous rock gets also depends on whether it cools within Earth (an intrusive or plutonic igneous rock), or whether it cools on the Earth’s surface after erupting from a volcano (an extrusive or volcanic igneous rock). For example, a felsic intrusive rock is called granite, whereas a felsic extrusive rock is called rhyolite. Granite and rhyolite have the same mineral composition, but their grain size gives each a distinct appearance.

The key difference between intrusive and extrusive igneous rocks—the size of crystals making them up—is related to how rapidly melted rock cools. The longer melted rock has to cool, the larger the crystals within it can become.  Magma cools much slower within Earth than on Earth’s surface because magma within Earth is insulated by surrounding rock.  Notice that in Figure 7.13, the intrusive rocks have crystals large enough that you can see individual crystals—either by identifying their boundaries, or seeing light reflecting from a crystal face.  A rock with individual crystals that are visible to the unaided eye has a phaneritic or coarse-grained texture. The extrusive rocks in the second row have much smaller crystals.  The crystals are so small that individual crystals cannot be distinguished, and the rock looks like a dull mass. A rock with crystals that are too small to see with the unaided eye has an aphanitic or fine-grained texture.  Table 7.1 summarizes the key differences between intrusive and extrusive igneous rocks.

Table 7.1 Comparison of Intrusive and Extrusive Igneous Rocks
Magma cools within Earth Lava cools on Earth’s surface
Terminology Intrusive/ plutonic Extrusive/ volcanic
Cooling rate Slow: surrounding rocks insulate the magma chamber. Rapid: heat is exchanged with the atmosphere.
Texture Phaneritic (coarse-grained): individual crystals are large enough to see without magnification. Aphanitic (fine-grained): crystals are too small to see without magnification.

What this means is that two igneous rocks comprised of exactly the same minerals, and in the exactly the same proportions, can have different names.  A rock of intermediate composition is diorite if it is course-grained, and andesite if it is fine-grained.  A mafic rock is gabbro if it is course-grained, and basalt if fine-grained. The course-grained version of an ultramafic rock is peridotite, and the fine-grained version is komatiite. It makes sense to use different names because rocks of different grain sizes form in different ways and in different geological settings.

Does This Mean an Igneous Rock Can Only Have One Grain Size?

No. Something interesting happens when there is a change in the rate at which melted rock is cooling.  If magma is cooling in a magma chamber, some minerals will begin to crystallize before others do.  If cooling is slow enough, those crystals can become quite large.

Now imagine the magma is suddenly heaved out of the magma chamber and erupted from a volcano.  The larger crystals will flow out with the lava. The lava will then cool rapidly, and the larger crystals will be surrounded by much smaller ones.  An igneous rock with crystals of distinctly different size (Figure 7.14) is said to have a porphyritic texture, or might be referred to as a porphyry.  The larger crystals are called phenocrysts, and the smaller ones are referred to as the groundmass.

Figure 7.14 Porphyritic rhyolite with quartz and potassium feldspar phenocrysts within a dark groundmass. Porphyritic texture (when different crystal sizes are present) is an indication that melted rock did not cool at a constant rate. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo by R. Weller/Cochise College (2011) view source

 

Exercise: Which Mineral Will the Phenocryst Be?

As a magma cools below 1300°C, minerals start to crystallize within it. If the magma is then erupted, the rest of the liquid will cool quickly to form a porphyritic texture. The rock will have some relatively large crystals (phenocrysts) of the minerals that crystallized early, and the rest will be very fine-grained or even glassy. Using the diagram shown here, predict what phenocrysts might be present where the magma cooled as far as line a. Which would be present where magma cooled to line b?

Figure 7.15 Bowen’s reaction series. Source: Steven Earle (2015) CC BY 4.0 view source

 

Classifying Igneous Rocks According to the Proportion of Dark Minerals

If you unsure of which minerals are present in an intrusive igneous rock, there is a quick way to approximate the composition of that rock.  In general, igneous rocks have an increasing proportion of dark minerals as they become more mafic (Figure 7.16).

Figure 7.16 Simplified igneous rock classification according to the proportion of light and dark (or ferromagnesian) minerals. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

The dark-coloured minerals are those higher in iron and magnesium (e.g., olivine, pyroxene, amphibole, biotite), and for that reason they are sometimes referred to collectively as ferromagnesian minerals. By estimating the proportion of light minerals to dark minerals in a sample, it is possible to place that sample in Figure 7.16.  Graphical scales are used to help visualize the proportions of light and dark minerals (Figure 7.17).

Figure 7.17 A guide for estimating the proportion of dark minerals in an igneous rock. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) view source

It is important to note that estimating the proportion of dark minerals is only approximate as a means for identifying igneous rocks. One problem is that plagioclase feldspar is light-coloured when it is sodium-rich, but can appear darker if it is calcium-rich. Plagioclase feldspar is not ferromagnesian, so it falls in the non-ferromagnesian (light minerals) region in Figure 7.16 even when it has a darker colour.

Exercise: Classifying Igneous Rocks by the Proportion of Dark Minerals

The four igneous rocks shown below have differing proportions of ferromagnesian silicates (dark minerals). Estimate the proportion of dark minerals using the guide in Figure 7.17, and then use Figure 7.16 to determine the likely rock name for each one.

Figure 7.18 Identify these rocks by estimating the proportion of dark minerals in each sample. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

 

 

Classifying Igneous Rocks When Individual Crystals Are Not Visible

The method of estimating the percentage of minerals works well for phaneritic igneous rocks, in which individual crystals are visible with little to no magnification. If an igneous rock is porphyritic but otherwise aphanitic (e.g., Figure 7.14), the minerals present as phenocrysts give clues to the identity of the rock. However, there are cases where mineral composition cannot be determined by looking at visible crystals. These include volcanic rocks without phenocrysts, and glassy igneous rocks.

Volcanic Rocks Without Phenocrysts

In the absence of visible crystals or phenocrysts, volcanic rocks are be classified on the basis of colour and other textural features. As you may have noticed in Figure 7.13, the colour of volcanic rocks goes from light to dark as the composition goes from felsic to mafic. Rhyolite is often a tan or pinkish colour, andesite is often grey, and basalt ranges from brown to dark green to black (Figure 7.19).

Figure 7.19 In volcanic igneous rocks, individual crystals are not visible. Colours change from light to dark as the composition of the rocks go from felsic to mafic. Vesicles and amygdules are common characteristics of basalt. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for links to photos.

Basalt often shows textural features related to lava freezing around gas bubbles. When magma is underground, pressure keeps gases dissolved, but once magma has erupted, the pressure is much lower. Gases dissolved in the lava are released, and bubbles can develop. When lava freezes around the bubbles, vesicles are formed (circular inset in 7.19). If the vesicles are later filled by other minerals, the filled vesicles are called amygdules (box inset in Figure 7.19).

Glassy Volcanic Rocks

Crystal size is a function of cooling rate. The faster magma or lava cools, the smaller the crystals it contains. It is possible for lava to cool so rapidly that no crystals can form. The result is called volcanic glass. Volcanic glass can be smooth like obsidian or vesicular like scoria (mafic) and pumice (felsic; Figure 7.20). Pumice can float on water because of its low-density felsic composition and enclosed vesicles.

Figure 7.20 Glassy volcanic rocks. Obsidian has a glassy lustre, but scoria and pumice are highly vesicular. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/Cochise College. Click the image for links to photos.

44

7.4 Intrusive Igneous Rocks

In most cases, a body of hot magma is less dense than the rock surrounding it, so it has a tendency to creep upward toward the surface. It does so in a few different ways:

When magma forces itself into cracks, breaks off pieces of rock, and then envelops them, this is called stoping.  The resulting fragments are xenolithsFrom the Greek words xenos, meaning "foreigner" or "stranger," and lithos for "stone.". Xenoliths may appear as dark patches within a rock (Figure 7.21).

Figure 7.21 Xenoliths of mafic rock in granite, Victoria, B.C. The fragments of dark rock have been broken off and incorporated into the light-coloured granite. Source: Steven Earle (2015) CC BY 4.0 view source

Some of the magma may reach the surface, resulting in volcanic eruptions, but most cools within the crust. The resulting body of rock is called a pluton.After Pluto was demoted from planet status, astronomers tried to come up with a name for objects like Pluto. For a while they considered "pluton" however geologists rightly objected that they had first claim on the word. In the end the International Astronomical Union settled on "dwarf planet" instead. Plutons can have different shapes and different relationships with the surrounding country rock (Figure 7.22). These characteristics determine what name the pluton is given.

Large, irregularly shaped plutons are called stocks or batholiths, depending upon their size. Tabular plutons are called dikes if they cut across existing structures, and sills if they are parallel to existing structures. Laccoliths are like sills, except they have caused the overlying rocks to bulge upward. Pipes are cylindrical conduits.

Figure 7.22 Plutons can have a variety of shapes, and be positioned in a variety of ways relative to the surrounding rocks. They are named according to these characteristics. Source: Karla Panchuk (2018) CC BY 4.0

Types of Plutons

Stocks and Batholiths

Large irregular-shaped plutons are called either stocks or batholiths, depending on their area. If an irregularly shaped body has an area greater than 100 km2, then it’s a batholith, otherwise it’s a stock. Note that our knowledge of the size of a body can be limited to what we see at the surface. A body with an area of less than 100 km exposed at the surface might in fact be much larger at depth. It might be classified as a stock initially, until someone is able to map out its true extent.

Batholiths are typically formed when a number of stocks coalesce beneath the surface to create one large body. One of the largest batholiths in the world is the Coast Range Plutonic ComplexAlso referred to as the Coast Range Batholith, which extends all the way from the Vancouver region to southeastern Alaska (Figure 7.23).

Figure 7.23 The Coast Range Plutonic Complex (also called the Coast Range Batholith) is the largest in the world. It is part of a chain of batholiths along the western coast of North America. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Bally (1989).

Tabular Intrusions

Tabular (sheet-like) plutons are classified according to whether or not they are concordant with (parallel to) existing layering (e.g., sedimentary bedding or metamorphic foliationSedimentary bedding refers to the layers in which sedimentary rocks form. Metamorphic foliation refers to the way minerals or other elements in a rock are aligned as a result of being deformed by heat and pressure. Bedding and foliation will be discussed in more detail in later chapters.) in the country rock. A sill is concordant with existing layering, and a dikeAlso spelled dyke. is discordant. If the country rock has no bedding or foliation, then any tabular body within it is a dike. Note that the sill-versus-dike designation is not determined simply by the orientation of the feature. A dike could be horizontal and a sill could be vertical- it all depends on the orientation of features in the surrounding rocks.

A laccolith is a sill-like body that has expanded upward by deforming the overlying rock. If a sill forms, but magma pools and sags downward, it creates a lopolith.

Pipes

A pipe, as the name suggests, is a cylindrical body with a circular, elliptical, or even irregular cross-section, that serves as a conduit (or pipeline) for the movement of magma from one location to another. Pipes may feed volcanoes, but pipes can also connect plutons.

Chilled Margins

As discussed already, plutons can interact with the rocks into which they are intruded. Partial melting of the country rock may occur, or stoping may form xenoliths. The heat from magma can even cause causing mineralogical and textural changes in country rock. However, country rock can also affect the magma.

The most obvious effect that country rock can have on magma is a chilled margin along the edges of the pluton (Figure 7.24). The country rock is much cooler than the magma, so magma that comes into contact with the country rock cools faster than magma toward the interior of the pluton. Rapid cooling leads to smaller crystals, so the texture along the edges of the pluton is different from that of the interior of the pluton, and the colour may be darker.

Figure 7.24 A mafic dike with chilled margins within basalt at Nanoose, B.C. The coin is 24 mm in diameter. The dike is about 25 cm across and the chilled margins are 2 cm wide. Source: Steven Earle (2015) CC BY 4.0 view source

Exercise: Pluton Problems

The diagram below is a cross-section through part of the crust showing a variety of intrusive igneous rocks. Indicate whether each of the plutons labelled a to e on the diagram below is a dike, a sill, a stock, or a batholith. (Note the trees for scale.)

Figure 7.25 A variety of igneous intrusions. Source: Steven Earle (2015) CC BY 4.0 view source

 References

Bally, A. W. (1989). Plate 10. Selected distribution maps, rate of accumulation maps, and lithofacies maps—Phanerozoic, North America. In A. W. Bally & A. R. Palmer (Eds.), The Geology of North America—An Overview: Volume A. Boulder: Geological Society of America.

45

Chapter 7 Summary

The topics covered in this chapter can be summarized as follows:

7.1 Magma and How It Forms

Magma is molten rock, and in most cases, it forms from partial melting of existing rock. The chemistry of magma depends on the source rock that is melting, as well as the degree of partial melting that occurs. Magma forms by decompression melting, flux-induced melting, and heat transfer. Magmas range in composition from ultramafic to felsic. Mafic rocks are rich in iron, magnesium, and calcium, and contain approximately 50% silica. Felsic rocks are richer in silica (~70%) and have lower levels of iron, magnesium, and calcium, and higher levels of sodium and potassium than mafic rocks.

7.2 Crystallization of Magma

As a body of magma starts to cool, the first process to take place is the polymerization of silica tetrahedra into chains. This increases the magma’s viscosity (makes it thicker) and because felsic magmas have more silica than mafic magmas, they tend to be more viscous. Bowen’s reaction series allows us to predict the order of crystallization of magma as it cools. Magma can be modified by fractional crystallization (separation of early-forming crystals), by mixing in material from the surrounding rocks by partial melting, and by mixing with magmas of differing chemistry.

7.3 Classification of Igneous Rocks

Igneous rocks are classified based on their mineral composition and texture. Felsic igneous rocks have less than 20% dark minerals (ferromagnesian silicates including amphibole and/or biotite) with varying amounts of quartz, both potassium and plagioclase feldspars, and sometimes muscovite. Mafic igneous rocks have more than 50% dark minerals (primarily pyroxene) plus plagioclase feldspar. Most intrusive igneous rocks are phaneritic (individual crystals are visible unmagnified). If there were two stages of cooling (slow then fast), the texture may be porphyritic (large crystals in a matrix of smaller crystals).

7.4 Intrusive Igneous Bodies

Magma intrudes into country rock by pushing it aside or melting through it. Intrusive igneous bodies tend to be irregular (stocks and batholiths), tabular (dikes and sills), or pipe-like. Batholiths have areas of 100 km2 or greater, while stocks are smaller. Sills are parallel to existing layering in the country rock, while dikes cut across layering. A pluton that intruded into cold rock is likely to have a chilled margin.

Review Questions

1. What is the significance of the term reaction in Bowen’s reaction series?

2. Why is it common for plagioclase crystals to be zoned from relatively calcium-rich in the middle to more sodium-rich toward the edge?

3. What must happen within a magma chamber for fractional crystallization to take place?

4. Explain the difference between aphanitic and phaneritic textures.

5. Name the following rocks:
(a) An extrusive rock with 40% Ca-rich plagioclase and 60% pyroxene
(b) An intrusive rock with 65% plagioclase, 25% amphibole, and 10% pyroxene
(c) An intrusive rock with 25% quartz, 20% potassium feldspar, 50% plagioclase feldspar, and minor amounts of biotite

6. What is the difference between a concordant tabular intrusion and a discordant tabular intrusion?

7. Why do dikes commonly have fine-grained margins?

8. What is the difference between a batholith and a stock?

9. Describe two ways in which batholiths intrude into existing rock.

10. Why is compositional layering a common feature of mafic plutons but not of felsic plutons?

46

Answers to Chapter 7 Review Questions

1. As the temperature decreases minerals that formed early (e.g., olivine) may react with the remaining magma to form new minerals (e.g., pyroxene).

2. Calcium-rich plagioclase forms early on in the cooling process of a magma, but as the temperature drops, a more sodium-rich variety forms around the existing crystals.

3. Some minerals must begin to form while melt is still present. Early-forming minerals, which are typically quite dense (e.g., olivine), will sink to the bottom of the magma chamber if the magma is not too viscous, thus becoming separated from the rest of the magma. The composition of the remaining magma will be more felsic than before.

4. Phaneritic texture means that individual crystals can be distinguished by the naked eye. Aphanitic texture means that individual crystals cannot be distinguished without a microscope. The dividing line is somewhere between 0.1 and 1 mm, depending on the minerals.

5. a) basalt; b) diorite; c) granite

6. A concordant body (a sill) is parallel to any pre-existing layering (e.g., bedding or foliation) in the country rock is. A discordant body (a dike) cuts across any pre-existing layering, or has intruded in country rock without layering (e.g., in granite).

7. When the hot magma intrudes into cold country rock, the margins cool quickly and small crystals form, whereas magma that is not in contact with the cool country rock will cool more slowly, and larger crystals will form. The chilled margin is the band of small crystals along the edge of the dike.

8. A batholith has an area of 100 km2 or greater, whereas a stock is smaller.

9. Batholiths (or stocks) intrude into existing rock by (a) melting through the country rock, or (b) causing the country rock to break and fall into the magma (stoping), or (c) pushing the country rock aside.

10. Compositional layering forms when early-crystallizing minerals sink toward the bottom of a magma chamber. This can only happen in non-viscous magma. Mafic magma is typically much less viscous than felsic magma.

VIII

Chapter 8. Weathering, Sediment, and Soil

Adapted by Karla Panchuk from Physical Geology by Steven Earle

 

Figure 8.1 The Hoodoos, near Drumheller, Alberta, have formed from the differential weathering (weaker rock weathering faster than stronger rock) of sedimentary rock. Source: Steven Earle (2015) CC BY 4.0 view source

 Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

What Is Weathering?

Weathering occurs when rock is exposed to the “weather” — to the forces and conditions that exist at Earth’s surface. Rocks that form deep within Earth experience relatively constant temperature, high pressure, have no contact with the atmosphere, and little or no interaction with moving water. Once overlying layers are eroded away and a rock is exposed at the surface, conditions change dramatically. Temperatures vary widely, and pressure is much lower. Reactive gases like oxygen and carbon dioxide are plentiful, and in many climates, water is abundant.

Weathering can be characterized as mechanical (or physical), and chemical. In mechanical weathering, physical processes break rock into smaller pieces. In chemical weathering, chemical reactions change minerals into forms that are less affected by chemical reactions that occur at Earth’s surface. Mechanical and chemical weathering reinforce each other, because mechanical weathering provides new fresh surfaces for attack by chemical processes, and chemical weathering weakens the rock so that it is more susceptible to mechanical weathering. Together, these processes create the particles and ions that can eventually become sedimentary rock.  They also create soil, which is necessary for our existence on Earth.

 

47

8.1 Mechanical Weathering

Intrusive igneous rocks form at depths of 100s of metres to 10s of kilometres. Most metamorphic rocks are formed at depths of kilometres to 10s of kilometres. Sediments are turned into sedimentary rocks only when they are buried by other sediments to depths in excess of several 100s of metres. Weathering cannot happen until these rocks are revealed at Earth’s surface by uplift and the erosion of overlying materials. Once the rock is exposed at the surface as an outcrop, weathering begins.

The agents of mechanical weathering can be broadly classified into two groups: those that cause the outer layers of a rock to expand, and those that act like wedges to force the rock apart.

Mechanical Weathering By Expansion

Some processes at Earth’s surface can cause a thin outer layer of a rock to expand. Deeper than the thin outer layer, the rock does not expand. The difference is accommodated by a crack developing between the outer and inner layers, breaking the outer layer off in slabs (Figures 8.2 and 8.3). When layers break off a rock in slabs or sheets, it is referred to as exfoliation.

Figure 8.2 Close-up view of exfoliation of a granite dome in the Enchanted Rock State Natural Area, Texas, USA. Source: Wing-Chi Poon (2005) CC BY-SA 2.5 view source
Figure 8.3 View of exfoliation at a distance (centre of image) in granite exposed on the west side of the Coquihalla Highway north of Hope, B.C. Source: Steven Earle (2015) CC BY view source

Granite tends to exfoliate parallel to the exposed surface because it does not have planes of weakness to determine how it breaks. In contrast, sedimentary rocks tend to exfoliate along the contacts between different sedimentary layers, and metamorphic rocks tend to exfoliate parallel to aligned minerals.

Reasons Rocks Expand

A rock within the Earth has pressure exerted upon it by other rocks sitting above it. This is called confining pressure. When the overlying mass is removed by weathering, the confining pressure decreases, allowing the rock to expand. The cracking that results is sometimes called pressure-release cracking.

Heating a rock can also cause it to expand. If the rock is heated rapidly, as during a wildfire, cracks can form. If it goes through large daily temperature swings (e.g., in the desert where it is very hot during the day but cold at night), cracking can also eventually result as the rock is weakened.

Mechanical Weathering by Wedging

In wedging, a pre-existing crack in a rock is made larger by forcing it open.

Frost Wedging

Frost wedging (or ice wedging) happens when water seeps into cracks, then expands upon freezing. The expansion enlarges the cracks (Figure 8.4). The effectiveness of frost wedging depends on how often freezing and thawing occur.  Frost wedging won’t be as important in warm areas where freezing is infrequent, in very cold areas where thawing is infrequent, or in very dry areas, where there is little water to seep into cracks.

Figure 8.4 A rock broken by ice wedging sits in a stream in Mount Revelstoke National Park, Canada. Rocks break apart when ice expands in pre-existing cracks. Source: Karla Panchuk (2018) CC BY 4.0.

Frost wedging is most effective in Canada’s climate, where for at least part of the year temperatures oscillate between warm and freezing. In many parts of Canada, the temperature swings between freezing at night and thawing in the day tens to hundreds of times a year. Even in warm coastal areas of southern British Columbia, freezing and thawing transitions are common at higher elevations. A common feature in areas of effective frost wedging is a talus slope — a fan-shaped deposit of fragments removed by frost wedging from the steep rocky slopes above (Figure 8.5).

Figure 8.5 An area with very effective frost wedging near Keremeos, BC. The fragments that were wedged away from the cliffs above have accumulated in a talus deposit at the base of the slope. The rocks in this area are variable in colour, which is reflected in the colours of the talus. Source: Steven Earle (2015) CC BY 4.0 view source

Salt Wedging

Salt wedging happens when saltwater seeps into rocks and then evaporates on a hot sunny day. Salt crystals grow within cracks and pores in the rock, and the growth of these crystals can push grains apart, causing the rock to weaken and break. There are many examples of this on the rocky shorelines of Vancouver Island and the Gulf Islands, where sandstone outcrops are common and salty seawater is readily available (Figure 8.6). The honeycomb structure of rounded holes, called tafoni, is related to the original roughness of the surface. Low spots collect salt water, causing the effect to be accentuated around existing holes.

Figure 8.6 Tafoni (Honeycomb weathering) in sandstone on Gabriola Island, British Columbia. The holes are caused by crystallization of salt within rock pores. Source: Steven Earle (2015) CC BY 4.0 view source

Plant and Animal Activity

The effects of plants are significant in mechanical weathering. Roots can force their way into even the tiniest cracks. They exert tremendous pressure on the rocks as they grow, widening the cracks and breaking the rock. This is called root wedging (Figure 8.7).

Figure 8.7 Root wedging along a quarry wall. Left: Rocks beneath the thick red beds have been split into sheets by tree roots. Right: A closer examination reveals that tree roots are working into vertical cracks as well. Source: Karla Panchuk (2018) CC BY 4.0

Although most animals do not normally burrow through solid rock, they can excavate and remove huge volumes of soil, and thus expose the rock to weathering by other mechanisms. Humans modify vast tracts of land by excavation, and have a profound effect on accelerating mechanical weathering.

Exercise: Mechanical Weathering

What mechanical weathering processes do you think take place on this mountain in British Columbia?

Figure 8.8 Granite at the top of Siy’ám’ Smánit (also known as Stawamus Chief Mountain), near Squamish, British Columbia. Source: Steven Earle (2015) CC BY 4.0 view source

 Erosion

Mechanical weathering is greatly facilitated by erosion.  Erosion is the removal of weathering products, such as fragments of rock. This exposes more rock to weathering, accelerating the process. A good example of weathering and erosion working together is the talus shown in Figure 8.5. The rock fragments forming the talus piles were broken off the steep rock faces at the top of the cliff by ice wedging, and then removed by gravity.

Gravity does not always work alone to remove weathering products. Other agents of erosion include water in streams, ice in glaciers, and waves on coasts.

 

48

8.2 Chemical Weathering

Chemical weathering results from chemical changes to minerals that become unstable when they are exposed to surface conditions. The kinds of changes that take place are specific to the mineral and the environmental conditions. Some minerals, like quartz, are virtually unaffected by chemical weathering. Others, like feldspar, are easily altered.

Types of Chemical Weathering Reactions

Dissolution

Dissolution reactions produce ions, but no minerals, and are reversible if the solvent is removed. A household example would be dissolving a teaspoon of table salt (the mineral halite) in a glass of water. The halite will separate into Na+ and Cl ions. If the water in the glass is allowed to evaporate, there will not be enough water molecules to hold the Na+ and Cl ions apart, and the ions will come together again to form halite. Gypsum and anhydrite are other minerals that will dissolve in water alone.

Other minerals, such as calcite, will dissolve in acidic water. Acidic water is common in nature, because carbon dioxide (CO2) in the atmosphere reacts with water vapour in the atmosphere, and with water on land and in the oceans to produce carbonic acid (Figure 8.9).

Figure 8.9 Calcite weathering by dissolution. Top: Carbon dioxide reacts with water to make acid. Bottom: Acid reacts with calcite and produces ions. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Modified after What-When-How. Molecules from JMSE Molecular Editor, Bienfait and Ertl (2013), with permission for CC BY-NC-SA use.

While rainwater and atmospheric CO2 can combine to create carbonic acid, the amount of CO2 in the air is enough to make only very weak carbonic acid. In contrast, biological processes acting in soil can result in a much higher concentration of CO2 within soil, as well as adding organic acids. Water that percolates through the soil can become significantly more acidic.

Calcite is a major component of the sedimentary rock called limestone (typically more than 95%). In the presence of acidic groundwater, limestone can dissolve underground.  Over time the dissolution can remove enough of the calcite to form caves.

If dissolution of limestone or other materials removes enough rock to undermine support near the surface, the surface may collapse, creating a sinkhole such as the one in Figure 8.10, downstream of the Mosul Dam in Iraq.

A large, deep, circular hole.
Figure 8.10 Sinkhole downstream of the Mosul Dam in Iraq. The sinkhole is a result of dissolution of gypsum and anhydrite layers. Source: U. S. Army Corps of Engineers (2007) Public Domain view source

Although the sinkhole in Figure 8.9 might appear minor, it indicates a serious problem. The dam itself is constructed on limestone supported by beds of gypsum and anhydrite. Gypsum and anhydrite are soluble in water, and the gypsum and anhydrite beneath the dam are rapidly dissolving away. This was the case prior to construction of the dam. However, once the dam was filled, the increased water pressure began to force water through the formations much faster, accelerating dissolution. Ongoing measures to fill gaps with grout are required, or else there is a grave risk of catastrophic failure, placing nearly 1.5 million people at risk.

Hydrolysis

The term hydrolysis combines the prefix hydro, referring to water, with lysis, which is derived from a Greek word meaning to loosen or dissolve. Thus, you can think of hydrolysis as a chemical reaction where water loosens the chemical bonds within a mineral. This might sound the same as dissolution but the difference is that hydrolysis produces a different mineral in addition to ions. An example of hydrolysis is when water reacts with potassium feldspar to produce clay minerals and ions. The results can be seen by comparing weathered and unweathered surfaces of the same sample of granite (Figure 8.11). On the recently broken unweathered surface (Figure 8.11, left) feldspar is visible as bright white crystals. On a weathered surface (right) the feldspar has been altered to the chalky-looking clay mineral kaolinite.

 

Potassium feldspar (formula KAlSi3O8) is broken down by water to produce kaolinite (a clay mineral, formula Al2Si2O5(OH)4), quartz (formula SiO2), and potassium and hydroxyl ions.
Figure 8.11 A piece of granite with unweathered (left) and weathered (right) surfaces. On the unweathered surfaces the feldspars are still fresh and glassy looking. On the weathered surface there are chalky white patches where feldspar has been altered to the clay mineral kaolinite. Source: Karla Panchuk (2018) CC BY 4.0. Photos by Steven Earle (2015) CC-BY 4.0 view source

Silicate minerals other than feldspar can undergo hydrolysis, but with different end results. For example, pyroxene can be converted to the clay minerals chlorite or smectite. Olivine can be converted to the clay mineral serpentine.

Hydration

Hydration reactions involve water being added to the chemical structure of a mineral. An example of a hydration reaction is when anhydrite (CaSO4) is transformed into gypsum (CaSO4·2H2O). A consequence of hydration is that the resulting mineral has a greater volume than the original mineral. In the case of the Mosul Dam, hydration of anhydrite has important consequences. The increase in volume applied force to an overlying limestone layer, breaking it into pieces. While unbroken limestone is a strong enough material upon which to build a foundation, broken limestone is too weak to provide a safe foundation.

Oxidation

Oxidation happens when free oxygen (i.e., oxygen not bound up in molecules with other elements) is involved in chemical reactions. Oxidation reactions provide valuable insight into Earth’s early surface conditions because there is a clear transition in the rock record from rocks containing no minerals that are products of oxidation reactions, to rocks containing abundant minerals produced by oxidation. This reflects a transition from an oxygen-free atmosphere to an oxygenated one.

In iron-rich minerals such as olivine, the oxidation reaction begins with taking iron out of the mineral and putting it into solution as an ion. Olivine reacts with carbonic acid, leaving dissolved iron, bicarbonate, and silicic acid:

Fe2SiO4 + 4H2CO→ 2Fe2+ +  4HCO3 +  H4SiO4

Iron and oxygen dissolved in water react in the presence of bicarbonate to produce hematite and carbonic acid:

2Fe2+  + ½ O2 + 2H2O + 4HCO3  → Fe2O+ 4H2CO3

When the olivine in basalt is oxidized, the basalt takes on a reddish colour that is distinct from the dark grey or black of unweathered basalt (Figure 8.12).

 

Figure 8.12 Basalt pillows in Andalusia, Spain, with reddish weathered surfaces. Where parts of the pillows have broken away, darker unweathered basalt is visible. Source:Ignacio Benvenuty Cabral (2011) CC BY-NC-SA view source

The oxidation reaction would be similar for other iron-containing silicate minerals such as pyroxene, amphibole, and biotite. Iron in sulphide minerals such as pyrite (FeS2) can also be oxidized in this way.

Hematite is not the only mineral that can result from oxidation. In fact, a wide range of iron oxide minerals that can form in this way, In granite, for example, biotite and amphibole can be altered to form the iron oxide and iron hydroxyoxide minerals that are referred to in combination as limonite (orange material in Figure 8.13).

Figure 8.13 Biotite and amphibole in this granite have been altered by oxidation to limonite (orange-yellow coating), which is a mixture of iron oxide and iron hydroxyoxide minerals. Source: Steven Earle (2015) CC-BY 4.0 view source

Oxidation Reactions and Acid Rock Drainage

Oxidation reactions can pose an environmental problem in areas where rocks have elevated levels of sulphide minerals such as pyrite. This is because when oxygen and water react with pyrite, sulphuric acid is produced:

2FeS2 + 7O2 + 2H2O → 2FeSO4 + 2H2SO4

The runoff from areas where this process is taking place is known as acid rock drainage (ARD), and even a rock with 1% or 2% pyrite can produce significant ARD. Some of the worst examples of ARD are at metal mine sites, especially where pyrite-bearing rock and waste material have been mined from deep underground, and then piled up and left exposed to water and oxygen. In these cases the problem is referred to as acid mine drainage. One example is the Mt. Washington Mine near Courtenay on Vancouver Island (Figure 8.12), but there are many similar sites across Canada and around the world.

Figure 8.14 Acid mine drainage. Left: Mine waste where exposed rocks undergo oxidation reactions and generate acid at the Washington Mine, BC. Right: An example of acid drainage downstream from the mine site. Source: Steven Earle (2015) CC BY 4.0 view source

At many ARD sites, the pH of the runoff water is less than 4 (very acidic). Under these conditions, metals such as copper, zinc, and lead easily dissolve in water, which can be toxic to aquatic life and other organisms. For many years, the river downstream from the Mt. Washington Mine had so much dissolved copper in it that it was toxic to salmon. Remediation work has since been carried out at the mine and the situation has improved.

 

Exercise: Chemical Weathering

For each of the following reactions, indicate which chemical weathering process—dissolution, hydrolysis, hydration, or oxidation—is the primary mechanism.

  1. Pyrite → hematite
  2. Calcite → calcium and bicarbonate ions
  3. Feldspar → clay
  4. Olivine → serpentine
  5. Pyroxene → iron oxide
  6. Anhydrite → gypsum

References

Bienfait, B., & Ertl P. (2013). JSME: a free molecule editor in JavaScript. Journal of Cheminformatics 5(24). https://doi.org/10.1186/1758-2946-5-24

49

8.3 Controls on Weathering Processes and Rates

Weathering does not happen at the same rate in all environments. The same types of weathering do not happen in all environments. There are a variety of factors that determine what kinds of weathering will occur, and how fast the processes will proceed.

Climate

Water and temperature are key factors controlling both weathering rates and the types of weathering that occur:

This means, for example, that chemical weathering will be faster in a tropical rainforest than in the Arctic, a cold desert. It means physical weathering will be the predominant form of weathering in the Arctic.

Oxygen and Carbon Dioxide

The presence and abundance of oxygen and carbon dioxide affect chemical weathering rates. Surface environments on Earth almost all have some free oxygen available, permitting oxidation reactions to take place. Exceptions are in settings such as deep lakes or swamps where oxygen cannot easily mix into the water, and where biological processes consume the oxygen rapidly.

Carbon dioxide, which acidifies water and contributes to chemical weathering, is more concentrated in some settings than others. For example, because of the activities of organisms, soils can have very high concentrations of carbon dioxide, whereas carbon dioxide concentrations will be lower on surfaces free of soils and exposed to the atmosphere.

Minerals

The minerals making up a rock will determine what kinds of chemical weathering reactions are possible, and how rapidly chemical weathering reactions occur. Under the same conditions, dissolution of the calcite making up limestone will occur more rapidly than hydrolysis reactions happening to feldspar in granite. Quartz is very resilient to chemical weathering, and will remain long after calcite and feldspar have been weathered away. A rock with grains cemented by calcite will weather faster than a rock with grains cemented by quartz.

In general, differences in the rates of chemical weathering among minerals can be broken down as follows:

Weathering Makes Weathering Go Faster

Weathering accelerates weathering. Physical weathering forms cracks and breaks rocks apart into smaller pieces. The smaller the pieces, the greater the surface area exposed to chemical weathering. When the newly exposed surfaces are exposed to chemical weathering, it weakens the rock even further, making it more susceptible to physical weathering processes.

Differential Weathering

When rocks in an outcrop weather at different rates, the result is called differential weathering. Differential weathering causes some beds in an outcrop to be recessed relative to the others, because beds that are slow to weather will take longer to recede than weaker beds (Figure 8.15).

Figure 8.15 Differential weathering in an outcrop along the Blaeberry River near Golden BC. The recessed beds within the outcrop are weathering faster than the surrounding beds. Source: Karla Panchuk (2009) CC BY 4.0

50

8.4 Weathering and Erosion Produce Sediments

The visible products of weathering and erosion are the unconsolidated materials that we find around us on slopes, beneath glaciers, in stream valleys, on beaches, and in deserts. The loose collection of material is referred to as sediment, and the individual pieces that make it up are called clasts.  Clasts can be of any size: sand-sized and smaller (in which case they might be referred to as particles or grains), or larger than a house.

Clasts can range widely in size and shape (Figure 8.16) depending on the processes involved in making and transporting them. If and when deposits like these are turned into sedimentary rocks, the mineralogy and textures of these rocks will vary significantly. Importantly, when we describe sedimentary rocks that formed millions of years in the past, we can study the mineralogy and textures to make inferences about the conditions that existed during the deposition of the sediment, and the later burial and formation of sedimentary rock.  The properties we look at are composition, grain size, sorting, rounding, and sphericity.

 

Example 1: Boundlers in a talus deposit at Keremeos. All are angular fragments from the same rock source. Example 2: Pebbles on a beach in Victoria. All are rounded fragments of rock from different sources. Example 3: Sand from a beach at Gabriola. Most are angular quartz grains, some are fragments of rock. Example 4: Sand from a due in Utah. All are rounded quartz grains.
Figure 8.16 Products of weathering and erosion formed under different conditions. Source: Steven Earle (2016) CC BY 4.0 view source

Composition

Composition refers to the mineral or minerals making up the clast. Small clasts might be single mineral grains, but larger ones can have several different mineral grains, or even several different pieces of rock within them.  The composition can tell us about what rock the sediments came from, and about the geological setting from which the sediment was derived.

Not all minerals have the same hardness and resistance to weathering, so as weathering and erosion proceed, some minerals become more abundant than others within sediments. Quartz is one example of a mineral that is more abundant. It is highly resistant to weathering by weak acids or reaction with oxygen. This makes it unique among the minerals that are common in igneous rocks. Quartz is also very hard, so it is resistant to mechanical weathering.

In contrast, feldspar and iron- and magnesium-bearing minerals are not as resistant to weathering. As weathering proceeds, they are likely to be broken into small pieces and converted into clay minerals and dissolved ions. Ultimately this means that quartz, clay minerals, iron oxides, aluminum oxides, and dissolved ions are the most common products of weathering.

Grain Size

Whether a grain is large or small tells us about its journey from its source to where it was deposited.  Mechanical weathering can break off large pieces from rock.  Large pieces carried along by streams will bump into each other, causing smaller pieces to break off.  Over time the grains get smaller and smaller still. If we find grains that are very small, we can conclude that they travelled over a long distance.

Geologists have a specific set of definitions to describe the size of grains (Figure 8.17).

Figure 8.17 Classification of grain sizes. Silt and clay are considered fine-grained particles, sand is medium-grained, and particles larger than sand are considered coarse-grained.. Source: Karla Panchuk (2016) CC BY 4.0 Click the image for a text version.

The scale has some of the grain sizes listed in microns (µm). There are 1000 µm in 1 mm. The particles classified as sand are what you would intuitively think of as being sand-sized, so an easy way to remember the scale is that anything smaller than sand is fine-grained, and anything larger is coarse-grained. Fine sand grains are still easily discernible with the naked eye. Silt grains are barely discernible in rocks, and silty rocks feel gritty when rubbed. Clay grains are invisible to the naked eye, and rocks comprised of clay feel smooth when rubbed.

One other thing to notice about this scale is that the finest-grained particle is referred to as clay. While a clay-sized particle could be composed of clay minerals (and often they are), it doesn’t have to be. Any particle of that size would be referred to as clay.

Grain Size and Transportation

The grain size of sediments is not just for purposes of description. It’s also a valuable clue to the processes that have acted on those sediments, because the size of the clast determines how much energy is required to move it.

Whether or not a medium such as water or air has the ability to move a clast of a particular size and keep it moving depends on the velocity of the flow.  For the most part, the faster the medium flows, the larger the clasts that can be moved. Figure 8.18 shows a stream bed that now contains only a trickle of water—barely enough to move particles of sand or cool puppy feet. But the velocity of water in the stream changes from season to season, as does the volume of water.  All of the clasts in the stream bed were transported there by water at some point.

 

Figure 8.18 Ruby looks upstream in a channel near Golden BC. For much of the year the only water in the stream is the trickle in which Ruby stands, but in the spring the water can flow rapidly enough to carry boulders. Source: Karla Panchuk (2009) CC BY 4.0

Very fine-grained particles are the exception to rule that the larger the clasts, the faster the water that is required to transport them. Clay and silt grains stick together, requiring higher water velocities to pick them up and move them than some larger particles. Water that flows fast enough to pick up sand would not be fast enough to pick up clay.

Sorting

Weathering can break off large fragments of rock, and erosion and transport can break these fragments down to smaller and smaller sizes. The extent to which the grains in sediment differ in size is described by sorting (Figure 8.19, top).

Figure 8.19 Top: Sorting of grains, ranging from well sorted where the grains are similar in size, to poorly sorted, where the grains vary greatly in size. Bottom: Rounding refers to how smooth or rough the edges of a clast are.  Clasts with sharp edges and corners are angular. Clasts with smooth surfaces are rounded.  Clasts that fall in between are sub-angular or sub-rounded. Source: Reagan et al. (2015) CC BY 3.0 view source

If the grains in a sample of sediment are the same size or very nearly so, the sediment is said to be well sorted.  If the grains vary substantially in size, the sediment is poorly sorted. Because grains become progressively smaller as they are transported, sorting improves the further the sediments are from their source.

Rounding

Rounding refers to whether clasts have sharp edges and corners or not (Figure 8.19, bottom). If the grains are rough, with lots of edges and corners, then they are referred to as angular. Grains with smooth surfaces are rounded. Grains in between can be sub-angular or sub-rounded. The farther sediments are transported, the rounder they become.

Sphericity

Sphericity describes whether a grain is elongate or not.  Grains that are longer than they are wide (like an ellipse) have low sphericity, whereas grains that have the same diameter no matter where you measure it (like a sphere) have high sphericity. In the bottom row of boxes in Figure 8.19 the grains at the top of each box exhibit high sphericity, and the grains at the bottom exhibit low sphericity.  Notice that a grain can be angular but still have high sphericity.  It can be rounded, but still have low sphericity. Sphericity also increases the further the sediments are from their source.

 

Exercise: Looking at Sand

Three samples of sand are shown below. Read the descriptions of what they contain and where they are from, then describe each sample in terms of grain size, sorting, rounding, and sphericity.

Figure 8.20 Three examples of sand grains. Source: Steven Earle (2016) CC BY 4.0 View sources: Sample A, Sample B, Sample C.

 

Sample A: Fragments of red coral, algae plates, and urchin needles from a shallow water area (~2 m depth) near a reef in Belize. The grains are between 0.1 and 1 mm across.

 

Sample B: Quartz and rock fragments from a glacial stream deposit near Osoyoos, BC. The grains are between 0.25 and 0.5 mm in diameter.

 

Sample C: Grains of olivine (green) and volcanic glass (black) from a beach on the big island of Hawai’i. The grains are approximately 1 mm in diameter.

References

Reagan, M.K., Pearce, J.A., Petronotis, K., and the Expedition 352 Scientists, (2015). Proceedings of the International Ocean Discovery Program, Volume 352, publications.iodp.org, doi:10.14379/iodp.proc.352.102.2015

51

8.5 Weathering and Soil Formation

Weathering is a key part of the process of soil formation, and soil is critical to our existence on Earth. In other words, we owe our existence to weathering, and we need to take care of the soil!

Many people refer to any loose material on Earth’s surface as soil, but to scientists soil is the material that includes organic matter, forms within the top few tens of centimetres of the surface, and is important for sustaining plant growth.

Soil is a complex mixture of minerals (~45%), organic matter (~5%), and empty space (~50%, filled to varying degrees with air and water). The mineral content of soil varies, but is dominated by clay minerals and quartz, along with minor amounts of feldspar and small fragments of rock.

The types of weathering that take place within a region have a major influence on soil composition and texture. For example, in a warm climate where chemical weathering dominates, soils tend to be richer in clay. Soil scientists describe soil texture in terms of the relative proportions of sand, silt, and clay (Figure 8.21). Sand and silt components are dominated by quartz, with lesser amounts of feldspar and rock fragments. The clay component is dominated by clay minerals.

Figure 8.21 Soil texture classification diagram. Textures are determined by the proportions of sand-, silt-, and clay-sized grains. Source: Mike Norton (2011) CC BY-SA 3.0 view source

Factors Affecting How Soil Forms

Soil forms through the mechanical and chemical weathering of rocks and sediments, and the accumulation and decay of organic matter. The factors that affect the nature of soil and the rate of its formation include:

Climate

Both the mechanical breakup of rocks and the chemical weathering of minerals contribute to soil formation.  The downward percolation of water brings dissolved ions and also facilitates chemical reactions. Soil forms most readily under temperate to tropical conditions, and moderate precipitation.  Temperature matters because chemical weathering reactions and those facilitated by organisms proceed fastest under warm conditions, and plant growth is enhanced in warm climates. Where the climate is cooler, the rates of chemical weathering reactions decrease, and when water is frozen, may cease entirely.

Although water is needed for chemical weathering to take place, too much water can lead to soils that lack nutrients. In rain forests, for example, high rainfall contributes so much water that important nutrients are leached away, and acidic soils are left behind. In humid and poorly drained regions, swampy conditions may prevail, producing soil that is dominated by organic matter, but low in inorganic nutrients.

Too little water (e.g., in deserts and semi-deserts) limits the rate of downward chemical transport, and it also means that salts and carbonate ions dissolved in upward-moving groundwater can precipitate and build up in sediments, hindering organic activity. These soils also lack organic matter (Figure 8.22).

Figure 8.22 Soil consisting of wind-blown silt (loess) and little organic matter in an arid part of north-eastern Washington state. Source: Steven Earle (2016) CC BY 4.0 view source

Parent Material

Parent material for soils can be any type of bedrock, and any type of unconsolidated sediment, such as glacial deposits and stream deposits. Soils are described as residual soils if they develop on bedrock, and transported soils if they develop on transported material such as glacial sediments. This doesn’t mean that the soils themselves have been transported, but that the soil developed on unconsolidated material rather than on bedrock.

Sandy soils develop from quartz-rich parent material, such as granite, sandstone, or loose sand. Quartz-poor material, such as shale or basalt, generates soils with little sand.

Parent materials provide important nutrients to residual soils. For example, a minor constituent of granitic rocks is the calcium-phosphate mineral apatite, which is a source of the important soil nutrient phosphorus. Basaltic parent material tends to generate very fertile soils because, in addition to phosphorus, it provides significant amounts of iron, magnesium, and calcium.  The iron, magnesium, and calcium come from minerals such as olivine ((Mg,Fe)2SiO4) and plagioclase feldspar (CaAl2Si2O8) in the basalt.

Some unconsolidated materials, such as river-flood deposits, make for especially good soils because they tend to be rich in clay minerals. Clay minerals have large surface areas with negative charges that are attractive to positively charged elements like calcium, magnesium, iron, and potassium — important nutrients for plant growth.

Slope

Soil can only develop where surface materials remain in place and are not frequently washed away or lost to mass wasting (landslides). Soils cannot develop where the rate of soil formation is lower than the rate of erosion, so steep slopes tend to have little or no soil.

Time

Even under ideal conditions, soil takes thousands of years to develop. Virtually all of southern Canada was covered with glaciers up until 14,000 years ago, and most of the central and northern parts of BC, the prairies, Ontario, and Quebec were still glaciated at 12,000 years ago. Glaciers remained in the central and northern parts of Canada until around 10,000 years ago, so conditions were still not ideal for soil development even in the southern regions. This means that soils in Canada, particularly in central and northern Canada, are relatively young and not well developed.

The same applies to soils that are forming on newly created surfaces, such as recent deltas or sand bars, in areas of mass wasting, or where an area has been resurfaced by volcanic deposits.

Because soil takes so long to form, human activities that damage soils have long-term consequences for ecosystems, and for the utility of the soil for food production.

Soil Horizons

When soils form, the downward movement of clay, water, and dissolved ions can lead to the development of chemically and texturally distinct layers known as soil horizons. In temperate climates, common soil horizons that develop are the following (Figure 8.23):

Figure 8.23 Typical horizons in a temperate soil, from Wales. Source: Karla Panchuk (2018) CC BY 4.0. Photograph: Richard Hartnup (2005) Public Domain view source

Although rare in Canada, another type of layer that develops in hot arid regions is known as caliche (pronounced ca-lee-chee). It forms from the downward (or in some cases upward) movement of calcium ions, and the precipitation of calcite within the soil. When well developed, caliche cements the surrounding material together to form a layer that has the consistency of concrete.

 

How Soil Is Lost

Like all geological materials, soil is subject to erosion.  Under natural conditions on gentle slopes, the rate of soil formation either balances or exceeds the rate of erosion. However, human practices related to forestry and agriculture have significantly upset this balance.

Soils are held in place by vegetation. When vegetation is removed, either through cutting trees or routinely harvesting crops and tilling the soil, this protection is lost. When soil is not protected, wind and water can easily erode it away.

Water erosion is accentuated on sloped surfaces because fast-flowing water has greater eroding power than still water. Raindrops can disaggregate exposed soil particles, putting clay into suspension in the water. Sheetwash—unchannelled flow across a surface—carries suspended material away, and channels erode right through the soil layer, removing both fine and coarse material (Figure 8.24).

Figure 8.24 Soil erosion by rain and unchanneled runoff in a field in Alberta. Source:Alberta Agriculture and Rural Development. Click the image for source information and terms of use.

Wind erosion is exacerbated by the removal of trees that act as wind breaks, and by agricultural practices that leave bare soil exposed (Figure 8.25).

Figure 8.25 Soil erosion by wind in Alberta. Source:Alberta Agriculture and Rural Development. Click the image for source information and terms of use.

Tillage is also a factor in soil erosion, especially on slopes, because each time the soil is lifted by a cultivator, it is moved a few centimetres down the slope.

52

8.6 Soils of Canada

Until the 1950s, the classification of soils in Canada was based on the system used in the United States. However, it was long recognized that the U.S system did not apply well to many parts of Canada because of climate and environmental differences. The Canadian System of Soil Classification was first outlined in 1955 and has been refined and modified numerous times since then. There are 10 orders of soil recognized in Canada (Table 8.1), and you can explore the distribution of soils using Agriculture and Agri-Food Canada’s interactive map (Figure 8.26). See the resources section at the bottom of the page for additional sources of information on Canadian soils, including videos.

Table 8.1 Canadian Soil Classification System
Order Brief Description Environment
Forests
Podzolic Well-developed A and B horizons Coniferous forests throughout Canada
Luvisolic Clay-rich B horizon Northern prairies and central BC, mostly on sedimentary rocks
Brunisolic Poorly developed or immature soil, that does not have the well-defined horizons of podsol or luvisol Boreal-forest soils in the discontinuous permafrost areas of central and western Canada, and also in southern BC.
Grasslands
Chernozemic High levels of organic matter and an A horizon at least 10 cm thick Southern prairies and parts of BC’s southern interior, in areas that experience summer water deficits
Solonetzic A clay-rich B horizon, commonly with a salt-bearing C horizon Southern prairies, in areas that experience water deficits during the summer
Glacial and tundra
Cryosolic Poorly developed soil, mostly C horizon Permafrost areas of northern Canada
Vertisolic Clay-rich soils associated with glacial lake deposits Southern prairies
Other
Organic Dominated by organic matter; mineral horizons are typically absent Wetland areas, especially along the western edge of Hudson Bay, and in the area between the prairies and the boreal forest
Regosolic Does not have a B horizon (i.e., no accumulation of leached minerals) Unstable sediments including steep slopes prone to landslides, shifting sand dunes, and floodplains where sediments are frequently moved by streams
Gleysolic Colour patterns related to the absence of oxygen Water-saturated soils
Figure 8.26 Distribution of soil orders in Canada. Click here to go to the interactive map. Source: Agrifood and Agriculture Canada. Contains information licensed under the Open Government License – Canada. Click the image for terms of use.

 

Processes of soil formation include downward transport of solid and dissolved materials, and the nature of those processes depends in large part on the climate. In Canada’s predominantly cool and humid climate—characteristic of most places other than the far north— podzolization is the norm. This involves downward transportation of hydrogen, iron, and aluminum from the upper part of the soil profile, and accumulation of clay, iron, and aluminum in the B-horizon. Most of the podzols, luvisols, and brunisols of Canada form through various types of podzolization.

In the grasslands of the dry southern parts of the prairie provinces and in some of the drier parts of southern BC, dark brown organic-rich chernozem soils are dominant. In some cases, weak calcification takes place when calcium is leached from the upper layers and accumulates in the B-horizon. Development of caliche layers is rare in Canada.

Organic soils form in areas with poor drainage and a rich supply of organic matter, such as in swamps. These soils have very little mineral matter.

In the permafrost regions of the north, where glacial retreat was most recent, the time available for soil formation has been short and the rate of soil formation slow. The soils are called cryosols (the cryo prefix is used to indicate extreme cold). In permafrost areas, the freeze-thaw process churns the soil, resulting in limited soil horizon development.

 

Exercise: Soils of Canada

Examine Figure 8.26 showing the distribution of soils in Canada, or use the interactive map by clicking on the figure. For each of the five soils types listed below, briefly describe the distribution. Explain the distribution based on what you know about the conditions under which the soil forms and the variations in climate and vegetation related to it.

Soil type Describe the Distribution Explain the Reason for This Distribution
Chernozem
Luvisol
Podzol
Brunisol
Organic

 

 

Resources

Soils of Canada (University of Saskatchewan)

 

Soil Classification: Soil Orders of Canada Watch videos about each soil order.

53

8.7 Weathering and Climate Change

Carbon cycling on Earth operates on different timescales depending on the components of the Earth system that are involved. Over the short term, biological processes are important. In particular, living organisms — mostly plants — consume carbon dioxide from the atmosphere to make their tissues.  After they die, the carbon is released back into the atmosphere over years to decades as the plant matter decays.

Over the longer term, geological processes drive the carbon cycle. Geological carbon-cycle processes operate very slowly, but they affect much more of Earth’s carbon than the biological component. Carbon can move from the biological cycle to the geological cycle if it is buried in sedimentary rocks. The biological carbon could be fragments of plant material or organic molecules that are preserved as coal or in organic-rich shale. It could also be calcium carbonate body parts of marine organisms that are preserved in limestone.

The geological component of the carbon cycle is shown in Figure 8.27. The various steps in the process (not necessarily in this order) are as follows:

a: Organic matter from plants is stored in peat, coal, and permafrost for thousands to millions of years.
b: Weathering of silicate minerals converts atmospheric carbon dioxide to dissolved bicarbonate, which is stored in the oceans for thousands to tens of thousands of years.
c: Dissolved carbon is converted by marine organisms to calcite, which is stored in carbonate rocks for tens of millions to hundreds of millions of years.
d: Organic carbon compounds are stored in sediments for tens to hundreds of millions of years; some end up in petroleum deposits.
e: Carbon-bearing sediments are transferred to the mantle, where the carbon may be stored for tens of millions to billions of years.
f: During volcanic eruptions, carbon dioxide is released back to the atmosphere, where it is stored for years to decades.
Figure 8.27 The geological component of the carbon cycle includes: (a) organic carbon in peat, coal and permafrost, (b) weathering of silicate minerals converts atmospheric carbon dioxide to dissolved bicarbonate, (c) marine organisms convert dissolved carbon to calcium carbonate, (d) carbon compounds are stored in sediments, (e) carbon-bearing sediments are transferred to longer-term storage in the mantle, and (f) carbon dioxide is released back to atmosphere during volcanic eruptions. Source: Steven Earle (2016) CC BY 4.0 view source

 

At some times in Earth’s history, the geological carbon cycle has been balanced, with carbon being released to the atmosphere by some processes at approximately the same rate as other processes store it. Under these conditions, the climate can remain relatively stable.

At other times, the balance is upset. Prolonged periods of greater than average volcanism can cause an imbalance. The eruption of the Siberian Traps at around 250 Ma warmed the climate significantly over a few million years, leading to a mass extinction.

Mountain-building events may also cause an imbalance. The formation of the Himalaya range between about 40 Ma and 10 Ma ago exposed rocks to weathering over a large region. The over-all rate of weathering on Earth increased because the mountains were so high, and the range was so extensive. The weathering of these rocks — most importantly the hydrolysis of feldspar — consumed atmospheric carbon dioxide and transferred carbon to the oceans and to ocean-floor carbonate minerals. Decreasing carbon dioxide levels contributed to climate cooling that culminated in the Pleistocene glaciations.

Today, burning fossil fuels is causing an imbalance in the carbon cycle. Burning coal, oil, and gas releases in a geological instant carbon that was stored by the biological carbon cycle over hundreds of millions of years. Scientists who study Earth’s past climate tell us that today carbon dioxide is being added to the atmosphere faster than during some of the most extreme climate change events in Earth history. Eventually, higher carbon dioxide levels will accelerate chemical weathering, and that will help to remove some of the carbon dioxide from the atmosphere. However, weathering is part of the geological carbon cycle, and operates over long timescales. If humans stopped burning all fossil fuels today, it could still take thousands of years for balance to be restored.

54

Chapter 8 Summary

The topics covered in this chapter can be summarized as follows:

8.1 Mechanical Weathering

Rocks weather when they are exposed to surface conditions. In most cases, conditions at Earth’s surface are very different from the conditions under which the rocks formed. Mechanical weathering processes include exfoliation, freeze-thaw, salt crystallization, and the wedging effects of plant growth.

8.2 Chemical Weathering

Chemical weathering takes place when minerals within rocks are not chemically stable in their existing environment. Chemical weathering processes include hydrolysis of silicate minerals to form clay minerals, oxidation of iron in silicate and other minerals to form iron oxide minerals, and dissolution of calcite.

8.3 Controls on Weathering Processes and Rates

Chemical weathering is faster when temperatures are warmer and moisture is present. Physical weathering is more important in regions with frequent freeze-thaw cycles. Weathering rates can depend on the abundance oxygen and carbon, and will vary with the mineral composition of a rock. Weathering accelerates weathering by exposing more surface area to chemical reactions.

8.4 Weathering and Erosion Produce Sediments

Quartz grains are one of main products of weathering and erosion because quartz is resistant to chemical and mechanical weathering. Clay minerals, iron oxide and iron hydroxide minerals, aluminum hydroxide minerals, and ions in solution are common products of chemical weathering. Particles produced by weathering can be described in terms of their composition, grain size, sorting, rounding, and sphericity.

8.5 Weathering and Soil Formation

Soil is a mixture of fine mineral fragments (including quartz and clay minerals), organic matter, and empty spaces that may be partially filled with water. Soil formation is controlled by climate (especially temperature and humidity), the nature of the parent material, the slope (because soil can’t accumulate on steep slopes), and the amount of time available. Typical soils have layers called horizons, which form because of differences in the conditions with depth.

8.6 Soils of Canada

Canada has a range of soil types related to our unique conditions. The main types of soil form in forested and grassland regions, but there are extensive wetlands in Canada that produce organic soils, and large areas where soil development is poor because of cold conditions.

8.7 Weathering and Climate Change

The geological component of the carbon cycle affects Earth’s climate over the long term by changing atmospheric carbon dioxide levels. Carbon is added to the atmosphere during volcanic eruptions. It is extracted from the atmosphere when silicate minerals are weathered, and when it is transformed into organic matter by plants. Organic matter can be stored in soil, permafrost, and rocks. Burning of fossil fuels involves moving carbon from geological reservoirs to the atmosphere on timescales much faster than the geological carbon cycle operates.

Review Questions

  1. What must happen to a body of rock before exfoliation can occur?
  2. Saskatchewan’s climate is consistently cold in the winter and consistently warm in the summer. What times of year would frost wedging to be an important weathering mechanism?
  3. What are the products of the hydrolysis of the feldspar albite (NaAlSi3O8)?
  4. Oxidation weathering of  pyrite (FeS2) can lead to acid rock drainage (ARD). What are the environmental impacts of ARD?
  5. Imagine that you and a friend encounter an old graveyard on a walk through the forest. You see a granite tombstone with the date 1705 carved into it. The tombstone next to it is too badly weathered to read the date. Your friend looks at the badly weathered stone and declares, “Look how weathered this stone is! It must be from way before 1705.” Do you agree?
  6. Many sand deposits are dominated by quartz, with very little feldspar. What weathering and erosion conditions are required to get feldspar-rich sand?
  7. What ultimately happens to most of the clay that forms during the hydrolysis of silicate minerals?
  8. Why are the slope and the parent materials important factors in soil formation?
  9. Which soil constituents move downward to produce the B-horizon of a soil?
  10. What are the main processes that lead to the erosion of soils in Canada?
  11. Where in Canada would you expect to find a chernozemic soil? What characteristics of this region produce this type of soil?
  12. Where are brunisolic soils found in Saskatchewan?
  13. Why does weathering of silicate minerals, especially feldspar, lead to consumption of atmospheric carbon dioxide? What eventually happens to the carbon that is involved in that process?

55

Answers to Chapter 8 Review Questions

  1. Rock must be exposed at surface. It has to be uplifted from where it formed deep in the crust, and the material on top has to be eroded.
  2. Frost wedging is most effective when temperatures swing between freezing and thawing from day to day. In Saskatchewan that happens consistently in the early spring and late fall.
  3. The feldspar albite (NaAlSi3O8) will be converted to a clay (such as kaolinite) and sodium ions in solution.
  4. Acid rock drainage (ARD) creates acidic stream runoff. It also increases the solubility of a wide range of metals, some of which are toxic to wildlife and humans.
  5. If the stones are both granite, then it would be reasonable to conclude that the badly weathered tombstone is much older than the other, because it has been exposed to weathering for much longer. On the other hand, if the badly weathered stone is a rock that is less resistant to weathering, like limestone, then the badly weathered stone could be the same age, or even younger than the granite one.
  6. Feldspar-rich sand is formed where granitic rocks are being weathered and where mechanical weathering predominates over chemical weathering. For a deposit of feldspar-rich sand to be preserved, the sand must be deposited close to its source to limit the opportunities for chemical weathering.
  7. Most of the clay that forms during hydrolysis of silicate minerals ends up in rivers and is washed out to the oceans. There it eventually settles to the sea floor.
  8. On a steep slope, gravity will remove materials, making it unlikely for soils to accumulate. The mineral composition of the parent rock or sediment will influence the composition of the resulting soil.
  9. Clay minerals and iron move downward to produce the B horizon of a soil.
  10. Wind and water are the main processes of soil erosion in Canada. Removal of vegetation makes it easier for erosion to happen.
  11. Chernozemic soils are common in the southern prairies, where organic matter from grasslands is added to soils
  12. Brunisolic soils are found in the northern half of Saskatchewan where forest cover is common.
  13. The weathering of feldspar to clay involves the conversion of atmospheric carbon dioxide to dissolved bicarbonate, which ends up in the ocean.

IX

Chapter 9. Sedimentary Rocks

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 9.1 Cretaceous sedimentary rocks exposed along a road near Drumheller, Alberta, Canada. Sedimentary rocks form in layers called beds, and the planar boundaries that separate each bed are called contacts. Each bed tells a story about the conditions in which it formed. In this picture the beds are indicating that sea level repeatedly rose and fell. The black layer about halfway up the picture is a coal seam. It tells us that the environment at that time was swampy. Source: Karla Panchuk (2008) CC BY 4.0

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

Sedimentary Rocks Form From the Products of Weathering and Erosion

Weathering and erosion (Chapter 8) are the first two steps in the transformation of pre-existing rocks into sedimentary rocks. The remaining steps in the formation of sedimentary rocks are transportation, deposition, burial, and lithification. These steps are shown on the right-hand side of the rock cycle diagram in Figure 9.2.

Figure 9.2 The rock cycle. Processes related to sedimentary rocks are shown on the right-hand side. Source: Steven Earle (2015) CC BY 4.0 view source

Transportation is the movement of sediments or dissolved ions from the site of erosion to a site of deposition. This can be by wind, flowing water, glacial ice, or mass movement down a slope. Deposition takes place where the conditions change enough so that the sediments can no longer be transported. This could happen if the current slows down.

Burial occurs when sediments are deposited upon existing sediments, covering and compacting them. Lithification is what happens when those compacted sediments become cemented together to form solid sedimentary rock. Lithification occurs at depths of hundreds to thousands of metres within Earth.

Four Types of Sedimentary Rocks

Sedimentary rocks can be divided into four main types: clastic, chemical, biochemical, andorganic. Clastic sedimentary rocks are composed mainly of material that is transported as solid fragments (called clasts), and then cemented together by minerals that precipitated from solution. Chemical sedimentary rocks are composed mainly of material that is transported as ions in solution. Biochemical sedimentary rocks also form from ions in solution, but organisms play an important role in converting those ions into calcium carbonate or silica body parts. Organic sedimentary rocks contain large amounts of organic matter, such as from plant leaves and tree bark.

56

9.1 Clastic Sedimentary Rocks

How Clastic Sediments Become Sedimentary Rocks

Lithification (Figure 9.3) is the process of converting sediments into solid rock. Compaction is the first step. Sediments that have been deposited are buried when more and more sediments accumulate above them. The weight of the overlying sediments pushes the clasts together, closing up some of the pore spaces (the gaps between grains) and forcing them together. Pore spaces often contain water (although they can also contain air or even hydrocarbons), so the water is squeezed out.

Figure 9.3 Lithification turns sediments into solid rock. Lithification involves the compaction of sediments and then the cementation of grains by minerals that precipitate from groundwater in the spaces between these grains. Source: Karla Panchuk (2016) CC BY 4.0

Cementation is the next step. Groundwater flowing through the remaining pore spaces contains ions, and these ions may precipitate, leaving behind minerals in the pore spaces. These minerals bind the grains together, and are referred to collectively as cement. Quartz and calcite are common cement minerals, but depending on pressure, temperature, and chemical conditions, cement might also include other minerals such as hematite and clay.

Figure 9.4 shows sandstone viewed under a microscope. The grains are all quartz but they appear different shades of grey because they are being viewed through cross-polarized light. It is difficult to tell the grains from the cement in this case because both are made of quartz, but in the image on the right the more obvious grain boundaries are marked with dashed lines.  Some of the cement is marked with blue shading. Using the image on the right, see if you can pick out the grain boundaries in the image on the left.

Figure 9.4 Sandstone under a microscope. Grains and cement are quartz. Left- Original image. Right- Visible grain boundaries are marked with dashed lines, and some of the cement is shaded in blue. Source: Karla Panchuk (2018) CC BY 4.0 modified after Woudloper, Public Domain view source

Types of Clastic Sedimentary Rocks

Clastic sedimentary rocks are named according to the characteristics of clasts (rock and mineral fragments) that comprise them. These characteristics include grain size, shape, and sorting. The different types of clastic sedimentary rocks are summarized in Figure 9.5.

Figure 9.5 Types of clastic sedimentary rocks. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0, Photos by James St. John and  R. Weller/ Cochise College. Click the image for more attributions.

Coarse-Grained Clastic Rocks

Clastic sedimentary rocks in which a significant proportion of the clasts are larger than 2 mm are known as conglomerate if the clasts are well rounded, or breccia if they are angular (Figure 9.5, top row). Conglomerates form in high-energy environments, such as fast-flowing rivers, where the particles can become rounded as they bump into each other while being carried along.  Breccias typically form where the particles are not transported a significant distance, such as in alluvial fans and talus slopes.

Medium-Grained Clastic Rocks

Sandstone (Figure 9.5, middle row) is a very common sedimentary rock, and there are many different kinds of sandstone. It is worth knowing something about the different types because they are organized according to characteristics that are useful for the detective work of figuring out what conditions led to the formation of a particular sandstone. Broadly, sandstones can be divided into two groups: arenite and wacke (rhymes with tacky).

Arenite is “clean” sandstone consisting mostly of sand-sized grains and cement, with less than 15% of fine-grained silt and clay in the matrix (the material between the sand-sized grains). Arenites are subdivided according to what the sand-sized grains are made of (Figure 9.6). If 90% or more of the grains are quartz, then the sandstone is called a quartz arenite (also called a quartz sandstone). If more than 10% of the grains are feldspar and more of the grains are feldspar than fragments of other rocks (lithic“Lithic” means “rock.” Lithic clasts are rock fragments (multimineralic fragments), as opposed to single-mineral fragments. fragments) then the sandstone is called an arkosic arenite, or just arkose. If the rock has more than 10% rock fragments, and more rock fragments than feldspar, it is lithic arenite.

Figure 9.6 A compositional triangle for arenite sandstones, with the three most common components of sand-sized grains: quartz, feldspar, and rock fragments. Arenites have less than 15% silt or clay. Source: Steven Earle (2015) CC BY 4.0 view source

 

Wacke is a “dirty” sandstone, containing 15-75% fine-grained particles (clay, silt) in its matrix.  A wacke can have more fine-grained particles than cement in its matrix, making for a crumbly rock.  Wackes are subdivided in the same way that arenites are: quartz wacke, feldspathic wacke, and lithic wacke. Another name for a lithic wacke is greywacke.

Figure 9.7 shows thin sectionsThin sections are slivers of rock sliced thinly enough so that light can pass through them, and they can be examined under a microscope. (microscopic views) of quartz arenite, arkose, and lithic wacke. In the images, quartz grains are marked Q, feldspar grains are marked F, and lithic fragments are marked L. Notice the relative abundances of each component in the three types of rocks.

Figure 9.7 Photos of thin sections of three types of sandstone. Some of the minerals are labelled: Q=quartz, F=feldspar and L= lithic (rock fragments). The quartz arenite and arkose have relatively little silt/clay matrix, while the lithic wacke has abundant matrix. Source: Steven Earle (2016) CC BY 4.0 view source

Fine-Grained Clastic Rocks

Rock composed of at least 75% silt- and clay-sized clasts is called mudrock (Figure 9.5, bottom row). If a mudrock shows evidence of fine layers (laminations) and breaks into sheets, it is called shale. Otherwise, it is siltstone (dominated by silt), mudstone (a mix of silt and clay), or claystone (dominated by clay). The fine-grained nature of mudrocks tells us that they form in very low energy environments, such as lakes, flood plains, and the deep ocean.

Exercise: Classifying Sandstones

Use Figures 9.6 and 9.7 to give the appropriate name to the sandstone in each of the magnified thin sections shown below.

Figure 9.8a Sandstone 1. Source: Steven Earle (2015) CC BY 4.0 view source
Figure 9.8b Sandstone 2. Source: Steven Earle (2015) CC BY 4.0 view source
Sandstone 1. Rounded sand-sized grains are approximately 99% quartz and 1% feldspar. Silt and clay make up less than 2% of the rock. Sandstone 2. Angular sand-sized grains are approximately 70% quartz, 20% lithic fragments, and 10% feldspar. Silt and clay make up ~20% of the rock.

Clastic sediments are deposited in a wide range of environments, including from melting glaciers, slope failures, rivers (both fast and slow flowing), lakes, deltas, and ocean environments (both shallow and deep). Depending on the grain size in particular, they may eventually form into rocks ranging from mudstone to breccia and conglomerate. By examining clastic sedimentary rocks it is possible to translate the classification you have just learned into an interpretation of the environment in which the rocks were deposited.

Sediment Maturity

Maturity in sediments refers to the extent to which sediment characteristics reflect prolonged weathering and transport. Prolonged weathering and transport cause clasts to become smaller, rounder, and more well-sorted. It removes minerals that are more susceptible to weathering, such as feldspar, leaving a sediment consisting predominantly of quartz or clay. On the spectrum of sediment maturity, quartz sandstone or shale would be mature sedimentary rocks, and wacke or conglomerate would be an immature rocks.

57

9.2 Chemical and Biochemical Sedimentary Rocks

Clastic sedimentary rocks are dominated by components that have been transported as solid clasts (clay, silt, sand, etc.). In contrast, chemical and biochemical sedimentary rocks are dominated by components that have been transported as ions in solution (e.g., Na+, Ca2+, HCO3, etc.). There is some overlap between the two because almost all clastic sedimentary rocks contain cement formed from dissolved ions, and many chemical sedimentary rocks include some clasts. The difference between chemical and biochemical sedimentary rocks is that in biochemical sedimentary rocks, organisms play a role in turning the ions into sediment. This means the presence and nature of biochemical sedimentary rocks are linked to the life requirements of the organisms that comprise them. In chemical sedimentary rocks, the process is inorganic, often resulting from a body of water evaporating and concentrating the ions.  It is possible for one type of sedimentary rock to form from both chemical (inorganic) and biochemical (organically mediated) processes.

Chemical and biochemical sedimentary rocks are classified based on the minerals they contain, and are frequently dominated by a single mineral. It is true that some clastic sedimentary rocks, such as quartz arenite, can also be dominated by a single mineral, but the reasons for this are different. A clastic sedimentary rock can contain whatever minerals were present in the parent rock.  The minerals the clastic rock ends up containing will depend on how much “processing” the sediments undergo by physical and chemical weathering, and transport, before the sediment was cemented. On the other hand, chemical sedimentary rocks are limited largely to those minerals that are highly soluble in water.  Because mineral content is a defining characteristic of chemical and biochemical sedimentary rocks, we will use it to organize our discussion of these rocks.

Carbonate Rocks

Carbonate rocks are those in which the dominant mineral contains the carbonate anion (CO32-).  The main carbonate minerals are calcite and aragonite. Both minerals have the formula CaCO3 but they have different crystal structures.  A less common carbonate mineral that is still important for forming carbonate rocks is dolomite, which has the formula CaMg(CO3)2. It is similar to calcite and aragonite, except that some of the calcium is replaced with magnesium. Dolomite is more common as a replacement mineral, which has replaced calcite in carbonate rocks.

Limestone

Limestone is comprised of calcite and aragonite. It can occur as a chemical sedimentary rock, forming inorganically due to precipitation, but most limestone is biochemical in origin.  In fact, limestone is by far the most common biochemical sedimentary rock.

Almost all limestone forms in marine (i.e., oceans or salty seas) environments, and most of that forms on the shallow continental shelves, especially in tropical regions with coral reefs. Today continental shelves are relatively narrow zones along the margins of continents, but for large parts of geologic history sea-level was much higher, and large parts of the interiors of continents were flooded.

Reefs are highly productive ecosystems populated by a wide range of organisms, many of which use calcium and bicarbonate ions from seawater to make carbonate minerals (especially calcite) for their shells and other structures. These include corals as well as green and red algae, urchins, sponges, molluscs, and crustaceans. Some of micro-organisms use CaCO3 to build tiny tests (shells) which accumulate on the ocean floor when these organisms die. Erosion can break all of these carbonate materials apart, scattering fragments throughout surrounding region (Figure 9.9).

Figure 9.9 Various corals and green algae on a reef at Ambergris, Belize. The light-coloured sand consists of carbonate fragments eroded from the reef organisms. Source: Steven Earle (2015) CC BY 4.0 view source

Figure 9.10 shows a cross-section through a typical reef environment in a tropical region (normally between 40° N and 40° S). Reefs tend to form near the edges of steep drop-offs because the reef organisms thrive on nutrient-rich upwelling currents. As the reef builds up, it is eroded by waves and currents to produce carbonate sediments that are transported into the steep offshore fore-reef area and the shallower inshore back-reef area. Reef-derived sediments are dominated by reef-type carbonate fragments of all sizes, including mud.

Figure 9.10 Cross-section through a typical tropical reef. Source: Steven Earle (2015) CC BY 4.0 view source

In many such areas, carbonate-rich sediments also accumulate in quiet lagoons, where mud and mollusc-shell fragments predominate (Figure 9.11, left) or in offshore areas with strong currents, where either foraminifera tests accumulate (Figure 9.11, middle) or calcite crystallizes inorganically to form ooids – spheres of calcite that form in shallow tropical ocean water with strong currents (Figure 9.11, right).

Figure 9.11 Carbonate rocks and sediments. Left- Mollusc-rich limestone formed in a lagoon area at Ambergris, Belize. Middle- Foraminifera-rich sediment from a submerged carbonate sandbar near to Ambergris, Belize . Right- Ooids from a beach at Joulters Cay, Bahamas. Sources: Left, Middle- Steven Earle (2015) CC BY 4.0 view source; Right- Wilson44691 (2010) Public Domain view source

Limestone also accumulates in deeper water, from the steady settling out of the carbonate shells of tiny organisms that lived near the ocean surface. Processes on the ocean floor cause the water in the deepest parts of the ocean to become more acidic. This puts a lower limit on how deep in the ocean calcite and aragonite can accumulate, because they dissolve under acidic conditions.

Tufa and Travertine

Calcite can form chemical sedimentary rocks on land in a number of environments. Tufa forms at springs. The tufa towers in Figure 9.12 formed where spring water discharged into lake water.

Figure 9.12 Tufa towers (made of calcium carbonate) in Mono Lake, California. Evaporation keeps the concentration of ions in the lake very high, allowing the calcium carbonate to precipitate. Source: Brocken Inaglory (2006) CC BY-SA 3.0 view source

Travertine (which is less porous) forms at hot springs. Similar material precipitates within limestone caves to form speleothems (mineral deposits in caves, Figure 9.13) such as stalactites and stalagmites.

Figure 9.13 Speleothems in Cave Nefza in Tunisia Source: Badreddine Besbes (2015) CC BY-SA 3.0 view source

Dolostone

Dolostone (also referred to as dolomite) is the carbonate rock made of the mineral dolomite (CaMg(CO3)2). Dolostone is quite common (there’s a whole Italian mountain range named after it), which is surprising because marine organisms do not precipitate dolomite. Dolomite forms through dolomitization, a process thought to involve chemical reactions between magnesium-rich water percolating through rocks, and sediments containing calcite.

Calcite and dolomite can be distinguished from one another by applying a drop of weak acid to the rock; calcite will react with weak acid, whereas dolomite will not. Also, when dolomite weathers, it tends to turn buff (tan) in colour, whereas calcite tends toward grey and white.

Chert

Chert is made of silica (SiO2). It has the same chemical formula as quartz, but is cryptocrystalline, meaning that the quartz crystals comprising chert are so small that it is difficult to see them even under a microscope.  Chert can be a chemical sedimentary rock, often forming as beds within limestone (Figure 9.14), or as irregular lenses or blobs (nodules). It can also be biochemical. Some tiny marine organisms (e.g., diatoms and radiolaria) make their tests from silica. When they die their tiny shells (or tests) settle slowly to the bottom of the lake or ocean, where they accumulate and are transformed into chert.

Figure 9.14 Chert (brown layers) interbedded with limestone, Triassic Quatsino Fm, Quadra Island, BC. All of the layers have been folded, and the chert, being more resistant to weathering than limestone, stands out. Source: Steven Earle (2015) CC BY 4.0 view source

Banded Iron Formations (BIFs)

Some ancient chert beds — most dating to between 1800 and 2400 Ma — are also part of a rock known as a  banded iron formation (BIF). It is a deep sea-floor deposit of iron oxide that is a common ore of iron. These rocks consist of alternating layers of dark iron oxide minerals (magnetite and hematite) and chert stained red by hematite (Figure 9.15).

Figure 9.15 An example of a banded iron formation with dark iron oxide layers interspersed with chert stained red by hematite. This rock is 2.1 billion years old. Source: Andre Karwath CC BY-SA (2005) view source
BIFs formed before Earth’s atmosphere was fully oxygenated.  At that time, seawater contained abundant soluble ferrous iron (Fe2+).  However, once cyanobacteria began releasing oxygen into the atmosphere as a byproduct of photosynthesis, the iron in the seawater reacted with the oxygen, turning it into insoluble ferric iron (Fe3+). The result was that iron oxide minerals precipitated and sank to the ocean floor. The prevalence of BIFs in rocks dating from 2400 to 1800 Ma reflects a time when free oxygen was being added to the atmosphere, but removed just as quickly by chemical reactions. After 1800 Ma, little dissolved iron was left in the oceans so no more BIFs formed.

Evaporites

In arid regions, lakes and inland seas typically have no stream outlet, and the water that flows into them is removed only by evaporation. Under these conditions, the water becomes increasingly concentrated with dissolved salts, and eventually some of these salts may reach saturation levels and start to crystallize (Figure 9.16).

Figure 9.16 Spotted Lake, near Osoyoos, BC. This photo was taken in May when the water was relatively fresh because of winter rains. By the end of the summer the surface of this lake is typically fully encrusted with salt deposits. Source: Steven Earle (2015) CC BY 4.0 view source

Although all evaporite deposits are unique because of differences in the chemistry of the water, in most cases minor amounts of carbonates start to precipitate when the solution is reduced to about 50% of its original volume. Gypsum (CaSO4·H2O) precipitates at about 20% of the original volume, and halite (NaCl) precipitates at 10%. Other important evaporite minerals include sylvite (KCl) and borax (Na2B4O7·10H2O). Sylvite is mined as potash at numerous locations across Saskatchewan from evaporites that formed during the Devonian (~385 Ma) when an inland sea occupied much of the region.

 

Exercise: Making Evaporite

This is an easy experiment that you can do at home. Pour about 50 mL (just less than 1/4 cup) of very hot water into a cup and add 2 teaspoons (10 mL) of salt. Stir until all or almost all of the salt has dissolved, then pour the salty water (leaving any undissolved salt behind) into a shallow wide dish or a small plate. Leave it to evaporate for a few days and observe the result. It may look a little like the Figure 9.17. These crystals are up to ~3 mm in diameter.

Figure 9.17 Salt crystals up to ~ 3 mm across. Source: Steven Earle (2015) CC BY 4.0 view source

 

58

9.3 Organic Sedimentary Rocks

Organic sedimentary rocks are those containing large quantities of organic molecules. Organic molecules contain carbon, but in this context we are referring specifically to molecules with carbon-hydrogen bonds, such as materials from the soft tissues of plants and animals. In other words, the carbon in calcite- CaCO3 wouldn’t make calcite an organic mineral because it isn’t bonded to hydrogen.

An important organic sedimentary rock is coal. Most coal forms in swampy land adjacent to rivers and within deltas, and where climates are humid and tropical to temperate. The vigorous growth of vegetation leads to an abundance of organic matter that accumulates within stagnant, acidic water. This limits decay and oxidation of the organic material. If this situation—where the dead organic matter is submerged in oxygen-poor water—is maintained for centuries to millennia, a thick layer of material can accumulate. Limited decay will transform this layer into peat (Figure 9.18a, Figure 9.19 upper left).

Figure 9.18 Formation of coal. (a) Accumulation of organic matter within a swampy area forms a layer of peat; (b) The organic matter is buried under sediment and is compressed; (c) With greater burial, lignite coal forms; (d) At even greater depths, bituminous and eventually anthracite coal form. Source: Steven Earle (2015) CC BY 4.0 view source

At some point the swamp deposit is covered with more sediment — typically because a river changes its course or sea level rises (Figure 9.18b). As more sediments are added, the organic matter is compressed and heated as temperatures increase with depth. This has the effect of concentrating the carbon within the coal. The amount of heating will determine how far this process progresses.

The further the process does progress, the more the coal will go from having obvious pieces of plant material within it, to being a black, shiny mass.  Low-grade lignite coal forms at depths between 100 m to 1,500 m and temperatures up to ~50°C (Figure 9.18c). This is still a relatively early stage in the coal formation process, so the lignite commonly displays plant fossils that have not yet been destroyed in the process of coalification (Figure 9.19 upper right).

At between 1,000 m to 5,000 m depth and temperatures up to 150°C m, bituminous coal forms (Figure 9.18d, 9.19 lower right). At depths beyond 5,000 m and temperatures over 150°C, anthracite coal forms (Figure 9.19 lower left). In fact, as temperatures rise, the lower-grade forms of coal are actually being transformed from sedimentary to metamorphic rocks.

Figure 9.19 The formation of coal begins when plant matter is prevented from decaying by accumulating in low-oxygen, acidic water. A layer of peat forms. Heating and compression of peat form lignite, bituminous coal, and finally anthracite, as pressure and temperature increases. Source: Karla Panchuk (2017) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College and U. S. Geological Survey. Click the image for more attributions and terms of use.

The transition from peat to anthracite results in a progressive increase in the carbon concentration, in hardness, and in the amount of energy available to be released upon combustion.

59

9.4 Depositional Environments and Sedimentary Basins

Sediments accumulate in a wide variety of environments, both on the continents and in the oceans. Some of the more important of these environments are illustrated in Figure 9.20.

Figure 9.20 Some of the important depositional environments for sediments and sedimentary rocks. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Mike Norton (2008) CC BY-SA 2.0 view source

Tables 9.1 and 9.2 provide a summary of the processes and sediment types that pertain to the various depositional environments illustrated in Figure 9.19. The types of sediments that accumulate in these environments are examined in more detail in the last section of this chapter.

Table 9.1 Terrestrial Depositional Environments
Environment Key Transport Processes Depositional Settings Typical Sediments
Glacial Gravity, moving ice, moving water Valleys, plains, streams, lakes Glacial till, gravel, sand, silt, clay
Alluvial Gravity, moving water Where steep-sided valleys meet plains Coarse angular fragments
Fluvial Moving water Streams Gravel, sand, silt, organic matter
Aeolian Wind Deserts and coastal regions Sand, silt
Lacustrine Moving Water Lakes Sand, silt, clay, organic matter
Evaporite Still water Lakes in arid regions Salts, clay
Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source.
Table 9.2 Marine Depositional Environments
Environment Key Transport Processes Depositional Settings Typical Sediments
Deltaic Moving water Deltas Sand, silt, clay, organic matter
Beach Waves, long-shore currents Beaches, spits, sand bars Gravel, sand
Tidal Tidal currents Tidal flats Fine-grained sand, silt, clay
Reef Waves, tidal currents Reefs and adjacent basins Carbonates
Shallow marine Waves, tidal currents Shelves, slopes, lagoons Carbonates in tropical climates; sand/silt/clay elsewhere.
Lagoonal Little transportation Lagoon bottom Carbonates in tropical climates, silt, clay
Submarine fan Underwater gravity flows Continental slopes, abyssal plains Gravel, sand, silt, clay
Deep water Ocean currents Deep-ocean abyssal plains Clay, carbonate mud, silica mud
Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source.

Most of the sediments that you might see around you, including talus on steep slopes, sand bars in streams, or gravel in road cuts, will never become sedimentary rocks. This is because they have only been deposited relatively recently — perhaps a few centuries or millennia ago — and will be re-eroded before they are buried deep enough beneath other sediments to be lithified. In order for sediments to be preserved long enough to be turned into rock (a process that takes millions or tens of millions of years) they need to have been deposited in a basin in which sediments can be preserved for that long. Most such basins are formed by plate tectonic processes (Figure 9.21).

Figure 9.21 Some types of tectonically produced basins: (a) trench basin, (b) forearc basin, (c) foreland basin, and (d) rift basin. Source: Steven Earle (2015) CC BY 4.0 view source

Trench basins form where a subducting oceanic plate dips beneath the overriding continental or oceanic lithosphere. They can be several kilometres deep, and in many cases, host thick sequences of sediments from nearby eroding coastal mountains. There is a well-developed trench basin off the west coast of Vancouver Island.

A forearc basin lies between the subduction zone and the volcanic arc, and may be formed in part by friction between the subducting plate and the overriding plate, which pulls part of the overriding plate down. The Strait of Georgia, the channel between Vancouver Island and the BC mainland, is a forearc basin.

A foreland basin is caused by the mass of a mountain range depressing the crust. A rift basin forms where continental crust is being pulled apart, and the crust on both sides the rift subsides. If rifting continues this will eventually becomes a narrow sea, and then an ocean basin. The East African rift basin represents an early stage in this process.

60

9.5 Sedimentary Structures and Fossils

Through careful observation over the past few centuries, geologists have discovered that the accumulation of sediments and sedimentary rocks takes place according to some important geological principles that can be summarized as follows:

These and other principles are discussed in more detail in Chapter 19.

In addition to these principles that apply to all sedimentary rocks, a number of other important characteristics of sedimentary processes lead to the development of distinctive sedimentary features in specific sedimentary environments. By understanding the origins of these features, we can make some very useful inferences about the processes and depositional environment that ultimately resulted in the rocks that we are studying.

Bedding refers to sedimentary layers that can be distinguished from one another on the basis of characteristics such as texture, composition, colour, or weathering characteristics (Figure 9.22). They may also be similar layers separated by partings, narrow regions marking weaker surfaces where erosion is enhanced. Bedding is an indication of changes in depositional processes that may be related to seasonal differences, changes in climate, changes in locations of rivers or deltas, or tectonic changes. Bedding can form in almost any depositional environment.

Figure 9.22 Beds in the Triassic Sulphur Mt. Formation near Exshaw, Alberta. Bedding is defined by differences in colour and texture, and also by partings (darker lines) between beds that may otherwise appear to be similar. Source: Steven Earle (2015) CC BY 4.0 view source

Cross-bedding is bedding that contains angled layers. It forms when sediments are deposited by flowing water or wind (Figure 9.23). Cross-beds in streams tend to be on the scale of cm to tens of cm, while those in aeolian (wind deposited) sediments can be on the scale of metres.

Figure 9.23 Cross-bedded Jurassic Navajo Formation aeolian sandstone at Zion National Park, Utah. In most of the layers the cross-beds dip down toward the right, implying wind direction from right to left during deposition. One bed dips in the opposite direction, implying a different wind direction. Source: Steven Earle (2015) CC BY 4.0 view source

Cross-beds form as sediments are deposited on the leading edge of an advancing ripple or dune. Each layer is related to a different ripple that advances in the flow direction, and is partially eroded by the following ripple (Figure 9.23). Cross-bedding is a very important sedimentary structure to recognize because it can provide information on the direction of current flows and, when analyzed in detail, on other features like the rate of flow and the amount of sediment available.

Figure 9.24 Formation of cross-beds as a series of ripples or dunes that migrate with the flow. Each ripple advances forward (right to left in this view) as more sediment is deposited on its leading face. Source: Steven Earle (2015) CC BY 4.0 view source

Ripples, which are associated with the formation of cross-bedding under unidirectional flow, may be preserved on the surfaces of sedimentary beds. Ripples formed in flowing water can also help to determine flow direction because they tend to have their steepest surface facing down-flow. Ripples can also form from back-and-forth flows, like at a beach, but these do not leave cross-beds, and are symmetrical, without one side steeper than the other.

Graded bedding is characterized by a change in grain size from bottom to top within a single bed. “Normal” graded beds are coarse at the bottom and become finer toward the top (Figure 9.25), a product of deposition from a slowing current. Some graded beds are reversed (coarser at the top), and this normally results from deposition by a fast-moving debris flow. Most graded beds form in a submarine fan environment, where sediment-rich flows descend periodically from a shallow marine shelf down a slope and onto the deeper sea floor.

Figure 9.25 Graded bedding going from pebbles at the bottom to sand at the top. Source: Cropped from James St. John (2018) CC BY 2.0 view source

In a stream environment, boulders, cobbles, and pebbles can become imbricated, meaning that they are generally tilted in the same direction. Clasts in streams tend to tilt with their upper ends pointing downstream, because this is the most stable position with respect to the stream flow (Figure 9.26).

Figure 9.26 Imbrication of clasts in a fluvial environment. Source: Steven Earle (2015) CC BY 4.0 view source

Mud cracks form when a shallow body of water (e.g., a tidal flat or pond), into which muddy sediments have been deposited, dries up and cracks (Figure 9.27). This happens because the clay in the upper mud layers shrinks upon drying.

Figure 9.27 Mud cracks in a tidal flat in England. Source: Alan Parkinson (2000) CC BY-SA 2.0 view source

The various structures described above are critical to understanding and interpreting the formation of sedimentary rocks. In addition to these structures, geologists also look very closely at sedimentary grains to determine their mineralogy or lithology (in order to make inferences about the type of source rock and the weathering processes), their degree of rounding, their sizes, and the extent to which they have been sorted by transportation and depositional processes.

A Note About Fossils

Fossils are not covered in detail in this book, but they are extremely important for understanding sedimentary rocks. Fossils can be used to date sedimentary rocks, but just as importantly, they tell us a great deal about the depositional environment of the sediments and the climate at the time: they can help to differentiate marine, aquatic, and terrestrial environments; estimate the depth of the water; detect the existence of currents; and estimate average temperature and precipitation.

 

Exercise: Interpreting Past Environments

Sedimentary rocks can tell us a great deal about the environmental conditions that existed during the time of their formation. For each of the following rocks, make some inferences about the following:

  • source rock
  • weathering
  • sediment transportation (medium of transport, transport distance)
  • depositional conditions

Quartz sandstone: no feldspar, well-sorted and well-rounded quartz grains, cross-bedded

Feldspathic sandstone and mudstone: feldspar, volcanic fragments, angular grains, repetitive graded bedding from sandstone upwards to mudstone

Conglomerate: well-rounded pebbles and cobbles of granite and basalt; imbrication

Breccia: poorly sorted, angular limestone fragments; orange-red matrix

61

9.6 Groups, Formations, and Members

Geologists who study sedimentary rocks need ways to divide them into manageable units, and they also need to give those units names so that they can easily be referred to and compared with other rocks deposited in other places. The International Commission on Stratigraphy (ICS) (http://www.stratigraphy.org/) has established a set of conventions for grouping, describing, and naming sedimentary rock units.

The main stratigraphic unit is a formation. A formation is a series of beds that is distinct from other beds above and below, and is thick enough to be shown on the geological maps that are widely used within the area in question. In most parts of the world, geological mapping is done at a relatively coarse scale, and so most formations are on the order of a few hundred metres thick. At that thickness, a typical formation would appear on a typical geological map as an area that is at least a few millimetres thick.

A series of formations can be classified together to define a group, which could be as much as a few thousand metres thick, and represents a series of rocks that were deposited within a single basin (or a series of related and adjacent basins) over millions to tens of millions of years.

In areas where detailed geological information is needed (for example, within a mining or petroleum district) a formation might be divided into members, where each member has a specific and distinctive lithology (rock type). For example, a formation that includes both shale and sandstone might be divided into members, one of which is shale, and the other sandstone. In some areas, where even more detail is required, members may be divided into beds, but this is only applicable to beds that have a special geological significance. Groups, formations, and members are typically named for the area where they are found.

The sedimentary rocks of the Nanaimo Group on Vancouver Island provide a useful example for understanding groups, formations, and members. During the latter part of the Cretaceous Period, from about 90 Ma to 65 Ma, a thick sequence of clastic rocks was deposited in a foreland basin between what is now Vancouver Island and the BC mainland (Figure 9.28). Nanaimo Group comprises a 5000 m thick sequence of conglomerate, sandstone, and mudstone layers. Coal was mined from the Nanaimo Group rocks from around 1850 to 1950 in the Nanaimo region, and is still being mined in the Campbell River area.

Figure 9.28 The distribution of the Upper Cretaceous Nanaimo Group rocks on Vancouver Island, the Gulf Islands, and in the Vancouver area. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Mustard (1994).

The Nanaimo Group is divided into 11 formations (Table 9.3). In general, the boundaries between formations are based on major lithological differences. A wide range of depositional environments existed during the accumulation of the Nanaimo Group rocks, from nearshore marine for the Comox and Haslam Formation; to fluvial and deltaic with backwater swampy environments for the coal-bearing Extension, Pender, and Protection Formations; to a deep-water submarine fan environment for the upper six formations. The differences in the depositional environments are probably a product of variations in tectonic-related uplift over time.

Table 9.3 Nanaimo Group Formations
Age (Ma) Formation Lithologies Depositional Environmemt
~65-66 Gabriola Sandstone with minor mudstone Submarine fan, high energy
~66-67 Spray Mudstone/ sandstone turbidites Submarine fan, low energy
~67-68 Geoffrey Sandstone and conglomerate Submarine fan, high energy
~68-70 Northumberland Mudstone turbidites Submarine fan, low energy
~70 De Courcy Sandstone Submarine fan, high energy
~70-72 Cedar District Mudstone turbidites Submarine fan, low energy
~72-75 Protection Sandstone and minor coal Nearshore marine and onshore deltaic and fluvial
~75-80 Pender Sandstone and minor coal Nearshore marine and onshore deltaic and fluvial
~80 Extension Conglomerate, with minor sandstone and some coal Nearshore marine and onshore deltaic and fluvial
~80-85 Haslam Mudstone and siltstone Shallow marine
~85-90 Comox Conglomerate, sandstone, mudstone (coal in the Campbell River area) Nearshore fluvial and marine
Source: Steven Earle, with data from Mustard (1994)

The five lower formations of the Nanaimo Group are all exposed in the Nanaimo area, and were well studied during the coal-mining era between 1850 and 1950. With the exception of the Haslam formation, they were divided into members, because that was useful for understanding the rocks in the areas where coal was mined.

There is much variety in the Nanaimo Group rocks, and it would take hundreds of photographs to illustrate all of the different types of rocks. Nevertheless, a few representative examples are shown in Figures 9.30-9.32.

Figure 9.30 Nanaimo Group, Spray Formation. Turbidite layers on Gabriola Island. Each turbidite set consists of a lower sandstone layer (light colour) that grades upward into siltstone, and then into mudstone. Source: Steven Earle (2015) CC BY 4.0 view source
Figure 9.31 Nanaimo Group, Pender Formation. Two separate layers of fluvial sandstone with a thin (approx. 75 cm) coal seam in between. Source: Steven Earle (2015) CC BY 4.0 view source
Figure 9.32 Nanaimo Group, Comox Formation. The metal object is the end of a rock hammer that is 3 cm wide. Almost all of the clasts in this view are well-rounded basalt pebbles cobbles eroded from the Triassic Karmutsen Formation which makes up a major part of Vancouver Island. Source: Steven Earle (2015) CC BY 4.0 view source

References

Mustard, P. (1994). The Upper Cretaceous Nanaimo Group, Georgia Basin. In J. Monger (Ed.), Geology and Geological Hazards of the Vancouver Region. Geological Survey of Canada Bulletin 481, 27-95.

62

Chapter 9 Summary

The topics covered in this chapter can be summarized as follows:

9.1 Clastic Sedimentary Rocks

Sedimentary clasts are classified based on their size, and variations in clast size have important implications for transportation and deposition. Clastic sedimentary rocks range from conglomerate to mudstone. Clast size, sorting, composition, and shape are important features that allow us to differentiate clastic rocks and understand the processes that took place during their deposition.

9.2 Chemical and Biochemical Sedimentary Rocks

Chemical and biochemical sedimentary rocks form from ions that were transported in solution, and then converted into minerals by chemical and/or biological processes. The most common biochemical rock, limestone, typically forms in shallow tropical marine environments, where biological activity is a very important factor. Chert and banded iron formations can be from deep-ocean environments. Evaporites form where the waters of lakes and inland seas become supersaturated due to evaporation.

9.3 Organic Sedimentary Rocks

Organic sedimentary rocks contain abundant organic carbon molecules (molecules with carbon-hydrogen bonds). An example is coal, which forms when dead plant material is preserved in stagnant swamp water, and later compressed and heated.

9.4 Depositional Environments and Sedimentary Basins

There is a wide range of depositional environments, both on land (including glaciers, lakes, and rivers) and in the ocean (including deltas, reefs, shelves, and the deep-ocean floor). In order to be preserved, sediments must accumulate in sedimentary basins, many of which form through plate-tectonic processes.

9.5 Sedimentary Structures and Fossils

The deposition of sedimentary rocks can be described in terms of a series of principles, including original horizontality, superposition, and faunal succession. Sedimentary rocks can also have distinctive structures that are important in determining their depositional environments. Fossils are useful for determining the age of a rock, the depositional environment, and the climate at the time of deposition.

9.6 Groups, Formations, and Members

Sedimentary sequences are classified into groups, formations, and members so that they can be mapped easily and without confusion.

Review Questions

  1. What are the minimum and maximum sizes of sand grains?
  2. The material that makes up a rock such as conglomerate cannot be deposited by a slow-flowing river. Why?
  3. Describe the two main processes of lithification.
  4. What is the difference between a lithic arenite and a lithic wacke?
  5. How does a feldspathic arenite differ from a quartz arenite?
  6. What can we say about the source area lithology, and the weathering and transportation history of a sandstone that is primarily composed of rounded quartz grains?
  7. What is the original source of the carbon that is present within carbonate deposits such as limestone?
  8. What long-term environmental change on Earth led to the deposition of banded iron formations?
  9. Name two important terrestrial depositional environments and two important marine ones.
  10. What is the origin of a foreland basin, and how does it differ from a forearc basin?
  11. Explain the origins of  (a) bedding, (b) cross-bedding, (c) graded bedding, and (d) mud cracks.
  12. Under what conditions will reverse-graded bedding form?
  13. What are the criteria for the application of a formation name to a series of sedimentary rocks?
  14. Explain why some of the Nanaimo Group formations have been divided into members, while others have not.

 

63

Answers to Chapter 9 Review Questions

1. Sand grains range in size from 1/16 mm to 2 mm.

2. Conglomerate cannot be deposited by a slow-flowing river because clasts larger than 2 mm are not transported by slow-moving water.

3. Sediments are buried beneath other sediments where, because of the increased pressure, they become compacted and water is forced out from between the grains. With additional burial they are warmed to the point where cementing minerals can form between the grains (less than 200˚C).

4. Lithic arenite has less than 15% silt- and clay-sized particles, while a lithic wacke has more than 15%. Both have more than 10% rock fragments and more rock fragments than feldspar.

5. Feldspathic arenite has more than 10% feldspar and more feldspar than rock fragments. Quartz arenite has less than 10% feldspar and less than 10% rock fragments. Both have less than 15% silt and clay.

6. Source area lithology: rock that contains quartz (such as granite or sandstone). Strong weathering is required to remove feldspar, and long fluvial transportation to round the grains.

7. The carbon within carbonate deposits such as limestone ultimately comes from the atmosphere.

8. Most of Earth’s banded iron formations formed during the initial oxygenation of the atmosphere between 2.4 and 1.8 Ga because iron that had been soluble in the anoxic oceans became insoluble in the oxidized oceans.

9. Terrestrial depositional environments: rivers, lakes, deltas, deserts, glaciers. Marine depositional environments: continental shelves, continental slopes, deep ocean.

10. A foreland basin forms in the vicinity of a large range of mountains where the weight of the mountains depresses the crust on either side. A forearc basin lies between a subduction zone and the related volcanic arc.

11. (a) Bedding forms where there is an interruption or change in the depositional process, or a change in the composition of the material being deposited. (b) Cross-bedding forms in fluvial or aeolian environments where sand-sized sediments are being moved and ripples or dunes are present. (c) Graded bedding forms when transport energy decreases, depositing finer and finer particles. (d) Mud cracks form where fine-grained sediments (silt or clay) are allowed to dry.

12. Reverse-graded bedding forms during gravity flows, such as debris flows.

13. A formation is a series of beds that is distinct from other beds above and below it, and is thick enough to be shown on the geological maps that are widely used within the area in question.

14. The Nanaimo Group was actively mined for coal for many decades. During that time the names were given to members and individual beds that were important to the coal miners.

X

Chapter 10. Metamorphism and Metamorphic Rocks

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 10.1 Grey and white striped metamorphic rocks (called gneiss) at Pemaquid Point were transformed by extreme heat and pressure during plate tectonic collisions. Source: Karla Panchuk (2018) CC BY 4.0. Photos by Joyce McBeth (2009) CC BY 4.0 view source left/ right. Map by Flappiefh (2013), derivative of Reisio (2005), Public Domain view source.

Learning Objectives

After reading this chapter and completing the review questions at the end, you should be able to:

Metamorphism Occurs Between Diagenesis And Melting

Metamorphism is the change that takes place within a body of rock as a result of it being subjected to high pressure and/or high temperature.  The parent rock or protolith is the rock that exists before metamorphism starts. New metamorphic rocks can form from old ones as pressure and temperature progressively increase. The term parent rock is typically applied to the initial unmetamorphosed rock, rather than referring to each metamorphic rock that formed as metamorphism progresses.  We don’t always know whether metamorphism occurred in an uninterrupted sequence or whether metamorphism stopped and started again for different reasons at different times.

Metamorphic rocks form under pressures and temperatures that are higher than those experienced by sedimentary rocks during diagenesis, but at temperatures lower than those that cause igneous rocks to melt.  Given that pressure and water content affect the temperature at which rocks melt, metamorphism can occur at higher temperatures for some kinds of rocks, whereas other rocks will begin to melt under these same conditions. Metamorphic rocks can have very different mineral assemblages and textures than their parent rocks (Figure 10.2), but their over-all chemical composition usually does not change very much.

Figure 10.2 Shale is the parent rock of gneiss (pronounced “nice”). These rocks look very different, but gneiss can form when the atoms contained within the shale are re-arranged into new mineral structures. Source: Karla Panchuk (2018) CC BY-NC-SA. Photos by R. Weller/Cochise College. Click the image for photo sources.

Most metamorphism results from the burial of igneous, sedimentary, or pre-existing metamorphic rocks, to the point where they experience different pressures and temperatures than those at which they formed. Metamorphism can also take place if cold rock near the surface is intruded and heated by a hot igneous body. Metamorphism usually involves temperatures above 150°C, but some types of metamorphism do occur at temperatures lower than those at which the parent rock formed.

64

10.1 Controls on Metamorphic Processes

The main factors that control metamorphic processes are:

Mineral composition

Parent rocks can be from any of the three rock types: sedimentary, igneous, or metamorphic.  The critical feature of the parent rock is its mineral composition.  This is because the stability of minerals—how they are influenced by changing conditions—is what determines which minerals form as metamorphism takes place. When a rock is subjected to increased temperatures and pressures, some minerals will undergo chemical reactions and turn into new minerals, while others might just change their size and shape.

Temperature

The temperature under which metamorphism occurs is a key variable in determining which metamorphic reactions happen. Mineral stability depends on temperature, pressure, and the presence of fluids. Minerals are stable over a specific range of temperatures. Quartz, for example, is stable from surface temperatures up to approximately 1800°C. If the pressure is higher, that upper limit will also be higher. If there is water present, it will be lower. Most other common minerals have upper limits between 150°C and 1000°C.

Some minerals will change their crystal structure depending on the temperature and pressure. Quartz has different polymorphs that are stable between 0°C and 1800°C. The minerals kyanite, andalusite, and sillimanite are polymorphs with the composition Al2SiO5. The fact that they are stable at different pressures and temperatures means that their presence can be used to determine the pressures and temperatures experienced by a metamorphic rock that contains one or more of the polymorphs (Figure 10.3).

Figure 10.3 The Al2SiO5 polymorphs andalusite, kyanite, and sillimanite, and their stability fields. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photos by Rob Lavinsky/ iRocks.com (pre-2010) CC BY-SA 3.0. View source for andalusite/ kyanite/ sillimanite.

Pressure

Pressure has implications for mineral stability, and therefore the mineral content of metamorphic rocks, but it also determines the texture of metamorphic rocks. When directed pressure (or directed stress) acts on a rock, it means the stress on the rock is much greater in one direction than another. In an experiment with cylinders of modeling clay stacked in a block (Figure 10.4, top), pushing down on the clay from above resulted in higher directed pressure in the up-down direction (larger arrows; downward from pushing on the clay, and upward from the force of the table beneath the clay) than in the sideways direction, where only air pressure was acting (small arrows). The clay cylinders became elongated in the direction of least pressure.

Figure 10.4 Modelling clay experiments showing the effects of pressure on textures. Top: Directed pressure- clay was set on a flat surface and pushed down on from above (large arrows). Cylinders making up the clay block became elongated in the direction of least stress. Bottom: Shear stress applied to the top and bottom of a block of clay caused the interior to stretch. Note white dashed reference circles and elongated ellipses. Source: Karla Panchuk (2018) CC BY 4.0

Rocks undergo shear stress when forces act parallel to surfaces. In another modelling-clay experiment, applying oppositely directed forces to the top and bottom of a block of clay (Figure 10.4, bottom) caused diagonal stretching within the block. Note the change in shape of the dashed white reference circles.

In both experiments, parts of the clay became elongated in a particular direction. When mineral grains within a rock become aligned like this, it produces a fabric called foliation. Foliation is described in more detail later in this chapter.

Fluids

Water is the main fluid present within rocks of the crust, and the only one considered here. The presence of water is important for two main reasons. First, water facilitates the transfer of ions between minerals and within minerals, and therefore increases the rates at which metamorphic reactions take place. This speeds the process up so metamorphism might occur more rapidly, or metamorphic processes that might not otherwise have had time to be completed are completed.

Secondly, water—especially hot water—can have elevated concentrations of dissolved substances, making it an important medium for moving ions from one place to another within the crust. Processes facilitated by hot water are called hydrothermal processes (hydro refers to water, and thermal refers to heat).

Time

Most metamorphic reactions occur very slowly. Estimates of the growth rates of new minerals within a rock during metamorphism suggest that new material is added to the outside of mineral crystals at a rate of approximately 1 mm per million years. Very slow reaction rates make it difficult to study metamorphic processes in a lab.

While the rate of metamorphism is slow, the tectonic processes that lead to metamorphism are also very slow, so there is a good chance that metamorphic reactions will be completed. For example, an important setting for metamorphism is many kilometres deep within the roots of mountain ranges. A mountain range takes tens of millions of years to form, and tens of millions of years more to be eroded to the extent that we can see the rocks that were metamorphosed deep beneath it.

Exercise: How Long Did It Take?

The large reddish crystals in Figure 10.5 are garnet, and the surrounding light coloured rock is dominated by muscovite mica. The Euro coin is 23 mm in diameter. Assume that the diameters of the garnets increased at a rate of 1 mm per million years. Based on the approximate average diameter of the garnets visible, estimate how long this metamorphic process might have taken.

Figure 10.5 Garnet-mica schist from the Greek island of Syros. Source: Graeme Churchard (2005) CC BY 2.0 view source

 

 

65

10.2 Foliation and Rock Cleavage

How Foliation Develops

When a rock is acted upon by pressure that is not the same in all directions, or by shear stress (forces acting to “smear” the rock), minerals can become elongated in the direction perpendicular to the main stress. The pattern of aligned crystals that results is called foliation.

Foliation can develop in a number of ways. Minerals can deform when they are squeezed (Figure 10.6), becoming narrower in one direction and longer in another.

Figure 10.6 Foliation that develops when minerals are squeezed and deform by lengthening in the direction perpendicular to the greatest stress (indicated by black arrows). Left- before squeezing. Right- after squeezing. Source: Steven Earle (2015) CC BY 4.0 view source

If a rock is both heated and squeezed during metamorphism, and the temperature change is enough for new minerals to form from existing ones, the new minerals can be forced to grow longer perpendicular to the direction of squeezing (Figure 10.7). If the original rock had bedding (represented by diagonal lines in Figure 10.7, right), foliation may obscure the bedding.

Figure 10.7 Effects of squeezing and aligned mineral growth during metamorphism. Left: Protolith with diagonal bedding. Right: Metamorphic rock derived from the protolith. Elongated mica crystals grew perpendicular to the main stress direction. The original bedding is obscured. Source: Steven Earle (2015) CC BY 4.0 view source

This is not always the case, however. The large boulder in Figure 10.8 in has strong foliation, oriented nearly horizontally in this view, but it also has bedding still visible as dark and light bands sloping steeply down to the right.

Figure 10.8 A geologists sits on a rock that has foliation (marked by the dashed line that is nearly horizontal), and still retains evidence of the original bedding (steeply dipping dashed line). The rock has undergone a relatively low degree of metamorphism, which is why the bedding is still visible. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

Foliation and Crystal Habit

Most foliation develops when new minerals are forced to grow perpendicular to the direction of greatest stress. This effect is especially strong if the new minerals grow in platy or elongated shapes. The rock in the upper left of Figure 10.9 is foliated, and the microscopic structure of the same type of foliated rock is shown in the photograph beneath it. Over all, the photomicrograph shows that the rock is dominated by elongated crystals aligned in bands running from the upper left to the lower right. The stress that produced this pattern was greatest in the direction indicated by the black arrows, at a right angle to the orientation of the minerals. The aligned minerals are mostly mica, which has a platy crystal habit, with plates stacked together like pages in a book.

Figure 10.9 A foliated metamorphic rock called phyllite (upper left). The satin sheen comes from the alignment of minerals. Lower left- a view of the same kind of rock under a microscope showing mica crystals (colourful under polarized light) aligned in bands. The region outlined in a red dashed line shows a lens of quartz crystals that do not display alignment. Upper right- stacks of platy mica crystals. Lower right- a blocky quartz crystal. Source: Karla Panchuk (2018) CC BY-SA 4.0. Click the image for photo sources.

The zone in the photomicrograph outlined with the red dashed line is different from the rest of the rock. Not only is the mineral composition different—it is quartz, not mica—but the crystals are not aligned. The quartz crystals were subjected to the same stress as the mica crystals, but because quartz grows in blocky shapes rather than elongated ones, the crystals could not be aligned in any one direction.

Even though the quartz crystals themselves are not aligned, the mass of quartz crystals forms a lens that does follow the general trend of alignment within the rock. This happens because the stress can cause some parts of the quartz crystals to dissolve, and the resulting ions flow away at right angles to the greatest stress before forming crystals again.

The effects of recrystallization in Figure 10.9 would not be visible with the unaided eye, but when larger crystals or large clasts are involved, the effects can be visible as “shadows” or “wings” around crystals and clasts. The rock in Figure 10.10 had a quartz-rich conglomerate as a parent rock. Differential stress has caused quartz pebbles within the rock to become elongated, and it has also caused wings to form around some of the pebbles (see the pebble in the dashed ellipse). The location of the wings depends on the distribution of stress on the rock (Figure 10.10, upper right).

Figure 10.10 Metaconglomerate with elongated of quartz pebbles. The pebbles have developed “wings” to varying degrees (e.g., white dashed ellipse). These are the result of quartz dissolving where stress is applied, and flowing away from the direction of maximum stress before recrystallizing (upper right sketch). Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo by R. Weller/ Cochise College view source. Click the image to view terms of use.

Foliation Controls How Rocks Break

Foliated metamorphic rocks have elongated crystals that are oriented in a preferred direction. This forms planes of weakness, and when these rocks break, they tend to break along surfaces that parallel the orientation of the aligned minerals (Figure 10.11). Breaks along planes of weakness within a rock that are caused by foliation are referred to as rock cleavage, or just cleavage.  This is distinct from cleavage in minerals because mineral cleavage happens between atoms within a mineral, but rock cleavage happens between minerals.

Figure 10.11 Close-up view of a metamorphic rock with aligned elongated crystals. The crystals control the shape of the break in the rock (black gap), resulting in breaks occurring along parallel surfaces. Source: Karla Panchuk (2018) CC BY 4.0

The mineral alignment in the metamorphic rock called slate is what causes it to break into flat pieces (Figure 10.12, left), and is why slate has been used as a roofing material (Figure 10.12, right). The tendency of slate to break into flat pieces is called slaty cleavage.

Figure 10.12 Rock cleavage in the fine-grained metamorphic rock called slate results in breaks along relatively flat surfaces (left). This is why slate has been used for roofing material (right). Source: Left- Roger Kidd (2008) CC BY-SA 2.0 view source; Right- Michael C. Rygel (2007) CC BY-SA 3.0 view source

Rock cleavage is what caused the boulder in Figure 10.8 to split from bedrock in a way that left the flat upper surface upon which the geologist is sitting.

66

10.3 Classification of Metamorphic Rocks

Metamorphic rocks are broadly classified as foliated or non-foliated. Non-foliated metamorphic rocks do not have aligned mineral crystals. Non-foliated rocks form when pressure is uniform, or near the surface where pressure is very low. They can also form when the parent rock consists of blocky minerals such as quartz and calcite, in which individual crystals do not align because they aren’t longer in any one dimension. This distinction breaks down in zones of intense deformation, where even minerals like quartz can be squeezed into long stringers, much like squeezing toothpaste out of a tube (Figure 10.13).

Figure 10.13 Rocks from the Western Carpathians mountain range without deformation (left) and after deformation (right). Scale bar: 1 mm. Left- An undeformed granitic rock containing the mica mineral biotite (Bt), plagioclase feldspar (Pl), potassium feldspar (Kfs), and quartz (Qtz). Right- A metamorphic rock (mylonite) resulting from extreme deformation of granitic rocks. Quartz crystals have been flattened and deformed. The other minerals have been crushed and deformed into a fine-grained matrix (Mtx). Source: Farkašovský et al. (2016) CC BY-NC-ND. Click the image to view the original figure captions and access the full text.

Types of Foliated Metamorphic Rocks

Four common types of foliated metamorphic rocks, listed in order of metamorphic grade or intensity of metamorphism are slate, phyllite, schist (pronounced “shist”), and gneiss (pronounced “nice”). Each of these has a characteristic type of foliation

Slate

Slate (Figure 10.14) forms from the low-grade metamorphism of shale. Slate has microscopic clay and mica crystals that have grown perpendicular to the maximum stress direction. Slate tends to break into flat sheets or plates, a property described as slaty cleavage.

Figure 10.14 Slate, a low-grade foliated metamorphic rock. Left- Slate fragments resulting from rock cleavage. Right- The same rock type in outcrop. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photos: Left- Vincent Anciaux (2005) CC BY-SA 3.0 view source; Right- Gretarsson (2006) CC BY-SA 3.0 view source

Phyllite

Phyllite (Figure 10.15) is similar to slate, but has typically been heated to a higher temperature. As a result, the micas have grown larger.  They still are not visible as individual crystals, but the larger size leads to a satiny sheen on the surface.  The cleavage of phyllite is slightly wavy compared to that of slate.

Figure 10.15 Phyllite, a fine-grained foliated metamorphic rock. Left- A hand sample showing a satin texture. Right- The same rock type in outcrop in the city of Sopron, Hungary. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photos: Left- Chadmull (2006) Public Domain view source; Right- Laszlovszky András (2008) CC BY-SA 2.5 view source

Schist

Schist (Figure 10.16) forms at higher temperatures and pressures and exhibits mica crystals that are large enough to see without magnification. Individual crystal faces may flash when the sample is turned in the light, making the rock appear to sparkle. Other minerals such as garnet might also be visible, but it is not unusual to find that schist consists predominantly of a single mineral.

Figure 10.16 Schist, a medium- to high-grade foliated metamorphic rock. Top- Hand sample showing light reflecting off of mica crystals. Bottom- Close-up view of mica crystals and garnet. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photos by R. Weller/ Cochise College. Click the image for photo sources and terms of use.

Gneiss

Gneiss (Figure 10.17) forms at the highest pressures and temperatures, and has crystals large enough to see with the unaided eye. Gneiss features minerals that have separated into bands of different colours. The bands of colours are what define foliation within gneiss. Sometimes the bands are very obvious and continuous (Figure 10.17, upper right), but sometimes they are more like lenses (upper left). Dark bands are largely amphibole while the light-coloured bands are feldspar and quartz. Most gneiss has little or no mica because it forms at temperatures higher than those under which micas are stable.

Figure 10.17 Gneiss, a coarse-grained, high grade metamorphic rock, is characterized by colour bands. Top- Hand samples showing that colour bands can be continuous (left) or less so (right). Bottom- Gneiss in outcrop at Belteviga Bay, Norway. Notice the light and dark stripes on the rock. Source: Karla Panchuk (2018) CC BY-SA 4.0. Click the image for more attributions.

While slate and phyllite typically form only from mudrock protoliths, schist and especially gneiss can form from a variety of parent rocks, including mudrock, sandstone, conglomerate, and a range of both volcanic and intrusive igneous rocks.

Schist and gneiss can be named on the basis of important minerals that are present: a schist derived from basalt is typically rich in the mineral chlorite, so we call it chlorite schist. One derived from shale may be a muscovite-biotite schist, or just a mica schist, or if there are garnets present it might be mica-garnet schist. Similarly, gneiss that originated as basalt and is dominated by amphibole, is an amphibole gneiss or amphibolite (Figure 10.18).

Figure 10.18 Amphibolite in thin section (2mm field of view), derived from metamorphism of a mafic igneous rock. Green crystals are the amphibole hornblende, and colourless crystals are plagioclase feldspar. Note horizontal crystal alignment. Source: D.J. Waters, University of Oxford view source/ view context. Click the image for original figure caption and terms of use.

Types of Non-foliated Metamorphic Rocks

Metamorphic rocks that form under low-pressure conditions or under the effects confining pressure, which is equal in all directions, do not become foliated. In most cases, this is because they are not buried deeply enough, and the heat for the metamorphism comes from a body of magma that has moved into the upper part of the crust. Metamorphism that happens because of proximity to magma is called contact metamorphism. Some examples of non-foliated metamorphic rocks are marble, quartzite, and hornfels.

Marble

Marble (Figure 10.19) is metamorphosed limestone. When it forms, the calcite crystals recrystallize (re-form into larger blocky calcite crystals), and any sedimentary textures and fossils that might have been present are destroyed. If the original limestone is pure calcite, then the marble will be white.  On the other hand, if it has impurities such as clay, silica, or magnesium, the marble could be “marbled” in appearance (Figure 10.19, bottom).

Figure 10.19 Marble is a non-foliated metamorphic rock with a limestone protolith. Left- Marble made of pure calcite is white. Upper right- microscope view of calcite crystals within marble that are blocky and not aligned. Lower right- A quarry wall showing the “marbling” that results when limestone contains components other than calcite. Source: Karla Panchuk (2018) CC BY-NC-SA. Click the image for more attributions.

Quartzite

Quartzite (Figure 10.20) is metamorphosed sandstone. It is dominated by quartz, and in many cases, the original quartz grains of the sandstone are welded together with additional silica. Sandstone often contains some clay minerals, feldspar or lithic fragments, so quartzite can also contain impurities.

Figure 10.20 Quartzite is a non-foliated metamorphic rock with a sandstone protolith. Left- Quartzite from the Baraboo Range, Wisconsin. Right- Photomicrograph showing quartz grains in quartzite from the Southern Appalachians. In the upper left half of the image, blocky quartz crystals show some evidence of alignment running from the upper right to the lower left. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photomicrograph: Geologian (2011) CC BY-SA 3.0 view source

Even if formed under directed pressure, quartzite is generally not foliated because quartz crystals do not normally align with the directional pressure. On the other hand, any clay present in the original sandstone is likely to be converted to mica during metamorphism, and any such mica is likely to align with the directional pressure.

Hornfels

Hornfels is another non-foliated metamorphic rock that normally forms during contact metamorphism of fine-grained rocks like mudstone or volcanic rocks. Hornfels have different elongated or platy minerals (e.g., micas, pyroxene, amphibole, and others) depending on the exact conditions and the parent rock, yet because the pressure wasn’t substantially higher in any particular direction, these crystals remain randomly oriented.

The hornfels in Figure 10.21 (left) appears to have gneiss-like bands, but these actually reflect the beds of alternating sandstone and shale that were in the protolith. They are not related to alignment of crystals due to metamorphism. On the right of Figure 10.21 is a microscopic view of another sample of hornfels, also from a sedimentary protolith. The dark band at the top is from the original bedding.  Here you can see that the brown mica crystals (biotite) are not aligned.

Figure 10.21 Hornfels, a non-foliated metamorphic rock formed from a fine-grained protolith. Left- Hornfels from the Novosibirsk region of Russia from a sedimentary protolith. Dark and light bands preserve the bedding of the original sedimentary rock. The rock has been recrystallized during contact metamorphism and does not display foliation. (scale in cm). Right- Hornfels in thin section from a sedimentary protolith. Note that the brown mica crystals are not aligned. The dark band at the top reflects the layering within the sedimentary parent rock, similar to the way those layers are preserved in the sample on the left. Source: Left- Fedor (2006) Public Domain view source; Right- D.J. Waters, University of Oxford view source/ view context. Click the image for terms of use.

What Happens When Different Rocks Undergo Metamorphism?

The nature of the parent rock controls the types of metamorphic rocks that can form from it under differing metamorphic conditions (temperature, pressure, fluids). The kinds of rocks that can be expected to form at different metamorphic grades from various parent rocks are listed in Table 10.1.

Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source. Click the table for a text version.

Some rocks, such as granite, do not change much at the lower metamorphic grades because their minerals are still stable up to several hundred degrees. Sandstone and limestone don’t change much either because their metamorphic forms (quartzite and marble, respectively) have the same mineral composition, but re-formed larger crystals.

On the other hand, some rocks can change substantially.  Mudrock (e.g., shale, mudstone) can start out as slate, then progress through phyllite, schist, and gneiss, with a variety of different minerals forming along the way.  Schist and gneiss can also form from sandstone, conglomerate, and a range of both volcanic and intrusive igneous rocks.

Migmatite: Both Metamorphic and Igneous

If a metamorphic rock is heated enough, it can begin to undergo partial melting in the same way that igneous rocks do.  The more felsic minerals (feldspar, quartz) will melt, while the more mafic minerals (biotite, hornblende) do not.  When the melt crystallizes again, the result is light-coloured igneous rock interspersed with dark-coloured metamorphic rock.  This mixed rock is called migmatite (Figure 10.22). Note that the foliation present in the metamorphic rock is no longer present in the igneous rock. Liquids cannot support a differential stress, so when the melt crystallizes, the foliation is gone.

Figure 10.22 Migmatite photographed near Geirangerfjord in Norway. Source: Siim Sepp (2006) CC BY-SA 3.0 view source

A fascinating characteristic of migmatites is ptygmatic (pronounced “tigmatic“) folding. These are folds look like they should be impossible because they are enveloped by rock which does not display the same complex deformation (Figure 10.23).  How could those wiggly folds get in there without the rest of the rock being folded in the same way?

Figure 10.23 Ptygmatic folding from Broken Hill, New South Wales, Australia. Ptygmatic folding happens when a stiff layer within a rock is surrounded by weaker layers. Folding causes the stiff layer to crinkle while the weaker layers deform around it. Source: Roberto Weinberg (http://users.monash.edu.au/~weinberg) view source. Click the image for terms of use.

The answer to the ptygmatic fold mystery is that the folded layer is much stiffer than the surrounding layers.  When squeezing forces act on the rock, the stiff layer buckles but the surrounding rock flows rather than buckling, because it isn’t strong enough to buckle.

Exercise: Naming Metamorphic Rocks

Which metamorphic rock is described in each of the following?

  1. A rock with visible minerals of mica and with small crystals of andalusite. The mica crystals are consistently parallel to one another.
  2. A very hard rock with a granular appearance and a glassy lustre. There is no evidence of foliation.
  3. A fine-grained rock that splits into wavy sheets. The surfaces of the sheets have a sheen to them.
  4. A rock that is dominated by aligned crystals of amphibole.

References

Farkašovský, R., Bónová, K., & Košuth, M. (2016). Microstructural, modal and geochemical changes as a result of granodiorite mylonitisation – a case study from the Rolovská shear zone (Čierna hora Mts, Western Carpathians, Slovakia). Geologos 22(3), 171-190. doi: 10.1515/logos-2016-0019 View full text

67

10.4 Types of Metamorphism and Where They Occur

The outcome of metamorphism depends on pressure, temperature, and the abundance of fluid involved, and there are many settings with unique combinations of these factors. Some types of metamorphism are characteristic of specific plate tectonic settings, but others are not.

Burial Metamorphism

Burial metamorphism occurs when sediments are buried deeply enough that the heat and pressure cause minerals to begin to recrystallize and new minerals to grow, but does not leave the rock with a foliated appearance. As metamorphic processes go, burial metamorphism takes place at relatively low temperatures (up to ~300 °C) and pressures (100s of m depth). To the unaided eye, metamorphic changes may not be apparent at all. Contrast the rock known commercially as Black Marinace Gold Granite (Figure 10.24)—but which is in fact a metaconglomerate—with the metaconglomerate in Figure 10.10. The metaconglomerate formed through burial metamorphism does not display any of the foliation that has developed in the metaconglomerate in Figure 10.10.

Figure 10.24 Metaconglomerate formed through burial metamorphism. The pebbles in this sample are not aligned and elongated as in the metaconglomerate in Figure 10.10. Source: James St. John (2014) CC BY 2.0 view source

A Note About Commercial Rock Names

Names given to rocks that are sold as building materials, especially for countertops, may not reflect the actual rock type. It is common to use the terms granite and marble to describe rocks that are neither. While these terms might not provide accurate information about the rock type, they generally do distinguish natural rock from synthetic materials. An example of a synthetic material is the one referred to as quartz, which includes ground-up quartz crystals as well as resin. If you happen to be in the market for stone countertops and are concerned about getting a natural product, it is best to ask lots of questions.

Regional Metamorphism

Regional metamorphism refers to large-scale metamorphism, such as what happens to continental crust along convergent tectonic margins (where plates collide).  The collisions result in the formation of long mountain ranges, like those along the western coast of North America.  The force of the collision causes rocks to be folded, broken, and stacked on each other, so not only is there the squeezing force from the collision, but from the weight of stacked rocks. The deeper rocks are within the stack, the higher the pressures and temperatures, and the higher the grade of metamorphism that occurs. Rocks that form from regional metamorphism are likely to be foliated because of the strong directional pressure of converging plates.

The Himalaya range is an example of where regional metamorphism is happening because two continents are colliding (Figure 10.25). Sedimentary rocks have been both thrust up to great heights—nearly 9 km above sea level—and also buried to great depths. Considering that the normal geothermal gradient (the rate of increase in temperature with depth) is around 30°C per kilometre in the crust, rock buried to 9 km below sea level in this situation could be close to 18 km below the surface of the ground, and it is reasonable to expect temperatures up to 500°C. Notice the sequence of rocks that from, beginning with slate higher up where pressures and temperatures are lower, and ending in migmatite at the bottom where temperatures are so high that some of the minerals start to melt. These rocks are all foliated because of the strong compressing force of the converging plates.

Figure 10.25 Regional metamorphism beneath a mountain range resulting from continent-continent collision. Arrows show the forces due to the collision. Dashed lines represent temperatures that would exist given a geothermal gradient of 30 ºC/km. A sequence of foliated metamorphic rocks of increasing metamorphic grade forms at increasing depths within the mountains. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

Seafloor (Hydrothermal) Metamorphism

At an oceanic spreading ridge, recently formed oceanic crust of gabbro and basalt is slowly moving away from the plate boundary (Figure 10.26). Water within the crust is forced to rise in the area close to the source of volcanic heat, drawing in more water from further away. This eventually creates a convective system where cold seawater is drawn into the crust, heated to 200 °C to 300 °C as it passes through the crust, and then released again onto the seafloor near the ridge.

Figure 10.26 Seafloor (hydrothermal) metamorphism of ocean crustal rock on either side of a spreading ridge. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

The passage of this water through the oceanic crust at these temperatuers promotes metamorphic reactions that change the original olivine and pyroxene minerals in the rock to chlorite ((Mg5Al)(AlSi3)O10(OH)8) and serpentine ((Mg, Fe)3Si2O5(OH)4). Chlorite and serpentine are both hydrated minerals, containing water in the form of OH in their crystal structures. When metamorphosed ocean crust is later subducted, the chlorite and serpentine are converted into new non-hydrous minerals (e.g., garnet and pyroxene) and the water that is released migrates into the overlying mantle, where it contributes to melting.

The low-grade metamorphism occurring at these relatively low pressures and temperatures can turn mafic igneous rocks in ocean crust into greenstone (Figure 10.27), a non-foliated metamorphic rock.

Figure 10.27 Greenstone from the metamorphism of seafloor basalt that took place 2.7 billion years ago. The sample is from the Upper Peninsula of Michigan, USA. Source: James St. John (2012) CC BY 2.0 view source

Subduction Zone Metamorphism

At subduction zones, where ocean lithosphere is forced down into the hot mantle, there is a unique combination of relatively low temperatures and very high pressures.  The high pressures are to be expected, given the force of collision between tectonic plates, and the increasing lithostatic pressure as the subducting slab is forced deeper and deeper into the mantle. The lower temperatures exist because even though the mantle is very hot, ocean lithosphere is relatively cool, and a poor conductor of heat. That means it will take a long time to heat up, can be several hundreds of degrees cooler than the surrounding mantle. In Figure 10.28, notice that the isotherms (lines of equal temperature, dashed lines) plunge deep into the mantle along with the subducting slab, showing that regions of relatively low temperature exist deeper in the mantle.

Figure 10.28 Regional metamorphism of oceanic crust at a subduction zone occurs at high pressure but relatively low temperatures. Source: Steven Earle (2015) CC BY 4.0 view source

A special type of metamorphism takes place under these very high-pressure but relatively low-temperature conditions, producing an amphibole mineral known as glaucophane (Na2(Mg3Al2)Si8O22(OH)2).  Glaucophane is blue, and the major component of a rock known as blueschist. If you have never seen or even heard of blueschist, that not surprising. What is surprising is that anyone has seen it! Most of the blueschist that forms in subduction zones continues to be subducted. It turns into eclogite at about 35 km depth, and then eventually sinks deep into the mantle, never to be seen again. In only a few places in the world, the subduction process was interrupted, and partially subducted blueschist returned to the surface. One such place is the area around San Francisco. The blueschist at this location is part of a set of rocks known as the Franciscan Complex (Figure 10.29).

Figure 10.29 Franciscan Complex blueschist exposed north of San Francisco. The blue colour of the rock is due to the presence of the amphibole mineral glaucophane. Source: Steven Earle (2015) CC BY 4.0 view source

Contact Metamorphism

Contact metamorphism happens when a body of magma intrudes into the upper part of the crust. Heat is important in contact metamorphism, but pressure is not a key factor, so contact metamorphism produces non-foliated metamorphic rocks such as hornfels, marble, and quartzite.

Any type of magma body can lead to contact metamorphism, from a thin dyke to a large stock. The type and intensity of the metamorphism, and width of the metamorphic aureole that develops around the magma body, will depend on a number of factors, including the type of country rock, the temperature of the intruding body, the size of the body, and the volatile compounds within the body (Figure 10.30). A large intrusion will contain more thermal energy and will cool much more slowly than a small one, and therefore will provide a longer time and more heat for metamorphism. This will allow the heat to extend farther into the country rock, creating a larger aureole. Volatiles may exsolve from the intruding melt and travel into the country rock, facilitating heating and carrying chemical constituents from the melt into the rock. Thus, aureoles that form around “wet” intrusions tend to be larger than those forming around their dry counterparts.

Figure 10.30 Schematic cross-section of the middle and upper crust showing two magma bodies. The upper body, which has intruded into cool unmetamorphosed rock, has created a zone of contact metamorphism. The lower body is surrounded by rock that is already hot (and probably already metamorphosed), and so it does not have a significant metamorphic aureole. Source: Steven Earle (2015) CC BY 4.0 view source

Contact metamorphic aureoles are typically quite small, from just a few centimetres around small dykes and sills, to as much as 100 m around a large stock. Contact metamorphism can take place over a wide range of temperatures—from around 300 °C to over 800 °C. Different minerals will form depending on the exact temperature and the nature of the country rock.

Although bodies of magma can form in a variety of settings, one place magma is produced in abundance, and where contact metamorphism can take place, is along convergent boundaries with subduction zones, where volcanic arcs form (Figure 10.31). Regional metamorphism also takes place in this setting, and because of the extra heat associated with the magmatic activity, the geothermal gradient is typically steeper in these settings (between ~40 and 50 °C/km). Under these conditions, higher grades of metamorphism can take place closer to surface than is the case in other areas.

Figure 10.31 Contact metamorphism (yellow rind) around a high-level crustal magma chamber, and regional metamorphism in a volcanic-arc related mountain range. Dashed lines show isotherms. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

Shock Metamorphism

When extraterrestrial objects hit Earth, the result is a shock wave.  Where the object hits, pressures and temperatures become very high in a fraction of a second.  A “gentle” impact can hit with 40 GPa and raise temperatures up to 500 °C. Pressures in the lower mantle start at 24 GPa (GigaPascals), and climb to 136 GPa at the core-mantle boundary, so the impact is like plunging the rock deep into the mantle and releasing it again within seconds.  The sudden change associated with shock metamorphism makes it very different from other types of metamorphism that can develop over hundreds of millions of years, starting and stopping as tectonic conditions change.

Two features of shock metamorphism are shocked quartz, and shatter cones.  Shocked quartz (Figure 10.32 left) refers to quartz crystals that display damage in the form of parallel lines throughout a crystal.  The quartz crystal in Figure 10.32 has two sets of these lines.  The lines are small amounts of glassy material within the quartz, formed from almost instantaneous melting and resolidification when the crystal was hit by a shock wave. Shatter cones are cone-shaped fractures within the rocks, also the result of a shock wave (Figure 10.32 right).  The fractures are nested together like a stack of ice-cream cones.

Figure 10.32 Shock metamorphism features. Left- Shocked quartz with lines of glassy material, from the Suvasvesi South impact structure in Finland. Right- Shatter cones from the Wells Creek impact crater in the USA. Sources: Left- Martin Schmieder CC BY 3.0 view source. Right- Zamphuor (2007) Public Domain view source.

Dynamic Metamorphism

Dynamic metamorphism is the result of very high shear stress, such as occurs along fault zones. Dynamic metamorphism occurs at relatively low temperatures compared to other types of metamorphism, and consists predominantly of the physical changes that happen to a rock experiencing shear stress. It affects a narrow region near the fault, and rocks nearby may appear unaffected.

At lower pressures and temperatures, dynamic metamorphism will have the effect of breaking and grinding rock, creating cataclastic rocks such as fault breccia (Figure 10.33). At higher pressures and temperatures, grains and crystals in the rock may deform without breaking into pieces (Figure 10.34, left). The outcome of prolonged dynamic metamorphism under these conditions is a rock called mylonite, in which crystals have been stretched into thin ribbons (Figure 10.34, right).

Figure 10.33 Fault breccia, created when shear stress along a fault breaks up rocks. Left- close-up view of fault breccia clearly showing dark angular fragments. Right- A fault-zone containing fragments broken from the adjacent walls (dashed lines). Note that the deformation does not extend far past the margins of the fault zone. Source: Karla Panchuk (2018) CC BY 4.0. Click the image for more attributions.

 

Figure 10.34 Mylonite, a rock formed by dynamic metamorphism. Left- An outcrop showing the early stages of mylonite development, called protomylonite. Notice that the deformation does not extend to the rock at the bottom of the photograph. Middle- Mylonite showing ribbons formed of drawn-out crystals. Right- Microscope view of mylonite with mica (colourful crystals) and quartz (grey and black crystals). This is a case where the shape of quartz crystals is controlled more by stress than by crystal habit. Source: Karla Panchuk (2018) CC BY-SA 4.0. Click the image for more attributions.

References

Bucher, K., & Grapes, R. (2011) Petrogenesis of Metamorphic Rocks, 8th Edition. Springer.

French, B.M. (1998). Traces of Catastrophe: A Handbook of Shock-Metamorphic Effects in Terrestrial Meteorite Impact Structures. Houston, TX: Lunar and Planetary Institute  Read full text

68

10.5 Metamorphic Facies and Index Minerals

Metamorphic Facies

In any given metamorphic setting there can be a variety of protolith types exposed to metamorphism.  While these rocks will be exposed to the same range of pressure and temperatures conditions within that setting, the metamorphic rock that results will depend on the protolith. A convenient way to indicate the range of possible metamorphic rocks in a particular setting is to group those possibilities into metamorphic facies. In other words, a given metamorphic facies groups together metamorphic rocks that form under the same pressure and temperature conditions, but which have different protoliths.

Figure 10.35 shows the different metamorphic facies as patches of different colours. The axes on the diagram are temperature and depth; the depth within the Earth will determine how much pressure a rock is under, so the vertical depth axis is also a pressure axis. Each patch of colour represents a range of temperature and pressure conditions where particular types of metamorphic rocks will form. Metamorphic facies are named for rocks that form under specific conditions (e.g., eclogite facies, amphibolite facies etc.), but those names don’t mean that the facies is limited to that one rock type.

Figure 10.35 Metamorphic facies and types of metamorphism shown in the context of depth and temperature. The metamorphic rocks formed from a mudrock protolith under regional metamorphism with a typical geothermal gradient are listed. Letters correspond to the types of metamorphism shown in Figure 10.36. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source

Another feature to notice in the diagram are the many dashed lines. The yellow, green, and blue dashed lines represent the geothermal gradients in different environments. Recall that the geothermal gradient describes how rapidly the temperature increases with depth in Earth. In most areas (green dashed line), the rate of increase in temperature with depth is 30 °C/km. In other words, if you go 1,000 m down into a mine, the temperature will be roughly 30 °C warmer than the average temperature at the surface.  In volcanic areas (yellow dashed line), the geothermal gradient is more like 40 to 50 °C/km, so the temperature rises much faster as you go down. Along subduction zones (blue dashed line), the cold ocean lithosphere keeps temperatures low, so the gradient is typically less than 10 °C/km.

The yellow, green, and blue dashed lines in Figure 10.35 tell you what metamorphic facies you will encounter for rocks from a given depth in that particular environment. A depth of 15 km in a volcanic region falls in the amphibolite facies.  Under more typical conditions, this is the greenschist facies, and in a subduction zone it is the blueschist facies. You can make the connection more directly between the metamorphic facies and the types of metamorphism discussed in the previous section by matching up the letters a through e in Figure 10.35 with the labels in Figure 10.36.

 

Figure 10.36 Environments of metamorphism in the context of plate tectonics: (a) regional metamorphism related to mountain building at a continent-continent convergent boundary, (b) seafloor (hydrothermal) metamorphism of oceanic crust in the area on either side of a spreading ridge, (c) metamorphism of oceanic crustal rocks within a subduction zone, (d) contact metamorphism adjacent to a magma body at a high level in the crust, and (e) regional metamorphism related to mountain building at a convergent boundary. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

One other line to notice in Figure 10.35 is the red dashed line on the right-hand side of the figure. This line represents temperatures and pressures where granite will begin to melt if there is water present. Migmatite is to the right of the line because it forms when some of the minerals in a metamorphic rock begin to melt, and then cool and crystallize again.

Exercise: Metamorphic Rocks In Areas with Higher Geothermal Gradients

Figure 10.35 shows the types of rock that might form from mudrock at various points along the curve of the “typical” geothermal gradient (dotted green line). Looking at the geothermal gradient for volcanic regions (dotted yellow line), estimate the depths at which you would expect to find each of those rocks forming from a mudrock parent.

Index Minerals

Some common minerals in metamorphic rocks are shown in Figure 10.37, arranged in order of the temperature ranges within which they tend to be stable. The upper and lower limits of the ranges are intentionally vague because these limits depend on a number of different factors, such as the pressure, the amount of water present, and the overall composition of the rock.

Figure 10.37 Metamorphic index minerals and approximate temperature ranges. Source: Steven Earle (2015) CC BY 4.0 view source

Even though the limits of the stability ranges are vague, the stability range of each mineral is still small enough that the minerals can be used as markers for those metamorphic conditions. Minerals that make good markers of specific ranges of metamorphic conditions are called index minerals.

The Meguma Terrane of Nova Scotia: An Example of How to Use Index Minerals

The southern and southwestern parts of Nova Scotia were regionally metamorphosed during the Devonian Acadian Orogeny (around 400 Ma), when a relatively small continental block—the Meguma Terrane (Figure 10.38, top )—collided with the existing eastern margin of North America. The clastic sedimentary rocks within this terrane were variably metamorphosed.

Figure 10.38 Regional metamorphic zones in the Meguma Terrane of southwestern Nova Scotia. Top- Map of metamorphic zones. Bottom- Stability ranges for minerals within the Meguma Terrane. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view source, Keppie & Muecke (1979) and White & Barr (2012).

Index minerals have been used to map areas of higher or lower metamorphic intensity, called metamorphic zones. A metamorphic zone is a region bounded by the first appearance of an index mineral. In the Meguma Terrane, the biotite zone (Figure 10.38, darker green) begins in the east and north with the first appearance of biotite. The biotite zone ends toward the south and west where garnet first appears. Because index minerals can have overlapping stability conditions, a lower-intensity index mineral can still be present in a higher-intensity metamorphic zone.

Knowledge of metamorphic zones makes it possible to draw conclusions about the geological conditions in which metamorphic rocks formed. The highest-intensity metamorphism—the sillimanite zone—is in the southwest. Progressively lower grades of metamorphism exist toward the east and north. The rocks of the sillimanite zone were likely heated to over 700 °C, and therefore must have been buried to depths between 20 km and 25 km. The surrounding lower-grade rocks were not buried as deeply, and the rocks within the peripheral chlorite zone were likely not buried to more than about 5 km depth.

A probable explanation for this pattern is that the area with the highest-grade rocks was buried beneath the central part of a mountain range formed by the collision of the Meguma Terrane with North America. The collision caused rocks to be folded, and to be faulted and stacked on top of each other. These mountain-building processes thickened Earth’s crust, and increased its mass locally as the mountains grew. The increased mass of the growing mountains caused the lithosphere to float lower in the mantle (Figure 10.39, left). As the mountains were eventually eroded over tens of millions of years, the crust floated higher and higher in the mantle, and erosion exposed metamorphic rocks that were deep within the mountains (Figure 10.39, right).

Figure 10.39 Schematic cross-section through the Meguma Terrane. Left- Metamorphic zones and temperatures when mountain-building processes thickened the crust. Right- The mountains have been eroded, exposing metamorphic rocks that formed deep within the mountains. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source left/ right.

Building a narrative for the metamorphism in Nova Scotia’s Meguma Terrane is just one example of how index minerals can be used.

Exercise: Scottish Metamorphic Zones

The map in Figure 10.40 shows part of western Scotland between the Great Glen Fault and the Highland Boundary Fault. The shaded areas are metamorphic rock, and the three metamorphic zones represented are garnet, chlorite, and biotite.

  1. Label the three coloured areas of the map with the appropriate zone names (garnet, chlorite, and biotite).
  2. Indicate which part of the region was likely to have been buried the deepest during metamorphism.

British Geologist George Barrow studied this area in the 1890s and was the first person anywhere to map metamorphic zones based on their mineral assemblages. This pattern of metamorphism is sometimes referred to as Barrovian metamorphism.

Figure 10.40 Metamorphic zones in Barrovian metamorphism. Source: Steven Earle (2015) CC BY 4.0 view source

References

Keppie, D., & Muecke, G. (1979). Metamorphic map of Nova Scotia. (Nova Scotia Department of Mines and Energy, Map 1979-006).

White, C. E., & Barr, S. M. (2012) Meguma Terrane Revisted: Stratigraphy, Metamorphism, Paleontology and Provenance. Geoscience Canada 39(1). Full text

69

10.6 Metamorphic Hydrothermal Processes and Metasomatism

The heat from a body of magma in the upper crust can create a very dynamic situation with geologically interesting and economically important implications. In the simplest cases, water does not play a big role, and the main process is heat transfer from the pluton to the surrounding rock, creating a zone of contact metamorphism (Figure 10.41a). In many cases, however, water is released from the magma body as crystallization takes place, and this water is dispersed along fractures in the surrounding rock (Figure 10.41b). The water released from a magma chamber is typically rich in dissolved minerals. As this water cools, it interacts with the surrounding rocks, changing both the chemistry of the water and the chemistry of the rocks. This can cause minerals to precipitate from the water. Minerals can also precipitate if the water boils because of a drop in pressure. The precipitated minerals form veins within fractures in the surrounding rock. Quartz veins are commonly formed in this situation, and can include other minerals such as pyrite, hematite, calcite, and even silver and gold.

Figure 10.41 Metamorphism and alteration around a pluton in the upper crust. (a) Thermal metamorphism only (within the purple zone); (b) Thermal metamorphism plus veining (white) related to dispersal of magmatic fluids into the overlying rock; (c) Thermal metamorphism plus veining from magmatic fluids plus alteration and possible formation of metallic minerals (hatched yellow areas) from convection of groundwater. Source: Steven Earle (2015) CC BY 4.0 view source

Heat from the magma body will cause surrounding groundwater to expand and then rise toward the surface. In some cases, this may initiate a convection system where groundwater circulates past the pluton. Such a system could operate for thousands of years, resulting in the circulation of millions of tonnes of groundwater from the surrounding region past the pluton.

Hot water circulating through the rocks and interacting chemically with them can lead to significant changes in the mineralogy of the rock, including alteration of feldspars to clays, and deposition of quartz, calcite, and other minerals in fractures and other open spaces (Figure 10.42). Chemical change in rocks due to interaction with hot water is called hydrothermal alteration.

Figure 10.42 White veins of calcite in limestone of the Comox Formation, Nanaimo BC. Quarter for scale. Source: Steven Earle (2016) CC BY 4.0 view source

Metamorphic reactions involve the release of fluids as minerals change, and chemical reactions with locally-derived fluids. However, if a large amount of externally-derived fluid—such that supplied by magma—is flushed through the system at the high pressures and temperatures characteristic of metamorphism, it can substantially alter the chemical composition of the rock. This type of hydrothermal alteration is called metasomatism.

A special type of metasomatism takes place where a hot pluton intrudes into carbonate rock such as limestone. Magmatic fluids rich in silica, calcium, magnesium, iron, and other elements can dramatically change the chemistry of the limestone, forming minerals that would not normally exist in either the igneous rock or limestone. A rock called skarn results, containing minerals such as garnet, epidote, magnetite, and pyroxene, among others (Figure 10.43).

Figure 10.43 Skarn from Mount Monzoni, Northern Italy, with recrystallized calcite (blue), garnet (brown), and pyroxene (green). The rock is 6 cm across. Source: Siim Sepp (2012) CC BY-SA 3.0 view source

 

Exercise: Contact Metamorphism and Metasomatism

A pluton that has intruded into a series of sedimentary rocks, including sandstone, mudstone, and limestone (Figure 10.44). What types of metamorphic rocks would you expect to see at locations a, b, and c?

Figure 10.44 Contact metamorphism and metasomatism of sedimentary rocks. Source: Steven Earle (2015) CC BY 4.0 view source

70

Chapter 10 Summary

The topics covered in this chapter can be summarized as follows:

10.1 Controls on Metamorphic Processes

Metamorphism is controlled by five main factors: the composition of the parent rock, the temperature to which the rock is heated, the amount and direction of pressure, the volumes and compositions of fluids that are present, and the amount of time available for metamorphic reactions to take place.

10.2 Foliation and Rock Cleavage

When the pressure acting on a rock is not uniform in all directions, foliation can develop. Foliation may occur in the form of platy or elongated mineral crystals that have grown at right angles to the maximum pressure, or it may develop when crystals or clasts within a rock are deformed. Foliation causes crystals or clasts within a rock to become aligned. When metamorphic rocks break parallel to the direction of foliation, rock cleavage results.

10.3 Classification of Metamorphic Rocks

Metamorphic rocks are classified on the basis of texture and mineral composition. Foliation is a key feature of metamorphic rocks formed under directed pressure; foliated metamorphic rocks include slate, phyllite, schist, and gneiss. Metamorphic rocks formed in environments without strong directed pressure include hornfels, marble, and quartzite.

10.4 Types of Metamorphism and Where They Occur

Almost all regions that experience metamorphism are being acted upon by plate-tectonic processes. Oceanic crustal rock can be metamorphosed near the spreading ridge where it was formed. Regional metamorphism takes place in areas where mountain ranges are forming, which are most common at convergent boundaries. Contact metamorphism takes place around magma bodies in the crust, which are also most common above convergent boundaries. Shock metamorphism happens when extraterrestrial bodies impact Earth, and is unusual among metamorphic processes because it occurs in seconds or minutes, rather than taking millions of years. Dynamic metamorphism occurs when shear stress is applied to rocks, such as along faults.

10.5 Metamorphic Facies and Index Minerals

Metamorphic facies are groups of metamorphic rocks that form under the same range of pressure and temperature conditions, but from different parent rocks. Geologists use index minerals such as chlorite, garnet, andalusite, and sillimanite to identify metamorphic zones. Index minerals tell us about the pressure and temperature conditions under which metamorphic rocks formed.

10.6 Metamorphic Hydrothermal Processes and Metasomatism

Contact metamorphism takes place around magma bodies that have intruded into cool rocks in the crust. Heat from magma is transferred to the surrounding country rock, resulting in mineralogical and textural changes. Hot water from a cooling body of magma, or from convection of groundwater driven by the heat of the pluton, can lead to hydrothermal alteration. When large volumes of fluid are flushed through rocks experiencing metamorphic pressures and temperatures, metasomatism results. Metasomatism can cause valuable metals to accumulate in the surrounding rocks.

Review Questions

  1. What are the two main agents of metamorphism, and what are their respective roles in producing metamorphic rocks?
  2. What types of metamorphic rocks will form if a mudrock experiences very low, low, medium, and high-grade metamorphism?
  3. Why doesn’t granite change very much at lower metamorphic grades?
  4. Describe the main process of foliation development in a metamorphic rock such as schist.
  5. What process contributes to metamorphism of oceanic crust at a spreading ridge?
  6. How do variations in the geothermal gradient affect the depth at which different metamorphic rocks form?
  7. Blueschist metamorphism takes place within subduction zones. What are the particular temperature and pressure characteristics of this geological setting?
  8. Rearrange the following minerals in order of increasing metamorphic grade: biotite, garnet, sillimanite, chlorite.
  9. What is the role of magmatic fluids in the metamorphism that takes place adjacent to a pluton?
  10. How does metasomatism differ from regional metamorphism?
  11. How does the presence of a hot pluton contribute to metasomatism?
  12. What determines whether metasomatism will produce skarn?
  13. For each of the following metamorphic rocks, indicate the likely parent rock and the grade and/or type of metamorphism: chlorite schist, slate, mica-garnet schist, amphibolite, marble.

71

Answers to Chapter 10 Review Questions

  1. Heat and pressure are the main agents of metamorphism. Heat leads to mineralogical changes in the rock. Pressure also influences those mineralogical changes, while directed pressure (greater pressure in one direction) leads to foliation.
  2. Very low grade: slate; low grade: phyllite; medium grade: schist; high grade: gneiss.
  3. Granite remains largely unchanged at lower metamorphic grades because its minerals are still stable at those lower temperatures.
  4. Foliation develops in schist when new platy minerals grow with their longest dimension at a right angle to the direction of greatest pressure.
  5. At a spreading ridge the heat from volcanism leads to the development of a groundwater convection system in the rock of the oceanic crust. Heated water rises in the hot regions and is expelled into the ocean, while cold ocean water is drawn into the crust to replace it. The heated water leads to the conversion of olivine and pyroxene into chlorite and serpentine.
  6. The geothermal gradient varies as a function of tectonic setting, being greatest in volcanic regions and lowest along subduction zones. As a result the depth at which specific metamorphic grades is achieved will vary: the depth will be greater when the gradient is lower.
  7. The geothermal gradient is low within subduction zones because the cold subducting oceanic crust takes a long time to heat up. Pressure increases with depth at the normal rate, but temperature does not.
  8. Order of increasing metamorphic grade: chlorite, biotite, garnet, sillimanite.
  9. Water from any source facilitates metamorphism. Magmatic fluids typically contain dissolved ions at higher concentrations than in regular groundwater (especially copper, zinc, silver, gold, lithium, beryllium, boron and fluorine), leading to the formation of a unique set of minerals.
  10. Metasomatism involves fluids from magmatic or groundwater sources that play an important role in transporting ions into the system, and leading to the formation of new minerals. Regional metamorphism takes place over a larger area, depends more on plate tectonic conditions, and does not involve flushing the system with large amounts of fluid.
  11. A hot pluton heats the surrounding water, causing groundwater to convect. This can result a great deal of water, in some cases with elevated levels of specific ions, passing through the rock. Water from magma within the pluton also contributes to metasomatism.
  12. Limestone must be present to produce skarn.
  13. Parent rocks and metamorphic grades and types:
    Metamorphic Rock Likely Parent Rock Grade and/or Type of Metamorphism
    Chlorite schist A rock enriched in ferromagnesian minerals, such as basalt Low-grade regional metamorphism
    Slate Mudrock (shale, mudstone) Very low grade regional metamorphism
    Mica-garnet schist A rock that is rich in aluminum, which includes most clay-bearing rocks Medium-grade regional metamorphism
    Amphibolite A rock enriched in ferromagnesian minerals, such as basalt Medium- to high-grade regional metamorphism
    Marble Limestone or dolomite Regional or contact metamorphism

 

XI

Chapter 11. Volcanism

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Mt. Garibaldi (in the background), near Squamish B.C., is one of Canada’s most recently active volcanoes, last erupting approximately 10,000 years ago. It is also one of the tallest, at 2,678 m in height. Source: Karla Panchuk (2017) CC BY-SA 4.0. Click the image for more attributions...
Figure 11.1 Mt. Garibaldi (in the background), near Squamish BC, is one of Canada’s most recently active volcanoes, last erupted approximately 10,000 years ago. It is also one of the tallest, at 2,678 m in height. Source: Karla Panchuk (2017) CC BY-SA 4.0. Photograph: Michael Scheltgen (2006) CC BY 2.0 view source Click the image for more attributions.

 

Learning Objectives

After reading this chapter and answering the Questions For Review at the end, you should be able to:

 

Why Study Volcanoes?

Volcanoes are awe-inspiring natural events. They have instilled fear and fascination with their red-hot lava flows, and cataclysmic explosions. In his painting The Eruption of Vesuvius (Figure 11.2), Pierre-Jacques Volaire captured the stunning spectacle of the eruption on Mt. Vesuvius on 14 May 1771. He also captured some stunningly casual spectating being done by tourists and their dog (lower left).

Painting: The Eruption of Vesuvius, by Pierre-Jacques Volaire (1771). Public Domain
Figure 11.2 The Eruption of Vesuvius, by Pierre-Jacques Volaire (1771). Public Domain.

As Volaire’s painting suggests, curiosity alone would be enough to make people want to learn why volcanoes happen and how they work. However, there are other reasons why we should know more about volcanoes. One reason is that studying volcanoes helps us understand the evolution of the Earth system- not just Earth’s geological features, but past changes in climate, and even the causes of mass extinctions. Another reason is the critical need to study the hazards posed by volcanoes to people and infrastructure. Over the past few decades, volcanologists have made great strides in their ability to forecast volcanic eruptions and predict the consequences, saving thousands of lives.

72

11.1 What Is A Volcano?

Volcanoes Are Where Magma Emerges

A volcano is a location where molten rock flows out, or erupts, onto Earth’s surface as lava. Volcanic eruptions can happen on land or underwater. Some volcanic eruptions flow from mountains (such as Mount Garibaldi in Figure 11.1), but others do not. Fissure eruptions (Figure 11.3) are volcanic eruptions flowing from long cracks in the Earth.

Figure 11.3 Kamoamoa fissure eruption on the flanks of the Hawai’ian Kīlauea Volcano in March of 2011. Source: Karla Panchuk (2017) CC BY-SA 4.0. Photograph: U. S. Geological Survey (2011) Public Domain view source. Click the image for more attributions.

Volcano Anatomy

The main parts of a volcano are shown in Figure 11.4. When volcanoes erupt, magma moves upward from a magma chamber and into a vent or conduit. It flows out from a crater at the top, or sometimes emerges at a secondary site on the side of the volcano resulting in a flank eruption. Erupted materials accumulate around the vent forming a volcanic mountain. The accumulated material might consist of layers of solidified lava, called lava flows, but it might also include fragments of various sizes that have been thrown from the volcano.

Figure 11.4 The parts of a volcano (not to scale). Source: Karla Panchuk (2017) CC BY 4.0

Crater or Caldera?

A crater is the basin above a volcano’s vent. Craters have diameters on the scale of 10s to 100s of metres. A caldera is a bowl-shaped structure that resembles a crater, but it is much larger (km in scale) and forms when a volcano collapses in on itself. The process is illustrated in Figure 11.5, going from left to right. It begins when an eruption occurs, and the magma chamber beneath the volcano is drained. If a significant part of a volcano’s mass is supported by magma within the chamber, then depleting the magma also reduces the support for the volcano. The loss of support causes part of the volcano to collapse into the void in the magma chamber, leaving behind a broad basin rimmed by the remnants of the volcano. Over time, the basin can fill with water. If there is still activity within the magma chamber, magma may force its way upward again, causing the floor of the caldera to be lifted, or erupting to form a new volcano within the caldera.

Formation of a caldera. Calderas are the result of a volcano collapsing into a drained magma chamber. Source: Karla Panchuk CC BY 4.0. Modified after U. S. Geological Survey (2002)
Figure 11.5 Formation of a caldera. Calderas are the result of a volcano collapsing into a drained magma chamber. Source: Karla Panchuk (2017) CC BY 4.0. Modified after U. S. Geological Survey (2002) Public Domain view source

The island of Santorini (Figure 11.6) is an example of a caldera. The island itself is the rim of the caldera, and the bay is the flooded basin. The two small islands in the middle of the bay formed from magma refilling the chamber that feeds the volcano, as in the far right of Figure 11.5. The caldera formed after an enormous eruption between 1627 and 1600 BCEFriedrich, W. L., Kromer, B., Friedrich, M., Heinemeier, J., Pfeiffer, T., & Talamo, S. (2006). Santorini Eruption Radiocarbon Dated 1627-1600 B.C. Science (312)5773, 548. doi: 10.1126/science.1125087. The eruption is thought to have contributed to the downfall of the Minoan civilization, and some speculate that it might also be the source of the myth of Atlantis, a story about a lost continent that sank beneath the sea after a natural disaster.

Figure 11.6 The Greek Island of Santorini. Left: Aerial view of the island forming a ring around a flooded caldera. Right: A view from the rim of the caldera. The other side of the rim is visible in the distance. Source: Karla Panchuk (2017) CC BY-SA 4.0. Satellite image: NASA/GSFC/MITI/ERSDAC/JAROS, and U.S./Japan ASTER Science Team (2000) Public Domain view source; Caldera photograph: Klearchos Kapoutsis (2010) CC BY 2.0 view source Click the image for more attributions.

 

73

11.2 Materials Produced by Volcanic Eruptions

Volcanic eruptions produce three types of materials: gas, lava, and fragmented debris called tephra.

Volcanic Gas

Magma contains gas. At high pressures, the gases are dissolved within magma. However, if the pressure decreases, the gas comes out of solution, forming bubbles. This process is analogous to what happens when a pop bottle is opened. Pop is bottled under pressure, forcing carbon dioxide gas to dissolve into the fluid. As a result, a bottle of pop that you find on the supermarket shelf will have few to no bubbles. If you open the bottle, you decrease the pressure within it. The pop will begin to fizz as carbon dioxide gas comes out of solution and forms bubbles.

The main component of volcanic gas emissions is water vapour, followed by carbon dioxide (CO2), sulphur dioxide (SO2), and hydrogen sulphide (H2S).

Volcanoes release gases when erupt, and through openings called fumaroles (Figure 11.7). They can also release gas into soil and groundwater.

A fumarole at Puʻu ʻŌʻō Crater. Hawaii. The yellow crust along the margin of the fumarole is made of sulphur crystals. The crystals form when sulphur vapour cools as it is released from the fumarole. Source: U. S. Geological Survey (2016) Public Domain
Figure 11.7 A fumarole at Puʻu ʻŌʻō Crater, Hawaii. The yellow crust along the margin of the fumarole is made of sulphur crystals. The crystals form when sulphur vapour cools as it is released from the fumarole. Source: U. S. Geological Survey (2016) Public Domain View source

Lava

The ease with which lava flows and the structures it forms depend on how much silica and gas the lava contains. The more silica, the more polymerization (formation of long molecules) occurs, stiffening the lava. The stiffness of lava is described in terms of viscosity– lava that flows easily has low viscosity, and lava that is sticky and stiff has high viscosity.

In general, high-silica lava contains more gas than low-silica lava. When the gas forms into bubbles, viscosity increases further. Consider the pop analogy again. If you were to shake the bottle vigorously and then open it, the pop would come gushing out in a thick, frothy flow. In contrast, if you took care to not shake the bottle before opening it, you could pour out a thin stream of fluid.

Chemical Composition Affects the Thickness and Shape of Lava Flows

The thickness and shape of a lava flow depends on its viscosity. The greater the viscosity, the thicker the flow, and the shorter the distance it travels before solidifying. Highly viscous lava might not flow very far at all, and simply accumulate as a bulge, called a lava dome, in a volcano’s crater. Figure 11.8 shows a dome formed from rhyolitic lava in the crater of Mt. St. Helens.

Lava dome in the crater of Mt. St. Helens. Source: Terry Feuerborn (2011) CC BY-NC 2.0
Figure 11.8 Lava dome in the crater of Mt. St. Helens. Source: Terry Feuerborn (2011) CC BY-NC 2.0 view source

Less viscous rhyolitic lava can travel further, as with the thick flow in Figure 11.9 (right). The left of Figure 11.9 shows thin streams of freely-flowing, low-silica, low-viscosity basaltic lava.

Lava flows. Left: A geologist collects a sample from a basaltic lava flow in Hawaii. Right: an andesitic lava flow from Kanaga Volcano in the Aleutian Islands. Source: Left- U. S. Geological Survey (2014) Public Domain; Right- Michelle Combs, U. S. Geological Survey (2015) Public Domain
Figure 11.9 Lava flows. Left: A geologist collects a sample from a basaltic lava flow in Hawaii. Right: an andesitic lava flow from Kanaga Volcano in the Aleutian Islands. Source: Left- U. S. Geological Survey (2014) Public Domain view source; Right- Michelle Combs, U. S. Geological Survey (2015) Public Domain view source

Low-viscosity basaltic lava flows may travel extended distances if they move through conduits called lava tubes. These are tunnels within older solidified lava flows. Figure 11.10 (top) shows a view into a lava tube through a hole in the overlying rock, called a skylight. Figure 11.10 (bottom) shows the interior of a lava tube, with a person for scale. Lava tubes form naturally and readily because flowing mafic lava preferentially cools near its margins, forming solid lava levées that eventually close over the top of the flow. Lava within tubes can flow for 10s of km because the tubes insulate the lava from the atmosphere and slow the rate at which the lava cools. The Hawai’ian volcanoes are riddled with thousands of old, drained lava tubes, some as long as 50 km.

Lava tubes. Top: An opening in the roof of a lava tube (called a skylight) permitting a view of lava flowing through the tube (Puʻu ʻŌʻō crater, Kīlauea). The opening is approximately 6 m across. Bottom: Inside a lava tube that channelled lava away from Mt. St. Helens in an eruption 1,895 years ago. Sources: Top: U. S. Geological Survey (2016) Public Domain. Bottom: Thomas Shahan (2013) CC BY-NC 2.0
Figure 11.10 Lava tubes. Top: An opening in the roof of a lava tube (called a skylight) permitting a view of lava flowing through the tube (Puʻu ʻŌʻō crater, Kīlauea). The opening is approximately 6 m across. Bottom: Inside a lava tube that channelled lava away from Mt. St. Helens in an eruption 1,895 years ago. Sources: Top: U. S. Geological Survey (2016) Public Domain. view source Bottom: Thomas Shahan (2013) CC BY-NC 2.0 view source

Lava Structures

Pahoehoe

Lava flowing on the surface can take on different shapes as it cools. Basaltic lava with an unfragmented surface, like that in Figure 11.9 (right), is called pahoehoe. (pronounced pa-hoy-hoy). Pahoehoe can be smooth and billowy. It can also develop a wrinkled texture, called ropy lava, as shown in Figure 11.11. Ropy lava forms when the outermost layer of the lava cools and develops a skin (visible as a dark layer in Figure 11.11, left), but the skin is still hot and thin enough to be flexible. The skin is stiffer than the lava beneath it, and is dragged by flowing lava and folded up into wrinkles. Figure 11.11 (right) is a close-up view after a cut has been made to show the internal structure of a wrinkled lava flow. Notice the many holes, or vesicles, within the lava, formed when the lava solidified around gas bubbles.

Ropy lava (pahoehoe) from Hawaii. Left: Ropy texture forming as a thin surface layer of lava cools and is wrinkled by the motion of lava flowing beneath it (near). Right: Cross-section view of ropy lava. Sources: Left: Z. T. Jackson (2005) CC BY NC-ND 2.0; Right: Fiddledydee (2011) CC BY-NC 2.0.
Figure 11.11 Ropy lava from Hawaii. Left: Ropy texture forming as a thin surface layer of lava cools and is wrinkled by the motion of lava flowing beneath it. Right: Cross-section view of ropy lava. Sources: Left: Z. T. Jackson (2005) CC BY NC-ND 2.0 view source; Right: Fiddledydee (2011) CC BY-NC 2.0 view source.

A’a and Blocky Lava

If the outer layer of the lava flow cannot accommodate the motion of lava beneath by deforming smoothly, the outer layer will break into fragments as lava moves beneath it. This could happen if the lava flow develops a thicker, more brittle outer layer, or if it moves faster. The result is a sharp and splintery rubble-like lava flow called a’a (pronounced like “lava” but without the l and v). Figure 11.12 (left) shows a close-up view of the advancing front of an a’a lava flow (the flow is moving toward the viewer). Figure 11.12 (right) shows an a’a lava flow viewed from the side. Compare the texture of the a’a flow with the texture of the lighter-grey pahoehoe lava in the foreground of the picture.

Aa lava flows. Left: Close-up view of aa forming during an eruption of Pacaya Volcano in Guatemala. Field of view approximately 1 m across. Right: Rubbly reddish-brown aa lava flow viewed from Chain of Craters Road, Hawai’i Volcanoes National Park. Pahoehoe is visible in the foreground. Sources: Photo of Hawaiian aa and pahoehoe: Roy Luck (2009) CC BY 2.0; Pacaya aa: Greg Willis (2008) CC BY-SA 2.0
Figure 11.12 Aa lava flows. Left: Close-up view of a’a forming during an eruption of Pacaya Volcano in Guatemala. Field of view approximately 1 m across. Right: Rubbly reddish-brown a’a lava flow viewed from Chain of Craters Road, Hawai’i Volcanoes National Park. Pahoehoe is visible in lighter grey in the foreground. Sources: Photo of Hawaiian aa and pahoehoe: Roy Luck (2009) CC BY 2.0 view source; Pacaya aa: Greg Willis (2008) CC BY-SA 2.0 (labels added) view source.

Higher viscosity andesitic lava flows also develop a fragmented surface, called blocky lava. This is visible in the toe of the andesitic lava flow from Figure 11.9 (right). The difference between a’a and the andesitic blocky lava is that the blocky lava has fragments with smoother surfaces and fewer vesicles.

Lava Pillows

When lava flows into water, the outside of the lava cools quickly, making a tube (Figure 11.13 (top left)). Blobs of lava develop at the end of the tube (Figure 11.13 (top right)), forming pillows. The bottom left of Figure 10.13 shows pillows covering the sea floor, and the bottom right shows the distinctive rounded shape of pillows in outcrop. Because pillows always form underwater, finding them in the rock record gives us information that the environment was underwater.

Pillow lavas. Top left: A tube of lava extruding underwater. Hot lava can be seen through cracks in the wall of the tube. The image is approximately 1 m across. (Pacific Ocean, near Fiji). Top right: The rounded end of a lava tube with cracks showing the lava within. (Pacific Ocean, near Fiji). Bottom left: sea floor covered with pillow lavas near the Galápagos Islands. Bottom right: A boulder made of 2.7 billion year old pillow lavas, derived from the Ely Greenstone in north-eastern Minnesota. Sources: Top left: NSF and NOAA (2010) CC BY 2.0; Top right: NSF and NOAA (2010) CC BY 2.0; Bottom left: NOAA Okeanos Explorer Program, Galápagos Rift Expedition 2011 (2011) CC BY 2.0; Bottom right: James St. John (2015) CC BY 2.0.
Figure 11.13 Pillow lavas. Top left: A tube of lava extruding underwater. Hot lava can be seen through cracks in the wall of the tube. The image is approximately 1 m across. (Pacific Ocean, near Fiji). Top right: The rounded end of a tube with cracks showing the lava within. (Pacific Ocean, near Fiji). Bottom left: sea floor near the Galápagos Islands covered with pillow lavas. Bottom right: A boulder made of 2.7 billion year old pillows derived from the Ely Greenstone in north-eastern Minnesota. Sources: Top left- NSF and NOAA (2010) CC BY 2.0 view source; Top right- NSF and NOAA (2010) CC BY 2.0 view source; Bottom left- NOAA Okeanos Explorer Program, Galápagos Rift Expedition 2011 (2011) CC BY 2.0 view source; Bottom right- James St. John (2015) CC BY 2.0 view source.

Columnar Joints

When lava flows cool and solidify, they shrink. Long vertical cracks, or joints, form within the brittle rock to allow for the shrinkage. Viewed from above, the joints form polygons with 5, 6, or 7- sides, and angles of approximately 120º between sides (Figure 11.14).

Columnar joints viewed from above. Source: Meg Stewart (2012) CC BY-SA 2.0
Figure 11.14 Columnar joints viewed from above, Giant’s Causeway, Northern Ireland. Source: Meg Stewart (2012) CC BY-SA 2.0 view source

Figure 11.15 shows a side view of columnar joints in a basaltic lava flow in Iceland.

Figure 11.15 Columnar joints in a basaltic lava flow, Svartifoss (Black Fall) Vatnajökull National Park, Iceland. Source: Ron Kroetz (2015) CC BY-ND 2.0. view source

Pyroclastic Materials

The pop bottle analogy illustrates another key point about gas bubbles in fluid, which is that the bubbles can propel fluid. In the same way that shaking a pop bottle to make more bubbles will cause pop to gush out when the bottle is opened, gas bubbles can violently propel lava and other materials from a volcano, creating an explosive eruption.

Collectively, loose material thrown from a volcano is referred to as tephra. Individual fragments are referred to in general terms as pyroclasts, so sometimes tephra is also referred to as pyroclastic debris. Pyroclasts are classified according to size.

Volcanic Ash

Particles less than 2 mm in diameter are called volcanic ash. Volcanic ash consists of small mineral grains and glass. Figure 11.16 shows volcanic ash on three scales: in the upper left is ash from the 2010 eruption of Eyjafjallajökull in Iceland. The image was taken with a scanning electron microscope at approximately 1000 times magnification. In the upper right is ash from the 1980 eruption of Mt. St. Helens, collected in Yakima, Washington, about 137 km northeast of Mt. St. Helens. Individual particles are under 1 mm in size. Figure 11.16 (bottom) shows a village near Mt. Merapi in Indonesia dusted in ash after an eruption 2010.

Figure 11.16 Volcanic ash. Upper left: Ash from 2010 eruption of Eyjafjallajökull in Iceland, magnified approximately 1000x. Upper right- Ash from the 1980 eruption of Mt. St. Helens, collected at Yakima, Washington. Bottom: Indonesian village after the eruption of Mt. Merapi in 2010. Sources: Upper left: Birgit Hartinger, AEC (2010) CC BY-NC-ND 2.0. view source Upper right: James St. John (2014) CC BY 2.0 (scale added) view source Bottom: AusAID/Jeong Park (2010) CC BY 2.0. view source  

Lapilli

Fragments with dimensions between 2 mm and 64 mm are classified as lapilli. Figure 11.17 (upper left) shows lapilli at the ancient city of Pompeii, which was buried when Mt. Vesuvius erupted in 79 C.E. Figure 11.17 (lower left) is a form of lapilli called Pele’s tears, named after the Hawai’ian diety Pele. Pele’s tears form when droplets of lava cool quickly as they are flung through the air. Rapidly moving through the air may draw the Pele’s tears out into long threads called Pele’s hair (Figure 11.17, right). The dark masses in Figure 11.17 (right) within the Pele’s hair are Pele’s tears.

Figure 11.17 Lapilli are pyroclasts ranging between 2 mm and 64 mm in size. Upper left: lapilli from the site of the ancient city of Pompeii. Lower left: Pele’s tears, a type of lapilli that forms when droplets of lava fly through the air. Right: Pele’s hair, which form when Pele’s tears are drawn out into thin threads as they fly. Sources: Upper left: Pauline (2009) CC BY-NC-ND 2.0 view source; Lower left: James St. John (2014) CC BY 2.0 (scale added) view source; Right: James St. John (2009) CC BY 2.0 (scale added) view source.

Blocks and Bombs

Fragments larger than 64 mm are classified as blocks or bombs, depending on their origin. Blocks are solid fragments of the volcano that form when an explosive eruption shatters the pre-existing rocks. Figure 11.18 shows one of many blocks from an explosive eruption at the Halema‘uma‘u crater at Kīlauea Volcano in May of 1924. The block has a mass of approximately 7 tonnes and landed 1 km from the crater.

Volcanic block weighing approximately 7 tonnes thrown 1 km from the Halema‘uma‘u crater at Kīlauea Volcano on May 18, 1924. Source: U. S. Geological Survey (1924) Public Domain
Figure 11.18 Volcanic block weighing approximately 7 tonnes thrown 1 km from the Halema‘uma‘u crater at Kīlauea Volcano on May 18, 1924. Source: U. S. Geological Survey (1924) Public Domain view source

Bombs form when lava is thrown from the volcano and cools as it travels through the air. Traveling through the air may cause the lava to take on a streamlined shape, as with the example in Figure 11.19.

Volcanic bomb with a streamlined shape. Source: James St. John (2016) CC BY 2.0
Figure 11.19 Volcanic bomb with a streamlined shape. Source: James St. John (2016) CC BY 2.0 (scale added) view source

Effects of Gas on Lapilli and Bombs

The presence of gas in erupting lava can cause lapilli and bombs to take on distinctive forms as the lava freezes around the gas bubbles, giving the rocks a vesicular (hole-filled) texture. Pumice (Figure 11.20) forms from gas-filled felsic lava. Figure 11.20 (right), shows a magnified view of the sample on the left. The dark patches in the photograph are mineral crystals that formed in the magma chamber before the lava erupted. Pumice floats on water because some of the holes are completely enclosed, and air-filled.

Lapilli-sided pumice fragment collected from the shores of Lake Atitlán in Guatemala by H. Herrmann. The lake is a flooded caldera, and is surrounded by active volcanoes. Right: magnified view showing vesicular structure and amphibole crystals (dark patches). Source: Karla Panchuk (2017) CC BY 4.0
Figure 11.20 Lapilli-sized pumice collected from the shores of Lake Atitlán in Guatemala by H. Herrmann. The lake is a flooded caldera, and is surrounded by active volcanoes. Right: Magnified view showing vesicular structure and amphibole crystals (dark patches). Source: Karla Panchuk (2017) CC BY 4.0

The mafic counterpart to pumice is scoria (Figure 11.21, left). Mafic lava can also form reticulite (Figure 11.21, right), a rare and fragile rock in which the walls surrounding the bubbles have all burst, leaving behind a delicate network of glass.

Mafic lapilli with vesicular textures. Left: Scoria from Mount Fuji, Japan. Scoria is the denser mafic counterpart to pumice. Right: Reticulite from Kīlauea Volcano. Reticulite is a delicate network of volcanic glass that forms when the walls separating gas bubbles pop. Sources: Left- James St. John (2014) CC BY 2.0 (scale added); Right- James St. John (2014) CC BY 4.0 (scale added)
Figure 11.21 Mafic lapilli with vesicular textures. Left: Scoria from Mount Fuji, Japan. Scoria is the denser mafic counterpart to pumice. Right: Reticulite from Kīlauea Volcano. Reticulite is a delicate network of volcanic glass that forms when the walls separating gas bubbles pop. Sources: Left- James St. John (2014) CC BY 2.0 (scale added) view source; Right- James St. John (2014) CC BY 4.0 (scale added) view source.

References

U. S. Geological Survey (2013) Mt. St. Helens National Volcanic Monument. Retrieved on 11 June 2017. Visit website

 

74

11.3 Types of Volcanoes

The products of volcanism that build volcanoes and leave lasting marks on the landscape include lava flows that vary in viscosity and gas content, and tephra ranging in size from less than a mm to blocks with masses of many tonnes. Individual volcanoes vary in the volcanic materials they produce, and this affects the size, shape, and structure of the volcano.

There are three types of volcanoes: cinder cones (also called spatter cones), composite volcanoes (also called stratovolcanoes), and shield volcanoes. Figure 11.22 illustrates the size and shape differences amongst these volcanoes.

Shield volcanoes, which get their name from their broad rounded shape, are the largest. Figure 11.22 shows the largest of all shield volcanoes- in fact, the largest of all volcanoes on Earth- Mauna Loa, which makes up a substantial part of the Island of Hawai‘i and has a diameter of nearly 200 km. The summit of Mauna Loa is presently 4,169 m above sea level, but this represents only a small part of the volcano. It rises up from the ocean floor at a depth of approximately 5,000 m. Furthermore, the great mass of the volcano has caused it to sag downward into the mantle by an additional 8,000 m. In total, Mauna Loa is a 17,170 m thick accumulation of rock.

Comparison of volcano sizes and shapes. Broad, rounded shield volcanoes are the largest, followed by cone-shaped composite volcanoes. Straight-sided cinder cones are the smallest.
Figure 11.22 Comparison of volcano sizes and shapes. Broad, rounded shield volcanoes are the largest, followed by cone-shaped composite volcanoes. Straight-sided cinder cones are the smallest, and barely visible in the scale of the diagram. Source: Karla Panchuk (2017) CC BY 4.0 modified after Steven Earle (2015) CC BY 4.0 view original

Kīlauea Volcano is also a shield volcano, albeit a much flatter one. Kīlauea Volcano rises only 18 m about the surrounding terrain, and is almost not visible in the scale of the diagram, however it still stretches over a distance of 125 km along the eastern side of the Island of Hawai‘i.

Composite volcanoes are the next largest. Mt. St. Helens is shown on the left of Figure 11.22. It rises 1,356 m above the surrounding terrain in the Cascade Range of the western United States, and has a diameter of approximately 6 km. Composite volcanoes tend to be no more than 10 km in diameter. Unlike shield volcanoes, composite volcanoes have a distinctly conical shape, with sides that steepen toward the summit.

Cinder cones are the smallest, and almost too small to see next to a volcano like Mauna Loa. Eve Cone is a cinder cone on the flanks of Mt. Edziza in northwestern British Columbia. It rises 172 m above the landscape, and has a diameter of under 500 m. Cinder cones have straight sides, unlike upward-steepening composite volcanoes, or rounded shield volcanoes.

Volcano Structure

Shield Volcanoes

Shield volcanoes, like the Sierra Negra volcano in the Galápagos Islands (Figure 11.23, top), get their gentle hill-like shape because they are built of successive flows of low-viscosity basaltic lava (Figure 11.23, bottom). The low viscosity of the lava means that it can flow for long distances, resulting in the greater size of shield volcanoes compared to composite volcanoes or cinder cones.

 

Figure 11.23 Shield volcano. Top: The Sierra Negra volcano in the Galápagos Islands exhibits the low, rounded shape characteristic of shield volcanoes. Bottom: Diagram of a shield volcano island, showing the build up of basaltic lava flows. Sources: Top- BRJ INC. (2012) CC BY-NC-ND 2.0 view source. Bottom- Karla Panchuk (2017) CC BY 4.0

Composite Volcanoes (Stratovolcanoes)

Composite volcanoes, like Cotopaxi in Figure 11.24 (top), consist of layers of lava alternating with layers of tephra (blocks, bombs, lapilli, and ash; Figure 11.24, bottom). The layers (strata) is where the alternative name, stratovolcano comes from. Cotopaxi displays the characteristic shape of composite volcanoes, which have slopes that get steeper near the top of the volcano. The change in the slope reflects the accumulation of tephra fragments near the volcano’s vent. Composite volcanoes typically erupt higher viscosity andesitic and rhyolitic lavas, which do not travel as far from the vent as basaltic lavas do. This results in volcanoes of smaller diameter than shield volcanoes. A notable exception is Mt. Fuji in Japan, which erupts basaltic lava.

Composite volcano. Cotopaxi in Ecuador exhibits the upward-steepening cone characteristic of composite volcanoes. Diagram of a composite volcano showing alternating layers of lava and tephra. <em>Sources: Top- Photo by Simon Matzinger (2014) CC BY 2.0. Bottom: Karla Panchuk (2017) CC BY 4.0.
Figure 11.24 Composite volcano. Top: Cotopaxi in Ecuador exhibits the upward-steepening cone characteristic of composite volcanoes. Bottom: Diagram of a composite volcano showing alternating layers of lava and tephra. Sources: Karla Panchuk (2017) CC BY 4.0; Top photo by Simon Matzinger (2014) CC BY 2.0 view source. Click the image for more attributions. 

From a geological perspective, composite volcanoes tend to form relatively quickly and do not last very long. If volcanic activity ceases, it might erode away within a few tens of thousands of years. This is largely because of the presence of pyroclastic eruptive material, which is not strong.

Cinder Cones (Spatter Cones)

Cinder cones, like Mt. Capulin in Figure 11.25, have straight sides and are typically less than 200 m high. Most are made up of fragments of scoria (vesicular rock from basaltic lava) that were expelled from the volcano as gas-rich magma erupted. Because cinder cones are made up almost exclusively of loose fragments, they have very little strength. They can be eroded away easily, and relatively quickly.

Cinder cone. These small, straight-sided volcanoes are made of volcanic fragments ejected when gas-rich basaltic lava erupts. Sources: Karla Panchuk (2017) CC BY 4.0, with photograph by R. D. Miller, U. S. Geological Survey (1980) Public Domain
Figure 11.25 Cinder cone. These small, straight-sided volcanoes are made of volcanic fragments ejected when gas-rich basaltic lava erupts. Sources: Karla Panchuk (2017) CC BY 4.0, with photograph by R. D. Miller, U. S. Geological Survey (1980) Public Domain view source. Click the image for more attributions.

 

References

Rubin, K. (n.d.) Mauna Loa Volcano. Retrieved 23 August 2017. Visit website

75

11.4 Types of Volcanic Eruptions

Volcanoes produce a variety of materials when they erupt. The characteristics of the eruptions themselves also vary from one volcano to the next, and sometimes from one eruption to the next for the same volcano. Eruptions are classified according how explosive they are, and the height of their eruption column– how high they blast material into the air.

Both the explosiveness of an eruption and the height of the eruption column are related in part to the composition of magma and the amount of gas it contains. In particular, magmas with more silica erupt more explosively. The higher the silica content, the greater the viscosity of the magma. This means more pressure can build up before the magma is forced out of the volcano. Magma with more silica also tends to contain more dissolved gas. The gas helps to propel magma out of the volcano, in the same way that the bubbles in a shaken bottle of pop cause the pop to foam out when the lid is removed.

There are four types of eruptions with properties determined mostly by the silica content of magma, and the amount of gas it contains. In order of increasing explosiveness, these are Hawai’ian, Strombolian, Vulcanian, and Plinian eruptions. Any composition of magma can have an explosive eruption if the magma suddenly encounters water. Hot magma contacting groundwater or seawater causes the water to flash to steam. Explosive eruptions driven by water are called hydrovolcanic (or phreatic) eruptions.

Hawai‘ian Eruptions

Hawai‘ian eruptions are named after the characteristic eruptions of volcanoes of the Hawai‘ian islands. Hawai‘ian eruptions are effusive (flowing) rather than explosive because they erupt low-viscosity basaltic lava. Hawai‘ian eruptions form shield volcanoes and can also take the form of fissure eruptions. Fissure eruptions occur when lava erupts from long cracks in the ground rather than from a central vent.

Figure 11.26 shows examples from two eruptions on of Hawai‘i. In the upper left and right are images from the November 1959 eruption of Kīlauea Iki Crater. The upper left shows a fissure eruption and effusive flow of lava. Burning trees appear as bright spots toward the bottom of the photo. Figure 11.26 (right) shows a lava fountain reaching 425 m above Kīlauea Iki Crater. U. S. Geological Survey scientists reported that volcanic bombs up to 60 cm across smashed the guard rail and dented the asphalt on the road. Figure 11.26 (lower left) shows Hawaiian Volcano Observatory (HVO) scientists making a quick getaway, with lava fountains from Mauna Loa Volcano in the background.

Hawaiian eruptions. Top left: Fissure eruption at Kīlauea Iki Crater in November of 1959. Bottom left: Lava fountains from an eruption of Mauna Loa Volcano in 1984. Right: Lava fountain from Kīlauea Iki Crater eruption in November of 1959.
Figure 11.26 Hawai‘ian eruptions. Top left: Fissure eruption at Kīlauea Iki Crater in November of 1959. Bottom left: Lava fountains from an eruption of Mauna Loa Volcano in 1984. Right: Lava fountain from Kīlauea Iki Crater eruption in November of 1959. Sources: Top left- U. S. Geological Survey (1959) Public Domain. view source Bottom left: R. B. Moore, U. S. Geological Survey (1984) Public Domain. view source Right- U. S. Geological Survey (1959) Public Domain. view source    

The photographs of the Kīlauea Iki Crater and Mauna Loa Volcano eruptions make the point that while Hawai‘ian eruptions are considered “gentle” eruptions, this is a relative term. “Gentle” eruptions range from lava flows that can be safely sampled by trained personnel, as in Figure 11.5, to lava fountains that soar hundreds of metres above the tree tops and rain large and dangerous rocks upon the surroundings.

Strombolian Eruptions

Strombolian eruptions, named for Mt. Stromboli in Italy, occur when basaltic lava has higher viscosity and higher gas content. The sticky lava is ejected in loud, violent, but short-lived spattery eruptions. Clumps of gas-rich lava thrown 10s to 100s of metres in the air accumulate as scoria in a pile around the vent, forming cinder cones. Figure 11.27 shows a strombolian eruption in the crater of Mt. Etna. A smaller cinder cone is forming around the vent as lava sputters out of it.

Strombolian eruption of Mt. Etna. Sputtering lava forms a smaller cinder cone around a vent within the crater of Etna.
Figure 11.27 Strombolian eruption of Mt. Etna. Sputtering lava forms a smaller cinder cone around a vent within the crater of Etna. Source: Karla Panchuk (2017) CC BY-SA 4.0. Photograph- Robin Wylie (2012) CC BY 2.0. view source Click the image for more attributions.  

Vulcanian Eruptions

Vulcanian eruptions get their name from the volcanic Italian island of Vulcano, which itself takes the name of the Roman god of fire, Vulcan. In Roman mythology, Vulcan was the maker of armour and weaponry for the gods, and volcanic eruptions were attributed to him working in his forge.

Vulcanian eruptions are far more explosive than Strombolian eruptions, and can blast tephra and gas to a height of 5 to 10 km. The explosiveness is related to a build-up of pressure as the higher viscosity of intermediate silica content lava restricts the escape of gas. Vulcanian eruptions produce large quantities of ash in addition to blocks and bombs.

The Vulcanian eruption of Mt. Pelée on the island of Martinique in 1902 resulted in the first detailed documentation by geologists of a devastating phenomenon that is now referred to as a pyroclastic flow (Figure 11.28). Volcanic debris from the collapse of a lava dome on Mt. Pelée combined with hot gas to form a searing avalanche that raced down the mountain, over the city of St. Pierre, and into the harbour.

A series of photos taken by Alfred Lacroix during the eruption of Mt. Pelée on May 8, 1902 showing the development of the pyroclastic flow that destroyed the city of St. Pierre and nearly 30,000 inhabitants.
Figure 11.28 A series of photos taken by Alfred Lacroix during the eruption of Mt. Pelée on May 8, 1902 showing the development of the pyroclastic flow that destroyed the city of St. Pierre and nearly 30,000 inhabitants. Source: Karla Panchuk (2017) CC BY 4.0. Photograph: Alfred Lacroix (1902) Public Domain. view original Click the image for more attributions.

The French geologist Alfred Lacroix described what he saw as a “nuée ardente,” or thick fiery cloud. The following first-hand account was published in Cosmopolitan Magazine in July of 1902, attributed to Ellery S. Scott, a sailor on the steamship Roraima:

“In idle interest, I turned my glass toward Mont Pelee. It was at that very moment that the whole top of the mountain seemed blown into the air. The sound that fol­lowed was deafening. A great mass of flames, seemingly a mile in diameter, with twisting giant wreaths of smoke, rolled thousands of feet into the air, and then overbalanced and came rolling down the seamed and cracked sides of the mountain.  Foot hills were overflowed by the onrush­ing mass. It was not mere flame and smoke. It was molten lava, giant blocks of stone and a hail of smaller stones, with a mass of scalding mud intermingled.
For one brief moment I saw the city of St. Pierre before me. Then it was blotted out by the overwhelming flood. There was no time for the people to flee. They had not even time to pray…. I had called to Carpenter Benson to start the windlass, but before he could move, the “Roraima” rolled almost on her port beam-ends, and then as suddenly went to starboard. The funnel, masts and boats went by the board in an instant. The decks were swept clean. The hatches were staved in. The next instant a hail of fire and red-hot stones was upon the ship. Then came the scalding mud. The saloon was ablaze. The ship seemed doomed. Men were struck down all around me by flaming masses of lava. From bright sun­light the air became dense as midnight. The smoke that rolled down from the cra­ter’s mouth had blotted the sun from our vision.”
Scott’s account vividly describes of the speed of the pyroclastic flow. In some cases, pyroclastic flows travel at speeds greater than 700 km/h. They are able to travel rapidly because they behave like a fluid, and can also ride on a cushion of hot gas. Scott says the city was “blotted out by a flood,” yet the lower parts of buildings remained (Figure 11.29), and human remains were found in streets and homes where they had fallen. The ruins of St. Pierre look as though the top of the city were shaved off, and that is effectively what happened as the pyroclastic flow rushed across it, buoyed by gas.
Two stereographs of the ruins of St. Pierre, published in 1902. Stereographs are viewed with a stereoscope to make an image appear three dimensional. Top- "St. Pierre, 'the city of dead,' Mt. Pelee smoking, Martinique"; Bottom- "Overlooking the mud-filled Roxelane River bed, and ash-covered ruins, to Mont Pelée, St. Pierre, Martinique."
Figure 11.29 Two stereographs of the ruins of St. Pierre, published in 1902. Stereographs are viewed with a stereoscope to make an image appear three dimensional. Top- “St. Pierre, ‘the city of dead,’ Mt. Pelee smoking, Martinique”; Bottom- “Overlooking the mud-filled Roxelane River bed, and ash-covered ruins, to Mont Pelée, St. Pierre, Martinique.” Source: Top- Boston Public Library (2013) CC BY 2.0 view source; Bottom- Boston Public Library (2013) CC BY 2.0 view source
The vast majority of fatalities from the eruption were caused by the heat of pyroclastic flow. Examination of the ruins of St. Pierre revealed that glass had melted, but copper had not, putting the temperature at between 700 ºC and 1000 ºC (1292 ºF to 1832 ºF).

Plinian Eruptions

Plinian eruptions are explosive eruptions of intermediate to felsic lava, and can form eruptive columns up to 45 km high. The origin of the name is the eruption of Vesuvius in 79 CE, which buried the towns of Pompeii and Herculaneum. The Roman admiral Gaius Plinius Secundus, also known as Pliny the Elder, attempted a rescue mission when he saw the column of ash and debris above Vesuvius, but died of unknown causes without being able to reach Herculaneum.

A more recent Plinian eruption was that of Mt. Redoubt on April 21, 1990, shown in Figure 11.30. Pyroclastic flows resulted, as did lahars, landslides that formed when glaciers melted and turned volcanic ash into mud. The shape of the eruptive column, with parts of the column appearing to spread out in flat layers at different levels, reflects differences in atmospheric characteristics.

Plinian eruption of Mt. Redoubt in Alaska on April 21, 1990.
Figure 11.30 Plinian eruption of Mt. Redoubt in Alaska on April 21, 1990. Source: Karla Panchuk (2017) CC BY 4.0. Photograph: R. Clucas, U. S. Geological Survey (1990) Public Domain. view source Click the image for more attributions.

Hydrovolcanic (Phreatic) Eruptions

Hydrovolcanic eruptions can be far more explosive than Plinian eruptions. They occur when water in the form of groundwater, seawater, or even melting glacial ice or snow comes into contact with magma. Heat from the magma changes water suddenly to steam, which can expand to more than a thousand times the original volume of water. The sudden expansion results in an explosive force that can blast a volcano to pieces and create large amounts of volcanic ash.

In April of 2010, activity by the Icelandic volcano Eyjafjallajökull (Figure 11.31) melted the glacier above it, releasing large quantities of water and triggering a hydrovolcanic eruption. Ash rose in a plume 10 km high, and was blown westward and into the skies over Europe. Volcanic ash can damage or destroy aircraft engines, so the precaution was taken to prohibit air travel for a 5-day period. The enormous economic impact of stopping flights has led to numerous studies about the best way to deal with similar events with volcanic ash in the future.

Figure 11-31 Hydrovolcanic eruption of Eyjafjallajökull in April of 2010. Left- Eruptive column with volcanic lightning. Volcanic lightning is caused by the static electricity generated by volcanic ash particles rubbing together. Right- Another view of the ash cloud, with westward winds carrying ash toward Europe where it would disrupt air traffic.
Figure 11.31 Hydrovolcanic eruption of Eyjafjallajökull in April of 2010. Left- Eruptive column with volcanic lightning. Volcanic lightning is caused by the static electricity generated by volcanic ash particles rubbing together. Right- Another view of the ash cloud, with westward winds carrying ash toward Europe where it would disrupt air traffic. Source: Karla Panchuk (2017) CC BY-SA 4.0. Left photograph: Terje Sørgjerd (2010) CC BY-SA 3.0 view source Right photograph: Henrik Thorburn (2010) CC BY 3.0 view source Click the image for more attributions.

References

Bressan, D (2012). Geology Scene Investigation: Death by Volcanic Fire. Visit website

British Geological Survey (n.d.). Eyjafjallajökull eruption, Iceland | April/May 2010. Visit website  

Digital History Project (2011). “Eyewitness Account to Eruption of Mont Pelee Matinique St Pierre Fort de France” By Ellery S. Scott. Visit website 

Rosen, J. (2015). Benchmarks: May 8, 1902: The deadly eruption of Mount Pelée. Visit website

U. S. Geological Survey (1997). Pyroclastic flows. Visit website

76

11.5 Plate Tectonics and Volcanism

Thus far volcanoes have been discussed in terms of the kinds of volcanic mountains they form, the materials they produce, and the style of eruption they have. All of these characteristics can be tied together into a big picture by considering the plate tectonic settings in which magma forms (Figure 11.32). The vast majority of volcanoes are present along plate tectonic boundaries.

Plate tectonic settings of volcanism. Volcanoes along subduction zones are the result of flux melting (lowering the melting point by adding water). Decompression melting produces volcanoes along divergent margins (ocean spreading centres and continental rift zones), as well as above mantle plumes. Contact between hot mafic partial melts and felsic rocks can trigger partial melting of the felsic rocks (melting from conduction)
Figure 11.32 Plate tectonic settings of volcanism. Volcanoes along subduction zones are the result of flux melting (lowering the melting point by adding water). Decompression melting produces volcanoes along divergent margins (ocean spreading centres and continental rift zones), as well as above mantle plumes. Contact between hot mafic partial melts and felsic rocks can trigger partial melting of the felsic rocks (melting from conduction). Source: Karla Panchuk (2017) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view original and U. S. Geological Survey (1999) Public Domain view original

There are four main scenarios to consider:

Decompression Causes Volcanism Along Spreading Centres and Rift Zones

At an ocean spreading ridge (centre of Figure 11.32), convection moves hot mantle rock slowly upward at rates of cm per year. At roughly 60 km below the surface, the mantle rocks have decompressed is enough to permit partial melting of approximately 10% of the ultramafic rock. Mafic magma is produced, and it moves up toward the surface. Magma fills vertical fractures produced by the spreading and spills out onto the sea floor making pillow lavas and lava flows. Spreading-ridge volcanism is taking place approximately 200 km offshore from the west coast of Vancouver Island.

In continental rift zones where continental crust is thinning (far right in Figure 11.32), a similar decompression process occurs, triggering partial melting of ultramafic mantle rocks. However, if the continental crust above the region where melting occurs has a lower melting temperature than the mafic melt that is produced, the continental crust will also melt.

Continental rift zones can have a range of volcano types. If mafic magma erupts, shield volcanoes, broad lava flows, and cinder cones result. However, if rocks of other compositions are melted and added in, or the mafic magma undergoes fractional crystallization before erupting, then composite volcanoes will also form.

Water Causes Partial Melting Along Subduction Zones

At an ocean-continent convergent boundary (Figure 11.32, right) or ocean-ocean convergent boundary (Figure 11.32, left), oceanic crust is pushed down into the mantle. Although temperatures are high, the slab is kept from melting by high pressures. However, under these conditions minerals in the slab release water from within their crystal structures. The water lowers the melting point of rock above the slab, and partial melting is triggered within the mantle. Mafic magma rises through the mantle to the base of the crust. There it contributes to partial melting of crustal rock, and more felsic material is added to the magma. The magma, now intermediate in composition, continues to rise and assimilate crustal material. In the upper part of the crust, it accumulates into plutons. Over time, fractional crystallization of magma within the pluton can make it even more silica-rich. From time to time, the magma from the plutons rises toward surface, leading to volcanic eruptions. 

Composite volcanoes with Vulcanian or Plinian eruption styles are characteristic of the volcanic arcs that form in subduction zones, although in the Trans-Mexico Volcanic Belt, Strombolian eruptions produce short-lived cinder cones. Where two margins of oceanic crust collide, the volcanic arc will be a chain of volcanic islands. Where continental and oceanic crust collide, there will be a volcanic arc on the continental crust.

Mt. St. Helens: A Composite Volcano in the Cascades Range Continental Volcanic Arc

On May 18, 1980 at 8:32 a.m., a M5.1 earthquake shook Mt. St. Helens, and marked the start of a 9-hour Plinian eruption (Figure 11.33) with a 24 km high eruption column and multiple pyroclastic flows. By the time the eruption was over, a large part of the volcano had been blasted away.

Figure 11-33 Eruption of composite subduction-zone volcano Mt. St. Helens on May 18, 1980. Top- Plinian eruption column. Bottom left- Mt. St. Helens before the eruption. Bottom right- The remains of Mt. St. Helens after the eruption.
Figure 11.33 Eruption of composite subduction-zone volcano Mt. St. Helens on May 18, 1980. Top- Plinian eruption column. Bottom left- Mt. St. Helens before the eruption. Bottom right- The remains of Mt. St. Helens after the eruption. Sources: Top- Karla Panchuk (2017) CC BY 4.0; Top- Photograph by NOAA (1980) Public Domain view source. Bottom left- R. Hoblitt, U. S. Geological Survey, Cascades Volcano Observatory (1979) Public Domain (label added) view source. Bottom right- Steven Earle (2015) CC BY 4.0 (label added) view source. Click the image for more attributions.

 

The explosive eruption was driven by gas-rich rhyolitic magma, however not all of Mt. St. Helens’ eruptions have been of felsic or intermediate material. The lava tube in Figure 11.10 (bottom) is from a time when Mt. St. Helens erupted basaltic lava. Data from the iMUSH (Imaging Magma Under St. Helens) project show that a magma chamber is present beneath Mt. St. Helens at between 5 and 14 km depth, but that a much larger magma chamber is present below it, which extends down to the mantle (Figure 11.34). Earthquakes in the 24 hours after the 1980 eruption suggested movement of magma within the smaller chamber (yellow arrows in Figure 11.34), but earthquakes from 1980 to 2005 indicate movement of magma within the deeper chamber as well (black arrows).

Magma chambers beneath Mt. St. Helens and Indian Heaven Volcanic Field, sketched from iMUSH (Imaging Magma Under St. Helens) project results. In a 24 hour period after the May 18, 1980 eruption, earthquakes in and around the smaller magma chamber suggested migration of magma (yellow arrows). Earthquakes recorded between 1980 and 2005 suggest migration of magma within a larger chamber that extends to the mantle (black arrows). The larger magma chamber might feed another smaller chamber beneath the Indian Heaven Volcanic Field.
Figure 11.34 Magma chambers beneath Mt. St. Helens and Indian Heaven Volcanic Field, sketched from iMUSH (Imaging Magma Under St. Helens) project results. In a 24 hour period after the May 18, 1980 eruption, earthquakes in and around the smaller magma chamber suggested migration of magma (yellow arrows). Earthquakes recorded between 1980 and 2005 suggest migration of magma within a larger chamber that extends to the mantle (black arrows). The larger magma chamber might feed another smaller chamber beneath the Indian Heaven Volcanic Field. Source: Karla Panchuk (2017) CC BY 4.0, based on Kiser et al. (2016), Figure 4B.

 

The complex history of Mt. St. Helens could reflect changes in the composition of magma within the small chamber over time, as fractionation proceeds, and the magma becomes more silica rich. However, movement of more mafic magma from the larger chamber could also contribute to eruptions with different chemical compositions. The larger magma chamber may be connected to a chamber feeding the nearby Indian Heaven Volcanic Field, which contains shield volcanoes and cinder cones, and for which basalt makes up 80% of erupted materials.

Mantle Plumes Can Cause Volcanism Away from Plate Boundaries

Mantle plumes are rising columns of hot solid rock. The column may be kilometres to 10s of kilometres across, but near the surface it spreads out to create a mushroom-like head that is 10s to over 100 kilometres across. Mantle plumes are different from the convection that normally occurs beneath ocean spreading centres: plumes rise approximately 10 times faster than mantle convection normally occurs, and may originate deep in the mantle, possibly just above the core-mantle boundary.

When the mantle plume rises to the base of the lithosphere, the pressure is low enough to permit partial melting of the plume material, producing mafic magma. Heat carried by the mantle plume may also melt rock adjacent to the plume. The magma rises and feeds hotspot volcanoes. The lithospheric plate above the mantle plume is moving across the plume, so a chain of hotspot volcanoes can result as existing hotspot volcanoes are slowly moved away from the mantle plume, and new volcanoes form in the lithosphere.

Many shield volcanoes are associated with mantle plumes, including those that make up the Hawai’ian islands. All of the Hawai’ian volcanoes are related to the mantle plume that currently lies beneath Mauna Loa, Kilauea, and Lōʻihi (Figure 11.35, top). There is evidence of crustal magma chambers beneath all three active Hawai’ian volcanoes. At Kīlauea, the magma chamber appears to be several kilometres in diameter, and is situated between 8 km and 11 km below surface (Lin et al., 2014). In this area, the Pacific Plate is moving northwest at a rate of about 7 cm/year. This means that the earlier formed — and now extinct — volcanoes have now moved well away from the mantle plume. The hotspot has in fact been present for at least 85 million years (Regelous et al., 2003), as evidenced by the long chain of eroded and submerged mountains stretching to the Aleutian Trench (Figure 11.35, bottom).

Hawai’ian hotspot volcanoes and volcanic chain. Top- A mantle plume beneath Hawai’i supplies magma to Mauna Loa Volcano, Kīlauea Volcano, and Lōʻihi Seamount. Volcanoes to the northwest are no longer active because they have moved away from the plume. Bottom- Bathymetric (depth) map showing the chain of islands stretching toward the Aleutian Trench, and marking the progress of the Pacific Plate over the mantle plume.
Figure 11.35 Hawai’ian hotspot volcanoes and volcanic chain. Top- A mantle plume beneath Hawai’i supplies magma to Mauna Loa Volcano, Kīlauea Volcano, and Lōʻihi Seamount. Volcanoes to the northwest are no longer active because they have moved away from the plume. Bottom- Bathymetric (depth) map showing the chain of islands stretching toward the Aleutian Trench, and marking the progress of the Pacific Plate over the mantle plume. Source: Top- J. E. Robinson, U. S. Geological Survey (2006) Public Domain view source. Bottom- National Geophysical Data Center/ U. S. Geological Survey (2006) Public Domain (labels added) view source.

Kīlauea Volcano is approximately 300 ka old, while neighbouring Mauna Loa Volcano is over 700 ka and Mauna Kea Volcano is over 1 Ma. If volcanism continues above the Hawaii mantle plume in the same manner that it has for the past 85 Ma, it is likely that Kīlauea Volcano will continue to erupt for at least another 500,000 years. By that time, its neighbour, Lōʻihi Seamount, will have emerged from the sea floor, and its other neighbours, Mauna Loa and Mauna Kea, will have become significantly eroded, like their cousins, the islands to the northwest.

Large Igneous Provinces (LIPs)

While the Hawaii mantle plume has produced a relatively low volume of magma for approximately 85 Ma, other mantle plumes are less consistent, and some generate massive volumes of magma over relatively short time periods. Although their origin is still controversial, it is thought that the volcanism leading to large igneous provinces (LIPs) is related to very high volume but relatively short duration bursts of magma from mantle plumes. An example of an LIP is the Columbia River Basalt Group, which extends across Washington, Oregon, and Idaho in the United States (Figure 11.36). This volcanism, which covered an area of about 160,000 km2 with basaltic rock up to several hundred metres thick, took place between 17 and 14 Ma.

Part of the Columbia River Basalt Group at Frenchman Coulee, eastern Washington, United States. All of the flows visible here have formed large (up to two metres in diameter) columnar basalts, a result of relatively slow cooling of flows that are tens of m thick. The inset map shows the approximate extent of the 17 to 14 Ma Columbia River Basalts, with the location of the photo shown as a star.
Figure 11.36 Part of the Columbia River Basalt Group at Frenchman Coulee, eastern Washington, United States. All of the flows visible here have formed large (up to two metres in diameter) columnar basalts, a result of relatively slow cooling of flows that are tens of m thick. The inset map shows the approximate extent of the 17 to 14 Ma Columbia River Basalts, with the location of the photo shown as a star. Source: Steven Earle (2015) CC BY 4.0 view source

The mantle plume that is assumed to be responsible for the Columbia River LIP is now situated beneath the Yellowstone area, where it leads to felsic volcanism. Over the past 2 Ma, three very large explosive eruptions at Yellowstone have yielded approximately 900 km3 of felsic magma. This is approximately 900 times the volume of the 1980 eruption of Mt. St. Helens, but only 5% of the volume of mafic magma in the Columbia River LIP.

Most other LIP eruptions are much bigger. The Siberian Traps (also basalt), which erupted at the end of the Permian period at 251 Ma, are estimated to have produced approximately 40 times as much lava as the Columbia River LIP. The largest known LIP is the Ontong Java Plateau, located in the southwest Pacific Ocean. It formed at around 122 Ma, and presently covers 1,500,000 km2 and has a volume of 5,000,000 km3. But this is only a small fraction of its original size. The majority of it has been subducted, and it may have been split into pieces that have been classified as separate LIPs.

Kimberlites

Kimberlite pipes are carrot-shaped cones of ultramafic rock. They form from the explosive eruption of mantle plumes originating at depths of 150 to 450 km in the mantle. The plume makes its way to the surface quickly (over hours to days), having little interaction with the surrounding rocks, and thus preserving a sample of the ultramafic mantle. As the plume nears the surface, a build-up of gas causes it to pick up speed, and by the time it reaches the surface it may be travelling faster than the speed of sound. The explosiveness of kimberlite eruptions means that they do not form volcanic mountains on the surface, but leave circular holes in the ground.

Kimberlite eruptions that originate at depths greater than 200 km beneath old, thick, continental crust travel through the region of the mantle where diamond is stable. In some cases, such as in Saskatchewan and the Northwest Territories, kimberlites bring diamond-bearing material to the surface. All of Earth’s diamond deposits are thought to have originated in this way.

Diamond mines in kimberlites, such as the Ekati Mine in the Northwest Territories, are easy to spot by the characteristic circular hole that develops as miners excavate the cone-shaped structure (Figure 11.37). The kimberlites at Ekati erupted between 45 and 60 Ma. Many kimberlites are older, and some much older. There have been no kimberlite eruptions in historic times. The youngest known kimberlites are in the Igwisi Hills in Tanzania and are only about 10,000 years old. The next youngest date to approximately 30 Ma.

Figure 11.37 The Ekati diamond mine in the Northwest Territories, part of the Lac de Gras kimberlite field. Source: Karla Panchuk (2017) CC BY-SA 4.0; Photograph by J. Pineau (2010) CC BY-SA 3.0 view source. Click the image for more attributions.

References

Kiser, E., Palomeras, I., Levander, A., Zelt, C., Harder, S., Schmandt, B., Hansen, S., Creager, K., & Ulberg, C. (2016). Magma reservoirs from the upper crust to the Moho inferred from high-resolution Vp and Vs models beneath Mount St. Helens, Washington State, USA. Geology (44)6, 411-414.

Lin, G, Amelung, F, Lavallee, Y, and Okubo, P. (2014). Seismic evidence for a crustal magma reservoir beneath the upper east rift zone of Kilauea volcano, Hawaii. Geology, 42(3), 187-190. DOI: 10.1130/G35001.1

Regelous, M., Hofmann, A. W., Abouchami, W., & Galer, S. J. G. (2003) Geochemistry of lavas from the Emperor Seamounts, and the geochemical evolution of Hawaiian magmatism from 85 to 42 Ma. Journal of Petrology 44(1), 113-140. DOI: 10.1093/petrology/44.1.113 view PDF

U. S. Geological Survey, Volcano Hazards Program (n.d.). Indian Heaven Volcanic Field visit website

U. S. Geological Survey, Volcano Hazards Program (n.d.). Mount St. Helens: 1980 Cataclysmic Eruption visit website

77

11.6 Volcanic Hazards

The basaltic lava flows produced by volcanoes on the island of Hawai’i are responsible for extensive damage to homes, infrastructure, and habitats. Figure 11.38 shows lava flows (in black) from the Puʻu ʻŌʻō crater of Kīlauea Volcano. The lava flow destroyed the house, and is encroaching on the transfer station. Smoke in the background marks locations where additional flows have broken out and are burning vegetation.

Lava flow from Kīlauea's Puʻu ʻŌʻō crater. Lava has destroyed a house and threatens a transfer station.
Figure 11.38 Lava flow from Kīlauea’s Puʻu ʻŌʻō crater. Lava (in black) has destroyed a house and threatens a transfer station. Source: U. S. Geological Survey (2014) Public Domain view source

In spite of the damage that lava flows can cause, they are not the volcanic hazard with the greatest impact on lives and infrastructure. Even the relatively free-flowing Hawai’ian basaltic lava moves slowly enough that it can be escaped on foot. Far more dangerous hazards are related to gases and volcanic debris. However, the largest impact and the greatest suffering are caused not by the immediate effects of volcanic eruptions, but by large-scale changes to climate and environments caused by volcanism. Indirect effects resulting in respiratory distress, toxicity, famine, and habitat destruction have accounted for approximately 8 million deaths during historical times, while direct effects have accounted for fewer than 200,000, or 2.5% of the total.

Volcanic Gas and Tephra Emissions

Large volumes of rock and gases are emitted during major Plinian eruptions at composite volcanoes, and a large volume of gas is released during some very high-volume effusive eruptions. Gases and fine particles of volcanic ash can cause respiratory distress and poisoning, and ash poses a risk for aircraft.

Most of the tephra from large explosive eruptions ascends high into the atmosphere, and some of it is distributed around Earth by high-altitude winds. The larger components (larger than 0.1 mm) fall closer to the volcano, and the accumulation of tephra from large eruptions can cause serious damage and casualties. When the large eruption of Mt. Pinatubo in the Philippines occurred in 1991, tens of centimetres of ash accumulated in fields and on rooftops in the surrounding populated region. Heavy typhoon rains hit the island at the same time and added to the weight of the tephra. The weight was too much for roofs to bear, and thousands of structures collapsed, causing at least 300 of the 700 deaths attributed to the eruption.

One of the long-term effects of adding volcanic particles and gases to the atmosphere is cooling. Over an eight-month period in 1783 and 1784, a massive effusive eruption took place at the Laki volcano in Iceland. Although there was relatively little volcanic ash involved, a massive amount of sulphur dioxide was released into the atmosphere, along with a significant volume of hydrofluoric acid (HF). The sulphur dioxide combined with water to make sulphate aerosols, which block incoming solar energy. The accumulation of sulphate aerosols over that 8 months led to dramatic cooling in the northern hemisphere. There were serious crop failures in Europe and North America, and a total of 6 million people are estimated to have died from famine and respiratory complications. In Iceland, poisoning from the HF resulted in the death of 80% of sheep, and 50% of cattle. The ensuing famine, along with HF poisoning, resulted in more than 10,000 human deaths, about 25% of the population.

Pyroclastic Flows

In a typical explosive eruption at a composite volcano, the tephra and gases are ejected with explosive force and sent high up into the atmosphere. As the eruption proceeds, and the amount of gas in the rising magma starts to decrease, and less gas is supplied to the eruption column. Parts of the column will become denser than air, leading the column to collapse and flow downward along the flanks of the volcano (Figure 11.39), picking up speed as it cools.

The Plinian eruption of Mt. Mayon, Philippines in 1984. Although most of the eruption column is ascending into the atmosphere, pyroclastic flows are traveling down the sides of the volcano in several places. Warnings were issued in time to evacuate 73,000 people.
Figure 11.39 The Plinian eruption of Mt. Mayon, Philippines in 1984. Although most of the eruption column is ascending into the atmosphere, pyroclastic flows are traveling down the sides of the volcano in several places. Warnings were issued in time to evacuate 73,000 people. Source: C. G. Newhall, U. S. Geological Survey (1984) Public Domain view source

Pyroclastic flows can travel over water, in some cases for many kilometres. In 1902 the pyroclastic flow from the eruption of Mt. Pelée traveled out into the harbour and destroyed several wooden ships anchored there. The pyroclastic flow from the 1883 eruption of Krakatau traveled 80 km across the Sunda Straits and claimed victims on the southwest coast of Sumatra. It also triggered a tsunami.

One of the most famous pyroclastic flows occurred when Mt. Vesuvius erupted in 79 CE. It buried the cities of Pompeii and Herculaneum, killing an estimated 18,000 people.

Lahar

A lahar is any mudflow or debris flow that is related to a volcano (Figure 11.40). Most are caused by melting snow and ice during an eruption, as was the case with the lahar that destroyed the Colombian town of Armero in 1985 when the volcano Nevado del Ruiz caused the ice dam on a glacial lake to fail. The resulting lahar killed 23,000 in Armero, about 50 km from the volcano.

Mud left behind from the lahar after the May 18, 1980 eruption of Mt. St. Helens. The lahar carried an enormous boulder to its present location.
Figure 11.40 Mud left behind from the lahar after the May 18, 1980 eruption of Mt. St. Helens. The lahar carried the boulder to its present location. Source: L. Topinka, U. S. Geological Survey (1980) Public Domain view source

Lahars can also happen when there is no volcanic eruption, because composite volcanoes tend to be weak and easily eroded. In October 1998, category 5 hurricane Mitch slammed into the coast of Central America. Damage was extensive and 19,000 people died. Fatalities were largely because of mudflows and debris flows triggered by intense rainfall — some regions received almost 2 m of rain over a few days.

At Casita Volcano in Nicaragua, the heavy rains weakened rock and volcanic debris on the upper slopes, resulting in a debris flow that rapidly built in volume as it raced down the steep slope. It struck the towns of El Porvenir and Rolando Rodriguez killing more than 2,000 people. El Porvenir and Rolando Rodriguez were new towns that had been built without planning approval in an area that was known to be at risk of lahars.

Sector Collapse and Debris Avalanche

In the context of volcanoes, sector collapse or flank collapse is the catastrophic failure of a significant part of an existing volcano, creating a large debris avalanche. This hazard was first recognized with the failure of the north side of Mt. St. Helens immediately prior to the large eruption on May 18, 1980.

In the weeks before the eruption, a large bulge had formed on the side of the volcano (Figure 11.41) as magma moved from depth into a magma chamber within the mountain itself. Early on the morning of May 18, a moderate earthquake struck and destabilized the bulge, leading to Earth’s largest observed landslide in historical times. The failure of this part of the volcano exposed the underlying magma chamber, causing it to explode sideways. This in turn exposed the conduit leading to the magma chamber below, resulting in a Plinian eruption lasting nine hours.

Bulge forming on the north side of Mt. St. Helens, April 27 1980.
Figure 11.41 Bulge forming on the north side of Mt. St. Helens, April 27 1980. Source: P. Lipman, U. S. Geological Survey (1980) Public Domain view source

78

11.7 Monitoring Volcanoes and Predicting Eruptions

In 2005 U. S. Geological Survey geologist Chris Newhall made a list of the six most important signs of an imminent volcanic eruption. They are:

  1. Gas leaks — the release of gases (mostly H2O, CO2, and SO2) from the magma into the atmosphere through cracks in the overlying rock
  2. Bulging — the deformation of part of the volcano, indicating that a magma chamber at depth is swelling or becoming more pressurized
  3. Seismicity — many (hundreds to thousands) of small earthquakes, indicating that magma is on the move. The quakes may be the result of the magma forcing the surrounding rocks to crack, or a harmonic vibration that is evidence of magmatic fluids moving underground.
  4. Seismicity ceases — a sudden decrease in the rate of earthquake activity. This may indicate that magma has stalled, and that\ something is about to give way
  5. Big bump — a pronounced bulge on the side of the volcano (like the one at Mt. St. Helens in 1980), which may indicate that magma has moved close to surface
  6. Steam — steam eruptions ( phreatic eruptions) that happen when magma near the surface heats groundwater to the boiling point. The water eventually explodes, sending fragments of the overlying rock far into the air.

With these signs in mind, it is possible to determine the necessary equipment to have and actions to take to monitor a volcano and predict when it might erupt. we can make a list of the equipment we should have and the actions we can take to monitor a volcano and predict when it might erupt.

Assessing Seismicity

The simplest and cheapest way to monitor a volcano is with seismometers, instruments that detect vibration. In an area with several volcanoes that have the potential to erupt (e.g., the Squamish-Pemberton area), a few well-placed seismometers can provide an early warning that something is changing beneath one of the volcanoes. There are currently enough seismometers in the Lower Mainland and on Vancouver Island to provide this information. You can view a map of Canadian National Seismograph Network here.

If there is seismic evidence that a volcano is coming to life, more seismometers should be placed in locations within a few tens of kilometres of the source of the activity (Figure 11.42). This will allow geologists to determine the exact location and depth of the seismic activity so that they can see where the magma is moving.

Three men stand before an array of solar panels and a satellite dish.
Figure 11.42 A seismometer installed in 2007 in the vicinity of the Nazco Cone, BC. Source: Cathie Hickson (n.d.) used with permission.

Detecting Gases

Water vapour quickly turns into clouds of liquid water droplets and is relatively easy to detect just by looking, but CO2 and SO2 are not as obvious. It’s important to be able to monitor changes in the composition of volcanic gases, and we need instruments to do that. Some can be monitored from a distance (from the ground or even from the air) using infrared devices, but to obtain more accurate data, we need to sample the air and do chemical analysis. This can be achieved with instruments placed on the ground close to the source of the gases, or by collecting samples (Figure 11.43) and analyzing them in a lab.

Figure 11.43 A geologist collects a gas sample from Sherman Crater, Mt. Baker, Washington. Gas is drawn through a titanium tube inserted in a fumarole, and collected in a glass vacuum flask. Source: D. Tucker, U. S. Geological Survey (2006) Public Domain view source

Measuring Deformation

There are two main ways to measure ground deformation at a volcano. One is known as a tiltmeter, which is a sensitive three-directional level that can sense small changes in the tilt of the ground at a specific location. Another is through the use of GPS (global positioning system) technology (Figure 11.27). GPS is more effective than a tiltmeter because it provides information on how far the ground has actually moved — east-west, north-south, and up-down.

Figure 11.44 A GPS unit installed at Hualālai Volcano, Hawaii. The dish-shaped antenna on the right is the GPS receiver. The antenna on the left is for communication with a base station. Source: U. S. Geological Survey (n.d.) Public Domain view source

Putting It All Together

By combining information from these types of sources, along with careful observations made on the ground and from the air, and a thorough knowledge of how volcanoes work, geologists can get a good idea of the potential for a volcano to erupt in the near future (months to weeks, but not days). They can then make recommendations to authorities about the need for evacuations and restricting transportation corridors.

Our ability to predict volcanic eruptions has increased dramatically in recent decades because of advances in our understanding of how volcanoes behave and in monitoring technology. Providing that careful work is done, there is no longer a large risk of surprise eruptions, and providing that public warnings are issued and heeded, it is less and less likely that thousands will die from sector collapse, pyroclastic flows, ash falls, or lahars. Indirect hazards are still very real, however, and we can expect the next eruption like the one at Laki in 1783 to take an even greater toll than it did then, especially since there are now roughly eight times as many people on Earth.

79

11.8 Volcanoes in Canada

Canada’s volcanically active regions are located in British Columbia and the Yukon Territory (Figure 11.45). At least 49 eruptions have occurred within these regions in the last 10,000 years. There are five volcanic regions associated with three types of plate tectonic settings: a subduction zone, a mantle plume, and a continental rift zone.

Figure 11.45 Canada’s volcanic regions are located in British Columbia and the Yukon Territory. Volcanism is associated with three tectonic settings: the subduction zone along the west coast (Garibaldi Volcanic Belt, Wrangell Volcanic Belt), a continental rift zone (Wells Gray-Clearwater Volcanic Field, Stikine Volcanic Belt), and a mantle plume (Anahim Volcanic Belt). Source: Volcanoes Canada, Canadian Hazards Information Service, Natural Resources Canada (n.d.) view source Click the image for copyright information.

Subduction Zone Volcanism: Wrangell and Garibaldi Volcanic Belts

The Wrangell Volcanic Belt is the result of subduction beneath the North American Plate. Volcanoes in the Canadian part of the Wrangell Volcanic Belt erupted between 17.8 and 10.4 million years ago. They were fed by lava that seeped up along a leaky transform fault.

Southwestern British Columbia is at the northern end of the Juan de Fuca subduction zone, and part of the Cascade Volcanic Arc that extends south through Washington and Oregon. The Canadian part of the Cascade Arc has had a lower rate and volume of volcanism than U. S. portions. One reason is that the northern part of the Juan de Fuca Plate is subducting more slowly than the rest of the plate, or else has stalled.

The Garibaldi Volcanic Belt has several volcanic centres , or regions where volcanism has caused multiple volcanoes to develop (Figure 11.46).

Figure 11.46 Volcanic centres within the Garibaldi Volcanic Belt. The most recent eruption was 2,350 years ago at Mt. Meager. Source: Sémhur (2007) CC BY-SA 4.0 view source. Click the image to enlarge.

The most recent volcanic activity in this area was 2,350 years ago at Mt. Meager. An explosive eruption similar in magnitude to that of Mt. St. Helens in 1980 spread ash as far east as Alberta. There was also significant volcanic activity at Mt. Price and Mt. Garibaldi approximately 10,000 years ago as glacial ice receded. In both cases, lava and tephra built up against glacial ice. The western side of Mt. Garibaldi failed by sector collapse when the ice melted, leaving rocks unsupported. Eruption beneath glacial ice resulted in the formation of a tuya– a steep-sided, flat-topped volcano- called The Table near Mt. Garibaldi (Figure 11.47).

Figure 11.47 The Table, a tuya near Mt. Garibaldi. Tuyas form when volcanoes erupt beneath ice, and their shape is determined by rapid cooling beneath the ice sheet. Source: Andre Charland (2004) CC BY 2.0 view source

Mantle Plume Volcanism: Anahim Volcanic Belt

The chain of volcanic complexes and cones extending from Milbanke Sound to Nazko Cone is interpreted as being related to a mantle plume currently situated close to the Nazko Cone, just west of Quesnel (Figure 11.48). The North American Plate is moving in a westerly direction at about 2 cm per year with respect to this plume, and the series of now partly eroded shield volcanoes between Nazco and the coast is interpreted to have been formed by the plume as the continent moved over it.

Figure 11.48 Anahim Volcanic Belt, the result of a mantle plume beneath the North American Plate. Source: Sémhur (2007) CC BY-SA 4.0 view source Click the image to enlarge.

The Rainbow Range, which formed at approximately 8 Ma, is the largest of these older volcanoes. It has a diameter of about 30 km and an elevation of 2,495 m (Figure 11.49). The name “Rainbow” refers to the bright colours displayed by some of the volcanic rocks as they weather.

Figure 11.49 Tsitsutl, the “painted mountain” within the Rainbow Range of the Anahim Volcanic Belt. The vibrant colours of the Rainbow Range are the result of chemical weathering. Source: Drew Brayshaw (2015) CC BY-NC 2.0 view source

Rift-Related Volcanism: Wells Gray-Clearwater Volcanic Field and Stikine Volcanic Belt

While British Columbia is not about to split into pieces, two areas of volcanism are related to rifting, or at least to stretching-related fractures that might extend through the crust. These are the Wells Gray-Clearwater volcanic field southeast of Quesnel (Figure 11.50), and the Stikine Volcanic Belt (also called the Northern Cordillera Volcanic Province), which ranges across the northwestern corner of the province.

Figure 11.50 Wells Gray-Clearwater Volcanic Field is the result of extension in the crust. Source: Sémhur (2007) CC BY-SA 4.0. view source Click the image to enlarge.

The Stikine Volcanic Belt includes Canada’s most recent volcanic eruption, a cinder cone and mafic lava flow that formed around 250 years ago at the Tseax River Cone in the Nass River area north of Terrace.  According to Nisga’a oral history, lava overran a village on the Nass River, and 2,000 people were lost. The region is now part of the Anhluut’ukwsim Laxmihl Angwinga’asanskwhl Nisga’a (Nisga’a Memorial Lava Bed Park).

The Mount Edziza Volcanic Field near the Stikine River is a large area of lava flows, sulphurous ridges, and cinder cones. The most recent eruption in this area was about 1,000 years ago. While most of the other volcanism in the Edziza region is mafic and involves lava flows and cinder cones, Mt. Edziza itself (Figure 11.51) is a composite volcano with rock compositions ranging from rhyolite to basalt. A possible explanation for the presence of composite volcanism in an area dominated by mafic flows and cinder cones is that there is a magma chamber beneath this area, within which magma differentiation is taking place.

Figure 11.51 Mount Edziza, in the Stikine Volcanic Belt, BC, with Eve Cone in the foreground. Source: NASS5518 (2008) CC BY 2.0 view source

 

References

Geological Survey of Canada (n.d.) Catalog of Canadian Volcanoes: Anahim volcanic belt Visit website

Geological Survey of Canada (n.d.) Catalog of Canadian Volcanoes: Garibaldi volcanic belt: Garibaldi Lake volcanic field Visit website

Skulski, T., Francis, D., & Ludden, J. (1991) Arc-transform magmatism in the Wrangell volcanic belt. Geology (19)1, 11-14. doi:10.1130/0091-7613(1991)019<0011:ATMITW>2.3.CO;2

Trop, J. M., Hart, W. K., Snyder, D., & Idleman, B. (2012). Miocene basin development and volcanism along a strike-slip to flat-slab subduction transition: Stratigraphy, geochemistry, and geochronology of the central Wrangell volcanic belt, Yakutat-North America collision zone. Geosphere (8)4, 805-834. doi:10.1130/GES00762.1

Volcanoes Canada, Canadian Hazards Information System, Natural Resources Canada (n.d.). Where Are Canada’s Volcanoes? Visit website

80

Chapter 11 Summary

The topics covered in this chapter can be summarized as follows:

11.1 What Is A Volcano?

Volcanoes are places where molten rock escapes to Earth’s surface. Some volcanoes are cone-shaped or hill-shaped mountains, and some eruptions happen along fissures. Eruptions are fed by a magma chamber beneath the volcano. Sometimes a volcano collapses into empty space in the magma chamber beneath, forming a caldera.

11.2 Materials Produced by Volcanic Eruptions

Volcanoes produce gas, lava flows, and debris called tephra. The characteristics of a lava flows depend on whether the lava is thin and runny (mafic with low gas content) or thick and sticky (felsic with high gas content). Tephra is classified according to size. Ash is less than 2 mm in diameter, lapilli is between 2 mm and 64 mm, and blocks and bombs are larger than 64 mm.

11.3 Types of Volcanoes

Cinder cones are relatively small straight-sided volcanoes that are composed mostly of mafic rock fragments. Composite volcanoes consist of alternating layers of lava flows and tephra. The tend to be intermediate to felsic in composition, and get steeper toward the top. Shield volcanoes are broad, low, hill-like volcanoes that form from layers of low-viscosity mafic lava.

11.4 Types of Volcanic Eruptions

Volcanic eruptions can be classified according to how explosive they are, and how high into the atmosphere they blast material. Hawai’ian eruptions are relatively gentle effusive eruptions of low-viscosity mafic lava, and form shield volcanoes. Strombolian eruptions are more vigorous eruptions of mafic tephra. They blast material hundreds of metres into the air. The tephra falls out of the atmosphere to form a cinder cone. Vulcanian eruptions are explosive eruptions of intermediate composition lava, producing pyroclastic flows and eruptive columns from 5 to 10 km high. Plinian eruptions are highly explosive eruptions of felsic lava, and can produce eruption columns up to 45 km high. Both Vulcanian and Plinian eruptions are associated with composite volcanoes. Hydrovolcanic eruptions are the explosive result of magma or lava interacting with water, and rapidly changing the water to steam.

11.5 Plate Tectonics and Volcanism

Volcanism is closely related to plate tectonics. Most volcanoes are associated with convergent plate boundaries (at subduction zones), but a great deal of volcanic activity also occurs at divergent boundaries and areas of continental rifting. At convergent boundaries magma is formed where water from a subducting plate acts as a flux to lower the melting temperature of the adjacent mantle rock. At divergent boundaries magma forms because of decompression melting. Decompression melting also takes place within a mantle plume.

11.6 Volcanic Hazards

Most direct volcanic hazards are related to volcanoes that erupt explosively, especially composite volcanoes. Pyroclastic flows, some as hot as 1000 ˚C, can move at hundreds of km/h and will kill anything in the way. Lahars, volcano-related mudflows, can be large enough to destroy entire towns.  Lava flows are also destructive, but tend to move slowly enough to permit people to get to safety. Indirect hazards claim far more lives than direct hazards, and include famine related to volcanically-induced climate cooling.

11.7 Monitoring Volcanoes and Predicting Eruptions

Clues that a volcanic eruption might soon occur include earthquakes, a change in the type and amount of gases released, and changes in the shape of the volcano as magma moves within it. Volcanoes are monitored using seismometers to detect earthquakes, volcanic gases are sampled and analyzed, and instruments are used to detect deformation of the volcano. These tools make it possible to assess the hazard posed by a given volcano, and the risk of eruption.

11.8 Volcanoes in British Columbia

British Columbia and the Yukon Territory include examples of volcanoes that form as a result of fluid-induced melting along a subduction zone (the Wrangell and Garibaldi volcanic belts) , as a result of decompression where the crust is thinning and stretching (Stikine Volcanic Belt and Wells Gray-Clearwater Volcanic Field), and because of mantle plume activity (Anahim Volcanic Belt).

Review Questions

  1. What are the three main tectonic settings for volcanism on Earth?
  2. What is the primary mechanism for partial melting at a convergent plate boundary?
  3. Why are the viscosity and gas content of a magma important in determining the type of volcanic rocks that will be formed when that magma is extruded?
  4. Why do the gases in magma not form gas bubbles when the magma is deep within the crust?
  5. Where and why do pillow lavas form?
  6. What two kinds of volcanic materials make up a composite volcano?
  7. What is a lahar, and why are lahars commonly associated with eruptions of composite volcanoes?
  8. Under what other circumstances might a lahar form?
  9. Why do shield volcanoes have gentle slopes?
  10. Which type of volcanic mountain would last longest: a shield volcano, a cinder cone, or a composite volcano?
  11. Why is weak seismic activity (small earthquakes) typically associated with the early stages of a volcanic eruption?
  12. How can GPS technology be used to help monitor a volcano for activity?
  13. What is the likely geological origin of the Nazko Cone?
  14. What might be the explanation for southwestern B.C. having much less subduction-related volcanism than adjacent Washington and Oregon?

81

Answers to Chapter 11 Review Questions

  1. The three main tectonic settings for volcanism are (1) subduction zones at convergent plate boundaries, (2) divergent plate boundaries, and (3) mantle plumes (a.k.a. hot spots).
  2. The primary mechanism for partial melting at a convergent plate boundary is the addition of water to hot mantle rock. The water reduces the melting temperature of the rock (flux melting).
  3. The explosiveness of a volcanic eruption depends on the pressure of the magma. Gases create that pressure, and if the magma is viscous those gases cannot escape easily. Felsic and intermediate magmas tend to have more gas than mafic magmas, and are also more viscous, trapping the gas in.
  4. When magma is deep within the crust the pressure is too high for the gases to bubble out of solution.
  5. Pillow lavas form where mafic lava erupts in water. When the magma oozes out into the water the outside cools first forming a hard skin that maintains the pillow shape.
  6. Composite volcanoes are formed of layers of lava flows and tephra (volcanic fragments ranging from fine ash to blocks and bombs) from explosive eruptions.
  7. A lahar is a mud flow or debris flow on a volcano. Lahars are common on composite volcanoes because they are steeper than shield volcanoes, they typically have ice and snow, and they are not as strong as shield volcanoes.
  8. Some lahars form during an eruption when snow and ice melt quickly, while others may form from heavy rain.
  9. The lava that forms shield volcanoes is typically low viscosity. It can flow easily and also tends to form lava tubes. As a result, it is able to travel a long way from the vent, forming a low broad shield.
  10. Cinder cones erode rapidly because they are mostly piles of tephra. Composite volcanoes are more resistant to weathering and erosion than cinder cones because lava flows help hold together tephra, but composite volcanoes still don’t last as long as shield volcanoes. Shield volcanoes are stronger because they consist more of lava flows than tephra.
  11. Weak seismic activity is associated with all stages of a volcanic eruption. In the early stages magma is moving at depth and pushing rock aside, creating small earthquakes. The flow of magma can also produce special type of seismic response known as a harmonic tremor.
  12. GPS technology is used to determine if there is any slow deformation of the flanks of a volcano related to movement of magma toward the surface.
  13. The Nazko Cone is thought to be related to a mantle plume.
  14. One hypothesis to explain the lower rate of volcanism in British Columbia than in adjacent Washington and Oregon is that the northern part of the Juan de Fuca Plateis not subducting as quickly as the rest of the plate.

XII

Chapter 12. Earthquakes

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Heavy equipment moves debris from a fallen structure.
Figure 12.1 Demolition of a structure damaged when an earthquake of magnitude 6.3 struck Christchurch, New Zealand on February 22, 2011. Many structures had already sustained damage from an earthquake that struck six months earlier, in September of 2010. Collapsing structures and falling debris accounted for most of the 185 deaths. Source: Karla Panchuk (2017) CC BY-SA 4.0. Photograph: Terry Philpott (2012) CC BY-NC 2.0 view source. Click the image for more attributions.

Learning Objectives

After reading this chapter, and answering the review questions at the end, you should be able to:

Why Study Earthquakes?

On the morning of June 23, 1946, a magnitude 7.3 earthquake struck Vancouver Island. It caused substantial damage to structures, including the school shown in Figure 12.2, and resulting in one fatality. The shaking was so violent that the seismograph measuring the earthquake in Victoria was broken a few seconds after the earthquake started. Even if the seismograph had survived, it was not sensitive enough to provide sufficiently detailed information about local earthquakes, and was unable to record some of them at all. There was only one monitoring station in British Columbia, so while it was possible to determine how far away the earthquake was from the station, geologists were not able to say in which direction. In 1955, Canadian seismologist W. G. Milne wrote that “the 1946 earthquake indicated in a very forceful manner the need for better instruments for the study of earthquakes in British Columbia.” A network of improved instruments was established as a direct result of the earthquake.

Figure 12.2 Damage to an elementary school in Courtenay, British Columbia after a magnitude 7.3 earthquake on Sunday, June 23, 1946. Left: A hole left after the chimney collapsed through the roof. Right: Damage inside the school. In addition to damaging structures, the earthquake triggered numerous slope failures. Source: Photographs courtesy of Earthquakes Canada. Click the image for image sources and terms of use.

Time and time again earthquakes have caused massive damage and many, many casualties. Recording earthquakes and determining their location of origin is important for establishing what geological conditions are responsible for the earthquakes, and understanding the risk they pose. After new, more sensitive instruments were installed at the Dominion Astrophysical Observatory in Victoria, geologists quickly learned that they had underestimated earthquake activity in the area. Between June of 1948 and August of 1951, 224 local earthquakes were recorded!

By studying earthquakes, geoscientists and engineers are making progress toward learning how to minimize earthquake damage, and how to reduce the number of people affected by earthquakes.  This knowledge can be communicated to governments, so they are aware of what is needed to keep the population safe. It can also be communicated to individuals, so they know what to expect and do in the event of an earthquake, and can be adequately prepared with emergency supplies.

Additional Resources

Lamontagne, M., Halchuk, S., Cassidy, J. F., and Rogers, G. C (2008) Significant Canadian Earthquakes of the Period 1600 – 2008. Seismological Research Letters 79(2), 211 – 223. doi: 10.1785/gssrl.79.2.211 Read the paper

Detailed description of the effects of the 1946 earthquake: Hodgson, E. A. (1946). British Columbia Earthquake, June 23, 1946. The Journal of the Royal Astronomical Society of Canada XL(8), 285 – 319. Read the paper

References

Milne, W. G. (1955). Seismology in British Columbia. The Journal of the Royal Astronomical Society of Canada XLIX(4), 141 – 150. Read the paper

Ministry for Culture and Heritage (2017). Christchurch earthquake kills 185. Visit website

82

12.1 What is an Earthquake?

Earthquake Shaking Comes from Elastic Deformation

Earthquakes occur when rock ruptures (breaks), causing rocks on one side of a fault to move relative to the rocks on the other side. Although motion along a fault is part of what happens when an earthquake occurs, rocks grinding past each other is not what creates the shaking. In fact, it could be said that the earthquake happens after rocks have undergone most of the displacement. Consider this: if rocks slide a few centimetres or even metres along a fault, would that motion alone explain the incredible damage caused by some earthquakes? If you were in a car that suddenly accelerated then stopped, you would feel a jolt. But earthquakes are not a single jolt. Buildings can swing back and forth until they shake themselves to pieces, train tracks can buckle and twist into s-shapes, and roads can roll up and down like waves on the ocean. During an earthquake, rock is not only slipping. It is also vibrating like a plucked guitar string.

Rocks might seem rigid, but when stress is applied they may stretch. If there hasn’t been too much stretching, a rock will snap back to its original shape once the stress is removed. Deformation that is reversible is called elastic deformation. Rocks that are stressed beyond their ability to stretch can rupture, allowing the rest of the rock to snap back to its original shape. The snapping back of the rock returning to its original shape causes the rock to vibrate, and this is what causes the shaking during an earthquake. The snapping back is called elastic rebound.

Figure 12.3 (top) shows this sequence of events. Stress is applied to a rock and deforms it. The deformed rock ruptures, forming a fault. After rupturing, the rock above and below the fault snaps back to the shape it had before deformation.

Figure 12.3 Elastic deformation, rupture, and elastic rebound. Top: Stress applied to a rock causes it to deform by stretching. When the stress becomes too much for the rock, it ruptures, forming a fault. The rock snaps back to its original shape in a process called elastic rebound. Bottom: On an existing fault, asperities keep rocks on either side of the fault from sliding. Stress deforms the rock until the asperities break, releasing the stress, and causing the rocks to spring back to their original shape. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view original.

Ruptures can also occur along pre-existing faults (Figure 12.3, bottom). The rocks on either side of the fault are locked together because bumps along the fault, called asperities, prevent the rocks from moving relative to each other. When the stress is great enough to break the asperities, the rocks on either side of the fault can slide again. While the rocks are locked together, stress can cause elastic deformation. When asperities break and release the stress, the rocks undergo elastic rebound and return to their original shape.

Rupture Surfaces Are Where the Action Happens

Images like 12.3 are useful for illustrating elastic deformation and rupture, but they can be misleading.  The rupture that happens doesn’t occur as in 12.3, with the block being ruptured through and through.  The rupture and displacement only happen along a subsection of a fault, called the rupture surface.  In Figure 12.4, the rupture surface is the dark pink patch.  It takes up only a part of the fault plane (lighter pink). The fault plane represents the surface where the fault exists, and where ruptures have happened in the past.  Although the fault plane is drawn as being flat in Figure 12.4, faults are not actually perfectly flat.

The location on the fault plane where the rupture happens is called the hypocentre or focus of the earthquake (Figure 12.4, right). The location on Earth’s surface immediately above the hypocentre is the epicentre of the earthquake.

Figure 12.4 Rupture surface (dark pink), on a fault plane (light pink). The diagram represents a part of the crust that may be tens or hundreds of kilometres long. The rupture surface is the part of the fault plane along which displacement occurred. Left: In this example, the near side of the fault is moving to the left, and the lengths of the arrows within the rupture surface represent relative amounts of displacement. Coloured arrows represent propagation of failure on a rupture surface. In this case, the failure starts at the dark blue heavy arrow and propagates outward, reaching the left side first (green arrows) and the right side last (yellow arrows). Right: An earthquake’s location can be described in terms of its hypocentre (or focus), the location on the fault plane where the rupture happens, or in terms of its epicentre (red star), the location above the hypocentre. Source: Left: Steven Earle (2015) CC BY 4.0 view source. Right: Karla Panchuk (2017) CC BY 4.0.

Within the rupture surface, the amount of displacement varies. In Figure 12.4, the larger arrows indicate where there has been more displacement, and the smaller arrows where there has been less.  Beyond the edge of the rupture surface there is no displacement at all.  Notice that this particular rupture surface doesn’t even extend to the land surface of the diagram.

The size of a rupture surface and the amount of displacement along it will depend on a number of factors, including the type and strength of the rock, and the degree to which the rock was stressed beforehand. The magnitude of an earthquake will depend on the size of the rupture surface and the amount of displacement.

A rupture doesn’t occur all at once along a rupture surface. It starts at a single point and spreads rapidly from there. Figure 12.4 illustrates a case where rupturing starts at the heavy blue arrow in the middle, then continues through the lighter blue arrows. The rupture spreads to the left side (green arrows), then the right (yellow arrows).

Depending on the extent of the rupture surface, the propagation of failures (incremental ruptures contributing to making the final rupture surface) from the point of initiation is typically completed within seconds to several tens of seconds. The initiation point isn’t necessarily in the centre of the rupture surface; it may be close to one end, near the top, or near the bottom.

Shifting Stress Causes Foreshocks and Aftershocks

Earthquakes don’t usually occur in isolation. There is often a sequence in which smaller earthquakes occur prior to a larger one, and then progressively smaller earthquakes occur after.  The largest earthquake in the series is the mainshock.  The smaller ones that come before are foreshocks, and the smaller ones that come after are aftershocks. These descriptions are relative, so it can be necessary to reclassify an earthquake.  For example, the strongest earthquake in a series is classified as the mainshock, but if another even bigger one comes after it, the bigger one is called the mainshock, and the earlier one is reclassified as an aftershock.

A rupture surface does not fail all at once.  A rupture in one place leads to another, which leads to another.  Aftershocks and foreshocks represent the same thing, except on a much larger scale. The rupture illustrated in Figure 12.4 reduced stress in one area, but in doing so, transferred stress to others (Figure 12.5). Imagine a frayed rope breaking strand by strand.  When a strand breaks, the tension on that strand is released, but the remaining strands must still hold up the same amount of weight. If another strand breaks under the increased burden, the remaining strands have an even greater burden than before.  In the same way that the stress causes one strand after another to fail, a rupture can trigger subsequent ruptures nearby.

Figure 12.5 Stress changes related to an earthquake. Stress decreases in the area of the rupture surface, but increases on adjacent parts of the fault. Source: Steven Earle (2015) CC BY 4.0 view source.

Numerous aftershocks were associated with the magnitude 7.8 earthquake that struck Haida Gwaii in October of 2012 (Figure 12.6; mainshock in red, aftershocks in white). Some of the stress released by the mainshock was transferred to other nearby parts of the fault, and contributed to a cascade of smaller ruptures. But stress transfer need not be restricted to the fault along which an earthquake happened. It will affect the rocks in general around the site of the earthquake and may lead to increased stress on other faults in the region. The aftershocks from the Haida Gwaii earthquake are scattered rather than located only on the main faults.

Figure 12.6 Magnitude 7.8 Haida Gwaii earthquake and aftershocks. Mainshock (red circle marks the epicentre) occurred on October 28th, 2012. Aftershocks are for the period from October 28th to November 10th of 2012. Although the epicentre is near a transform boundary, the rupture was influenced more by compression related to the subduction zone. Source: Karla Panchuk (2017) CC BY 4.0. Base map with epicentres from the U. S. Geological Survey Latest Earthquakes tool view interactive map. Subduction zone after Wang et al. (2015). Click the image for more attributions.

The effects of stress transfer may not show immediately. Aftershocks can be delayed for hours, days, weeks, or even years. Because stress transfer affects a region, not just a single fault, and because there can be delays between the event that transferred stress and the one that was triggered by the transfer, it can sometimes be hard to be know whether one earthquake is actually associated with another, and whether a foreshock or aftershock should be assigned to a particular mainshock.

Episodic Tremor and Slip

Episodic tremor and slip (ETS) is periodic slow sliding along part of a subduction boundary. It does not produce recognizable earthquakes, but does produce seismic tremor (observed as rapid seismic vibrations on instruments). It was first discovered on the Vancouver Island part of the Cascadia subduction zone by Geological Survey of Canada geologists Herb Dragert and Gary Rogers.Rogers, G. and Dragert, H. (2003). Episodic tremor and slip on the Cascadia subduction zone: The chatter of silent slip. Science, 300, 1942-1943.

The boundary between the subducting Juan de Fuca plate and the North America plate can be divided into three segments (Figure 12.7). The cold upper part of the boundary is the locked zone. There the plates are stuck together for long periods of time. When slip does occur, it generates very large earthquakes. The last time the locked zone along Canada’s west coast slipped was January 26, 1700. It caused an earthquake of magnitude 9. The warm lower part of the boundary, called the continuous slip zone, is sliding continuously because the warm rock is weaker. The central part of the boundary, the ETS zone, isn’t cold enough to be stuck, but isn’t warm enough to slide continuously. Instead it slips episodically approximately every 14 months for about 2 weeks, moving a few centimetres each time.

Figure 12.7 Episodic tremor and slip along the Cascadia subduction zone. The Juan de Fuca plate is locked to the North American plate at the top of the subduction zone, but lower down it is slipping continuously. In the intermediate (ETS) zone, the plate alternately sticks and slips on a regular schedule. Source: Steven Earle (2015) CC BY 4.0 view source

It might seem that periodic slip along this part of the plate helps to reduce tension, and thus reduce the risk of a large earthquake. In fact, the opposite is likely the case. The movement along the ETS part of the plate boundary transfers stress to the adjacent locked part of the plate. During the two-week ETS period, the transfer of stress means an increased chance of a large earthquake.

Since 2003, ETS processes have also been observed in subduction zones in Mexico and Japan.

Additional Resources

IRIS Teachable Moment slides for the October 2012 Haida Gwaii earthquake

References

Wang, K., Jiangheng, H., Schulzeck, F., Hyndman, R. D., and Riedel, M. (2015). Thermal Condition of the 27 October 2012 Mw 7.8 Haida Gwaii Subduction Earthquake at the Obliquely Convergent Queen Charlotte Margin. Bulletin of the Seismological Society of America, 105(2B), 1290–1300. doi: 10.1785/0120140183

 

83

12.2 Measuring Earthquakes

The shaking from an earthquake travels away from the rupture in the form of seismic waves. Seismic waves are measured to determine the location of the earthquake, and to estimate the amount of energy released by the earthquake (its magnitude).

Types of Seismic Waves

Seismic waves are classified according to where they travel, and how they move particles.

Body Waves

Seismic waves that travel through Earth’s interior are called body waves. P-waves are body waves that move by alternately compressing and stretching materials in the direction the wave moves. For this reason, P-waves are also called compression waves. The “P” in P-wave stands for primary, because P-waves are the fastest of the seismic waves. They are the first to be detected when an earthquake happens.

A P-wave can be simulated by fixing one end of a spring to a solid surface, then giving the other end a sharp push toward the surface (Figure 12.8, top).  The compression will propagate (travel) along the length of the spring. Some parts of the spring will be stretched, and others compressed. Any one point on the spring will jiggle forward and backward as the compression travels along the spring.

Figure 12.8 Seismic waves simulated using a spring and rope attached to a fixed surface. Top: P-waves travel as pulses of compression. Bottom: S-waves move particles at right angles to the direction of motion. Source: Karla Panchuk (2018) CC BY 4.0 modified after Steven Earle (2015) CC BY 4.0 view original.

S-waves are body waves that move with a shearing motion, shaking particles from side to side. S-waves can be simulated by fixing one end of a rope to a solid surface, then giving the other end a flick (Figure 12.8, bottom). Any one point on the rope will move from side to side at a right angle to the direction in which the snaking motion is traveling. The “S” in S-wave stands for secondary, because S-waves are slower than P-waves, and are detected after the P-waves are measured. S-waves cannot travel through liquids.

P-waves and S-waves can travel rapidly through geological materials, at speeds many times the speed of sound in air.

Surface Waves

When body waves reach Earth’s surface, some of their energy is transformed into surface waves, which travel along Earth’s surface. Two types of surface waves are Rayleigh waves and Love waves (Figure 12.9). Rayleigh waves (R-waves) are characterized by vertical motion of the ground surface, like waves rolling on water. Love waves (L-waves) are characterized by side-to-side motion. Notice that the effects of both kinds of surface waves diminish with depth in Figure 12.9.

Surface waves are slower than body waves, and are detected after the body waves. Surface waves typically cause more ground motion than body waves, and therefore do more damage than body waves.

Figure 12.9 Surface waves travel along Earth’s surface and have a diminished impact with depth. Rayleigh waves (left) cause a rolling motion, and Love waves (right) cause the ground to shift from side to side. Source: Steven Earle (2015) CC BY 4.0 view source. Click the image for more attributions.

Recording Seismic Waves Using a Seismograph

A seismometer is an instrument that detects seismic waves. An instrument that combines a seismometer with a device for recording the waves is called a seismograph. The graphical output from a seismograph is called a seismogram. Figure 12.10 (right) shows how a seismograph works. The instrument consists of a frame or housing that is firmly anchored to the ground. A mass is suspended from the housing, and can move freely on a spring. When the ground shakes, the housing shakes with it, but the mass remains fixed. A pen attached to the mass moves up and down on a rotating drum of paper, drawing a wavy line, the seismogram. The seismograph in Figure 12.10 (right) is oriented to measure vertical ground motion. The photo on the left shows a seismograph oriented to record horizontal ground motion.

Figure 12.10 How a seismograph records earthquakes. Source: Left- Karla Panchuk (2018) CC BY-NC-SA 4.0 modified after IRIS (2012) “How Does a Seismometer Work?” view source; Right: Karla Panchuk (2018) CC BY-SA 4.0, photo by Z22 (2014) CC BY-SA 3.0 view source. Click the image for more attributions.

The pen and drum of a mechanical seismograph record the motion of the ground relative to the mass. However, unless an earthquake causes a large amount of ground motion directly beneath the seismograph, the height of the wave recorded on paper might be very small, making the seismogram difficult to analyze. The seismograph on the right has a device to amplify the ground motion, drawing larger waves that are easier to study.

Modern seismographs record shaking as electrical signals, and are able to transmit those signals. This means seismologists need not return to the instrument to collect recordings before the records can be examined.

Finding The Location of an Earthquake

P-waves travel faster than S-waves. As the waves travel away from the location of an earthquake, the P-wave gets farther and father ahead of the S-wave. Therefore, the farther a seismograph is from the location of an earthquake, the longer the delay between when the P-wave arrival is recorded, and the S-wave arrival is recorded. The delay between the P-wave and S-wave arrival appears as a widening gap in a diagram of P-wave and S-wave travel times (Figure 12.11, grey lines).

P-wave and S-wave arrival times can be identified on seismograms. In the three seismograms in Figure 12.11, the arrivals of the P-waves and S-waves are marked with arrows, and the interval in minutes between the P-wave and S-wave arrivals are noted. The seismograms were recorded at three different seismic stations (earthquake monitoring locations equipped with seismographs). The distance of each station from the earthquake is determined by finding the distance along the graph where the gap between the P-wave and S-wave travel-time curves matches the delay between P-wave and S-wave arrivals on the seismogram.

Figure 12.11 Using P-wave and S-wave travel times to determine how far seismic waves have travelled. Grey curves show the distance travelled by P-waves and S-waves after an earthquake occurs. P-waves are faster than S-waves, and the gap between them increases with time and distance. The delays between P-wave and S-wave arrivals on seismograms are matched to the curve to find the distances of seismic stations from the source of the seismic waves. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0 modified after IRIS (n.d.) “How Are Earthquakes Located?” view source

 

The delay between the P-wave and S-wave arrival at a seismic station can indicate how far the station is from the source of the earthquake, but not the direction from which the seismic waves travelled. The possible locations of the earthquake can be represented on a map by drawing a circle around the seismic station, with the radius of the circle being the distance determined from the P-wave and S-wave travel times (Figure 12.12). If this is done for at least three seismic stations, the circles will intersect at the origin of the earthquake.

Figure 12.12 Locating earthquakes by drawing three circles with radii of lengths determined from P-wave and S-wave travel times. Station names (SOCO, TEIG, SSPA) correspond to seismograms in Figure 12.11. Source: IRIS (n.d.) “How Are Earthquakes Located?” view source Click the image for terms of use.

How Big Was It?

Earthquakes can be described in terms of their magnitude, which reflects the amount of energy released by the shaking. They can also be described in terms of intensity, which characterizes the impact of the shaking on people and their surroundings.

Earthquake Magnitude

Earthquake magnitudes are determined by measuring the amplitudes of seismic waves. The amplitude is the height of the wave relative to the baseline (Figure 12.13). Wave amplitude depends on the amount of energy carried by the wave. The amplitudes of seismic waves reflect the amount of energy released by earthquakes.

Figure 12.13 Seismogram for a small earthquake that occurred near Vancouver Island in 1997. The maximum amplitude of the S-wave is indicated. Source: Karla Panchuk (2018) CC BY 4.0 modified after Steven Earle (2015) CC BY 4.0 view source

The Richter magnitude scale uses the amplitudes of S-waves, and corrects for the decrease in amplitude that happens as the waves travel away from their source. The correction depends on how seismic waves interact with the specific rock types through which they travel, and therefore on local conditions, so the Richter magnitude is also referred to as the local magnitude.

While news reports about earthquakes might still refer to the “Richter scale” when describing magnitudes, the number they report is most likely the moment magnitude. The moment magnitude is calculated from the seismic moment of an earthquake. The seismic moment takes into account the surface area of the region that ruptured, how much displacement occurred, and the stiffness of the rocks. Moment magnitude can capture the difference between short earthquakes and longer ones resulting from larger ruptures, even of both types of earthquakes generate the same amplitude of waves. The moment magnitude scale is also better for earthquakes that are far from the seismic station. Seismic wave measurements are still used to determine the moment magnitude, however different waves are used than for the local magnitude scale.

The magnitude scale is a logarithmic one rather than a linear one- an increase of one unit of magnitude corresponds to a 32 times increase in energy release (Figure 12.14). There are far more low-magnitude earthquakes than high-magnitude earthquakes. In 2017 there were 7 earthquakes of M7 (magnitude 7) or greater, but millions of tiny earthquakes.

Figure 12.14 Earthquake magnitude and corresponding energy release. Energy release increases by approximately 32 times for each unit change in magnitude. Source: IRIS (n.d.) “How Often Do Earthquakes Occur?” view source

Earthquake Intensity

Intensity scales were first used in the late 19th century, and then adapted in the early 20th century by Giuseppe Mercalli and modified later by others to form what we now call the Modified Mercalli Intensity Scale (Table 12.1). To determine the intensity of an earthquake, reports are collected about what people felt and how much damage was done. The reports are then used to assign intensity ratings to regions where the earthquake was felt.

Table 12.1 Modified Mercalli Intensity Scale
I Not felt Not felt except by a very few under especially favourable conditions
II Weak Felt only by a few persons at rest, especially on upper floors of buildings
III Weak Felt quite noticeably by persons indoors, especially on upper floors of buildings; many people do not recognize it as an earthquake; standing motor cars may rock slightly; vibrations similar to the passing of a truck; duration estimated
IV Light Felt indoors by many, outdoors by few during the day; at night, some awakened; dishes, windows, doors disturbed; walls make cracking sound; sensation like heavy truck striking building; standing motor cars rocked noticeably
V Moderate Felt by nearly everyone; many awakened; some dishes, windows broken; unstable objects overturned; pendulum clocks may stop
VI Strong Felt by all, many frightened; some heavy furniture moved; a few instances of fallen plaster; damage slight
VII Very Strong Damage negligible in buildings of good design and construction; slight to moderate in well-built ordinary structures; considerable damage in poorly built or badly designed structures; some chimneys broken
VIII Severe Damage slight in specially designed structures; considerable damage in ordinary substantial buildings with partial collapse; damage great in poorly built structures; fall of chimneys, factory stacks, columns, monuments, walls; heavy furniture overturned
IX Violent Damage considerable in specially designed structures; well-designed frame structures thrown out of plumb; damage great in substantial buildings, with partial collapse; buildings shifted off foundations
X Extreme Some well-built wooden structures destroyed; most masonry and frame structures destroyed with foundations; rails bent
XI Extreme Few, if any (masonry), structures remain standing; bridges destroyed; broad fissures in ground; underground pipelines completely out of service; earth slumps and land slips in soft ground; rails bent greatly
XII Extreme Damage total; waves seen on ground surfaces; lines of sight and level distorted; objects thrown upward into the air
Source: U. S. Geological Survey (1989). The Severity of an Earthquake. USGS General Interest Publication 1989-288-913 view source

Intensity values are assigned to locations, rather than to the earthquake itself. This means that intensity can vary for a given earthquake, depending on the proximity to the epicentre and local conditions. For the 1946 M7.3 Vancouver Island earthquake, intensity was greatest in the central island region (Figure 12.15). In some communities within this region, chimneys were damaged on more than 75% of buildings. Some roads were made impassable, and a major rock slide occurred. The earthquake was felt as far north as Prince Rupert, as far south as Portland Oregon, and as far east as the Rockies, but with less intensity.

Figure 12.15 Intensity map for the M7.3 Vancouver Island earthquake on June 23, 1946. Source: Earthquakes Canada, Natural Resources Canada (2016) view source. Click the image for terms of use.

Intensity estimates are important as a way to identify regions that are especially prone to strong shaking. A key factor is the nature of the underlying geological materials. The weaker the underlying materials, the more likely it is that there will be strong shaking. Areas underlain by strong solid bedrock tend to experience far less shaking than those underlain by unconsolidated river or lake sediments.

An example of this effect is the 1985 M8 earthquake that struck the Michoacán region of western Mexico, southwest of Mexico City. There was relatively little damage near the epicentre, but 350 km away in heavily populated Mexico City there was tremendous damage and approximately 5,000 deaths. The reason is that Mexico City was built largely on the unconsolidated and water-saturated sediment of former Lake Texcoco. These sediments resonate at a frequency of about two seconds, which was similar to the frequency of the body waves that reached the city. Consequently, the shaking was amplified.  Survivors of the disaster recounted that the ground in some areas moved up and down by approximately 20 cm every two seconds for over two minutes. Damage was greatest to buildings between 5 and 15 storeys tall, because they also resonated at around two seconds, which amplified the shaking.

References

Ammon, C. J. (2001). Earthquake Size Visit website

84

12.3 Earthquakes and Plate Tectonics

Bands of earthquakes trace out plate boundaries (coloured dots, Figure 12.16). The depths of earthquakes, and the width of the band, depend on the type of plate boundary. Mid-ocean ridges and transform margins have shallow earthquakes (usually less than 30 km deep), in narrow bands close to plate margins. Subduction zones have earthquakes at a range of depths, including some more than 700 km deep. Bands of earthquakes are wider along subduction zones because they take place throughout the subducting slab that extends beneath the opposing plate. Wide swaths of scattered earthquakes may correspond to continent-continent collision zones, such as between the Eurasian plate and the African, Arabian, and Indian plates to the south. Wide swaths of scattered earthquakes may also correspond to continental rift zones, such as in eastern Africa.

Figure 12.16 Earthquakes greater than magnitude 5, from 2000 to 2008. Bands of earthquakes mark tectonic plates. Narrow bands with shallow earthquakes (marked in red) indicate transform boundaries or mid-ocean ridge divergent boundaries. Wider bands with earthquakes at a range of depths are subduction zones. Wide bands of scattered earthquakes mark continent-continent convergent margins (e.g., between the Indian and Eurasian plates), or continental rift zones (e.g., in eastern Africa). Source: Lisa Christiansen, Caltech Tectonics Observatory (2008) view source. Plate and ocean basin labels added. Click the image for terms of use.

 

Earthquakes are also relatively common at a few locations away from plate boundaries. Some are related to the buildup of stress due to continental rifting or the transfer of stress from other regions, and some are not well understood. Locations include the Great Rift Valley area of Africa, the Lake Baikal area of Russia, and Tibet.

Earthquakes at Divergent and Transform Plate Boundaries

Earthquakes along divergent and transform plate margins are shallow (usually less than 30 km deep) because below those depths, rock is too hot and weak to avoid being permanently deformed by the stresses in those settings. If deformation is permanent, then removing the stress does not result in the rocks snapping back to their original shape. No snapping back means no shaking.

Mid-ocean ridge divergent plate margins are offset by numerous transform faults (Figure 12.17). The locations of earthquakes along mid-ocean ridges, and the mechanisms for causing them, depend on how rapidly the mid-ocean ridges are spreading. The Pacific-Antarctic Ridge (left) is spreading relatively rapidly at 42 to 94 mm/year, depending on the location along the ridge.  Rapid spreading causes rocks near the axis of the spreading centre to be hot and weak. As a result, most of the earthquakes (white dots) are located along transform faults, where rocks are cooler and stronger. Along rapidly spreading ridges, new ocean crust is bent upward into wide, high ridges. As spreading proceeds and crust moves away from the ridge, the bend is relaxed, and the crust stretches and breaks. This triggers earthquakes many kilometres away from the ridge.

Figure 12.17 Locations of earthquakes of magnitude 4 and greater from 1990 to 2010 along two mid-ocean ridges. Plate boundaries are marked in red. Arrows show the direction of plate motion. Left: Rapidly spreading Pacific-Antarctic ridge with earthquakes concentrated along transform faults. Right: Slowly spreading Southwest Indian Ridge, with earthquakes along both spreading segments and transform faults. Source: Karla Panchuk (2017) CC BY 4.0. Base maps with epicentres generated using the U. S. Geological Survey Latest Earthquakes website. Visit Latest Earthquakes

The Southwest Indian Ridge (right) spreads very slowly, at approximately 14 mm/year. Rocks are cooler and stronger along the slowly spreading ridge than along the rapidly spreading one. In the slow-spreading environment, earthquakes are generated when rocks along the ridge axis stretch and break. Earthquakes are more evenly distributed between divergent and transform segments of the boundary than they are along fast-spreading ridges.

Earthquakes in continental rift zones are also shallow, but scattered more broadly than those along mid-ocean ridges. Lake Baikal (Figure 12.28), the world’s oldest, deepest, and largest freshwater lake, formed 25 million years ago because of continental rifting. Note the scale in Figure 12.18, and compare how widely the earthquakes (blue dots) are spread in the Lake Baikal region, versus along the mid-ocean ridges in Figure 12.17.

Figure 12.18 Blue circles mark the locations of earthquakes of M4 and greater from 1990 to 2010 along the Lake Baikal rift zone. White lines show some of the faults in the region. White lines with tick marks are normal faults related to spreading. Arrows show the direction of spreading. White lines without tick marks are transform faults. The Siberian Craton (shaded region) is strong 2 billion year old crust. Source: Karla Panchuk (2017) CC BY 4.0. Faults after U. S. Geological Survey (see references). Base maps (inset and rift views) with epicentres generated using the U. S. Geological Survey Latest Earthquakes website. Visit Latest Earthquakes

One reason for the difference in earthquake distribution in continental rift zones is that the rifts are only beginning to form. Faulting is “disorganized” within the continental crust. There is no well-established spreading centre, unlike mid-ocean ridges. Another reason is that the locations of faults, and thus earthquakes, in continental rift zones are affected by pre-existing geological structures within continental crust. In the case of the Lake Baikal rift, the strong, ancient crust of the Siberian Craton influences the orientation of the faults forming the rift. Faults run parallel to the craton near Lake Baikal. As rifting extends to the east, the part of the craton in the upper right of Figure 12.18 may deflect rifting southward.

Earthquakes at Convergent Boundaries

Subduction Zones

Along convergent plate margins with subduction zones, earthquakes range from shallow to depths of up to 700 km. Earthquakes occur where the two plates are in contact, as well as in zones of deformation on the overriding plate, and along the subducting slab deeper within the mantle. The result is that epicentres of earthquakes farther to the interior of the overriding plate will correspond to increasingly deep earthquakes.

Where the Pacific plate subducts beneath the North American plate, forming the Aleutian volcanic arc (Figure 12.19), earthquakes increase in depth moving northward, following the Pacific plate into the mantle. Earthquakes between 0 and 33 km deep (red circles) occur closest to the subduction zone (red line; teeth point in the direction of the subducting slab). While there is some overlap, earthquakes between 33 and 70 km deep (white circles) occur in a band that reaches farther the north. Farthest north are the epicentres for earthquakes between 70 and 300 km deep (green dots). The deepest earthquake during the seven year interval shown in Figure 12.19 is represented by the large green dot farthest to the north. It occurred at a depth of 265 km.

Figure 12.19 Earthquakes of M4.5 and greater from 2010 to 2017 along the Aleutian Trench subduction zone (red line; teeth point in the direction of the subducting slab). White arrows show the directions of plate movement. Circle colours indicate the depths of earthquakes (see legend, lower left). Earthquakes become deeper moving north from the subduction zone. Source: Karla Panchuk (2017) CC BY 4.0. Base maps with epicentres generated using the U. S. Geological Survey Latest Earthquakes website. Visit Latest Earthquakes

Earthquakes occur in subduction zones for a variety of reasons. Stresses associated with the collision of two plates cause deformation in the overriding plate, and thus shallow earthquakes. Shallow earthquakes also happen on the subducting slab when a locked zone (orange line, Figure 12.20) ruptures. The locked zone is where the largest earthquakes on Earth, called megathrust earthquakes, occur. There is the potential for a wider rupture zone on a gently dipping subduction zone boundary compared to other boundaries.

Figure 12.20 Factors contributing to earthquakes in subduction zones. Not all factors shown here are present in all subduction zones. Stress from bending, flexing, and stretching may cause ruptures. Changes in the mechanical properties of the mantle may affect how subucting slabs move, contributing additional stresses. The histogram at right shows the global average number of earthquakes at depth, for earthquakes greater than M5. The increase in earthquakes beneath 480 km may be caused in part by weakening as olivine transforms into high pressure/temperature polymorphs. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Green et al. (2010).

If subduction is rapid, the subducting plate will bend more as it enters the mantle (slab A in Figure 12.20), causing the upper edge of the plate to stretch, and the interior and lower edge to be compressed. Stress from bending can cause shallow to intermediate earthquakes on these plates. Even without bending, the subducting slab can become stretched by its own weight as it falls into the mantle.

The 410 km and 660 km discontinuities in Figure 12.20 mark boundaries where minerals transform into other, denser minerals that are stable at higher pressures and temperatures. When the subducting slab reaches the 660 km discontinuity (the top of the lower mantle), the increase in density in the surrounding mantle may slow down the leading edge of the sinking slab. Earthquakes can be generated when the slab is compressed by the lower mantle resisting its motion at the same time that the upper part of the slab continues to fall.

Slower rates of subduction mean that the subducting slab will enter the mantle at a lower angle (slab B in Figure 12.20). These slabs might not have earthquakes from being bent downward into the mantle, as with slab A, but earthquakes may be triggered by changes in stress if the plate relaxes and unbends.

The bar chart on the right of Figure 12.20 shows global average number of earthquakes that occur at different depths. Earthquakes are most abundant at the surface, and then fall to a minimum at 300 km. The number of earthquakes remains low until almost 500 km depth, and reaches a second peak around 600 km depth. The second peak might be explained by interactions between the subducting plate and the dense mantle beneath the 660 km discontinuity, but another hypothesis is that it is related to delayed mineral transformations. The subducting slab warms as it goes deeper into the mantle, but the warming is not uniform. The outer edges of the slab will warm before the interior does. The 410 km discontinuity is where olivine is transformed into the minerals wadsleyite and ringwoodite under normal mantle pressure and temperature conditions. However, if the interior of the subducting slab is still too cool at that depth, olivine will be retained to depths below 410 km. Olivine weakens prior to transforming into the high pressure minerals, and the weakening may make it easier for the slab to rupture.

Continent-Continent Convergence Zones

Where continents collide, earthquakes are scattered over a much wider area compared to earthquakes along mid-ocean ridges, transform margins, or subduction zones. An example is where the Indian plate collides with the Eurasian plate (Figure 12.21). At one time, India was a separate continent, and ocean crust separated India from the Eurasian plate. For a time, a subduction zone existed where ocean lithosphere from the Indian plate subducted beneath the Eurasian plate. But when the two land masses finally met, they became locked together and the subduction zone was closed off. Today the Indian plate is still pushing against the Eurasian plate in the regions indicated by the red arrows in Figure 12.21. The collision is accommodated by transform boundaries along the Indian plate. Regions of overall transform motion are indicated in Figure 12.21 with blue arrows.

Figure 12.21 Earthquakes of M4.5 and greater from 1990 to 2017 along the collision zone between the Indian and Eurasian plates. Red lines- plate boundaries; red arrows- collision zones; blue arrows- transform zones. Source: Karla Panchuk (2017) CC BY 4.0. Base maps with epicentres generated using the U. S. Geological Survey Latest Earthquakes website. Visit Latest Earthquakes

The majority of earthquakes in Figure 12.21 occur at depths less than 70 km, however they are still abundant down to 150 km, and extend to more than 300 km depth at some locations. Deeper earthquakes may be caused by continued northwestward subduction of part of the Indian plate beneath the Eurasian plate in this area. Even though the area is no longer a subduction zone, the subducted slab still remains, and is subject to stresses that can trigger earthquakes.

Some of the earthquakes in Figure 12.21 are related to the transform faults on either side of the Indian plate, and most of the others are related to the squeezing caused by the continued convergence of the Indian and Eurasian plates. That squeezing has caused the Eurasian plate to be thrust over the Indian plate, building the Himalayas and the Tibet Plateau to enormous heights. Most of the earthquakes of Figure 12.21 are related to the thrust faults shown in Figure 12.22 (and to hundreds of other similar ones that cannot be shown at this scale). The southernmost thrust fault in Figure 12.22 (the Main Boundary Fault) is equivalent to the convergent boundary in Figure 12.21.

Figure 12.22 Schematic diagram of the India-Asia convergent boundary, showing examples of the types of faults along which earthquakes are focused. Source: Steven Earle (2015) CC BY 4.0 view source after D. Vuichard (Figure 2.3) in Ives and Messerli (1989).

Intraplate Earthquakes

Intraplate earthquakes (within-plate earthquakes) are those that occur away from plate boundaries. Some intraplate earthquakes are related to human activities. When humans trigger earthquakes it is referred to as induced seismicity. In Saskatchewan there have been 20 earthquakes since 1985 (all less than magnitude 4), and the majority occurred near potash mines. Excavation changes the stress in surrounding rocks, so earthquakes may occur in the rocks above excavated parts of the mine. In Alberta, induced seismicity is triggered by hydraulic fracturing operations when water pressure increases along existing faults, causing them to slip.

Intraplate earthquakes not related to human activities often occur along ancient rift zones. In eastern Canada, the Charlevoix seismic zone (approximately 100 km northeast of Québec City; Figure 12.23), is associated with a rift-zone faults that developed when an ancient ocean basin began to form more than 500 million years ago. Coincidentally, the rocks of the Charlevoix Seismic Zone are also fragmented because of a meteorite impact (the crater margin is indicated by the blue circle in Figure 12.23), weakening them further. While the Charlevoix zone is far from any boundary of the North American plate, tectonic forces acting on plate boundaries are still transmitted to the interior of the continent, contributing to the stress that causes the faults along the rift zone to rupture.

Figure 12.23 Charlevoix seismic zone, site of intraplate earthquakes. The location of the Charlevoix seismic zone is indicated by the star on the map of Canada. Dots indicate earthquake epicentres. The size of the dots indicates magnitude. White lines indicate fault segments. The dashed circle marks the edge of a crater formed by a meteorite impact 342 million years ago. Source: Karla Panchuk (2017) CC BY-SA 4.0. Epicentre data from Earthquakes Canada. Click the image for more attributions and data sources.

Intraplate earthquakes can be large earthquakes. The Charlevoix seismic zone has had five earthquakes of magnitudes between 6 and 7 since 1663. The New Madrid seismic zone in the Mississippi River Valley had a series of four earthquakes with magnitudes between 7 and 8 in the winter of 1811-1812. The population of the region was sparse at the time, but today there are major cities in the New Madrid seismic zone, including Memphis, Tennessee, and St. Louis, Missouri.

References

Bao, X., and Eaton, D. W. (2016). Fault activation by hydraulic fracturing in western Canada. Science 354(6318), 1406-1409. doi: 10.1126/science.aag2583

Buck, W. R., Lavier, L. L., & Poliakov, A. N. B. (2005). Modes of faulting at mid-ocean ridges. Nature 434(7034), 719-23. doi: 10.1038/nature03358

Coastal and Marine Geology Program, U. S. Geological Survey (n.d.). Lake Baikal – A Touchstone for Global Change and Rift Studies. Visit website

DeMets, C., Gordon, R. G., & Argus, D. F. (2010). Geologically current plate motions. Geophysical Journal International 181, 1-80. doi: 10.1111/j.1365-246X.2009.04491.x

Earthquakes Canada, Natural Resources Canada (2016) Earthquake zones in Eastern Canada Visit website

Gendzwill, D. J., Horner, R. B., & Hasegawa, H. S. (1982). Induced earthquakes at a potash mine near Saskatoon, Canada. Canadian Journal of Earth Science 19, 466-475. doi: 10.1139/e82-038
Green, H. W. II, Chen, W.-P., & Brudzinski, M. R. (2010). Seismic evidence of negligible water carried below 400-km depth in subducting lithosphere. Nature 467, 828-830. doi:10.1038/nature09401
Hongyu, Y., Liu, Y., Harrington, R. M., & Lamontagne, M. (2016). Seismicity along St. Lawrence Paleorift Faults Overprinted by a Meteorite Impact Structure in Charlevoix, Québec, Eastern Canada. Bulletin of the Seismological Society of America 106(6), 2663-2673. doi:   https://doi.org/10.1785/0120160036
Ives, J. D., Messerli, B. (1989). The Himalayan Dilemma: Reconciling development and conservation. New York: Routledge. Access this book courtesy of the New Zealand Digital Library, University of Waikato
Myhill, R., & Warren, L. M. (2012). Fault plane orientations of deep earthquakes in the Izu-Bonin-Marianas subduction zone. Journal of Geophysical Research 117, B06307. doi:10.1029/2011JB009047
Nishikawa, T., & Ide, S. (2015). Background seismicity rate at subduction zones linked to slab-bending-related hydration. Geophysical Research Letters 42, 7081-7089. doi: 10.1002/2015GL064578
Sloan, R. A., Jackson, J. A., McKenzie, D., & Priestley, K. (2011). Earthquake depth distributions in central Asia, and their relations with lithosphere thickness, shortening and extension. Geophysics Journal International 185, 1-29. doi: 10.1111/j.1365-246X.2010.04882.x

85

12.4 The Impacts of Earthquakes

Earthquakes can have direct impacts, such as structural damage to buildings from shaking, and secondary impacts, such as triggering landslides, fires, and tsunami. The types and extent of impacts will depend on local conditions where the earthquake strikes. The geological materials in the area matter, as does the type of terrain, and whether the region is near the coast or not. The extent of impact and type of damage will depend on whether the area is predominantly urban or rural, densely or sparsely populated, highly developed or underdeveloped. It will depend on whether the infrastructure has been designed to withstand shaking.

Damage to Structures from Shaking

As with the example of the 1985 Mexico earthquake, the geological foundations on which structures are built will affect the amount of shaking that occurs. Earthquakes produce seismic waves that vibrate at different rates, or frequencies. Waves with rapid vibrations have a high frequency, and waves with slower vibrations have lower frequencies.

The energy of the higher frequency waves tends to be absorbed by solid rock. Lower frequency waves pass through solid rock without being absorbed, but are absorbed and amplified by soft sediments. It is therefore very common to see much worse earthquake damage in areas underlain by soft sediments than in areas of solid rock. During the 1989 Loma Prieta earthquake, parts of a two-layer highway in the Oakland area near San Francisco collapsed where they were built on soft sediments (Figure 12.24).

Figure 12.24 A collapsed section of the Cypress Freeway in Oakland California. Source: U. S. Geological Survey (1989) Public Domain view source

Building damage is also greatest in areas of soft sediments, and multi-storey buildings tend to be more seriously damaged than smaller ones. Buildings can be designed to withstand most earthquakes, and this practice is increasingly applied in earthquake-prone regions. Turkey is one such region, but even though Turkey had a relatively strong building code in the 1990s, adherence to the code was poor. Builders did whatever they could to save costs, including using inappropriate materials in concrete, and reducing the amount of steel reinforcing. The result was more than 17,000 deaths in the 1999 M7.6 Izmit earthquake (Figure 12.25). After two devastating earthquakes that year, Turkish authorities strengthened the building code further, but the new code has been applied only in a few regions, and enforcement of the code is still weak, as revealed by the amount of damage from a M7.1 earthquake in eastern Turkey in 2011.

Figure 12.25 Damage from the 1999 M7.6 Izmit, Turkey earthquake. Source: Left; USGS (1999) Public Domain view source; Right: USGS (1999) Public Domain view source

Structures underlain by sediments may be at risk of another hazard, called liquefaction, in which sediment is transformed into a fluid. When water-saturated sediments are shaken, the grains may lose contact with each other, and no longer support one another. Water between the grains holds them apart, causing the sediment to turn to mud and flow. The loss of support can lead to the collapse of buildings or other structures that might otherwise have sustained little damage. During the 1964 M7.6 earthquake in Niigata, Japan, liquefaction caused buildings to sink into the sediments (Figure 12.26).

Figure 12.26 Collapsed apartment buildings in the Niigata area of Japan. The material beneath the buildings was liquefied by the 1964 earthquake. Source: DOC/NOAA/NESDIS/NCEI (1964) Public Domain view source

Parts of the Fraser River delta are also prone to liquefaction-related damage. The region is characterized by a 2 m to 3 m thick layer of fluvial silt and clay above a layer of water-saturated fluvial sand that is at least 10 m thick (Figure 12.27). Under these conditions, seismic shaking can be amplified, and the sandy sediments will liquefy. This could lead to subsidence and tilting of buildings. Liquefaction can also contribute to slope failures and to fountains of sandy mud (sand volcanoes) in areas where there is loose saturated sand beneath a layer of more cohesive clay. Current building-code regulations in the Fraser delta area require that measures be taken to strengthen the ground beneath multi-story buildings prior to construction.

Figure 12.27 Recent unconsolidated sedimentary layers in the Fraser River delta area (top) and the potential consequences in the event of a damaging earthquake. Source: Steven Earle (2015) CC BY 4.0 view source

 

Experiments with Liquefaction

To see liquefaction for yourself, go to a sandy beach and find a place near the water’s edge where the sand is wet. While standing in one place on a wet part of the beach, start moving your feet up and down at a frequency of about once per second. Within a few seconds the previously firm sand will start to lose strength, and you’ll gradually sink in up to your ankles.

If you can’t get to a beach, put some sand into a small container, saturate it with water, and then pour the excess water off. Shake the container gently to get the water to separate and then pour the excess water away. You may have to do this more than once. Place a small rock on the surface of the sand. It should sit there for hours without sinking in. Now, holding the container in one hand gently thump the side or the bottom with your other hand, about twice a second. The rock should gradually sink in as the sand around it becomes liquefied.

Figure 12.28 Fine, damp sand before shaking (left) and after (right). Source: Steven Earle (2015) CC BY 4.0 view source

As you were moving your feet up and down or thumping the container, it’s likely that you soon discovered the most effective rate for getting the sand to liquefy. Stepping up and down as fast as you can (several times per second) on the wet beach would not have been effective, nor would you have achieved much by stepping once every several seconds. The body of sand vibrates most readily in response to shaking that is close to its natural harmonic frequency, and liquefaction is also most likely to occur at that frequency.

Slope Failure

Ground shaking during an earthquake can be enough to weaken rock and loose materials to the point of failure. Earthquakes can also trigger failures on slopes that are already weak. In January of 2011 a M7.6 earthquake offshore of El Salvador triggered slope failures that killed nearly 600 people (Figure 12.29).

Figure 12.29 A slope gives way in a suburb of San Salvador after the January 2001 earthquake offshore of El Salvador. This is one of hundreds of slope failures resulting from the earthquake. Source: U. S. Geological Survey (2001) Public Domain view source

Fires

Fires are commonly associated with earthquakes because gas lines rupture and electrical lines are damaged when the ground shakes. Most of the damage in the great 1906 San Francisco earthquake was caused by massive fires in the downtown area of the city (Figure 12.30). Some 25,000 buildings were destroyed by those fires, which were fueled by gas leaking from broken pipes. Fighting the fires was difficult because water mains had also ruptured. Today the risk of fires can be reduced through P-wave early warning systems if utility operators can decrease pipeline pressure and break electrical circuits.

Figure 12.30 Fires in San Francisco following the 1906 earthquake. Source: Pillsbury Picture Co. (1906) Public Domain, courtesy of the Library of Congress Prints and Photographs Division view source

Tsunami

Large earthquakes that take place beneath the ocean have the potential to displace large volumes of water. In a subduction zone, for example, the overriding plate becomes distorted by elastic deformation. It is squeezed laterally and pushed up (Figure 12.31 top). When an earthquake happens, the plate rebounds over an area of thousands of square kilometres, generating waves- a tsunami (Figure 12.31, middle). The waves spread across the ocean at velocities of several hundred kilometres per hour. Tsunami can make it to the far side of an ocean in about the same time as a passenger jet.

Figure 12.31 Tsunami triggered along a subduction zone. Top: The overriding crust is deformed because it is locked to the subducting slab. Middle: When the locked zone ruptures, the crust rebounds, and waves are created. Bottom: Tsunami waves have small amplitudes in the deep ocean, but once in shallow water, they slow down, causing the waves to become taller and closer together. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Top and middle modified after Steven Earle (2015) CC BY 4.0 view source. Bottom modified after COMET/UCAR (1997-2017) view source. Click the image for COMET/UCAR attribution and terms of use.

Tsunami waves gain their height as they travel through shallower waters. In the deep ocean, the waves may be so small as to go undetected by ships, but when they are slowed down by interacting with the ocean floor, they can become much taller (Figure 12.31, bottom). In the tsunami following the 2004 Sumatra earthquake, the tallest waves were more than 30 m high.

Subduction earthquakes must be large to generate significant tsunami. Earthquakes with magnitude less than 7 do not typically generate significant tsunami because the amount of vertical displacement of the sea floor is minimal. Sea-floor transform earthquakes, even large ones (M7 to M8), don’t typically generate tsunami either, because the motion is mostly side to side, not vertical.

Tsunami can have an impact across an entire ocean basin. They spread across the ocean at velocities of several hundred kilometres per hour, and can make it to the far side of an ocean in about the same time as a passenger jet. In 1700 a rupture along the Cascadia thrust fault running from Vancouver Island to northern California resulted in a M9 earthquake. It generated a tsunami that travelled across the Pacific Ocean, and was recorded in Japan nine hours later. A computer simulation of the tsunami (Figure 12.32) shows how long it took tsunami waves from the Cascadia earthquake to travel across the Pacific Ocean, and how high the waves were. Notice that over all, the waves decrease in height moving away from the rupture, but they increase in height again as they reach the opposite side of the Pacific Ocean.

Figure 12.32 Computer simulation of the tsunami from the 1700 M9 Cascadia earthquake. Colours show open-ocean wave heights. Contours show travel time in hours. Tsunami wave heights increase as the tsunami reached the western margin of the Pacific ocean. Source: NOAA/PMEL/Center for Tsunami Research (2011) Public Domain view source / view context

References

Earthquakes Canada, Natural Resources Canada (2016). The M9 Cascadia Megathrust Earthquake of January 26, 1700. Visit website

University Corporation for Atmospheric Research (2010). Propagation. Tsunamis. Visit website

U. S. Geological Survey (2014). Indian Ocean Tsunami Remembered — Scientists reflect on the 2004 Indian Ocean that killed thousands. Visit website

86

12.5 Forecasting Earthquakes and Minimizing Impacts

It has long been a dream of seismologists, geologists, and public safety officials to be able to accurately predict the location, magnitude, and timing of earthquakes on time scales that would be useful for minimizing danger to the public and damage to infrastructure (e.g., weeks, days, hours). Many methods of prediction that have been explored. These include using observations of warning foreshocks, changes in magnetic fields, episodic tremor and slip, changing groundwater levels, strange animal behaviour, patterns in the timing between earthquakes, and how stress is transferred after a rupture. So far, none of these has provided a reliable method. Although there are some reports of successful earthquake predictions, they are rare, and many are surrounded by doubtful circumstances.

The Parkfield Prediction Experiment

There was great hope for earthquake predictions late in the 1980s when attention was focused on part of the San Andreas Fault at Parkfield, approximately 200 km south of San Francisco, California. Between 1881 and 1965 there were five earthquakes at Parkfield. They were spaced at approximately 20-year intervals, all confined to the same 20 km-long segment of the fault, and all very close to M6 (Figure 12.33). Both the 1934 and 1966 earthquakes were preceded by small foreshocks exactly 17 minutes before the main quake.

Figure 12.33 Earthquakes on the Parkfield segment of the San Andreas Fault between 1881 and 2004. Source: Steven Earle (2015) CC BY 4.0 view source

The U. S. Geological Survey recognized this as an excellent opportunity to understand earthquakes and earthquake prediction, so they armed the Parkfield area with a huge array of geophysical instruments and waited. The next earthquake was expected to happen around 1987, but nothing happened! The “1987 Parkfield earthquake” finally struck in September 2004. Fortunately, all of the equipment was still there to record the earthquake, but it was no help from the perspective of earthquake prediction. There were no significant precursors to the 2004 Parkfield earthquake in any of the parameters measured, including tremors, changes in rock deformation, the magnetic field, the electrical conductivity of the rock, and creep (motion along the fault that is not accompanied by earthquakes). There was no foreshock. In other words, even though every available technique was used to monitor it, the 2004 earthquake came with no warning whatsoever.

Earthquake Probabilities

To be useful to the public and governments, earthquake predictions must be accurate most of the time, not just some of the time. If a prediction method is only accurate 10% of the time (and even that isn’t possible with the current state of knowledge), the public will lose faith in the process very quickly, and then will ignore all of the predictions. The hope for earthquake prediction is not dead, but it was hit hard by the Parkfield experiment.

Today the focus of efforts in earthquake-prone regions is to provide forecasts of earthquake probability. Earthquake probabilities express the likelihood that an earthquake of a given magnitude will occur at a location within a given period of time. An example of this approach for the San Francisco Bay region of California is shown in Figure 12.34. Based on a wide range of information, including past earthquake history, accumulated stress from plate movement, and known stress transfer, seismologists and geologists have predicted the likelihood of a M6.7 or greater earthquake on each of eight major faults that cut through the region. The greatest probabilities are on the San Andreas, Rogers Creek/Hayward faults, and Calaveras/Paicines faults. There is a 72% chance that a major and damaging earthquake will take place somewhere in the region prior to 2043.

Figure 12.34 Earthquake outlook for 2014-2047. Probabilities for individual faults are marked on the faults. Source: U. S. Geological Survey (2014) Public Domain view report

Using Earthquake Probability Information

Decision makers can use forecasts of earthquake probability to assist with educating the public about earthquake risks, and to determine what action is necessary to make infrastructure earthquake-safe. Building safe infrastructure requires strong building codes, and enforcement of those codes. Building code compliance is robust in most developed countries, but is inadequate in many developing countries.

New buildings are not the only ones requiring attention. Existing buildings — especially schools and hospitals — and other structures such as bridges and dams, must also be made safe. British Columbia began a multi-billion-dollar program in 2004 to make public schools safer for students. The program is focused on older public schools, because, according to the government, those built since 1992 already comply with modern seismic codes. Some schools would require too much work to make upgrading economically feasible and they are replaced. Where upgrading is feasible, the school is assessed carefully before any upgrade work is initiated. The seismic mitigation program identified 346 schools as being at high risk and in need of upgrades. As of December 2017, upgrades were completed at 168 schools, 28 schools were under construction or had approval to proceed with construction, and 150 did not yet have plans in place for upgrades.

The program in British Columbia illustrates a challenge with seismic upgrades of public buildings, which is that governments must make adequate funds available for the upgrades to be done in a timely manner. The priority allocated to funding those projects will depend on how urgent the need for upgrades is considered to be, given other demands on public funds.

Earthquake Preparedness

Earthquake preparedness involves the formulation of public emergency plans, including escape routes, medical facilities, shelters, and food and water supplies. It also includes personal planning, such as emergency supplies (food, water, shelter, and warmth), escape routes from houses and offices, and communication strategies (with a focus on ones that don’t involve the cellular network).

References

Finn, W. D. L., & Dexter, A. (2012). Risk Management Plan for School Seismic Mitigation Program. View report

Province of British Columbia (n.d.). Seismic Mitigation Program. Visit website

Province of British Columbia (2017). Seismic Mitigation Program Progress Report. View report

87

Chapter 12 Summary

The topics covered in this chapter can be summarized as follows:

12.1 What Is an Earthquake?

An earthquake is the shaking that results when a deformed body of rock snaps back to its original shape. The rupture is initiated at a point but quickly spreads across the area of a fault, with aftershocks initiated by stress transfer. Episodic tremor and slip is a periodic slow movement, accompanied by harmonic tremors, along the middle part of a subduction zone boundary.

12.2 Measuring Earthquakes

Earthquakes produce seismic waves that can be measured by a seismograph. The amplitudes of seismic waves are used to determine the amount of energy released by an earthquake- its magnitude. For the moment magnitude scale used today, the amount of energy released by an earthquake is proportional to the size of the rupture surface, the amount of displacement, and the strength of the rock. Intensity is a measure of the amount of shaking that occurs, and damage done at locations that experience an earthquake. Intensity will vary depending on the distance to the epicentre, the depth of the earthquake, and the type of geological materials present.

12.3 Earthquakes and Plate Tectonics

Most earthquakes happen at or near plate boundaries. Along divergent and transform boundaries earthquakes are shallow (less than 30 km depth), but at convergent boundaries they can be hundreds of kilometers beneath the surface. The largest earthquakes happen when a broad segment of the locked zone of a subduction zone ruptures. Intraplate earthquakes happen away from plate boundaries. They can be caused by human activities, or renewed motion on ancient faults.

12.4 The Impacts of Earthquakes

Damage to buildings is the most serious consequence of most large earthquakes. The amount of damage is related to the type and size of buildings, how they are constructed, and the nature of the material on which they are built. Other important consequences are fires, damage to bridges and highways, slope failures, liquefaction, and tsunami.

12.5 Forecasting Earthquakes and Minimizing Impacts

There is no reliable technology for predicting earthquakes, but the probability of one happening within a certain time period can be forecast. We can minimize earthquake impacts by ensuring that the public is aware of the risk, that building codes are enforced, that existing buildings like schools and hospitals are seismically sound, and that both public and personal emergency plans are in place.

Review Questions

  1. What causes the shaking during an earthquake?
  2. What is a rupture surface, and how does the area of a rupture surface relate to earthquake magnitude?
  3. What is an aftershock and how are aftershocks related to stress transfer?
  4. Episodic slip on the middle part of the Cascadia subduction zone is thought to increase the stress on the locked zone. Why?
  5. What is the difference between the magnitude of an earthquake and its intensity?
  6. How much more energy is released by a magnitude 7 earthquake compared to a magnitude 5 earthquake?
  7. The images below show earthquake locations for three regions of ocean lithosphere. The colour scheme indicates the depths of earthquakes. a) The images show a subduction zone, a slowly spreading mid-ocean ridge, and a rapidly spreading mid-ocean ridge. Which is which? b) In the image with the subduction zone, which side is the subducting plate, and which is the overriding plate?
    Figure 12.35 Earthquakes along plate margins. Dots are colour coded according to depth. Source: Map details from Lisa Christiansen, Caltech Tectonics Observatory (2008) view source. Click the image for terms of use.
  8. Why is earthquake damage likely to be more severe for buildings built on unconsolidated sediments as opposed to on solid rock?
  9. Why are fires common during earthquakes?
  10. What type of earthquake is likely to lead to a tsunami?
  11. What did we learn about earthquake prediction from the 2004 Parkfield earthquake?
  12. What are some of the things we should know about an area in order to help minimize the impacts of an earthquake?
  13. What is the difference between earthquake prediction and forecasting earthquake probability?

88

Answers to Chapter 12 Review Questions

1. Rocks under stress can deform elastically. When rocks break, or a rupture occurs along an existing fault, the deformed rocks snap back to their original shape, causing vibrations. The vibrations are the shaking felt during earthquakes.

2. The rupture surface is the surface over which there is displacement of rock during an earthquake. The magnitude of an earthquake is proportional to the area of the rupture surface.

3. An aftershock is an earthquake that is triggered when a rupture from a previous earthquake has transferred too much stress to rocks, causing further ruptures.

4. Episodic slip on the middle part of the Cascadia subduction zone decreases stress within that area, but some of that stress is transferred to the locked zone along the plate boundary. This increases the amount of stress on the locked part.

5. Magnitude is the amount of energy released by an earthquake. Each earthquake has only one magnitude, although there are different ways of measuring it, and they may give slightly different results. Intensity is a measure of the amount of damage done or what people felt. Intensity varies depending on the distance to the epicentre and the type of rock or sediment underlying an area.

6. Each unit increase in magnitude corresponds to a 32x increase in energy. The difference between M5 and M7 is two units of magnitude. An earthquake that is two magnitude units larger would release 32 x 32 = 1,024 times as much energy.

7.  a) A: Slowly spreading mid-ocean ridge- earthquakes are shallow and evenly distributed; B: Subduction zone- earthquakes range in depth; C: Rapidly-spreading mid-ocean ridge- earthquakes are shallow, and occur in patches, corresponding to a greater likelihood of earthquakes where transform faults offset ridges.

b) The plate to the right of the chain of earthquakes is the subducting plate. The earthquakes increase in depth moving from right to left, indicating that the subducting slab is getting deeper farther to the left.

8. Unconsolidated sediments, especially if they are saturated with water, can lose strength when subjected to earthquake shaking. This can cause buildings to subside or tilt. Unconsolidated sediments can also amplify the vibrations of an earthquake.

9. Gas lines and electrical transmission wires are typically damaged during an earthquake, and this can lead to serious fires.

10. A large subduction-zone earthquake (greater than M7.5) can generate a tsunami because those earthquakes result in sufficient vertical displacement of the sea floor.

11. The 2004 Parkfield earthquake showed that we cannot rely on foreshocks to predict earthquakes, or on any of the many other parameters that were being carefully measured around Parkfield in the years leading up to the quake.

12. We should know about the history of past large earthquakes, the typical locations of small earthquakes, the types of geological materials beneath the surface (especially soft water-saturated sediments), the types of infrastructure that is present, and the various ways that people can be evacuated from an area or assistance can be brought in.

13. Forecasting involves estimating the risk of an earthquake happening in a region within a period of time. Prediction involves stating that an earthquake is likely to happen at a certain location on a specific day or month or year in the future. With our current state of knowledge of earthquakes, prediction is not possible.

XIII

Chapter 13. Geological Structures and Mountain Building

 Adapted by Karla Panchuk from Physical Geology by Steven Earle

Figure 13.1 Folded rocks in the Cariboo Mountains of BC. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photograph by Drew Brayshaw (2009) CC BY-NC 2.0 view source. Click the image for more attributions.

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

Folds, like those in the centre of Figure 13.1, are a common feature of mountain belts. Have you ever wondered how something as hard as rock could flex and bend to make folds, and what forces are required? Geologists have.

Observing and analysing geological structures helps us to understand the kinds of forces that affect rocks, both on small scales, and on scales as large as tectonic plates. With that knowledge we can understand how plate tectonic processes change the shape of Earth’s lithosphere. We can also piece together a history of past changes, including how mountain belts are formed. Structural geologists make careful observations of the orientations of breaks and bends in rocks, and can compile those measurements into maps of geological structures. These maps can be valuable tools for finding mineral resources.

89

13.1 Stress and Strain

Plate collisions and the accumulated weight of overlying rocks exert forces on rocks at depth. While the size of the force is important, it also matters whether the force is distributed over a wide region, or tightly focused on a small area. The same force will have a greater effect when acting over a small area than when acting over a larger area. If you have ever used snowshoes to walk across a snow bank without sinking in, you have taken advantage of the effects of distributing force (your mass acted upon by gravity) over a wider area (the area of your snowshoes rather than the soles of your boots). Stress is force adjusted for the area over which it is distributed. Strain is the change in shape that happens when rocks are deformed by stress.

Types of Stress

Stresses fall into two categories: normal stress acts at right angles to a surface, and shear stress acts parallel to a surface (Figure 13.2).  Normal stress is subdivided into compression, when the stresses are squeezing a rock, and tension, when stress is pulling it apart. Rocks undergo compression in regions where plates are colliding, or where they are being buried beneath other rocks. Rocks experience tension where divergence is happening, such as when a continent is beginning the rifting process.  Shear stress is characteristic of transform plate boundaries, where plates are moving side by side.

Figure 13.2 Rocks can be affected by normal stress (compression and tension) or shear stress. Source: Karla Panchuk (2016) CC BY 4.0

Although Figure 13.2 shows only one set of stress arrows for each scenario, rocks within the Earth are subject to stress from all directions.  The relative size of the stresses in different directions will determine the response of the rock. Consider a deeply buried rock being stretched as a continent breaks apart (Figure 13.3). It is also being compressed by the weight of overlying sediments and rocks, but the stress from compression is relatively small compared to the tension from rifting. The net effect of stress acting on the rock will be determined more by the tension from rifting than by the compression from overlying rocks.

Figure 13.3 A rock layer (dark brown) is compressed by the weight of rocks above, and stretched by rifting. The sizes of the arrows indicate the relative sizes of the stresses. Source: Karla Panchuk (2018) CC BY 4.0

Rocks experience stress from all directions, but it is possible to break down stresses into three directions, just like a graph with x, y, and z axes. In diagrams showing these three directions, the sizes of arrows representing each direction will indicate the relative size of stresses, as they do in Figure 13.3.  Analyzing stress in this way makes it much easier to describe the stresses operating on a rock, and to understand what their net effect will be.

Types of Strain

How a rock responds to stress depends on many factors. The “how” is not simply a matter of how much strain a rock will undergo, but what type of strain will occur. Is the deformation permanent or temporary? Does the rock break or does it deform without breaking?

Elastic Strain

Elastic strain is reversible strain. You can think of elastic strain as what happens to the elastic waistband of your favourite sweatpants when you put them on. The elastic stretches to let you into your pants, and once you’re in them, it shrinks to keep them from falling down. When you take the pants off again, the elastic goes back to its original shape. Similarly, rocks undergoing elastic strain will snap back to their original shape once the stress is removed. Rocks snapping back to their original shape undergo elastic rebound. Elastic rebound of rocks on a large scale can have profound consequences, because the energy released causes the Earth to vibrate. We experience those vibrations as earthquakes.

Plastic Strain

If enough stress is applied, the changes that a material undergoes to accommodate the stress will leave it permanently deformed. When the stress is removed, the material does not go back to its original shape. The permanent deformation is called plastic strain.

Ductile or Brittle?

Ductile deformation refers to deformation that happens by flowing or stretching. The marble monument in Figure 13.4 is undergoing ductile deformation as it sags beneath its own weight.

Figure 13.4 A marble monument at the Westminster Hall and Burying Ground in Baltimore, Maryland. The horizontal surface is undergoing slow ductile deformation as it sags beneath its own weight. The author Edgar Allan Poe is interred nearby. Source: Ray Pennisi (2007) CC BY-NC 2.0 view source

When a material breaks, it has undergone brittle deformation (Figure 13.5). The stone cylinders in Figure 13.5 are part of an experiment to test the strength of the rock. The cylinder on the right looked like the cylinder on the left before it was compressed, with force applied to the top and bottom. Strain gauges have been glued on to measure the amount of deformation lengthwise and across the cylinders.

Figure 13.5 Cylinders of rock used to test the strength of rock under compression. The cylinder on the left has been equipped with strain gauges to measure the amount of deformation. The cylinder on the right has undergone brittle deformation after being compressed. Source: Karla Panchuk (2016) CC BY 4.0

A material can undergo more than one kind of deformation when stress is applied. The barrel-shaped cylinder of potash in Figure 13.6 (right) originally looked like the cylinder on the left. The cylinder was compressed, with stress applied from the top and bottom. Initially, it underwent ductile deformation and thickened in the middle, creating the barrel shape. But as more stress was applied, the cylinder eventually underwent brittle deformation, resulting in the crack across the middle.

Figure 13.6 Cylinders of potash before and after deformation. The potash underwent ductile deformation before it finally broke. Source: Karla Panchuk (2018) CC BY 4.0

Factors That Determine How A Rock Will Deform

A rock is not limited to exclusively brittle deformation, or exclusively ductile deformation. Even the deformed rock in Figure 13.5, which has clearly undergone brittle deformation, shows a slight curvature on the right side, near the top. This indicates that a small amount of ductile deformation occurred before brittle failure.

For a given rock, deformation will be different depending on the amount of stress applied. Up to a point, rocks undergo elastic deformation, and will spring back to their original shape after the stress is removed. If more stress is applied, the rock may deform in a ductile manner. If stress increases further, the rock may fracture. The amount of stress required in each case will depend on the type of rock, as well as conditions such as pressure and temperature.

Composition

In general, sedimentary rocks will be more likely to undergo ductile deformation than igneous or metamorphic rocks under the same conditions. Rocks within each group will also deform differently.

Boudinage structures (Figure 13.7) highlight the effect of composition on how rocks deform. These structures occur when a stronger rock more prone to brittle deformation is surrounded by weaker rocks prone to ductile deformation. The stronger rock will fracture into segments, called boudins, and the weaker rock will flow into the spaces between. In Figure 13.7 (top), the white layer reached the stage of pinching off, just before separating into segments. The surrounding black layer flowed in to fill the gap where the pinch was happening. Remarkably, the white layer itself contains a dark layer that has fragmented into boudins. Not all boudins break into blocky segments. Some display more ductile deformation (Figure 13.7, bottom).

Figure 13.7 Two examples of boudinage structure. Top- The white layer has pinched off into segments, and the surrounding black layers have flowed into the gap forming between segments. Within the white layer is a thinner black layer that has also broken into segments. Bottom- Boudins displaying ductile deformation. Source: Top- Marek Cichanski (2012) CC BY-NC 2.0 view source. Bottom- Joyce McBeth (n.d.) CC BY 4.0 

Temperature and Pressure

At higher temperatures, and under higher confining pressures, rocks are more likely to undergo ductile deformation. Confining pressure is the stress that a material experiences uniformly from all sides as a result of the weight of material above and around it. The pressure that a diver feels deep in the ocean is confining pressure due to the weight of water above and around the diver. This kind of confining pressure is called hydrostatic pressure.  Within Earth, the confining pressure is due to the weight of overlying rocks. Confining pressure due to the weight of rocks is called lithostatic pressure.

The rocks in Figures 13.5 and 13.6 experienced confining pressure from the atmosphere, and temperatures comfortable for the humans working in the lab. Under those conditions the rocks ultimately underwent brittle failure when they were compressed in the lab. Deep within the crust, the temperatures and confining pressures are far greater. Deep enough within the crust, both samples would undergo only ductile deformation if the same amount of stress were applied as in the experiment. The depth at which temperatures and confining pressures are high enough for rocks to go from brittle deformation to ductile deformation is called the brittle-ductile transition zone.

The brittle-ductile transition zone occurs between approximately 10 km and 30 km depth, corresponding to temperatures around 300 ºC and greater. The depth at which temperatures reach 300 ºC at any particular location will depend on heat flow at that location. In continental crust, rocks at 300 ºC are deeper than in ocean crust. The change in pressure with depth also varies, depending on the mass and density of rocks. If depths are measured relative to sea level, the pressure at 10 km measured beneath a tall mountain belt will be greater than the pressure at 10 km measured within ocean crust.

Experiments like those shown in Figures 13.5 and 13.6 can be used to determine where the brittle-ductile transition zone will be for a particular rock type. Experimenters apply stress to sample of a rock for a range of temperatures and confining pressures. They note the conditions under which the rock breaks or deforms in a ductile manner, and plot those on a graph (Figure 13.8). The results in Figure 13.8 are from experiments on limestone. The vertical axis is pressure. The more pressure, the deeper the rock would have to be within the Earth to experience that pressure. The white line represents the brittle-ductile transition zone. Above the white line are pressures and temperatures under which the limestone would fracture. Below the white line in the tan area are pressures and temperatures where the limestone would deform by flowing. Notice that the higher the temperatures, the less confining pressure is required for ductile deformation.

Figure 13.8 Experimental results on limestone with tension applied (left) and compression applied (right). Source: Karla Panchuk (2018) CC BY 4.0 modified after Heard (1960)

How Stress Is Applied

The limestone experiments were performed by applying stress as tension (Figure 13.8 left) and again by applying stress as compression (right). When tension was applied, temperature and confining pressure had to be much higher before ductile deformation occurred. Under compression, ductile deformation was possible with far less confining pressure, and at lower temperatures.

Strain rate, the rate at which deformation occurs, also makes a difference. If stress is applied at a rate that causes rapid deformation, the rock will be more likely to fracture than if deformation happens slowly. The marble slab in Figure 13.4 is a good example of this. It has sagged rather than broken because the rate of deformation has been very slow, at millimetres per decade.

Fluids

When rocks are under pressure, fluids trapped within the pore spaces of rocks- the gaps between grains- are also under pressure. Higher confining pressure is required for deformation to be ductile rather than brittle, but pressure from fluids, called pore pressure, resists the confining pressure. The result is that the effective confining pressure is lower than it would be without the fluids. Depending on the amount of pore pressure, and how close the rock is to the brittle-ductile transition zone, pore pressure could cause brittle failure in a rock that would otherwise undergo ductile deformation.

Stress and Geological Structures

Many different geologic structures can form when stress is applied to rocks. Structures form as a result of fracturing, tilting, folding, stretching, and squeezing (Figure 13.9). Some structures, like the fractures that make basalt columns (Figure 13.9, upper left), happen when rocks shrink due to cooling, but others are a consequence of plate tectonic forces. The types of structures that form depend on the plate tectonic setting and other geological conditions, making them valuable tools for understanding what happened to the rocks. The following sections address the different kinds of structures that form, and what information we can gather from these structures to learn more about the tectonic environment and regional geology.

Figure 13.9 Structures resulting from deformation. Upper left- Fracturing in basalt near Whistler, British Columbia. Upper right- Tilting of sedimentary rock near Exshaw, Alberta. Lower left- Stretched limestone (light grey) and chert (dark grey) from Quadra Island, British Columbia. Lower right- Faulted shale near Cache Creek, British Columbia. Rocks above the fault moved up relative to those below. Source: Karla Panchuk (2018) CC BY 4.0. Photographs by Steven Leslie (2015) CC BY 4.0 view source

References

Heard, H. C. (1960). Transition from Brittle Fracture to Ductile Flow in Solenhofen Limestone as a Function of Temperature, Confining Pressure, and Interstitial Fluid Pressure. In D. Griggs & D. Handin (Eds.), Rock Deformation (A Symposium): Geological Society of America Memoir 79 (pp. 193-226). https://doi.org/10.1130/MEM79

90

13.2 Folds

Folds are a type of ductile deformation. They form when rocks bend in response to stress. The sides of a fold are its limbs (Figure 13.10). The limbs meet in a region of curvature called the hinge zone. A fold’s axial surface is an imaginary surface that runs along the hinge zone and cuts the fold in half. The line that forms when the axial surface intersects another surface, such as the top of a bed, is called the axial trace. Axial traces are sometimes marked on geological maps to show the location of the fold’s hinge zone.

Figure 13.10 The parts of a fold. A fold consists of limbs that meet at the hinge zone. An axial surface bisects the fold along the hinge zone. The axial trace is where the axial surface intersects another surface, such as the top of a bed. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo: Ron Schott (2009) CC BY-NC-SA 2.0 view source

 

Fold Classification

Synclines and Anticlines

Folds can be classified according to the whether the limbs slope toward or away from the hinge zone. If the limbs slope toward the hinge zone (i.e., the hinge zone points downward), as in the fold in the left of Figure 13.11, the fold is called a syncline. If the limbs slope away from the hinge zone  (i.e., the hinge zone points upward), the fold is called an anticline.  There is an anticline on the right side of Figure 13.11. The fold in Figure 13.10 is also an anticline. Sometimes an anticline or a syncline will occur by itself, but they can also occur in a series of alternating synclines and anticlines, similar to the way the anticline and syncline share a limb in Figure 13.11. A sequence of linked anticlines and synclines is called a fold train.

Figure 13.11 An asymmetrical syncline linked to an anticline on a beach in Cornwall, United Kingdom. The beds slope toward the hinge at different angles on either side of the axial surface. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo: Harry Soar (2014) CC BY-NC-SA 2.0 view source

Symmetrical, Asymmetrical, Overturned, and Recumbent

In a symmetrical fold, the limbs slope at approximately the same angle on either side of the axial surface. The fold in Figure 13.10 is symmetrical. In an asymmetrical fold, the limbs slope at different angles on either side of the axial surface. The syncline in Figure 13.11 is asymmetrical. The limb on the left side of the syncline slopes toward the hinge at a steeper angle than the limb on the right.

If the fold is sufficiently tilted that the beds on one side have been tilted past vertical, and are sloping in the same direction, the fold is overturned (Figure 13.12).

Figure 13.12 Overturned folds in Andalusia in southern Spain. Some limbs have been overturned far enough to be sloping in the same direction on either side of the axial trace. Source: Karla Panchuk (2018) CC BY-NC-SA 2.0. Photo: Ignacio Benvenuty Cabral (2017) CC BY-NC-SA 4.0 view source

It is possible for rocks to be folded so tightly that the fold limbs are nearly parallel. Folds with parallel limbs are called isoclinal folds. A recumbent fold is an isoclinal fold that has been overturned to the extent that the limbs are horizontal (Figure 13.13).

Figure 13.13 A recumbent fold has limbs that are nearly parallel, and an axial trace that is nearly horizontal. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Photo: Ignacio Benvenuty Cabral (2017) CC BY-NC-SA 4.0 view source

Folds in the Landscape

Folds can be of any size, and it’s very common to have smaller folds within larger folds (Figure 13.14).  Large folds can extend over 10s of kilometres, and very small ones might only be visible under a microscope.

Figure 13.14 Folded limestone (grey) and chert (rust-coloured) in rocks of the Triassic Quatsino Formation on Quadra Island, British Columbia.  The image is about 1 m across. Source: Steven Earle (2015) CC BY 4.0 view source

When folded rocks are weathered and eroded, they can alter the landscape by forming long ridges and valleys (Figure 13.15). Ridges and valleys curve into V-shapes if the hinge of the fold is not horizontal. A fold with a hinge that slopes downward is called a plunging fold (Figure 13.16).

Figure 13.15 Ridges and valleys in central Pennsylvania formed from weathered and eroded folds. The V-shapes indicate the folds are plunging. Source: NASA on the Commons (2001) Public Domain view source

 

Figure 13.16 Plunging folds have sloping hinges. Plunging folds are described in terms of the plunge angle, the angle the hinge makes with a horizontal line. Inset- When a plunging fold intersects a surface, the result is a V-shaped pattern. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photo- Dieter Mueller (2004) CC BY-SA 3.0 view source

Folds can create landforms, but anticlines are not necessarily expressed as ridges in the terrain. Likewise, synclines do not necessarily appear as valleys. When folded rocks erode, the landform that results depends how resistant individual layers are to erosion. For example, if the rocks in the interior of an anticline are more resistant to weathering than the surrounding rocks, a ridge will result (e.g., the low hill represented by units 4 and 5 in Figure 13.17, top). On the other hand, if rocks in the interior of the anticline are weaker, a valley will result (Figure 13.17, bottom, units d1 and d2). Similarly, a syncline with stronger rocks in the interior will weather to form a ridge, and a syncline with weaker rocks in the interior will weather to form a valley.

Figure 13.17 Cross-sections of eroded folds expressed as hills and valleys, from an early study on the geology of Wales, Devon, and Cornwall. Top- An anticline in Shropshire, England. Beds in the interior of the anticline form a gentle hill. Bottom- An anticline in Herefordshire, England in which beds in the interior of an anticline weathered to form a valley. Source: Symonds (1872) Public Domain. View source: TopBottom

 

Exercise: Fold Types

What kind of folds are shown here? If you are finding it difficult to see the folds, follow the trace line formed by the white beds through the outcrop.

Figure 13.18 Folds in the Rocky Mountains near Golden, British Columbia. Source: Steven Earle (2015) CC BY 4.0 view source

References

Symonds, W. S. (1872). Records of the rocks; or, Notes on the geology, natural history, and antiquities of North & South Wales, Devon, & Cornwall. London: J. Murray  Read the book

 

91

13.3 Fractures, Joints, and Faults

When rocks break in response to stress, the resulting break is called a fracture. If rocks on one side of the break shift relative to rocks on the other side, then the fracture is a fault. If there is no movement of one side relative to the other, and if there are many other fractures with the same orientation, then the fractures are called joints. Joints with a common orientation make up a joint set (Figure 13.19).

Figure 13.19 Joint sets have broken these siltstone and shale beds into long rectangular planks. Source: Michael C. Rygel (2008) CC BY-SA 3.0 view source

Jointing

Most joints form when the overall stress regime is one of tension (pulling apart) rather than compression. The tension can be from a rock contracting, such as during the cooling of volcanic rock (Figure 13.9, upper left). It can also be from a body of rock expanding. Exfoliation joints, which make the rock appear to be flaking off in sheets (Figure 13.20), occur when a body of rock expands in response to reduced pressure, such as when overlying rocks have been removed by erosion.

Figure 13.20 Half Dome at Yosemite National Park is an exposed granite batholith that displays exfoliation joints, causing sheets of rock to break off. Source: HylgeriaK (2010) CC BY-SA 3.0 view source

Nevertheless, it is possible for joints to develop where the overall regime is one of compression. Joints can develop where rocks are being folded, because the hinge zone of the fold is under tension as it stretches to accommodate the bending (Figure 13.21).

Figure 13.21 Joints developed in the hinge zone of folded rocks. Source: Steven Earle (2015) CC BY 4.0 view source

Joints can also develop in a rock a rock under compression as a way to accommodate the change in shape (Figure 13.22).  The joints accommodate the larger compression stress  (larger red arrows) by allowing the rock to stretch in the up-down direction (along the green arrows).

Figure 13.22 Joints developing to accommodate the larger horizontal component of compression (large red arrows). Source: Steven Earle CC BY 4.0 view source


Faulting

A fault is a boundary between two bodies of rock along which there has been relative motion (e.g., Figure 13.23). Some large faults, like the San Andreas fault in California or the Tintina fault, extending from northern British Columbia through central Yukon and into Alaska, show evidence of hundreds of kilometres of motion. Other faults show only centimetres of movement. In order to estimate the amount of motion on a fault, it is necessary to find a feature that shows up on both sides of the fault, and has been offset by the fault. This could be the edge of a bed or dike as in Figure 13.23, or it could be a landscape feature, such as a fence or a stream.

Figure 13.23 View looking down on a fault (white dashed line) in intrusive rocks on Quadra Island, British Columbia. The pink dyke has been offset approximately 10 cm by the fault (length of the white arrow). Source: Steven Earle (2015) CC BY 4.0 view source

Types of Faults

Different kinds of faults develop under different stress conditions. We describe faults in terms of how the rocks on one side of the fault move relative to the other.

Dip-Slip Faults

Dip-slip faults are so named because the dominant motion involves moving up or down the dipping (tilting) fault plane. In dip-slip faults we identify rock above the fault as the hanging wall, (or headwall) and the rock beneath as the footwall. These terms were originally used by miners to describe the rocks above and below an ore body (Figure 13.24).

Figure 13.24 The hanging wall (or headwall) of a fault is the rock above the fault. The footwall is the rock below. These terms were originally used by miners to describe the rocks above and below an ore body. Source: Photo- Gold Hill Mine, Yukon Territory, by Eric A. Hegg (1898) Public Domain view source. Diagram- Karla Panchuk (2018) CC BY 4.0

Tension produces normal faults, in which the crust undergoes extension. This permits the hanging wall to slide down the footwall in response to gravity (Figure 13.25, left). Compression produces reverse faults, pushing the hanging wall up relative to the footwall. Reverse faults shorten and thicken the crust (Figure 13.25, right).

Figure 13.25 Dip slip faults. Normal faults are caused by tension, while reverse faults happen during compression. Source: Karla Panchuk (2018) CC BY-SA 4.0, modifed after Woudloper (2010) CC BY-SA 3.0 view source

Strike-Slip Faults

Faults where the motion is mostly horizontal and along the “strike” or the length of the fault are called strike-slip faults (Figure 13.26 bottom). These happen where shear stress causes bodies of rock to slide sideways with respect to each other, as is the case along a transform boundary. If the far side moves to the right, as in Figures 13.23 and 13.26 (right), it is a right-handed, right-lateral,or dextral strike-slip fault. If the far side moves to the left it is a left-handed, left-lateral, or sinistral strike-slip fault.

Figure 13.26 Strike-slip faults. Rocks on either side of the fault move parallel to the fault. In dextral strike-slip faults the far side moves to the right of the observer. In sinistral strike-slip faults the far side moves to the left of the observer. Source: Karla Panchuk (2018) CC BY 4.0

Different Tectonic Settings Have Distinct Types of Faults

Horst and Graben Structure

In areas that are characterized by extensional tectonics, and with many normal faults arranged side-by-side, some blocks may subside (settle downward) relative to neighbouring parts. This is typical in areas of continental rifting, such as the Great Rift Valley of East Africa or in parts of Iceland. In such situations, blocks that move down relative to the other blocks are graben, and elevated blocks with graben on either side are called horsts. There are many horsts and grabens in the Basin and Range area of the western United States, especially in Nevada. Part of the Fraser Valley region of British Columbia, in the area around Sumas Prairie, is a graben.

Figure 12.14  Depiction of graben and horst structures that form in extensional situations.  All of the faults are normal faults.  [SE]
Figure 13.27  Graben and horst structures form where extension is happening. All of the faults are normal faults. Source: Steven Earle (2015) CC BY 4.0 view source

Thrust Faults

Thrust faults are a type of reverse fault with a very low-angle fault plane. The fault planes of thrust faults typically slope at less than 30°. Thrust faults are relatively common in mountain belts that were created by continent-continent collisions. Some represent tens of kilometres of thrusting, where thick sheets of sedimentary rock have been pushed up and over other layers of rock (Figure 13.28).

Figure 12.15 Depiction a thrust fault.  Top: prior to faulting.  Bottom: after significant fault offset. [SE]
Figure 13.28 A thrust fault. Top: prior to faulting. Bottom: after significant fault offset. Source: Steven Earle (2015) CC BY 4.0 view source

There are numerous thrust faults in the Rocky Mountains, and a well-known example is the McConnell Thrust, along which a sequence of sedimentary rocks about 800 m thick has been pushed for about 40 km from west to east over underlying rock (Figure 13.29). The thrusted rocks range in age from Cambrian to Cretaceous, so in the area around Mt. Yamnuska Cambrian-aged rock (around 500 Ma) has been thrust over, and now lies on top of Cretaceous-aged rock (around 75 Ma) (Figure 13.30).

Figure 12.16  Depiction of the McConnell Thrust in the eastern part of the Rockies.  The rock within the faded area has been eroded. [SE]
Figure 13.29  The McConnell Thrust in the eastern part of the Rockies. The rock within the faded area has been eroded. Source: Steven Earle (2015) CC BY 4.0 view source
Figure 13.30 The McConnell Thrust at Mt. Yamnuska near Exshaw, Alberta. Cambrian limestones have been thrust over top of Cretaceous mudstone. Source: Steven Earle (2015) CC BY 4.0 view source

Exercise: Fault Types

 

Figure 13.31 A dip-slip fault. Source: Steven Earle (2015) CC BY 4.0 view source

https://h5p.org/h5p/embed/190965

https://h5p.org/h5p/embed/190967

 

 

 

92

13.4 Mountain Building

Some of Earth’s mountains are entirely or almost entirely the result of volcanic activity. These include volcanic islands like the Hawai’ian hotspot volcanoes, and newly formed volcanic island arcs along subduction zones. But the majority of mountain building on Earth is the result of plate tectonic forces that deform Earth’s crust through faulting and folding. Mountain building through folding and faulting may or may not be supplemented by volcanic activity.

Mountain Building Along Convergent Margins

Mountain building along convergent margins is referred to as orogeny, and the mountains that are built are called orogens.

Ocean-Continent Collision

In ocean-continent collision zones, folding and faulting of rocks combines with volcanism to build mountains. An example of mountains built this way is the Sierra Nevada mountain range in Utah and Nevada. The orogeny that formed the Sierra Nevada range began around 140 million years ago.

The mountain range was built up by igneous intrusions and volcanic eruptions along a continental volcanic arc (Figure 13.32). The terrain was altered further inland as well. Sheets of rock were thrust on top of each other, and pushed inland along a detachment fault, similar to the example of the McConnell Thrust in Figure 13.29.

Figure 13.32 Orogeny in an ocean-continent collision zone. Mountains form from subduction zone volcanism, and from large sheets of rock that are thrust inland and folded. Materials accumulating on the leading edge of the continent in an accretionary wedge are eventually smashed onto the continent, adding to continental crust. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Ron Blakey, NAU Geology (n.d.) view source. Click the image for terms of use.

Continental crust flexed downward because of the weight of the mountains, and this formed a fore arc basin seaward of the new mountain range. Sediments accumulated within that basin. The leading edge of the continent also collected sediments and igneous rock scraped off the subducting plate, forming an accretionary wedge. Over time, the force of the collision would smash the basin sediments and the accretionary wedge against the continent, turning it into new continental crust.

Continent-Continent Collision

When two continents collide, it means the closure of a subduction zone, and an end to volcanism. The Alleghenian Orogeny, which brought together North America and Africa, helping to form Pangea, is an example of mountain building in a continent-continent collision zone. Before the continents came into contact with each other, mountain building on the eastern coast of North America would have involved deformation from an ocean-continent collision, as with Figure 13.32. But as subduction proceeded, the subducting plate drew Africa closer and closer to North America. The gap between the two continents began to close, and fill with sediments (Figure 13.33, top).

Figure 13.33 Orogeny by continent-continent collision. The formation of Pangea included the merging of Africa and North America. This closed an ocean basin and stopped subduction along the coast of North America. Volcanism ended with the closure of the ocean basin, but mountains continued to grow through folding and faulting. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Ron Blakey, NAU Geology (n.d.) view source. Click the image for terms of use.

While a subduction zone existed, the addition of water to the mantle permitted partial melting of mantle rocks, and thus volcanic activity. However, when the two continents collided, the subduction zone closed off and volcanism was no longer possible. As the continents smashed together, deep faults formed and stacked blocks of crust on top of each other. Old faults were reactivated. Rocks also began to shift along the boundary between an earlier orogen, the Taconic Orogen, and North America (Figure 13.33, bottom).  When the continents had finally merged, Africa met North America along a suture zone with remnants of a continental volcanic arc on one side, and folded and faulted sedimentary rocks on the other.

Mountain Building Along Divergent Margins

When continents begin to split apart, normal faults form. This can lead to large blocks of crust that are tilted, raised, or lowered compared to adjacent blocks. Blocks that are elevated compared to adjacent blocks can form another type of mountain, called a fault-block mountain. Fault block mountains formed in eastern North America when Pangea began to split up, and Africa pulled away from North America (Figure 13.34).

Figure 13.34 Fault-block mountains formed in a rift zone. Magma can move up along normal faults, resulting in igneous intrusions, or volcanic eruptions. Over time, valleys between elevated blocks will fill with sediment as the blocks erode. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Ron Blakey, NAU Geology (n.d.) view source. Click the image for terms of use.

Over time, elevated blocks erode, filling up valleys with sediment. The thinning of continental crust that occurs with rifting can decrease the pressure on mantle rocks enough to trigger partial melting. Magma can move up along the normal faults, forming igneous intrusions, or feeding volcanoes. The Palisades Sill in New York and New Jersey is a result of rift-zone magmatism. It is a cliff-like feature resulting from erosion that exposed the tip of a structure like the sills in Figure 13.34.

93

13.5 Measuring Geological Structures

Documenting the characteristics of geological structures is used to understand the geological history of a region. One of the key features to measure is the orientation, or attitude, of bedding. We know that sedimentary beds are deposited in horizontal layers, so if the layers are no longer horizontal, then we can infer that tectonic forces have folded or tilted them.

The orientation of a planar feature, such as a bed of sedimentary rock, can be described with two values. The strike of the bed is the compass orientation of a horizontal line on the surface of the bed. The dip is the angle at which the surface tilts down from the horizontal (Figure 13.35). The dip is measured perpendicular to strike, otherwise the dip angle that is measured will be smaller than the actual tilt of the bed.

Figure 13.35 Strike and dip for tilted sedimentary beds. Water provides a horizontal surface. The strike and dip symbol is a T with the long horizontal bar representing the strike direction, and the small tick mark indicating the dip direction. The dip angle is written next to the tick mark. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view source

It may help to imagine a vertical surface, such as a wall in your house. The strike is the compass orientation of the wall and the dip is 90˚ from horizontal. If you could push the wall so it is leaning over, but still attached to the floor, the strike direction would be the same, but the dip angle would be less than 90˚. If you pushed the wall over completely so it was lying on the floor, it would no longer have a strike direction because you could draw a horizontal line in any of an infinite number of directions on the horizontal surface of the wall. Its dip would be 0˚.

When reporting the dip, include the direction. For example, if the strike runs north-south and the dip is 30˚, it would be necessary to specify “to the west” or “to the east.”  Similarly if the strike is northeast-southwest and the dip is 60˚, it would be necessary to say “to the northwest” or “to the southeast.” In the case of the vertical wall with a dip angle of 90˚, there is no dip direction. The dip points straight down, not toward any compass direction.

Measurement of geological features is done with a special compass that has a built-in clinometer, which is a device for measuring vertical angles. The strike is measured by aligning the compass along a horizontal line on the surface of the feature (Figure 13.36, left). The dip is measured by turning the compass on its side and aligning it along the dip direction (Figure 13.36, right).

Figure 13.36 Measurement of strike (left) and dip (right) using a geological compass with a clinometer.  Source: Steven Earle (2015) CC BY 4.0 view source left/ right

Strike and dip are used to describe any other planar features, including joints, faults, dykes, sills, and even the foliation planes in metamorphic rocks. Figure 13.37 shows an example of how we would depict the beds that make up an anticline on a map. The beds on the west (left) side of the map are dipping at various angles to the west. The beds on the east side are dipping to the east. The beds in the middle are horizontal; this is denoted by a cross within a circle on the map. The dyke is dipping at 80˚ to the west. The hinge line of the fold is denoted with a dashed line on the map, with two arrows pointing away from it, indicating the general dip directions of the limbs.  If it were a syncline, the arrows would point inward toward the line.

Figure 13.37 A depiction of an anticline and a dyke in cross-section (looking from the side) and in map view (or plan view) with the appropriate strike-dip and anticline symbols. Source: Steven Earle (2015) CC BY 4.0 view source

 

Exercise: Putting Strike and Dip on a Map

This cross-section shows seven tilted sedimentary layers (a to g), a fault, and a steeply dipping dyke.

Figure 13.38 Practice with strike and dip symbols. Source: Steven Earle (2015) CC BY 4.0 view source
  1. Place strike and dip symbols on the map to indicate the orientations of the beds shown, the fault, and the dyke.
  2. What type of fault is shown?
  3. What kind of stress created the fault?

 

 

94

Chapter 13 Summary

The topics covered in this chapter can be summarized as follows:

13.1 Stress and Strain

Stress within rocks, which includes compression, extension and shearing, originates from plate tectonic processes and the weight of overlying rocks. Rock that is stressed responds with either elastic or plastic strain, and may eventually break. The way a rock responds to stress depends on its composition and structure, the rate at which strain is applied, and also on the temperature, pressure, and the presence of fluid within the rock.

13.2 Folding

Folding is generally a ductile response to compression, although some brittle behaviour can happen during folding. A fold with a hinge that points upward is an anticline. A fold with a hinge that points downward is a syncline. The axial surface of a fold can be vertical, inclined, or even horizontal. The landforms produced by folds will depend on the resistance to weathering of rock layers within the fold.

13.3 Fractures, Joints, and Faults

Joints typically form during extension, but can also form during compression. Faulting, which involves the displacement of rock, can take place during compression or extension, as well as during shearing at transform boundaries.  Thrust faults are a type of reverse fault with a fault plane tilted at a low angle. Thrust faults are common in mountain belts formed by plate collisions.

13.4 Mountain Building

Mountain building in zones of plate collision is called orogeny. The mountains that form are orogens, and consist of crust thickened and deformed by folding and faulting, as well as the intrusion of igneous rocks. Orogens in ocean-continent collision zones have volcanoes. Mountains formed in rift zones are the result of tilting of normal-faulted blocks, or some normal-faulted blocks subsiding while others remain elevated.

13.5 Measuring Geological Structures

The strike and dip of planar surfaces, such as a bedding planes, fractures or faults are measured to help understand the geological history of a region.  Special symbols are used to show the orientation of structural features on geological maps.

Review Questions

1. What types of plate boundaries are most likely to contribute to (a) compression, (b) extension, and (c) shearing?

2. Explain the difference between elastic strain and plastic strain.

3. List some of the factors that influence whether a rock will undergo ductile deformation or break when placed under stress.

4. Draw in the axial traces of the folds in Figure 13.39, and label each with the appropriate type (e.g., overturned syncline).

Figure 13.39 A cross-section showing folds. Source: Steven Earle (2015) CC BY 4.0 view source

5. Explain why fractures are common in volcanic rocks.

6. What is the difference between a normal fault and a reverse fault, and under what circumstances would you expect these to form?

7. What type of fault would you expect to see at a transform plate boundary?

8. Figure 13.40 is a map of the geology of a region. The coloured areas represent sedimentary beds.

(i) Describe in words the general attitude (strike and dip) of these beds.

(ii) What is “a” and what is its attitude?

(iii) What is “b” and what is its attitude?

(iv) Is “b” left handed or right handed?

Figure 13.40 Geological map. Source: Steven Earle (2015) CC BY 4.0 view source

95

Answers to Chapter 13 Review Questions

1. Convergent plate boundaries are the most likely to contribute to compression, divergent boundaries to extension, and transform boundaries to shearing. However, all of these stress regimes can exist at any one of these boundaries.

2. When elastic strain takes place the rock can rebound to its original shape when stress is removed. Plastic strain is permanent.

3. Ductile deformation is more likely under higher temperatures and confining pressures. It is more likely when rocks are deformed slowly, and by compression. It is more likely for sedimentary rocks, and for rocks without fluids.

4. Axial traces are shown with dashed red lines.

Figure 13.41 Folds with labels. Source: Steven Earle (2015) CC BY 4.0 view source

5. Volcanic rocks cool quickly at surface and the resulting reduction in volume can easily lead to fracturing.

6. In a normal fault the rock above the fault (hanging wall or headwall) moves down with respect to the lower rock (footwall). This normally indicates extension. In a reverse fault the hanging wall is pushed up, which indicates compression.

7. Most faults near transform boundaries are strike-slip faults, meaning that there is horizontal motion along the fault.

8.(i) The beds are dipping at about 30˚ to the northwest. (ii) “a” is a dyke and it is dipping steeply to the northeast. (iii) “b” is a fault and it is dipping steeply to the southeast. (iv) The motion on fault “b” appears to be left handed.

Figure 13.42 Geological map with strike and dip symbols. Source: Steven Earle (2015) CC BY 4.0 view source

 

XIV

Chapter 14. Streams and Floods

Adapted by Joyce M. McBeth, University of Saskatchewan
from Physical Geology by Steven Earle

Learning Objectives

After reading this chapter, completing the exercises within it, and answering the questions at the end, you should be able to:

Figure 14.1 A small waterfall on Johnston Creek in Johnston Canyon, Banff National Park, AB Source: Steven Earle (2015) CC BY 4.0 view source

Why Study Streams?

Streams are the most important agents of erosion and transportation of sediments on Earth’s surface at this time in Earth’s history. They are responsible for generating much of the topography on that land surfaces that we see around us. They are places of beauty and tranquility, and they provide much of the water that is essential to our existence. But streams are not always peaceful and soothing. During large storms and rapid snowmelts, they can become raging torrents capable of moving cars and houses, and destroying roads and bridges. When they spill over their banks, they can flood huge areas, devastating populations and infrastructure. Over the past century, many of the most damaging natural disasters in Canada have been floods.

96

14.2 Drainage Basins

A stream is a body of flowing surface water of any size, ranging from a tiny trickle to a mighty river. The area from which the water flows to form a stream is known as its drainage basin or watershed. All of the precipitation (rain or snow) that falls within a drainage basin eventually flows into its stream, unless some of this water is able to cross into an adjacent drainage basin via groundwater flow. An example of a drainage basin is shown in Figure 14.5.

Figure 13.4 Cawston Creek near Keremeos, B.C. The blue line shows the extent of the drainage basin. The dashed red line is the drainage basin of one of its tributaries. [SE]
Figure 14.5 A schematic diagram of the drainage basin of Cawston Creek near Keremeos, BC. The blue line shows the extent of the drainage basin. The dashed red line is the drainage basin of one of its tributaries.  Source: Steven Earle (2015) CC BY 4.0 view source 

An important characteristic of streams is their gradient, the rate of change in elevation with distance along the stream. A steep gradient has a rapid change in elevation with horizontal distance, and a shallow gradient has a slow change in elevation with horizontal distance. Cawston Creek drainage basin in BC is approximately 25 km2 and is a typical small drainage basin within a very steep glaciated valley. As shown in Figure 14.6, the upper and middle parts of the creek have steep gradients averaging about 200 m/km but ranging from 100 to 350 m/km, and the lower part, within the valley of the Similkameen River, is relatively flat at <5 m/km.

The shape of the valley has changed through time to result in the shape we see now. First, there was tectonic uplift (related to tectonic plate convergence). Then the landscape changed due to stream erosion and mass wasting (landslides). This was followed by several episodes of glacial erosion. Finally, there was post-glacial stream erosion up to the present time. The lowest elevation of Cawston Creek (275 m, where the creek flows into the Similkameen River) is its base level. Cawston Creek cannot erode below this level unless the Similkameen River erodes deeper into its flood plain (the area that is inundated during a flood). Base level is the elevation where a stream will no longer erode deeper into the bedrock or sediments it flows through, because the elevation of the stream does not drop below this level, and further erosion can only occur if there is an elevation drop to propel the water deeper due to the force of gravity.

The ocean is the ultimate base level, but lakes and other rivers act as base levels  for many smaller streams.

Figure 13.5 Profile of the main stem of Cawston Creek near Keremeos, B.C. The maximum elevation of the drainage basin is about 1,840 m, near Mount Kobau. The base level is 275 m, at the Similkameen River. As shown, the gradient of the stream can be determined by dividing the change in elevation between any two points (rise) by the distance between those two points (run). [SE]
Figure 14.6 Profile of the main portion of Cawston Creek near Keremeos, BC. The maximum elevation of the drainage basin is about 1,840 m, near Mount Kobau. The base level is 275 m, at the Similkameen River. As shown, the gradient of the stream can be determined by dividing the change in elevation between any two points (rise) by the distance between those two points (run). Source: Steven Earle (2015) CC BY 4.0 view source

The water supply for the Vancouver, BC metropolitan area comes from three large drainage basins on the north shore of Burrard Inlet, as shown in Figure 14.7. This map illustrates the concept of a drainage basin divide. The boundary between two drainage basins is the ridge of land between them. A drop of rain falling on the boundary between the Capilano and Seymour drainage basins, for example, could flow into either basin. Rain falling on the Capilano basin side cannot flow into the Seymour drainage basin, because of the drainage basin divide.

Figure 14.7 The three drainage basins that supply water to the metropolitan Vancouver, BC area. Source: Wikimedia user “Alaidlaw” (2016) CC BY-SA 2.0. view source

The pattern of tributaries within a drainage basin depends largely upon the type of underlying rock, and on structures within that rock such as folds, fractures, and faults. Three types of drainage patterns are illustrated in Figure 14.8. Dendritic patterns, which are by far the most common, develop in areas where the rock (or unconsolidated material) beneath the stream does not have structures that control the stream flow patterns such as folds and joints; the materials can be eroded by the stream equally easily in all directions. Most areas of British Columbia have dendritic patterns, as do most areas of the prairies and the Canadian Shield.

Trellis drainage patterns typically develop where sedimentary rocks have been folded or tilted, and then eroded to varying degrees depending on their resistance to erosion. The Rocky Mountains of BC and AB have some fine examples of trellis drainage.

Rectangular patterns develop in areas that have very little topography and a system of bedding planes, fractures, or faults that form a rectangular network. Rectangular drainage patterns are rare in Canada

Figure 13.7 Typical dendritic, trellis, and rectangular stream drainage patterns. [SE]
Figure 14.8 Typical dendritic, trellis, and rectangular stream drainage patterns. Source: Steven Earle (2015) CC BY 4.0 view source

In many parts of Canada, especially relatively flat areas with thick glacial sediments, and throughout much of Canadian Shield in eastern and central Canada, drainage patterns are chaotic , also known as deranged (Figure 14.9, left). Lakes and wetlands are common in this type of environment.

Radial drainage (Figure 14.9, right) is a pattern that forms around isolated mountains (such as volcanoes) or hills. The individual streams that radiate out from the hill typically have dendritic drainage patterns.

Figure 13.8 Left: a typical deranged pattern; right: a typical radial drainage pattern developed around a mountain or hill. [SE]
Figure 14.9 Left: a typical deranged pattern; right: a typical radial drainage pattern developed around a mountain or hill.  Source: Steven Earle (2015) CC BY 4.0 view source

The process of a stream eroding downward into bedrock is called downcutting. Over geological time, and during tectonic quiescence, a stream will erode its drainage basin into a smooth profile similar to that shown in Figure 14.10. This is called a graded stream. Graded streams are steepest in their headwaters and their gradient gradually decreases toward their mouths. Ungraded streams are still in the process of rapidly eroding and downcutting their drainage basin, they have steep sections at various points, and typically have rapids and waterfalls at numerous locations along their lengths (e.g., Cawston Creek, Figure 14.6).

Figure 13.9 The topographic profile of a typical graded stream. [SE]
Figure 14.10 The topographic profile of a typical graded stream. Source: Steven Earle (2015) CC BY 4.0 view source

A graded stream can become ungraded if there is renewed tectonic uplift, or if there is a change in the base level. Base level changes can occur due to tectonic uplift or some other reason such as construction of a dam downstream. As stated earlier, the base level of Cawston Creek is defined by the level of the Similkameen River, but this can change, and has done so in the past. Figure 14.11 shows the valley of the Similkameen River in the Keremeos area. The river channel is just beyond the row of trees. The green field in the distance is underlain by material eroded from the hills behind and deposited by a small creek (not Cawston Creek) adjacent to the Similkameen River when its level was higher than it is now. Some time in the past several centuries, the Similkameen River eroded down through these deposits (forming the steep bank on the other side of the river), and the base level of the small creek was lowered by about 10 m. Over the next few centuries, this creek will erode through the sediments and eventually become graded again.

Figure 13.10 An example of a change in the base level of a small stream that flows into the Similkameen river near Keremeos. The previous base level was near the top of the sandy bank. The current base level is the river. [SE]
Figure 14.11 An example of a change in the base level of a small stream that flows into the Similkameen River near Keremeos. The previous base level was near the top of the sandy bank. The current base level is the Similkameen River, located on the far side of the line of trees. Source: Steven Earle (2015) CC BY 4.0 view source

Another example of a change in base level can be seen along the Juan de Fuca Trail on southwestern Vancouver Island. As shown in Figure 14.12, many of the small streams along this part of the coast flow into the ocean as waterfalls. It is evident that the land in this area has risen by about 5 m in the past few thousand years, probably in response to deglaciation. The streams that used to flow directly into the ocean now have a lot of downcutting to do before they will be a graded stream again.

Juan de Fuca Trail
Figure 14.12 Two streams with a lowered base level on the Juan de Fuca Trail, southwestern Vancouver Island. Source: Steven Earle (2015) CC BY 4.0 view source

Sediments accumulate within the channel and flood plain of a stream, and then, if the base level changes, or if there is less sediment supplied to the stream to deposit, the stream may cut down through these existing sediments to form terraces. A terrace on the Similkameen River is shown in Figure 14.11 and terraces on the Fraser River are shown in Figure 14 .13.

Figure 14.13 Terraces on the Fraser River north of Lillooet, BC (above the river on the left-hand side of the image). Source: Wikimedia user “China Crisis” (2007) CC BY-SA 3.0. view source

In the late 19th century, American geologist William Davis proposed that streams and the surrounding terrain develop in a cycle of erosion (Figure 14.14). Following tectonic uplift , the stream patterns are immature. Streams erode quickly, developing deep V-shaped valleys that tend to follow relatively straight paths. Gradients are high, and profiles are ungraded. Rapids and waterfalls are common. As the landscape matures, the streams erode wider valleys and deposited thick sediment layers. Even after maturity, gradients are slowly reduced and grading increases. In old age, streams are surrounded by rolling hills, and they occupy wide sediment-filled valleys. Meandering patterns are common, and erosion now is focussed towards the channel walls, with little downcutting.

Figure 13.13 A depiction of the Davis cycle of erosion: a: initial stage, b: youthful stage, c: mature stage, and d: old age. [SE]
Figure 14.14 A depiction of the Davis cycle of erosion: a: initial stage, b: youthful stage, c: mature stage, and d: old age . Source: Steven Earle (2015) CC BY 4.0 view source

Davis’s work was done long before the idea of plate tectonics, and he was also not familiar with the impacts of glacial erosion on streams and their environments. While some parts of his idea are out of date, it is still a useful way to understand streams and their evolution.  Plate tectonic activity and other processes such as isostatic rebound after glaciation results in uplift that alters stream gradients, so streams are constantly adjusting due to these changing conditions. It would be relatively rare to find a stream that is able to mature through all of these stages without interruption.

Exercise 14.2 The Effect of a Dam on Base Level

Revelstoke Dam and Revelstoke Lake on the Columbia River at Revelstoke, BC [SE]
Revelstoke Dam and Revelstoke Lake on the Columbia River at Revelstoke, BC. Source: Steven Earle (2015) CC BY 4.0 view source

When a dam is built on a stream, a reservoir (artificial lake) forms behind the dam. The dam reservoir temporarily (for many decades at least) creates a new base level for the part of the stream above the reservoir. How does the formation of a dam reservoir affect the stream where it enters the reservoir, and what happens to the sediment it was carrying? The water leaving the dam has no sediment in it. Why? How does this affect the stream below the dam?

97

14.3 Stream Erosion and Deposition

Stream Velocity Depends on the Shape and Size of the Channel

Flowing water is a very important mechanism for both erosion and deposition. Water flow in a stream is primarily related to the stream’s gradient, but it is also controlled by the geometry of the stream channel. As shown in Figure 14.15, water flow velocity decreases due to friction along the stream bed. The stream is thus slowest at the bottom and edges and fastest near the surface and in the middle of the stream (where there is the least amount of friction). The velocity just below the surface of the water is typically a little higher than right at the surface because of friction between the water and the air. On a curved section of a stream, flow is fastest on the outside of the curve and slowest on the inside of the curve.

Figure 13.14 The relative velocity of stream flow depending on whether the stream channel is straight or curved (left), and with respect to the water depth (right). [SE]
Figure 14.15 The relative velocity of stream flow depending on whether the stream channel is straight or curved (left). (Right) it is also dependent on the water depth. The length of each of the arrows indicates the relative velocity of the stream at that position in the channel. Shorter arrows mean slower flow . Source: Steven Earle (2015) CC BY 4.0 view source

Another important factor influencing stream-water velocity is the discharge, or volume of water passing a point in a unit of time (e.g., m3/second). Water levels rise during a flood and due to the higher discharge of water the stream flow velocity increases. The higher discharge also increases the cross-sectional area of the stream, so it fills up the channel. In a flood it may overflow the banks. Another factor that affects stream-water velocity is the size of sediments on the stream bed. Large particles tend to slow the flow more than small ones.

Sediment Transport Depends on Stream Velocity and Turbulence

If you drop a piece of gravel into a glass of water, it will sink quickly to the bottom. If you drop a grain of sand into the same glass, it will sink more slowly. A grain of silt will take longer yet to get to the bottom, and a particle of fine clay will take a long time settle out. The rate of settling is determined by the balance between gravity and friction, as shown in Figure 14.16.

How quickly a grain settles to the bottom of a stream depends on its mass (affecting the force of gravity acting on it), and the friction between the grain and the water which slows the fall of the grain. [SE]
Figure 14.16 How quickly a grain settles to the bottom of a stream depends on its mass (affecting the force of gravity acting on it), and the friction between the grain and the water, which slows the fall of the grain.  Source: Steven Earle (2015) CC BY 4.0 view source

One of the key principles of sedimentary geology is that the ability of a moving medium (air or water) to move sedimentary particles and keep them moving is dependent on the velocity of flow. The faster the medium flows, the larger the particles it can move. As you probably know from intuition and from experience, streams that flow rapidly tend to be turbulent (flow paths are chaotic and the water surface appears rough) and the water may be muddy. In contrast, streams that flow more slowly tend to have laminar flow (straight-line flow and a smooth water surface) and clearer water. Turbulent flow is more effective than laminar flow at keeping sediments suspended within the water.

Particles within a stream are transported in different ways depending on their size (Figure 14.17). Large particles which rest on the stream bed are known as the bedload. The bedload may only be transported when the flow rate is rapid and under flood conditions. They are transported by saltation (bouncing along, and colliding with other particles) and by traction (being pushed along by the force of the flow).

Smaller particles may rest on the bottom occasionally, where they can be transported by saltation and traction, but they can also be held in suspension in the flowing water (the suspended load), especially at higher flow velocities.

Stream water also has a dissolved load, which represents (on average) about 15% of the mass of material transported, and includes ions such as calcium (Ca+2) and chloride (Cl) in solution. The solubility of these ions is not affected by flow velocity.

Figure 13.15 Modes of transportation of sediments and dissolved ions (represented by red dots with + and – signs) in a stream. [SE]
Figure 14.17 Modes of transportation of sediments and dissolved ions (represented by red dots with + and – signs) in a stream . Source: Steven Earle (2015) CC BY 4.0 view source

If you look at a typical stream, there are always some sediments being deposited, some staying where they are, and some being eroded and transported. This changes over time as the discharge of the river changes in response to changing weather conditions.

The Hjulström-Sundborg Diagram Summarizes What Happens to Grains of Different Sizes at Different Stream Velocities

The faster water is flowing, the larger the particles that can be kept in suspension and transported within the flowing water. However, as Swedish geographer Filip Hjulström discovered in the 1940s, the relationship between grain size and the likelihood of a grain being eroded, transported, or deposited is not as simple as one might imagine (Figure 14.18). Consider, for example, a 1 mm grain of sand. If it is resting on the bottom of the stream, it will remain there until the flow velocity is high enough to erode it (ca 20 cm/s). But once it is in suspension, that same 1 mm particle will remain in suspension as long as the velocity doesn’t drop below 10 cm/s. For a 10 mm gravel grain, the velocity is 105 cm/s to be eroded from the bed but only 80 cm/s to remain in suspension.

Figure 13.16 The Hjulström-Sundborg diagram showing the relationships between particle size and the tendency to be eroded, transported, or deposited at different current velocities
Figure 14.18 The Hjulström-Sundborg diagram showing the relationships between particle size and the tendency to be eroded, transported, or deposited, at different current velocities Source: Steven Earle (2015) CC BY 4.0 view source

On the other hand, a 0.01 mm silt particle only needs a velocity of 0.1 cm/s to remain in suspension, but requires 60 cm/s to be eroded. In other words, a tiny silt grain requires a greater velocity to be eroded than a grain of sand that is 100 times larger! For clay-sized particles, the discrepancy is even greater. In a stream, the most easily eroded particles are small sand grains between 0.2 mm and 0.5 mm. Anything smaller or larger requires a higher water velocity to be eroded and entrained in the flow. The reason for this is that small particles, especially tiny grains of clay, possess a net surface charge, hence experience a strong tendency to stick together, and so are difficult to erode from the stream bed.

It is important to be aware that a stream can both erode and deposit sediments at the same time. At 100 cm/s, for example, silt, sand, and medium gravel will be eroded from the stream bed and transported in suspension, coarse gravel will be transported by saltation and traction, pebbles will be both transported by saltation and traction, and will also be deposited, and cobbles and boulders will remain stationary on the stream bed.

Exercise 14.3 Understanding the Hjulström-Sundborg Diagram

Refer to the Hjulström-Sundborg diagram in Figure 14.18 to answer these questions.

  1. A fine sand grain (0.1 mm) is resting on the bottom of a stream bed.
  2. a) What stream velocity will it take to get this sand grain into suspension?
  3. b) Once the particle is in suspension, the velocity starts to drop. At what velocity will it finally come back to rest on the stream bed?
  4. A stream is flowing at 10 cm/s (which means it takes 10 s to travel 1 m).
    1. What size of particles can be eroded at 10 cm/s?
    2. What is the largest particle that, once in suspension, will remain in suspension at 10 cm/s?

Natural Levees Form Because of Changes in Stream Velocity

A stream typically reaches its greatest velocity when it is close to flooding over its banks. This is known as the bank-full stage, as shown in Figure 14.19. When the flooding stream overtops its banks and occupies the wide area of its flood plain, the water has a much larger area to flow through and the velocity drops dramatically. As water flows from the channel out across the flood plain, it slows down and starts to deposit its sediment load. This forms an elevated bank known as a levee along the edges of the channel. The coarsest and thickest sediments are deposited near the channel banks, with particle size and thickness decreasing as you move further into the flood plain. People also build levees as flood control measures; the idea for this engineered solution to floods came from the naturally-build levees that form during floods.

Figure 13.17 The development of natural levées during flooding of a stream. The sediments of the levée become increasingly fine away from the stream channel, and even finer sediments — clay, silt, and fine sand — are deposited across most of the flood plain. [SE]
Figure 14.19 The development of natural levees during flooding of a stream. The sediments of the levee become increasingly fine away from the stream channel, and even finer sediments — clay, silt, and very fine sand — are deposited across most of the flood plain .  Source: Steven Earle (2015) CC BY 4.0 view source

98

14.4 Stream Types

Stream channels can be straight or curved, deep or shallow, cleared or filled with coarse sediments. The cycle of erosion has some influence on the nature of a stream, but there are several other factors that are important.

Youthful streams that are actively downcutting their channels tend to be relatively straight and are typically ungraded (meaning that rapids and waterfalls are common). As shown in Figures 14.1 and 14.20, youthful streams commonly have a step-pool morphology, meaning that the stream consists of a series of pools connected by rapids and waterfalls. They also have steep gradients, and steep and narrow V-shaped valleys. In some cases these valley walls are steep enough to be called canyons.

Figure 13.18 The Cascade Falls area of the Kettle River, near Christina Lake, B.C. This stream has a step-pool morphology and a deep bedrock channel. [SE]
Figure 14.20 The Cascade Falls area of the Kettle River, near Christina Lake, BC. This stream has a step-pool morphology and a deep bedrock channel.  Source: Steven Earle (2015) CC BY 4.0 view source

In mountainous terrain, such as that in western AB and BC, steep youthful streams typically flow into wide and relatively low-gradient U-shaped glaciated valleys. The youthful streams have high sediment loads, and when they flow into the lower-gradient glacial valleys where the velocity is no longer high enough to carry all of the sediment, braided stream patterns develop, characterized by a series of narrow channels separated by gravel bars (Figure 14.21).

Figure 13.19 The braided channel of the Kicking Horse River at Field, B.C. [SE]
Figure 14.21 The braided channel of the Kicking Horse River at Field, BC. Source: Steven Earle (2015) CC BY 4.0 view source

Braided streams can develop anywhere where there is more sediment than a stream is able to transport. One such environment is in volcanic regions, where explosive eruptions produce large amounts of unconsolidated material that gets washed into streams. The Coldwater River next to Mt. St. Helens in Washington State is a good example of such a braided stream (Figure 14.22).

Figure 13.20 The braided Coldwater River, Mount St. Helens, Washington. [SE]
Figure 14.22 The braided Coldwater River, Mt. St. Helens, Washington. Source: Steven Earle (2015) CC BY 4.0 view source

A stream that occupies a wide, flat flood plain with a low gradient typically carries only sand-sized and finer sediments and develops a sinuous flow pattern. As you saw in Figure 14.15, when a stream flows around a bend, the main current of the stream flows near the outside portion of the bend. This leads to erosion of the banks on the outside of the bend, and deposition of a point bar on the inside of the bend (Figure 14.23). Over time, the sinuosity of the stream becomes increasingly exaggerated, and the channel migrates throughout its flood plain, forming a meandering pattern.

Figure 13.21 The meandering channel of the Bonnell Creek, Nanoose, B.C. The stream is flowing toward the viewer. The sand and gravel point bar must have formed when the creek was higher and the flow faster than it was when the photo was taken. [SE]
Figure 14.23 The meandering channel of the Bonnell Creek, Nanoose, BC. The stream is flowing toward the viewer. The sand and gravel point bar must have formed when the creek was higher and the flow faster than it was when the photo was taken, as the current stream velocity is too low to carry such coarse sediments. Most erosion and deposition take place during flooding events. Source: Steven Earle (2015) CC BY 4.0 view source

A well-developed meandering river is shown in Figure 14.24. The meander in the middle of the photo has reached the point where the thin neck of land between two parts of the channel is about to be eroded through. When this happens, an oxbow lake will form. These are small cut off bends from earlier curves in the river; several are visible outside the path of the main stream in Figure 14.24.

The meandering channel of the Nowitna River, Alaska. Numerous oxbow lakes are present and another meander cutoff will soon take place. [Oliver Kumis CC-BY-SA http://bit.ly/1SmQL7B]
Figure 14.24 The meandering channel of the Nowitna River, Alaska. Numerous oxbow lakes are present, and another meander cutoff will soon take place. Source: Oliver Kumis (2002) CC-BY-SA 2.0 view source

At the point where a stream enters a body of water such as a lake or the ocean, the flow rates drops dramatically, and sediment is deposited. Over time, as more and more sediments are deposited, the sediments form a distinctive triangular shape (with the bottom broad part of the triangle facing the ocean or lake and the point of the triangle facing upstream). This is called a delta; these are named after the Greek letter delta which is in the shape of a triangle. The Fraser River has created a large delta in BC where the river meets the Strait of Georgia (Figure 14.25). Much of the Fraser delta is very young in geological terms. Shortly after the end of the last glaciation (10,000 years ago), the delta did not extend past New Westminster. Since that time, all of the land that makes up Richmond, Delta, and parts of New Westminster and south Surrey has formed from sediment depositing from the Fraser River. You can see a more detailed description of the Fraser delta on the Geoscape Vancouver website: http://www.cgenarchive.org/vancouver-fraserdelta.html

The delta of the Fraser River and the plume of sediment that extends across the Strait of Georgia. The land outlined in red has formed over the past 10,000 years. [September 2011, SE after NASA http://bit.ly/FrasR]
Figure 14.25 The delta of the Fraser River and the plume of sediment that extends across the Strait of Georgia. The land outlined in red has formed over the past 10,000 years. Source: Steven Earle (2015) CC BY 4.0 view source after NASA, September 2011 view source

Exercise 14.4 Calculating Stream Gradients

Stream Gradients
Source: Steven Earle (2015) CC BY 4.0 view source

 

The gradient is the key factor controlling stream velocity, and stream velocity controls sediment erosion and deposition. This map shows the elevations of Priest Creek in the Kelowna area. The length of the creek between 1,600 m and 1,300 m elevation is 2.4 km, so the gradient is (1,600 m – 1,300 m)/2.4  km = 125 m/km.

  1. Use the scale bar to estimate the distance between 1,300 m and 600 m and calculate the gradient between these two elevations.
  2. Estimate the gradient between 600 and 400 m.
  3. Estimate the gradient between 400 m on Priest Creek and the point where Mission Creek enters Okanagan Lake.

99

14.5 Flooding

The discharge levels of streams are highly variable depending on the time of year and on variations in the weather from one year to the next. In Canada, most streams show discharge variability similar to that of the Stikine River in northwestern BC, illustrated in Figure 14.26. The Stikine River has its lowest discharge levels in the depths of winter when freezing conditions persist throughout most of its drainage basin. Discharge starts to rise slowly in May, and then rises dramatically through the late spring and early summer as the winter snow melts. For the year shown, the minimum discharge of the Stikine River was 56 m3/s in March, and the maximum was 37 times higher at 2,470 m3/s in May.

Figure 13.24 Variations in discharge of the Stikine River during 2013. [SE from data at Water Survey of Canada, Environment Canada, http://www.ec.gc.ca/rhc-wsc/]
Figure 14.26 Variations in discharge of the Stikine River during 2013. Source: Steven Earle (2015) CC BY 4.0 view source from data at Water Survey of Canada, Environment Canada view source

Streams in coastal areas of southern British Columbia show a very different pattern from those in most of the rest of the country. In this region, the drainage basins receive a lot of rain (rather than snow) during the winter and also do not remain entirely frozen throughout the winter. The Qualicum River on Vancouver Island typically has its highest discharge levels in January or February and its lowest levels in late summer (Figure 14.27). In 2013, the minimum discharge of the Qualicum River was 1.6 m3/s in August, and the maximum was 34 times higher at 53 m3/s in March.

Figure 13.25 Variations in discharge of the Qualicum River during 2013. [SE from data at Water Survey of Canada, Environment Canada, http://www.ec.gc.ca/rhc-wsc/]
Figure 14.27 Variations in discharge of the Qualicum River during 2013. Source: Steven Earle (2015) CC BY 4.0 view source, from data at Water Survey of Canada, Environment Canada view source

When a stream’s discharge increases, both the water level (stage) and the velocity increase as well. Rapidly flowing streams become muddy, and large volumes of sediment are transported both in suspension and along the stream bed. In extreme situations, the water level reaches the top of the stream’s banks (the bank-full stage, see Figure 14.19), and if it rises further, it will overflow the banks and floods the surrounding terrain. In the case of mature or old-age streams, this could include a vast area of relatively flat ground known as a flood plain, which is the area that is typically covered with water during a major flood. Since fine river sediments are deposited on flood plains, they are ideally suited for agriculture, and thus are typically occupied by farms and residences, and in many cases, by towns or cities. Such infrastructure is highly vulnerable to damage from flooding, and the people that live and work there are at risk.

Most streams in Canada have the greatest risk of flooding in the late spring and early summer when stream discharges rise in response to melting snow. In some cases, this is exacerbated by spring storms. In years when melting is especially fast and/or spring storms are particularly intense, flooding can be very severe.

One of the worst floods in Canadian history took place in the Fraser Valley of BC in late May and early June of 1948. The early spring of that year had been cold, and a large snow pack in the interior was slow to melt. In mid-May, temperatures rose quickly and melting was accelerated by rainfall. Fraser River discharge levels rose rapidly over several days during late May, and the dykes built to protect the valley were breached in a dozen places. Approximately one-third of the flood plain was inundated, and many homes and other buildings were destroyed, but there were no deaths.

The Fraser River flood of 1948, which was the worst flood in the Fraser Valley in the past century, was followed by very high river levels in 1950 and 1972, and by relatively high levels several times since then, the most recent being 2007 (Table 13.1). In the years following 1948, millions of dollars were spent repairing and raising the existing dykes and building new ones. Since then damage from flooding in the Fraser Valley has been relatively limited.

Rank Year Month Date Stage (m) Discharge (m3/s)
1 1948 May 31 11.0 15,200
2 1972 Jun 16 10.1 12,900
3 1950 Jun 20 9.9 12,500
4 1964 Jun 21 9.6 11,600
5 1997 Jun 5 9.5 11,300
6 1955 Jun 29 9.4 11,300
7 1999 Jun 22 9.4 11,000
8 2007 Jun 10 9.3 10,850
9 1974 Jun 22 9.3 10,800
10 2002 Jun 21 9.2 10,600

Table 14.1 Ranking of the maximum stage and discharge values for the Fraser River at Hope between 1948 and 2008. Typical discharge levels are ~1,000 m3/s. Source: Data from Mannerstrom (2008) Comprehensive Review of Fraser River at Hope Flood Hydrology and Flows Scoping Study, Report prepared for the B.C. Ministry of the Environment. view source

Serious flooding occurred in July 1996 in the Saguenay-Lac St. Jean region of Quebec. In this case, the floods were caused by two weeks of heavy rainfall followed by one day of exceptional rainfall. On July 19, 1996 there was 270 mm of rain, equivalent to the region’s normal rainfall for the entire month of July. Ten deaths were attributed to the Saguenay floods, and the economic toll was estimated at $1.5 billion.

Just a year after the Saguenay floods, the Red River in Minnesota, North Dakota, and Manitoba reached its highest level since 1826. As is typical for the Red River, the 1997 flooding was due to rapid snowmelt. Due to the south to north flow of the river, the flooding starts in Minnesota and North Dakota, where melting begins earlier, then extends northwards. The residents of Manitoba had plenty of warning that the 1997 flood was coming because there was severe flooding at several locations on the U.S. side of the border.

After the 1950 Red River flood, the Manitoba government built a channel around the city of Winnipeg to reduce the potential of flooding in the city (Figure 14.28). Known as the Red River Floodway, the channel was completed in 1964 at a cost of $63 million. Since then it has been used many times to alleviate flooding in Winnipeg, and is estimated to have saved many billions of dollars in flood damage. The massive 1997 flood (Figure 14.28, right side) was almost too much for the floodway; in fact the amount of water diverted was greater than the designed capacity. The floodway has recently been expanded so that it can be used to divert even more of the Red River’s flow away from Winnipeg.

Figure 13.26 Map of the Red river Floodway around Winnipeg, Manitoba (left), and aerial view of the southern (inlet) end of the floodway (right). [Map from http://en.wikipedia.org/wiki/1997_Red_River_Flood#/media/File:Rednorthfloodwaymap.png and photo from Natural Resources Canada 2012, courtesy of the Geological Survey of Canada (Photo 2000-118 by G.R. Brooks).]
Figure 14.28 (left) map of the Red River Floodway around Winnipeg, Manitoba; (right) aerial view of the southern (inlet) end of the floodway during the 1997 Red River flood. Sources: (left) Wikimedia user “Kmusser” (2007) CC BY 2.5 view source, (right) Natural Resources Canada 2012, courtesy of the Geological Survey of Canada (Photo 2000-118 by G.R. Brooks ).

Canada’s most costly flood ever was the June 2013 flood in southern Alberta. The flooding was initiated by snowmelt and worsened by heavy rains in the Rockies due to an anomalous flow of moist air from the Pacific and the Caribbean. At Canmore, AB rainfall amounts exceeded 200 mm in 36 hours, and at High River, AB 325 mm of rain fell in 48 hours.

Figure 13.27 Map of the communities most affected by the 2013 Alberta floods (in orange) [SE]
Figure 14.29 Map of the communities most affected by the 2013 Alberta floods  (in orange) Source: Steven Earle (2015) CC BY 4.0 view source

In late June and early July, the discharges of several rivers in the area, including the Bow River in Banff, Canmore, and Exshaw, the Bow and Elbow Rivers in Calgary, the Sheep River in Okotoks, and the Highwood River in High River, reached levels that were 5 to 10 times higher than normal for that time of year (see Exercise 14.5). Large areas of Calgary, Okotoks, and High River were flooded, and five people died (see Figures 14.29 and 14.30). The cost of the 2013 flood is estimated to be approximately $5 billion.

Figure 13.28 Flooding in Calgary (June 21, left) and Okotoks (June 20, right) during the 2013 southern Alberta flood [http://upload.wikimedia.org/wikipedia/commons/6/6a/Riverfront_Ave_Calgary_Flood_2013.jpg http://upload.wikimedia.org/wikipedia/en/9/9b/Okotoks_-_June_20%2C_2013_-_Flood_waters_in_local_campground_playground-03.JPG]
Figure 14.30 Flooding in Calgary (June 21, left) and Okotoks (June 20, right) during the 2013 southern Alberta flood. Sources: (left) Wikimedia user Ryan L.C. Quan (2013) CC BY-SA 3.0 view source (right) Wikimedia user “Stephanie N. Jones” (2013) CC BY-SA 3.0 view source 

One of the things that the 2013 flood of the Bow River teaches us is that we cannot predict when a flood will occur nor how big it will be, so in order to minimize damage and casualties we need to be prepared. Some ways of preparing include:

  • Mapping flood plains and not building within them
  • Building dykes or dams where necessary
  • Monitoring the winter snowpack, the weather, and stream discharges
  • Creating emergency plans
  • Educating the public on how to prepare for and respond to the threat of flooding

Exercise 14.5 Flood Probability on the Bow River

The graph below shows the highest discharge per year between 1915 and 2014 of the Bow River in Calgary. Using this data set, we can calculate the recurrence interval (Ri) for any particular flood magnitude using the equation: Ri = (n+1)/r (where n is the number of floods in the record being considered, and r is the rank of the particular flood). There are a few years missing in this record, so the actual number of floods is 95.

The largest flood recorded along the Bow River over that period of time was the one in 2013, reaching a peak flow rate of 1,840 m3/s on June 21. Ri for this flood is (95+1)/1 = 96 years. The probability of such a flood in any future year is 1/Ri x 100%, which is 1%. The fifth largest flood was just a few years earlier in 2005, at 791 m3/s. Ri for this flood is (95+1)/5 = 19.2 years. The recurrence probability of a flood of this magnitude is thus 5%.

  1. Calculate the recurrence interval for the second largest flood (1932, 1,520 m3/s).
  2. What is the probability that a flood of 1,520 m3/s will happen next year?
  3. Examine the 100-year trend for floods along the Bow River. If you ignore the major floods  (the labelled ones), what is the general trend of peak discharges over this time?
Water Surveys of Canada
Source: Steven Earle (2015) CC BY 4.0 view source, from data at Water Surveys of Canada, Environment Canada view source

100

Chapter 14 Summary

The topics covered in this chapter can be summarized as follows:

14.1 The Hydrological Cycle

Water is stored in the oceans, glacial ice, the ground, lakes, rivers, and the atmosphere. Its movement is powered by solar energy and gravity.

14.2 Drainage Basins

All of the precipitation that falls within a drainage basin flows into the stream that drains that area. Stream drainage patterns are determined by the type of rock within the basin. Over geological time, streams change the landscape that they flow within, and eventually they become graded, meaning their profile becomes a smooth curve. A stream can lose that gradation if there is renewed uplift or if their base level changes for some other reason such as construction of a dam downstream.

14.3 Stream Erosion and Deposition

The processes of erosion and deposition of particles within streams are primarily driven by the velocity of the stream water. Erosion and deposition of different-sized particles can happen simultaneously in a stream. Some particles are moved along the bottom of a river while others are carried in suspension. It takes a greater velocity of water to erode a particle from a stream bed than it does to keep it in suspension. Ions are also transported in solution. When a stream rises and then occupies its flood plain, the velocity of water over the flood plains slows and natural levees form along the edges of the stream channel.

14.4 Stream Types

Youthful streams in steep areas erode most rapidly downwards, and they tend to have steep, rocky, and relatively straight channels. Where sediment-rich streams empty into areas with lower gradients, braided streams can form. Meandering streams are common in areas with even lower gradients where silt and sand are the dominant sediments. Meandering streams erode the walls of their channels more rapidly than the channel base. Deltas form where streams flow into standing water.

14.5 Flooding

Most streams in Canada have their highest discharge rates in spring and early summer, although the highest discharge in many of BC’s coastal streams is in the winter. Floods happen when a stream rises high enough to spill over its banks and spread across its flood plain. Some of the more significant floods in Canada include the Fraser River flood of 1948, the Saguenay River flood of 1996, the Red River flood of 1997, and the Alberta floods of 2013. We can estimate the probability of a specific flood level based on the record of past floods, and we can take steps to minimize the impacts of flooding such as building floodways to divert excess water and not building in flood-prone areas.

Questions for Review

  1. What is the proportion of liquid fresh water on Earth expressed as a percentage of all water on Earth?
  2. What percentage of this fresh water is groundwater?
  3. What type of rock, and what processes, can lead to the formation of a trellis drainage pattern?
  4. Why do many of the streams in the southwestern part of Vancouver Island empty into the ocean over waterfalls?
  5. Where would you expect to find the fastest water flow along a straight stretch of a stream?
  6. Sand grains can be moved by traction and saltation. What minimum stream velocity is required to move 1 mm sand grains?
  7. If the flow velocity of a stream is 1 cm/s, what sizes of particles can be eroded, what sizes can be transported if they are already in suspension, and what sizes of particles cannot be moved at all?
  8. Under what circumstances might a braided stream develop?
  9. How would the gradient of a stream be affected if a meander is bypassed?
  10. The elevation of the Fraser River at Hope is 41 m. From there it flows approximately 147 km to the sea. What is the average gradient of the river (m/km) over this distance?
  11. How do BC’s coastal streams differ from most of the rest of the streams in Canada in terms of their annual flow patterns? Why?
  12. Why do most serious floods in Canada happen in late May, June, or early July?
  13. There is a 65-year record of peak annual discharges along the Ashnola River near Princeton, BC. During this time, the second highest discharge recorded was 175 m3/s. Based on this information, what is the recurrence interval (Ri) for this discharge level, and what is the probability that there will be a similar peak discharge next year?

101

Answers to Chapter 14 Review Questions

1. Approximately 1% of the Earth’s water is liquid fresh water.

2. Approxmately 30% of the Earth’s fresh water is groundwater.

3. A trellis drainage pattern typically forms on sedimentary rock that has been tilted and eroded.

4. Many of the streams in the southwestern part of Vancouver Island flow to the ocean as waterfalls because the land has been uplifted relative to sea level over the past several thousand years.

5. The fastest water flow on a straight stretch of a stream will be in the middle of the stream near the surface.

6. 1 mm sand grains will be eroded if the velocity if over 20 cm/s and will be kept in suspension as long as the velocity is over 10 cm/s.

7. If the flow velocity is 1 cm/s particles less than 0.1 mm (fine sand or finer) can be transported, while those larger than 0.1 mm cannot. At this velocity no particles can be eroded.

8. A braided stream can develop where there is more sediment available than can be carried in the amount of water present at the rate at which that water is flowing. This may happen where the gradient drops suddenly, or where there is a dramatic increase in the amount of sediment available (e.g., following an explosive volcanic eruption).

9. If a meander is cut off it reduces the length of a stream so it increases the gradient.

10. The average gradient of the Fraser River between Hope and the Pacific Ocean is 0.28 m/km (or 28 cm/km).

11. In coastal regions of B.C. the highest levels of precipitation are in the winter, and large parts of most drainage basins are not frozen solid. As a result stream discharges tend to be greatest in the winter.

12. In most parts of Canada winter precipitation is locked up in snow until the melt season begins, and depending on the year and the location that happens in late spring or early summer. If the thaw is delayed because of a cold spring, and then happens very quickly, flooding is likely. Some regions also receive heavy rainfall during this period of the year.

13. Ri = (n+1)/r (where n is the length of the record) and r is the rank of the flood in question. In the Ashnola River case Ri = (65+1)/2 = 33. The probability of such a flood next year is 1/Ri, or 1/33 which is 0.03 or 3%.

XV

Chapter 15. Mass Wasting

Adapted by Joyce M. McBeth, University of Saskatchewan
from Physical Geology by Steven Earle

Learning Objectives

After reading this chapter, completing the exercises within it, and answering the questions at the end, you should be able to:

Mass wasting is the failure and downslope movement of rock or unconsolidated materials in response to gravity. Mass wasting is an agent of erosion. The term “landslide” is used synonymously with the term mass wasting, but they are not the same thing. Some people reserve the term “landslide” for relatively rapid slope failures, while others do not. Due to this ambiguity, we avoid the use of the term “landslide” in this textbook.

The Hope Slide: a Historic Canadian Example of a Mass Wasting Event

Early in the morning of January 9, 1965, 47 million cubic metres of rock broke away from the steep upper slopes of Johnson Peak (16 km southeast of Hope, B.C.) and tumbled 2,000 m down the mountain, gouging out the contents of a small lake at the bottom, and continuing a few hundred metres up the other side of the valley (Figure 15.1). Four people were killed who had been stopped on the highway by a snow avalanche. Many more people might have become victims, except that a Greyhound bus driver, en route to Vancouver, turned his bus around upon seeing the avalanche. The rock failed along foliation planes of the metamorphic rock on Johnson Peak, in an area that had been eroded into a steep slope by glacial ice. There is no evidence that it was triggered by any specific event, and there was no warning that it was about to happen. Even if there had been warning, nothing could have been done to prevent it. There are hundreds of similar sites throughout mountainous regions of British Columbia and elsewhere in Canada where large mass wasting events could occur.

Photograph of the site of the 1965 Hope Slide as seen in 2014. The initial failure is thought to have taken place along the foliation planes and sill within the area shown in the inset. [SE]
Figure 15.1 | The site of the 1965 Hope Slide, photographed in 2014. The initial failure is thought to have taken place along foliation planes in the rock and a sill.  Source: Steven Earle (2015) CC BY 4.0.View Source

What can we learn from the Hope Slide? In general, we cannot prevent most mass wasting events, and significant effort is required if an event is to be predicted with any level of certainty. Understanding the geology is critical to understanding mass wasting. Although slope failures are inevitable in a region with steep slopes, larger slope failures happen less frequently than smaller ones. The consequences of a large mass wasting event also vary depending on the downslope conditions, such as the presence of people, buildings, roads, or fish-bearing streams.

An important reason for learning about mass wasting is to understand the nature of how and why materials fail in mass wasting events. If we understand this better, we can use this knowledge to help minimize the risks from similar events in the future.

102

15.1 Factors That Control Slope Stability

Slope Angle and the Forces Acting On A Slope

A block of rock situated on a rock slope is pulled by gravity toward Earth’s centre (vertically down, Figure 15.2). We can split the vertical gravitational force into two components (vectors) relative to the slope: one pulling the block down parallel to the slope (the shear force), and the other pulling the block directly into (i.e., perpendicular) to the slope (the normal force).

The shear force pulls the block down the slope, but the block does not move unless the shear force overcomes (is greater than) the friction between the block and the slope. This friction holding the block on the slope may be quite weak if the block has split away from the main body of rock, or may be very strong if the block is still connected to the other rock on the slope. The strength of the relationship between the block and the slope is called the shear strength. For example, in Figure 15.2a, the shear strength is greater than the shear force, so the block will not move. In Figure 15.2b the slope is steeper, and the shear force is approximately equal to the shear strength. The block may or may not move under these circumstances. In Figure 15.2c, the slope is steeper still, so the shear force is considerably greater than the shear strength, and the block will move.

Figure 15.2 Differences in the shear and normal components of the gravitational force on slopes with differing steepness. The gravitational force is the same in all three cases. In (a) the shear force is substantially less than the shear strength, so the block should be stable. In (b) the shear force and shear strength are about equal, so the block may or may not move. In (c) the shear force is substantially greater than the shear strength, so the block is very likely to move. [SE]
Figure 15.2 | Differences in the shear and normal components of the gravitational force on slopes with differing steepness. The total gravitational force is the same in all three cases. In (a) the shear force (red line aligned with slope) is substantially less than the shear strength (green arrow), so the block is stable. In (b) the shear force and shear strength are nearly equal, so the block may or may not move. In (c) the shear force is greater than the shear strength, so the block will move. Source: Steven Earle (2015) CC BY 4.0. view source

Slopes are created by uplift in the Earth’s crust and modified by erosion. In areas with relatively recent uplift (such as most of British Columbia and the western part of Alberta in Canada), slopes tend to be quite steep. This is especially true where glaciation has taken place because glaciers in mountainous terrain create steep-sided U-shaped valleys. In areas without recent uplift (such as central Canada), slopes are less steep because they have been subjected to erosion, including mass wasting, for long periods of time. Note that mass wasting can happen even on relatively gentle slopes if the shear stress acting on the materials is greater than the materials’ shear strength.

Slope Strength

The strength of the materials on slopes can vary widely. Solid rocks tend to be strong, but rock strength varies widely, so this is not always the case. If we consider just the strength of the rocks and ignore issues such as fracturing and layering, then most crystalline rocks (e.g., granite, basalt, or gneiss) are very strong, while some metamorphic rocks (e.g., schist) are only moderately strong. Sedimentary rocks have variable strength. Dolostone and some limestone are strong, most sandstone and conglomerate are moderately strong, and some sandstone and all mudstones are quite weak.

Fractures, metamorphic foliation (excluding gneissosity and banding), or bedding can significantly reduce the strength of rock. In the context of mass wasting, this is most critical if the planes of weakness are parallel to the slope and least critical if they are perpendicular to the slope. This is illustrated in Figure 15.3. At locations A and B the bedding is nearly perpendicular to the slope and the layers of rock are relatively stable. At location D the bedding is nearly parallel to the slope and the layers of rock are relatively unstable. At location C the bedding is nearly horizontal, and the stability is intermediate between the two extremes.

Relative stability of slopes as a function of the orientation of weaknesses (in this case bedding planes) relative to the slope orientations. [SE]
Figure 15.3 | Relative stability of slopes. The stability is as a function of the orientation of planes of weakness (in this case bedding planes) relative to the slope orientations. Source: Steven Earle (2015) CC BY 4.0. View source

Internal variations in the composition and structure of rocks can significantly affect their strength. Schist, for example, may have layers that are rich in sheet silicates (micas) and these will tend to form weak layers. Some minerals tend to be more susceptible to weathering than others, and the weathered products are commonly quite weak (e.g., clay formed from feldspar). The side of Johnson Peak that failed in 1965 (Hope Slide) is made up of chlorite schist (metamorphosed sea-floor basalt) that has feldspar-bearing sills within it. The foliation and the sills are parallel to the steep slope. The schist is relatively weak to begin with, and the feldspar in the sills, which has been altered to clay, makes it even weaker.

Unconsolidated sediments are generally weaker than sedimentary rocks because they are not cemented and, in most cases, have not been significantly compressed by overlying materials. Unconsolidated sediments can still bind together, and the strength of that binding is called cohesion. A cohesive sediment binds together strongly and if you picked it up with a shovel it would stick together in a lump (e.g., sand mixed with clay, clay). A sediment that is not very cohesive is weakly bound and would probably fall apart if you picked it up with a shovel (e.g., sand, silt). The deposits that make up the cliffs at Point Grey, Vancouver, B.C. include sand, silt, and clay, overlain by sand. The finer deposits at Point Grey are relatively cohesive (they maintain a steep slope, Figure 15.4 left).  The overlying sand is not very cohesive (relatively weak) and has a shallower slope because there are many slope failures in the sand deposit.

Left: Glacial outwash deposits at Point Grey, in Vancouver. The dark lower layer is made up of sand, silt, and clay. The light upper layer is well-sorted sand. Right: Glacial till on Quadra Island, B.C. The till is strong enough to have formed a near-vertical slope. [SE]
Figure 15.4 Left: Glacial outwash deposits at Point Grey, Vancouver, B.C. The dark lower layer is made up of sand, silt, and clay. The light upper layer is well-sorted sand, which has experienced slope failure and formed a cone of talus. Right: Glacial till on Quadra Island, B.C. The till is strong enough to have formed a near-vertical slope. Source: Steven Earle (2015) CC BY 4.0. View source

In contrast to poorly cohesive sediment deposits, glacial till can be as strong as some sedimentary rock. Glacial till is typically a mixture of clay, silt, sand, gravel, and larger clasts and forms and is compressed beneath tens to thousands of metres of glacial ice (Figure 15.4, right).

Apart from the type of material on a slope, the amount of water that the material contains is the most important factor controlling its strength. This is especially true for unconsolidated materials (e.g., Figure 15.4), but it also applies to bodies of rock. Granular sediments, like the sand at Point Grey, have lots of pore spaces between the grains. These spaces may be completely dry (filled only with air), moist (some spaces are water filled), or completely saturated (Figure 15.5).

Unconsolidated sediments tend to be strongest when they are moist because the small amounts of water at grain boundaries holds the grains together due to surface tension. Surface tension is the tension at the surface of a fluid that allows the liquid to resist an external force. Liquids always tend to acquire the lowest surface area possible; this happens because molecules at the surface of the fluid are attracted to the molecules below the surface). This is the property of liquid water that allows insects to walk over it.  Dry sediments are held together only by the friction between grains, and if they are well sorted or well rounded, or both, this cohesion is weak, due to minimal grain contact. Saturated sediments tend to be the weakest of all because the water pushes the grains apart, decreasing friction between grains. Water will also reduce the strength of solid rock, if the rock has porosity, fractures, bedding planes, and/or clay-bearing zones, especially when the rock is saturated with water (saturated conditions).

Depiction of dry, moist, and saturated sand [SE]
Figure 15.5 | Depiction of dry, moist, and saturated sand. Source: Steven Earle (2015) CC BY 4.0. View source

Water pressure is an important factor in slope failure. As you move deeper in saturated sediment, the pressure of the water goes up due to gravity acting on the column of water above it; this pressure is called hydrostatic pressure. The greater the depth below the surface of the water table (the point where the rock or sediments are saturated), the greater the water pressure acting on the materials. Holes are often drilled into rocks in road cuts to allow water to drain and relieve this water pressure. One of the hypotheses advanced to explain the 1965 Hope Slide is that cold conditions that winter caused small springs in the lower part of the slope to freeze, preventing water from flowing out. It is possible that water pressure gradually built up within the slope, weakening the rock mass to the extent that the shear strength was no longer greater than the shear force.

Water also has an interesting effect on clay-bearing materials. All clay minerals will absorb a small quantity of water, which reduces the strength of the clay. The smectite clays (such as the bentonite used in cat litter) can absorb a lot of water, and this water pushes the clay sheets apart at a molecular level, which makes the clay swell. Smectite that has expanded in this way has almost no strength; it is extremely slippery. Thus, slopes containing smectite clay are more likely to experience slope failure when they are saturated.

Water can increase the mass of the material on a slope, because the mass of the water is a component of the overall mass of the slope material. This increases the gravitational force pulling the slope materials down. A water saturated body of sediment with 25% porosity weighs approximately 13% more than it does when it is completely dry, so the gravitational shear force is also 13% higher. In the situation shown in Figure 15.2b, a 13% increase in the shear force is enough to overcome the shear strength, and the block would move down the slope.

Exercise 15.1 Sand and Water

Sand and Water

Source: Steven Earle (2015) CC BY 4.0.View source

If you have ever been to the beach, you already know that sand behaves differently when it is dry than it does when it is wet. The following experiment will demonstrate the strength of sand when it is dry, moist and saturated.

Find approximately half a cup of clean, dry sand (or get some wet sand and dry it out), and then pour it from your hand onto a piece of paper. You should be able to make a cone-shaped pile that has a slope of ~30°. If you pour more sand onto the pile, it will get bigger, but the slope should remain the same.

Now add some water to the sand so that it is moist. One way to do this is to add enough water to saturate the sand, then let the water drain away for a minute. You should be able to form this moist sand into a steep pile (with slopes of ~80°).

Finally, put some sand into a cup and fill the cup with water so the sand is just covered. Swirl it around so that the sand remains in suspension, and then quickly tip it out onto a flat surface. It should spread out over a wide area, forming a pile with a slope of only a few degrees.

Triggers of Mass-Wasting

In the previous section, we discussed the shear force and the shear strength of materials on slopes, and factors that can decrease the shear strength. Shear force is primarily related to slope angle, and once a slope angle is set, the shear force is constant. But shear strength can change quickly for a variety of reasons. Events that lead to a rapid decrease in shear strength are triggers for mass wasting.

An increase in water content is the most common trigger of mass wasting. This can result from rapid melting of snow or ice, heavy rain, or other events that change the pattern of water flow on and through the material. Rapid melting can be caused by a dramatic increase in temperature (e.g., in spring or early summer), or by a volcanic eruption. Heavy rains are typically related to storms. An example of a major slope failure caused by an increase in water content was the Oso landslide that occurred in Washington State, USA in 2014 (Figure 15.6). The flow buried the community of Steelhead Haven and killed 43 people.

Figure 15.6 | The Oso landslide, a flow that occurred in Washington State, USA 22 March 2014. Source: Matthew Sissel (2014) Public Domain. View source

Changes in water flow patterns can be caused by earthquakes, dammed streams from previous slope, or human structures that interfere with runoff (e.g., buildings, roads, or parking lots). An example of this is a deadly 2005 debris flow in North Vancouver, B.C. This slope failure took place in an area where there had been previous slope failures. A report written in 1980 recommended that the municipal authorities and residents take steps to address surface and slope drainage issues. Unfortunately, little was done to improve the situation or to take steps that could have prevented the 2005 slope failure. The failure happened during a rainy period but was likely triggered by excess runoff related to the roads at the top of this slope and by landscape features including addition of fill to backyards in the area above the failure.

In some cases, a decrease in water content can lead to failure. This is most common with clean sand deposits (e.g., the upper layer in Figure 15.4 (left)), which loses some of its strength when most of the water around the grains is removed (i.e., as the sand water content drops the surface tension decreases).

Freezing and thawing can also trigger some forms of mass wasting e.g., thawing can release a block of rock that was frozen to a slope by a film of ice.

One other process that can weaken a body of rock or sediment is shaking. The most obvious source of shaking is an earthquake. Shaking from highway traffic, construction, or mining can also weaken rock. Several deadly mass wasting events (including snow avalanches) were triggered by the M7.8 earthquake in Nepal in April 2015.

103

15.2 Classification of Mass Wasting

While we do not classify slope failures by the shape of the rupture surface, it is nevertheless an important feature used to describe mass wasting processes. The rupture surface is the boundary between the slope and the sliding material. Slope failures with curved rupture surfaces are called rotational slope failures, and slope failures with planar rupture surfaces are called translational slope failures.

Unfortunately, classifying slope failure is not as simple as the classification scheme above suggests. Many slope failures involve more than one type of motion, and often it is not easy to tell how the material moved. The types of slope failure that are covered in this chapter are summarized in Table 15.1.

Failure Type Type of Material Type of Motion Rate of Motion
Rock fall Rock fragments Vertical or near-vertical fall (plus bouncing in many cases) Very fast (>10s m/s)
Rock slide A large rock body Motion as a unit along a planar surface (translational sliding) Typically very slow (mm/y to cm/y), but some can be faster
Rock avalanche A large rock body that slides and then breaks into small fragments Flow (at high speeds, the mass of rock fragments is suspended on a cushion of air) Very fast (>10s m/s)
Creep or solifluction Soil or other overburden; in some cases, mixed with ice Flow (although sliding motion may also occur) Very slow (mm/y to cm/y)
Slump Thick deposits (m to 10s of m) of unconsolidated sediment Motion as a unit along a curved surface (rotational sliding) Slow (cm/y to m/y)
Mudflow Loose sediment with a significant component of silt and clay Flow (a mixture of sediment and water moves down a channel) Moderate to fast (cm/s to m/s)
Debris flow Sand, gravel, and larger fragments Flow (similar to a mudflow, but typically faster) Fast (m/s)

Table 15.1 | Classification of slope failures based on type of material and motion.  Source: Steven Earle (2015) CC BY 4.0. View source

Rock Fall

Rock fragments can break off relatively easily from steep bedrock slopes, most commonly due to frost-wedging. If you ever hike along a steep mountain trail on a cool morning, you will probably hear the occasional fall of rock fragments onto a talus slope. Water in the cracks in the rock freezes and expands overnight, and this can break the rock apart. When the ice thaws in the morning sun, some of these broken fragments will fall to the slope below (Figure 15.7). Talus slopes form from this fallen rock, forming slopes at the angle of repose below high rock walls (Figure 15. 8, left). They are also known as scree slopes.

Figure 15.7 The contribution of freeze-thaw to rock fall [SE]
Figure 15.7 | The contribution of freeze-thaw to a rock fall. Source: Steven Earle (2015) CC BY 4.0. View source

A typical talus slope, near Keremeos in southern BC, is shown in Figure 15.8. In December 2014, a large block of rock split away from a cliff in this same area. It broke into smaller pieces that tumbled down the slope and crashed into the road, smashing the concrete barriers and gouging out large parts of the pavement. Luckily no one was hurt.

Figure 15.8 Left: A talus slope near Keremeos, B.C., formed by rock fall from the cliffs above. Right: The results of a rock fall onto a highway west of Keremeos in December 2014. [SE]
Figure 15.8 | Left: A talus slope near Keremeos, B.C., formed by rock falls from the cliffs above. Right: The results of a rock fall onto a highway west of Keremeos in December 2014. Source: Steven Earle (2015) CC BY 4.0. View source

Rock Slide

A rock slide is a large body of rock that is sliding along a sloping surface. In most cases, the movement is parallel to a fracture, bedding plane, or metamorphic foliation plane; thus, most rock slides are translational slope failures.

The speed of the movement can range from very slow to moderately fast. The word sackung describes the very slow motion of a block of rock (mm/y to cm/y) on a slope. A good example is the Downie Slide, a translational slide north of Revelstoke, BC, which is shown in Figure 15.9. In this case, a massive body of rock is very slowly sliding down a steep slope along a plane of weakness that is caused by the foliation in the rock. The foliation is approximately parallel to the slope.

The Downie Slide was recognized prior to the construction of the Revelstoke Dam, and was moving very slowly at the time of dam construction (a few cm/year). Geological engineers were concerned that the presence of water in the reservoir (visible in Figure 15.9) could further weaken the plane of failure, leading to an acceleration of the motion. The result would have been a catastrophic failure into the reservoir that would have sent a wall of water over the dam and into the community of Revelstoke.

Figure 15.9 | The Downie Slide, a sackung, on the shore of the Revelstoke Reservoir (above the Revelstoke Dam). The head scarp is visible at the top and a side-scarp along the left side. Source: Joyce McBeth (2018) CC BY 4.0, image © 2018 Google Earth, Data Google CNES / Airbus Data LDEO-Columbia, NSF, NOAA Data SIO, NOAA, U.S. Navy, NGA, GEBCO DigitalGlobe Landsat / Copernicus Province of BC.

During the construction of the dam, the engineers tunnelled into the rock at the base of the slide and drilled hundreds of drainage holes upward into the plane of failure. This allowed water to drain out more efficiently so that the hydrostatic pressure was decreased, which decreased the rate of movement of the sliding block. BC Hydro monitors this site continuously. The slide block is currently moving more slowly than it was prior to the construction of the dam.

In the summer of 2008, a large block of rock slid rapidly from a steep slope above Highway 99 near Porteau Cove, BC (between Horseshoe Bay and Squamish). The block crashed into the highway and adjacent railway and broke into many pieces, and the highway was closed for several days. Engineers and geoscientists took steps to stabilize the slope to decrease the risk of future rock falls. Rock bolts (long metal rods) were installed to anchor the blocks of rocks and prevent them falling Drainage holes were installed to drain water from the slope and decrease the water pressure. As shown in Figure 15.10, the rock along the slope is fractured parallel to the slope, and this almost certainly contributed to the failure. However, it is not actually known what triggered this event as the weather was dry and warm during the preceding weeks, and there was no significant earthquake in the region.

Figure 15.10 Site of the 2008 rock slide at Porteau Cove. Notice the prominent fracture set parallel to the surface of the slope. The slope has been stabilized with rock bolts (top arrow) and holes have been drilled into the rock to improve drainage (tube from drainage hole indicated with bottom arrow). Risk to passing vehicles from rock fall has been reduced by hanging mesh curtains (background), which secures loose material to the slope. Source: Joyce McBeth (2018) CC BY 4.0 after Steven Earle (2015) CC BY 4.0. View source

Rock Avalanche

If a rock slides and then starts moving quickly (m/s), the rock is likely to break into many small pieces. At this point it is considered to be a rock avalanche, in which the large and small fragments of rock move in a fluid manner supported by a cushion of air within and beneath the moving mass. The 1965 Hope Slide (Figure 15.1) was a rock avalanche, as was the famous 1903 Frank Slide in southwestern Alberta. The 2010 slide at Mt. Meager (west of Lillooet) was also a rock avalanche and rivals the Hope Slide as the largest slope failure in Canada during historical times (Figure 15.11).

Figure 15.11 The 2010 Mt. Meager rock avalanche, showing where the slide originated (arrow, 4 km upstream). It then raced down a steep narrow valley, and out into the wider valley in the foreground. [Mika McKinnon photo, http://www.geomika.com/blog/2011/01/05/the-trouble-with-landslides/ Used with permission.]
Figure 15.11 | The 2010 Mt. Meager landslide, showing where the slide originated (arrow, 4 km upstream). It then raced down a steep narrow valley and out into the wider valley in the foreground.  Source: Mika McKinnon (2011) CC BY-SA-NC, View source

Creep or Solifluction

The very slow — mm/y to cm/y — movement of soil or other unconsolidated material down slope is known as creep. Creep, which normally only affects the upper several centimetres of loose material, is typically a type of very slow flow. In some cases, sliding may take place too.

Creep can be facilitated by freezing and thawing because, as shown in Figure 15.12, particles are lifted perpendicular to the surface by the growth of ice crystals within the soil, and then move downwards vertically due to gravity when the ice melts. The same effect can be produced by frequent wetting and drying of the soil. In cold environments, solifluction is a more intense form of freeze-thaw-triggered creep.

Figure 15.12 A depiction of the contribution of freeze-thaw to creep. The blue arrows represent uplift caused by freezing in the wet soil underneath, while the red arrows represent depression by gravity during thawing. The uplift is perpendicular to the slope, while the drop is vertical. [SE]
Figure 15.12 | A depiction of the contribution of freeze-thaw to creep. The blue arrows represent uplift caused by freezing in the wet soil underneath, while the red arrows represent depression by gravity during thawing. The uplift is perpendicular to the slope, while the drop is vertical.  Source: Steven Earle (2015) CC BY 4.0. View source

Creep is most noticeable on moderate-to-steep slopes where trees or fence posts are consistently leaning in a downhill direction. In the case of trees, they try to correct their lean by growing upright, and this leads to a curved lower trunk known as a “pistol butt.” Creep can also generate terracettes, horizontal and repeating ridges on slopes (Figure 15.13). Historically, people thought terracettes formed where livestock or wild animals regularly travelled along slopes. While animals can accentuate terracettes, the primary reason terracettes form is creep.

Figure 15.13 | Evidence of creep-generated terracettes on the Peace River hills in northeastern B.C. Source: Joyce McBeth (2018) CC BY 4.0.

Slump

A slide is a mass movement where the material moves as a coherent mass. A slump is a type of slide that takes place within thick unconsolidated deposits (typically thicker than 10 m). Slumps involve movement along one or more curved failure surfaces, and are thus rotational slope failures. Slumps have downward motion near the top and outward motion toward the bottom (Figure 15.14). They are typically caused by high water pressure within these materials on a steep slope.

Figure 15.14 A depiction of the motion of unconsolidated sediments in an area of slumping [SE]
Figure 15.14 | The motion of unconsolidated sediments in an area of slumping.  Source: Steven Earle (2015) CC BY 4.0. View source

An example of a slump in the Lethbridge area of Alberta is shown in Figure 15.15. This feature has likely been active for many decades and moves a little more whenever there are heavy spring rains and snowmelt runoff. The toe of the slump is being eroded by the small stream at the bottom. The erosion contributes to continued slumping. The basal material (material at the toe of the slope) forms the support for the overlying mass of material in the slope and if this support is eroded away slumping will likely continue.

Figure 15.15 A slump along the banks of a small coulee near Lethbridge, Alberta. The main head-scarp is clearly visible at the top, and a second smaller one is visible about one-quarter of the way down. The toe of the slump is being eroded by the seasonal stream that created the coulee. [SE 2005]
Figure 15.15 | A slump along the banks of a small coulee near Lethbridge, Alberta. The main head-scarp is clearly visible at the top, and a second smaller one is visible about a quarter of the way down the slope. The toe of the slump is being eroded by the seasonal stream that created the coulee. Source: Steven Earle (2015) CC BY 4.0. View source

Mudflows and Debris Flows

As you saw in Exercise 15.1, when a mass of sediment becomes completely saturated with water, the mass loses strength, to the extent that the grains may be pushed apart and may flow, even on a gentle slope. This can happen during rapid spring snowmelt or heavy rains and is also relatively common during volcanic eruptions because of rapid melting of snow and ice. If the material involved is primarily sand-sized or smaller, it is known as a mudflow, such as the one shown in Figure 15.16. A mudflow or debris flow on a volcano or during a volcanic eruption is called a lahar.

If the material involved is gravel sized or larger, it is known as a debris flow. Since it takes more gravitational force to overcome friction and move larger particles, debris flows typically form in areas with steeper slopes and higher water pressure. In many cases, a debris flow takes place within a steep stream channel and is triggered by the collapse of bank material into the stream. This may create a temporary dam followed by a major flow of water and debris when the dam finally bursts. This is the situation that led to the fatal debris flow at Johnsons Landing, BC, in 2012.

Figure 15.16 A slump (left) and an associated mudflow (centre) at the same location as Figure 15.15, near Lethbridge, Alberta. [SE 2005]
Figure 15.16 | A slump (left) and an associated mudflow (centre) at the same location as Figure 15.15, near Lethbridge, Alberta. Source: Steven Earle (2015) CC BY 4.0. View source

A typical west-coast debris flow is shown in Figure 15.17. This event took place in November 2006 in response to very heavy rainfall. There was enough energy in the flow to move large boulders and to knock over large trees.

Figure 15.17 The lower part of debris flow within a steep stream channel near Buttle Lake, B.C., in November 2006. [SE]
Figure 15.17 The lower part of debris flow within a steep stream channel near Buttle Lake, B.C., in November 2006.  Note the trees along the edges of the stream that have been damaged by the rocks in the debris flow. Source: Steven Earle (2015) CC BY 4.0. View source

 

Exercise 15.2 Classifying Slope Failures

These four photos show some of the different types of slope failures described above. Try to identify each and provide some criteria to support your choice.

 Classifying Slope Failures1 Classifying Slope Failures2
 Classifying Slope Failures3  Classifying Slope Failures4

Source of images: Steven Earle (2015) CC BY 4.0. View source

104

15.3 Preventing, Delaying, Monitoring, and Mitigating Mass Wasting

We cannot prevent mass wasting, however, in many situations there are actions we can take to reduce or mitigate the damaging effects of mass wasting on people and infrastructure. Where we can neither delay nor mitigate mass wasting, we may consider trying to initiate the slope failure in a controlled manner. In areas prone to mass wasting that cannot be controlled or mitigated, we can minimize risk by not building in these areas at all.

Preventing and Delaying Mass Wasting

It is comforting to think that we can prevent some effects of mass wasting by mechanical means. For example, the rock bolts in the road cut at Porteau Cove on the Sea-toSky highway in BC (Figure 15.10) or the drill holes used to drain water out of the slope at the Downie Slide (Figure 15.9), or the building of physical barriers, such as retaining walls along highway roadcuts. These preventative measures are not permanent though, they are subject to degradation over time. The rock bolts in the road cut at Porteau Cove will slowly start to corrode after a few years, and within a few decades many of them will begin to lose their strength. Unless they are replaced, they will no longer support the slope. Likewise, drainage holes at the Downie Slide will eventually become plugged with sediment and chemical precipitates, and unless they are periodically unplugged, their effectiveness will decrease. Eventually, unless new holes are drilled, the drainage will be compromised, and the slide will start to move again. This is why careful slope monitoring by geological and geotechnical engineers is important at major mass wasting sites such as the Downie Slide and along the Sea-to-Sky highway. Our efforts to control mass wasting are only as good as our efforts to maintain the preventive measures.

Delaying mass wasting is a worthy endeavour because during the time that the measures are still effective, they can save lives and reduce damage to property and infrastructure such as homes and roads. But we must be careful to avoid activities that could make mass wasting more likely. One of the most common anthropogenic causes of mass wasting is road construction, and this applies both to remote gravel roads built for forestry and mining, and large urban and regional highways.

Figure 15.18 An example of a road constructed by cutting into a steep slope and the use of the cut material as fill. [SE]
Figure 15.18 | An example of a road constructed by cutting into a steep slope and the use of the cut material as fill. Source: Steven Earle (2015) CC BY 4.0. View source

 

Road construction is a potential problem for two reasons. First, creating a flat road surface on a slope inevitably involves creating a cut bank that is steeper than the original slope. This might also involve creating a filled bank that is both steeper and weaker than the original slope (Figure 15.18). Second, roadways typically cut across natural drainage features, and unless great care is taken to reroute the runoff water, oversaturation of slope material can occur, contributing to mass wasting.

Apart from saturation and water pressure considerations, engineers building roads and other infrastructure on bedrock slopes have to carefully consider the geology, and especially any weaknesses or discontinuities in the rock related to bedding, fracturing, or foliation. If possible, situations like that at Porteau Cove (Figure 15.10) should be avoided — by building somewhere else — rather than trying to stitch the slope back together with rock bolts.

It is widely believed that construction of buildings above steep slopes can contribute to the instability of the slope. This is likely true, but probably not because of the weight of the building. As you will determine by completing Exercise 15.3 below, a typical house is not heavier than the excavated ground that was removed to build the house. A more likely contributor to instability of the slopes below buildings is changes to the water drainage and to the saturation of the slope (by watering gardens, for example).

Exercise 15.3 How Much Does a House Weigh and Can It Contribute to Slope Failure?

It is commonly believed that building a house (or some other building) at the top of a slope will add a lot of extra weight to the slope, which could contribute to slope failure. But what does a house weigh compared to the material removed when you build it? A typical 150 m2 (approximately 1,600 ft2) wood-frame house with a basement and a concrete foundation weighs about 145 t (metric tonnes). But most houses are built on foundations that are excavated into the ground. This involves digging a hole and taking that material away, so we need to subtract what that excavated material weighs. Assuming our 150 m2 house required an excavation that was 15 m by 11 m by 1 m deep, which is 165 m3 of material. Unconsolidated sediments have densities ranging from about 0.8 to 1.7 t per m3.

For this exercise, consider a sand with a dry density of 1.2 t per m3 for this calculation. Calculate the weight of the materials that were removed and compare that with the weight of the house and its foundation.

If you are thinking that building a bigger building is going to add more weight, consider that bigger buildings need bigger and deeper excavations, and in many cases the excavations may be into solid rock, which is denser than surficial materials.

Consider how a building might change the drainage on a slope. Water can be collected by roofs, go into downspouts, and form concentrated flows that are directed onto or into the slope. Likewise, drainage from nearby access roads, lawn irrigation, leaking pools, and septic systems can all alter the surface and groundwater flow in a slope. Soil excavated from a basement.

How Much Does a House Weigh and Can It Contribute to a Slope Failure
Source: Steven Earle (2015) CC BY 4.0. View source

Monitoring Mass Wasting

Warning systems are helpful in some areas where there is a risk of mass wasting. They let us know if conditions have changed at a known slide area, or if a rapid failure, such as a debris flow, is on its way downslope. The Downie Slide above the Revelstoke Reservoir is continuously-monitored with a range of devices, such as inclinometers (slope-change detectors), bore-hole motion sensors, and GPS survey instruments. A simple mechanical device for monitoring the nearby Checkerboard Slide (which is also above the Revelstoke Reservoir) is shown in Figure 15.19. Both of these slides are very slow-moving, but it is important to be able to detect changes in their rates of motion. A rapid failure would result in large bodies of rock plunging into the reservoir and sending a wall of water over the Revelstoke Dam, potentially destroying the nearby town of Revelstoke.

Figure 15.19 Part of a motion-monitoring device at the Checkerboard Slide near Revelstoke, B.C. The lower end of the cable is attached to a block of rock that is unstable. Any incremental motion of that block will move the cable and this will be detectable on this device. [SE]
Figure 15.19 | Part of a motion-monitoring device at the Checkerboard Slide near Revelstoke, BC. The lower end of the cable (extending out from the top of the device to the right) is attached to a block of rock that is unstable. Any incremental motion of this block will move the cable, which will be detectable by this device.  Source: Steven Earle (2015) CC BY 4.0. View source

Mt. Rainier, a glacier-covered volcano in Washington State (15.20), could produce massive mudflows or debris flows (lahars) with or without a volcanic eruption. Over 100,000 people in the Tacoma, Puyallup, and Sumner areas are at risk because they currently reside on deposits from past lahars and future lahars would likely also follow these paths (Figure 15.21). In 1998, a network of acoustic monitors was established around Mt. Rainier. The monitors are embedded in the ground adjacent to expected lahar paths. These monitors will provide warnings to emergency officials in the event of a lahar. When a lahar is detected, the residents of the area will have between 40 minutes and three hours to get to safe ground.

Figure 15.20 | Mt. Rainier from Seattle, WA, USA. Source: Flickr user “accozzaglia dot ca” (2010) CC-BY-NC-ND 2.0. View source

 

Figure 15.21 | Major pathways of Mt Rainier lahars over the past 10,000 years, Washington State, USA. Source: USGS (2005) Public Domain view source, modified from Driedger et al (2005) Public Domain. View source

Mitigating the Impacts of Mass Wasting

In situations where we cannot predict, prevent, or delay mass-wasting hazards, some effective measures can be taken to minimize the associated risk. For example, many highways in BC and western Alberta have avalanche shelters like the one shown in Figure 15.22. In some parts of the world, similar structures have been built to protect infrastructure from other types of mass wasting.

Figure 15.21 A snow avalanche shelter on the Coquihalla Highway. The expected path of the avalanche is the steep untreed slope above. [SE]
Figure 15.22 | A snow avalanche shelter on the Coquihalla Highway (bottom centre of the image). The expected path of the avalanche is the steep and treeless slope above.  Source: Steven Earle (2015) CC BY 4.0. View source

Debris flows are inevitable, unpreventable, and unpredictable in many parts of BC, but nowhere more so than along the Sea-to-Sky Highway between Horseshoe Bay and Squamish. The results have been deadly and expensive many times in the past. It would be very expensive to develop a new route in this region, so provincial authorities have taken steps to protect residents, and traffic on the highway and railway. Debris flow defensive structures have been constructed in several drainage basins, as shown in Figure 15.23. One strategy is to allow the debris flow to flow quickly through to the ocean along a smooth channel. Another is to capture the debris within a constructed basin that allows the excess water to continue through.

Figure 15.22 Two strategies for mitigating debris flows on the Sea-to-Sky Highway. Left: A concrete –lined channel on Alberta Creek allows debris to flow quickly through to the ocean. Right: A debris-flow catchment basin on Charles Creek. In 2010 a debris flow filled the basin to the level of the dotted white line. [SE]
Figure 15.23 | Two strategies for mitigating debris flows on the Sea-to-Sky Highway. Left: A concrete –lined channel on Alberta Creek allows debris to flow quickly through to the ocean. Right: A debris flow catchment basin on Charles Creek. In 2010, a debris flow filled the basin to the level of the dotted white line.  Source: Steven Earle (2015) CC BY 4.0. View source

 

Finally, in situations where we cannot do anything to delay, predict, contain, or mitigate slope failures, the responsible and ethical thing to do is to avoid building in or using the risky area. Sometimes this may require relocating a community after discovering a previously-unidentified risk. There is a famous example of this in BC at a site known as Garibaldi, 25 km south of Whistler. In the early 1980s the village of Garibaldi had a population of about 100, with construction underway on some new homes, and plans for many more. In the months that followed the deadly 1980 eruption of Mt. St. Helens in Washington State, the BC Ministry of Transportation commissioned a geological study to assess risks along their highways.  The study revealed that a steep cliff known as The Barrier (Figure 15.24) had collapsed in 1855, leading to a large rock avalanche, and that it was likely to collapse again unpredictably, putting the village of Garibaldi at extreme risk. In an ensuing court case, it was ruled that the Garibaldi site was not a safe place for people to live. Those who already had homes there were compensated, and everyone was ordered to leave.

Figure 15.23 The Barrier, south of Whistler, B.C., was the site of a huge rock avalanche in 1855, which extended from the cliff visible here 4 km down the valley and across the current location of the Sea-to-Sky Highway and the Cheakamus River. [SE]
Figure 15.24 | The Barrier, south of Whistler, BC, was the site of a huge rock avalanche in 1855, which extended from the cliff visible here 4 km down the valley and across the current location of the Sea-to-Sky Highway and the Cheakamus River.  Source: Steven Earle (2015) CC BY 4.0. View source

105

Chapter 15 Summary

The topics covered in this chapter can be summarized as follows:

15.1 Factors That Control Stability on Slopes

Slope stability is controlled by the slope angle and the strength of the material on the slope. Slopes are a product of tectonic uplift, and their strength is determined by the type of material on the slope and its water content. Rock strength varies widely and is determined by internal planes of weakness and their orientation with respect to the slope. In general, the more water contained by the slope material, the greater the likelihood of failure. This is especially true for unconsolidated sediments, where excess water pushes against the grains. Addition of water is the most common trigger of mass wasting and can come from storms or rapid snow melt.

15.2 Classification of Mass Wasting

The key criteria for classifying mass wasting are the nature of the movement that takes place, the type of material, and the speed of the material movement. Mass wasting events can be a precipitous fall of rock through the air, material sliding as a solid mass along either a plane or a curved surface, or internal flow of material as a viscous fluid. The type of material influences the mass movement, specifically whether it is solid rock or unconsolidated sediments. Slope failures can have translational (planar) or rotational (curved) rupture surfaces. The important types of mass wasting are creep, slump, slide, fall, and debris flow or mudflow.

15.3 Preventing, Delaying, and Mitigating Mass Wasting

We cannot prevent mass wasting, but we can delay it through efforts to strengthen the materials on slopes. Strategies include adding mechanical devices such as rock bolts or ensuring that water in the slope materials can easily drain away. Such measures are never permanent but may be effective for decades or even centuries. We can also avoid practices that make matters worse, such as cutting into steep slopes or impeding proper drainage. In some situations, the best approach is to mitigate the risks associated with mass wasting by constructing shelters or diversionary channels. In other cases, where slope failure is inevitable, we should simply avoid building in that location.

Questions for Review

gravitational force on the unconsolidated sediment
Source: Steven Earle (2015) CC BY 4.0 View source

1. In the scenario shown here, the gravitational force on the unconsolidated sediment overlying the point marked with an X is depicted by the black arrow. The red arrow in the diagram depicts the shear strength of the sediment.

  1. Draw in the two arrows that show how this force can be resolved into the shear force (along the slope) and the normal force (perpendicular to the slope).
  2. Assuming that the relative lengths of the shear force arrow (which you drew in question 1), and the shear strength arrow are indicative of the likelihood of failure, predict whether this material is likely to fail or not.
  3. After several days of steady rain, the sediment becomes saturated with water and its shear strength is reduced by 25%. What are the likely implications for the stability of this slope?
  4. Did you consider the affect of the additional weight of the water on the gravitational force acting on the slope in your answer to (c)? Does this change your answer?

2. In the diagrams shown here, a road cut is constructed in sedimentary rock with well-developed bedding. On the left, draw in the orientation of the bedding that would represent the greatest likelihood of slope failure. On the right, show the orientation that would represent the least likelihood of slope failure.

a road cut a road cut

Source: Steven Earle (2015) CC BY 4.0 View source

3. Explain why moist sand is typically stronger than either dry sand or saturated sand.

4. In the context of mass wasting, how does a flow differ from a slide?

5. If a large rock slide starts moving at a rate of several metres per second, what is likely to happen to the rock, and what would the resulting failure be called?

6. In what ways does a debris flow differ from a mudflow?

7. In the situation described in the chapter regarding lahar warnings at Mt. Rainier, the residents of the affected regions have to assume some responsibility and take precautions for their own safety. What sort of preparation should the residents make to ensure that they can respond appropriately when they hear lahar warnings? What other considerations do officials have to make in their emergency plan, other than just sounding a warning?

8. What factors are likely to be important when considering the construction of a house near the crest of a slope that is underlain by glacial sediments?

106

Answers to Chapter 15 Review Questions

1. (a) The shear force and normal force vectors are shown on the left-hand diagram:

vectoris

(b) Based on the relative lengths of the arrows it appears that this material is stable, and unlikely to fail.
(c) If the shear strength was reduced by 25% (right-hand diagram) the material would be much closer to failure, but the strength (based on the length of the arrows) still appears to be greater than the shear force.
2. slope

3. In moist sand the grains are each surrounded by an envelope of water, and the water envelopes overlap. The attractive surface tension of the water holds the grains together.

4. In a the material moves like a fluid (individual particles move independently). In a the mass moves as an intact unit, with little or no relative motion between grains or clasts.

5. If a large rock slide starts moving at a rate of several metres per second, the rock is very likely to break into smaller pieces. If the pieces are small and numerous enough that the material can flow, then it becomes a rock avalanche.

6. A debris flow is composed mostly of sand-sized and larger clasts, while a mudflow is composed mostly of sand-sized and smaller clasts.

7. Residents at risk from Mt. Rainier lahars need to know what the warnings mean and roughly how much time they have between receiving a warning and being in actual danger. They need to create a plan to exit their residence quickly, and they need to know which way to go to get to safety as efficiently as possible.

8. Some of the important factors include:

XVI

Chapter 16. Earth-System Change

Adapted by Karla Panchuk from Physical Geology by Steven Earle

Sea ice floating in dark blue ocean water along the Antarctic Peninsula
Figure 16.1 Antarctic Peninsula. Antarctica was not always covered by ice. A change in the Earth system triggered the onset of Antarctic glaciation approximately 40 million years ago. Source: Karla Panchuk (2018) CC BY-SA 4.0. Photo: Liam Quinn (2011) view source. Click the image for more attributions.

Learning Objectives

After reading this chapter, completing the exercises within it, and answering the questions at the end, you should be able to:

The Only Constant Is Change

If one thing has been constant about the Earth system over geological time, it is unceasing change. In the geological record of climate, sedimentary deposits provide evidence of glaciations in the distant past; the chemical characteristics of sea-floor sediments tell about periods of extreme warmth. The Earth-system, and thus Earth’s climate, has not only changed frequently, but also with large temperature fluctuations. Today’s mean global temperature is approximately 16°C. During Snowball Earth episodes more than 600 million years ago, when Earth’s surface was frozen from pole to pole (or nearly so), the global mean was as cold as -50°C. At various other times in Earth history, it has been close to 30°C.

Part of this chapter addresses natural processes of climate change, how they work, and how we know what Earth’s past climate was like. Geologists study those natural climate-change processes to understand how human-caused, or anthropogenic, changes to the Earth system might affect the climate in the future, and how much the climate has changed over the time that humans increased their influence on the Earth-system. The rest of the chapter addresses what has been learned by asking those questions.

 

107

16.1 What Is the Earth System?

Earth can be characterized in terms of its “spheres.” The atmosphere is the envelope of gas surrounding the planet. The hydrosphere is the water on the planet, whether in oceans, rivers, glaciers, or the ground. The biosphere comprises living organisms. The lithosphere is the rocky outer shell of the planet.

Components of these spheres interact constantly, with processes occurring in one sphere having an impact in other spheres. Cycles such as the water cycle or the carbon cycle constantly move matter and energy between spheres. Taking an Earth-system approach—looking at how the spheres are connected—is a way to account for the web of interactions responsible for the “big picture” of the Earth that we know.

The climate change related to the opening of the Drake Passage (Figure 16.2) is a good example of why a system of interactions is needed to understand how Earth works. The Drake Passage (bottom left map) is the gap between the southern-most tip of South America, and Antarctica.

Prior to 40 million years ago, the Drake Passage did not exist (top left map), and neither did the Antarctic ice cap. The arm of land connecting South America and Antarctica allowed warm ocean currents (red arrows) to carry heat from the equator to Antarctica. When the gap opened up, a new cold-water current formed (blue arrows) that blocked warm water from reaching Antarctica. Without the warm current, Antarctica froze over.

Figure 16.2 An example of Earth-system interactions. The opening of the Drake Passage changed how ocean currents moved heat around the planet, and cooled Antarctica until it froze over. Source: Karla Panchuk (2018) CC BY-NC-SA 4.0. Map: Modified after Woods Hole Oceanographic Institution (view source). Click the image for more attributions and terms of use.

There were many interconnecting processes within the Earth-system (Figure 16.2, right) that drove glaciation in Antarctica. First, heat energy within Earth drove plate tectonics (lithosphere), making it possible for South America to separate from Antarctica. This impacted ocean currents (hydrosphere), and ultimately how water was stored on Antarctica (hydrosphere) by changing the climate of Antarctica.

In the Earth system, nothing happens in isolation. The change in the climate of Antarctica had a global impact. The ice cap on Antarctica increased Earth’s albedo, the reflectiveness of Earth’s surface. The more reflective Earth’s surface, the more of the sun’s light is reflected back into space without heating Earth. This caused even more ice to form, and cooled the planet as a whole.

When Earth cools, the change in temperature has a cascade of effects including changing precipitation patterns (hydrosphere), and changing the characteristics of habitats (biosphere). When habitats cool, organisms needing more warmth will migrate closer to the equator. This is true of plants as well as other forms of life.

Ice is not the only type of land cover that affects albedo—forests do as well. Forests also increase local atmospheric moisture levels through transpiration, when they release water vapour into the atmosphere. Local temperature and moisture differences also affect rainfall patterns on top of larger-scale changes resulting from cooling.

The chain of events in summarized in Figure 16.2 is only a broad overview of all of the consequences of opening a gap between South America and Antarctica. For example, it does not include the effects of what a change in the types of plants in a location does to local weathering and erosion. Trees can accelerate weathering, releasing more nutrients from rocks into runoff, which can affect algae blooms in water bodies, which in turn reduces oxygen levels in the water, which affects organisms living in the water that rely on oxygen.

Trees growing along a river can also slow the rate of erosion, reducing the amount of sediment in the river, and ultimately the rate of development of a delta at the river’s mouth. Deltas undergo subsidence as accumulated sediments are compressed, so if the sediment supply is reduced, parts of the delta may become flooded, changing the extent of wetlands. Wetlands with waters depleted in oxygen can prevent plant material from decaying and releasing their carbon back into the carbon cycle as carbon dioxide. Changing atmospheric carbon dioxide levels alters the way energy moves through Earth’s atmosphere, and affects Earth’s surface temperatures.

The short version of why it’s important to look at Earth as a system is that everything is connected, so that a change in one part of the system can ripple through the rest of the system and have effects well beyond any one location or time.

Feedbacks Amplify or Diminish Earth-System Change

The web of interactions in the Earth system is complex, but there is yet another level of complication. Sometimes a change in the Earth system can trigger other changes that have the effect of amplifying the original change, or diminishing it. The series of interactions that amplify or diminish a change are called feedbacks. A feedback that amplifies change is called a positive feedback. A feedback that diminishes the size of a change is called a negative feedback.

In the events related to the glaciation of Antarctica, the formation of ice is an example of a positive feedback. Ice formation was caused by cooling, but it triggered even more cooling by reflecting sunlight away from Earth’s surface. This is called ice albedo feedback.  An example of a negative feedback is plant growth. Plants need CO2 to make food, so as long as the plants have enough nutrients and water, and temperatures are still suitable, increasing CO2 in the atmosphere could increase plant growth. Plant growth would draw down atmospheric CO2, so that there would be less warming than would otherwise be expected from the initial rise in atmospheric CO2 levels.

Misconceptions About Feedbacks

There are two common misconceptions about feedbacks. One misconception is that positive feedbacks result in changes that are good, and negative feedbacks result in changes that are bad. In fact, whether a feedback is positive or negative is unrelated to whether or not the change would be considered a good thing. For example, if a feedback accelerates warming and makes an ecosystems uninhabitable for animals that used to live there, it would still be a positive feedback even though it had a negative impact on the animals in that ecosystem. A feedback that slowed the rate of warming and gave the animals time to adapt would still be considered a negative feedback even though it helped the animals to survive.

Another misconception is that a positive feedback always results in some value increasing (e.g., a rise in temperature), and a negative feedback results in a decrease in that value. Positive feedbacks can cause a value to decrease (e.g., as ice forms more sunlight is reflected, leading to decreased temperatures), and negative feedbacks can cause a value to increase. What matters is whether the initial change is amplified or reduced, not which way the numbers are changing.

Feedbacks and Instability in the Earth System

The potential for sudden extreme changes in the Earth system depends on what feedbacks are available. At times when Earth’s climate was much warmer than today, no glaciers were present. When the climate is much cooler, a relatively small decrease in temperature could be enough to start the formation of ice and trigger the ice albedo feedback. However, if the climate is much warmer, the same decrease in temperature would not cool Earth enough to trigger the ice albedo feedback, and further climate cooling would be avoided. The reverse is also true- if warming occurs in a climate that is cold enough for glaciers to form, some of that ice might melt, reducing the albedo of Earth’s surface, and permitting even more warming. On the other hand, if the climate is already too warm for ice to exist, a small amount of warming won’t be amplified in the same way.

The albedo effect is not the only feedback that can make cooler climates less stable. Melting of permafrost (sediment that remains frozen year round) can also have an impact. Frozen soil contains trapped organic matter that is converted by micro-organisms to CO2 and methane (CH4) when the soil thaws. Both these gases contribute to warming when they accumulate in the atmosphere. Additional warming can cause even more permafrost to melt, permitting even more activity by micro-organisms, and releasing more CO2 and CH4.

Either of these feedbacks is enough on their own to accelerate climate change, but when they are both present together, the effect is even stronger. What this means is that the conditions in the Earth system before a change happens—called the initial conditions—play an important role in determining the impact of any changes that occur. A change that would have little impact under one set of initial conditions could have far reaching effects under another. Thinking of Earth as a system is a way to factor in the initial conditions. Otherwise we would be very puzzled why a small rise in global temperatures at one time in Earth history could have almost no discernible effect, but the same rise in temperatures at another time could lead to profound change.

 

 

108

16.2 Causes of Climate Change

What Is Climate?

Our day-to-day experience of the Earth system is in the form of the conditions we experience at Earth’s surface. The daily conditions that we think of as weather—the temperature, presence or absence of precipitation, winds, humidity, and so on—are a snapshot of the state of the Earth system at a particular instant in time and in a particular location. The weather that we get is variable, but in Saskatchewan most people would not be surprised to experience summer days with temperatures of 20 °C to 30 °C, and winter days with temperatures between –20 °C to –30 °C. Our notion of what summers and winters are generally like reflects our understanding of Saskatchewan’s climate. If we get a day in July with a daytime high of 10 °C, that would seem like unusually cold weather because we know it is uncharacteristic of the climate over all.

We characterize the climate by collecting data about the weather every day, and then calculating the average conditions over a period of decades. The Government of Canada provides averages for the periods 1961 to 1990, 1971 to 2000, and 1981 to 2010 in an online database that is searchable by geographic location or station. Data measured at Saskatoon’s Diefenbaker International Airport show that the average annual temperature from 1981 to 2010 is 0.6 °C higher than the annual average from 1961 to 1990, due warmer conditions in the winter and early spring (Figure 16.3).

Figure 16.3 Average temperatures for the periods 1961 to 1990, and 1981 to 2010 measured at Saskatoon’s Diefenbaker International Airport (YXE). Source: Karla Panchuk (2018) CC BY 4.0. View YXE data for 1961 to 1990. View YXE data for 1981 to 2010.

The climate as represented by the 1961 to 1990 interval was slightly cooler than the climate represented by the 1981 to 2010 interval. People who lived in Saskatoon between 1961 and 2010 may or may not have a sense that the weather they experienced from day to day was different for those intervals. In fact, some may have the record high of 35.3 °C on September 4, 1978 seared into their memory, and feel that Septembers just aren’t as hot as they used to be. They would be correct that as of 2017, there are no September temperatures recorded at the Diefenbaker International Airport weather station with a daytime high greater than 35.3 °C. But if that gave them the impression that Septembers are cooler on average today than in the past, that would not be consistent with the data.

Climate-Forcing Mechanisms

A climate-forcing mechanism is a process that causes climate to change. Climate forcings work by initiating changes in how heat energy moves into, through, and out of the Earth system. When we discuss a particular climate change event, the climate-forcing mechanism is what initiated the change. Feedbacks also alter climate, but we want to know what triggered the feedbacks in the first place.

Climate Forcing by Changes in Insolation

Insolation, or incoming solar radiation, refers to how much of the sun’s energy reaches Earth’s surface in a given period of time. Insolation is measured in Watts per square meter (W/m2).

Long-term Solar Evolution

Over the long term (billions of years), stars like our sun become larger, brighter, and hotter (Figure 16.4). Earth receives 40% more heat from the sun today than it did 4.5 billion years ago. In Figure 16.4, the blue Now arrow shows the sun’s current point in its life history. Although the blue arrow appears to indicate an instant in time, the time interval reflecting the duration of human existence on Earth is but a tiny fraction of the width of the line. As far as human experience is concerned, the long-term evolution of the sun is so slow that it has made no difference at all on insolation for the entire time humans have existed.

Figure 16.4 The life history of a star, from condensation of a nebula, to expansion to a red giant, and ending as a white dwarf. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Oliver Beatson (2009) Public Domain view source

Orbital Cycles

Insolation is also affected by cyclical changes in Earth’s orbit and rotation. Over intervals of approximately 100,000 years, the eccentricity of Earth’s orbit changes. Eccentricity is a measure of how elliptical a circle is. Higher eccentricity means that the orbit is more elliptical (Figure 16.5, left, blue orbit), whereas lower eccentricity means the orbit is more circular (Figure 16.5, left, red orbit). Eccentricity is important because when it is high, the Earth-sun distance varies more from season to season than it does when eccentricity is low.

 

Figure 16.5 Cycles in Earth’s orbit. Left: The shape of Earth’s orbit (its eccentricity) changes over 100,000 year cycles from more circular to more elliptical. Middle: Over 41,000 year periods, Earth’s axis of rotation nods toward and away from the sun. Right: Over 21,000 year cycles, Earth wobbles on its axis of rotation. Source: Karla Panchuk (2017) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 view source. Click the image for more attributions.

Over intervals of approximately 41,000 years, the obliquity of Earth’s axis of rotation changes (Figure 16.5, middle). This results in a nodding motion that alters how directly the sun shines on Earth’s poles. When the angle is at its maximum (24.5°), Earth’s seasonal differences are accentuated. When the angle is at its minimum (22.1°), seasonal differences are minimized.

Cycles of precession happen over intervals of approximately 20,000 years, causing Earth’s axis of rotation to wobble (Figure 16.5, right). This means that although the North Pole is presently pointing to the star Polaris (the pole star), in 10,000 years it will point to the star Vega.

The importance of eccentricity, tilt, and precession to Earth’s climate cycles (now known as Milanković Cycles) was first pointed out by Yugoslavian engineer and mathematician Milutin Milanković in the early 1900s. Milanković recognized that although the variations in the orbital cycles did not affect the total amount of insolation that Earth received, it did affect where on Earth that energy was strongest. Glaciations are most sensitive to the insolation received at latitudes of approximately 65°. As continents are configured today, this is most significant at 65° N, because there is almost no land at 65° S.

The most important factors are whether the northern hemisphere is pointing toward or away from the sun at its closest or farthest approach, and how eccentric the sun’s position is in Earth’s orbit. For example, if the northern hemisphere is at it farthest distance from the sun during summer (Figure 16.6, top), this means cooler summers. If the northern hemisphere is at its closest distance to the sun during summer (Figure 16.6, bottom), this means hotter summers. Cool summers — as opposed to cold winters — are the key factor in the accumulation of glacial ice, so the upper scenario in Figure 16.6 is the one that promotes glaciation. This factor is greatest when eccentricity is high.

 

Figure 16.6 Effect of precession on insolation in the northern hemisphere summers. In (a) the northern hemisphere summer takes place at greatest Earth-sun distance, so summers are cooler. In (b) (10,000 years or one-half precession cycle later) the opposite is the case, so summers are hotter. The red dashed line represents Earth’s path around the Sun. Source: Steven Earle (2015) CC BY 4.0 view source

The effects of all three cycles are evident in geochemical climate data. Figure 16.7 shows the “signals” for obliquity (A), eccentricity (B), and precession (C) over a period from 800,000 years in the past, to 800,000 years in the future. The vertical black line running down the middle of the diagram marks the present day. When the insolation from all three signals is determined, the result is a more complex waveform (D) with times of low variation in insolation, and times with higher variation in insolation.

Figure 16.7 Comparison of orbital cycles, insolation, and climate data for a 1.6 million year period. Source: Karla Panchuk (2017) CC BY-SA 4.0, modified after Incredio (2009) CC BY 3.0 view source. Click the image for more attributions.

The graphs E and F are climate information measured in microfossils dwelling at the ocean floor (E) and in water from ice cores (F). Peaks in temperature in F correspond to peaks in the oxygen isotope record in E, which indicate that it was warmer, there was less ice, and sea level was higher. Troughs are times when Earth was deep within an ice age. It was cooler, there was ice on land, and sea level was lower.

The vertical dashed lines on the left-hand side of Figure 16.7 mark the times of peak warm temperatures and allow for comparison of the timing of the temperature peaks with the timing of the orbital cycles. Peak temperature events are approximately 100,000 years apart, suggesting that the eccentricity cycle might be the most important contributor. Indeed, in B most (but not all) of the peak temperature events correspond to a time when Earth’s orbit was at or near peak eccentricity for that cycle.

It is tempting to conclude that eccentricity is the most important orbital cycle for climate change over all. However, this pattern only began a little over 1 million years ago. For 1.5 million years before that, the 41,000-year obliquity cycle seems to dominate insolation cycles.

In general, times of warmest or coolest temperatures don’t line up perfectly with orbital cycles. There is no one orbital cycle that is most important for all of Earth history. It is also the case that changes in insolation due to orbital cycles are not sufficient to cause temperatures to change as much as the geological record says they have; feedbacks must be factored in to explain the observed temperature changes.

Sunspot Cycles

Sunspots are dark patches that appear on the surface of the sun as a result of intense local disturbances in the sun’s magnetic field (Figure 16.8, left). Loops of plasma (gas with electrical charge, Figure 16.8, right) follow along magnetic field lines from one sunspot to another.

 

Figure 16.8 Sunspots. Left: Photograph of sunspots with dots representing the size of Earth and Jupiter for scale. Right: Plasma loops viewed in x-ray wavelengths jumping from one sunspot to another on the sun’s surface. Source: Left- NASA/Solar Dynamics Observatory (2012) Public Domain view source. Right- NASA/Solar Dynamics Observatory (2015) Public Domain view source.

Sunspots appear dark because they are lower-temperature regions on the sun’s surface. For that reason you might think that more sunspots means a reduction in insolation. In fact, just the opposite is true, because sunspots are a side-effect of increased solar activity. Peaks in the number of sunspots counted annually since approximately 1870 (Figure 16.9, blue), coincide with peaks in measurements of solar energy output from the same time period (Figure 16.9, pink).

Figure 16.9 Sunspot cycles. Peaks in the number of sunspots (blue) occur approximately every 11 years, and these correspond to peaks in solar energy output (pink). The influence of sunspot cycles is too small to have a clear impact on global average temperatures (grey). Source: Karla Panchuk (2017) CC BY-SA 4.0. Sunspot records modified after D. Bice (n.d.) CC BY-SA 3.0 view source. Global average temperature modified after Met Office (2015) Contains public sector information licensed under the Open Government Licence v1.0. view source

Sunspot cycles happen over approximately 11 year intervals, and the changes in insolation that occur during these cycles are relatively small. In the end the effect of sunspot cycles on climate can be lost amidst other factors. In Figure 16.9 there is no clear relationship between the sunspot cycles and the global average temperatures (in grey) reported for the same period.

Be Aware of Graph Scales

Figure 16.9 shows three kinds of data: temperatures, sunspot numbers, and solar energy flow. Each of these data sets is a different type of information, so each needs its own vertical axis. The vertical axes are scaled so that the data fill the area of the graph as much as possible. Stretching the vertical scale to fit the full plotting area makes it easier to see how well the peaks and troughs in each record line up with each other. Unfortunately, this can also skew our impression of the data. For example, in the period from 1880 to 1920, all three records have a similar vertical distance from peak to trough. In other words, all three records have approximately the same size of wiggles. This does not mean that the change in insolation from sunspot cycles was big enough to cause all of the variation in the temperature record. From this graph alone, there is no way to tell how much the change in insolation due to sunspot cycles mattered to global temperatures during the period 1880 to 1920.

 

Climate Forcing by Changes in Heat Transport

The ocean transports large amounts of heat around the Earth through a conveyor-belt-like system of currents. The ocean has surface currents that are driven by wind, but it also has deeper currents that are not wind-driven. The deeper currents behave like stacked rivers because they are different temperatures and have different salt contents, and therefore different densities. The differences in density between these water masses are what drive circulation. Circulation that is driven by density is called thermohaline circulation; thermo refers to heat and haline refers to salt.

To see how this works, consider the warm and saline Gulf Stream current (Figure 16.10, top). It flows northward past Britain and Iceland into the Norwegian Sea, and cools as it moves north, becoming denser. Its high salinity contributes to its density, and it sinks, or downwells, deep beneath the surrounding water, forming the North Atlantic Deep Water (NADW) current that flows south. Meanwhile, at the southern extreme of the Atlantic, very cold water adjacent to Antarctica also sinks to the bottom to become the Antarctic Bottom Water (AABW) current. The AABW flows north, beneath the NADW.

 

Figure 16.10 A simplified north-south cross-section through the Atlantic Ocean basin showing the different current layers. Source: Steven Earle (2015) CC BY 4.0 view source

 

The water that sinks in the areas of deep water formation in the Norwegian Sea and adjacent to Antarctica moves very slowly at depth. It eventually resurfaces, or upwells, in the Indian Ocean between Africa and India, and in the Pacific Ocean, north of the equator (Figure 16.11).

Figure 16.11 Global thermohaline circulation patterns. Red lines are surface currents, and blue lines are deep currents. Source: NASA Earth Observatory (2008) Public Domain view source

 

Some ocean currents move warm water from the equator toward the poles. As in the example of the Drake Passage, the path of warm currents can have a significant impact on the climate of a region, and potentially of the planet as a whole. Processes that disrupt the density of seawater can slow or stop currents, preventing warm water from reaching higher latitudes. The recovery from the last ice age is characterized by sudden returns to glacial conditions over as little as 3 years. This is thought to be the result of enormous glacial lakes forming on continents as the glaciers melted, then being suddenly released into the ocean by a burst ice dam. The glacial water would be very cold, but it would also be fresh, making it less dense than the ocean water. The fresh glacial water would form a cap and slow the downwelling conveyor belt at high latitudes.

Scientists are trying to determine the current and past state of the Atlantic-basin system of circulation, called the Atlantic Meridional Overturning Circulation (AMOC), to tell whether it is changing in response to warming and adding fresh water from melting ice sheets. The AMOC varies considerably on decadal cycles because of cycles in the wind patterns in the Atlantic, so it is important to distinguish these cyclical changes from any longer-term underlying changes.

Because of these studies, we have an idea of what the physical properties of the Atlantic Ocean look like when circulation is stronger or weaker. When circulation slows, the density is lower in the downwelling regions (Figure 16.12, blue patch in the Labrador Sea). The density is higher south of this region, along the eastern coast of the United States and southern Canada (Figure 16.12, orange patch). Model simulations are used to confirm that the changes in density we observe are consistent with how we understand the circulation system to work.

Measurements of the actual flow rate at depth in the Atlantic Ocean (Figure 16.12, right, blue dots) confirm that density decreases in downwelling regions (black and red lines) when circulation slows. A more recent study (Caesar et al., 2018) has shown that observed and model temperatures follow the same pattern, with cooling where the blue patches are in Figure 16.12, and warming in the region of the orange patches. This is to be expected because the slowdown in circulation affects how heat is moved northward.

Figure 16.12 Using changing density to track circulation in the Atlantic Ocean. Left- Density calculated from measurements of temperature and salinity in a layer between 1,000 m and 2,500 m depth in the Labrador Sea (black box). Middle- Model results used to see what change in density can be expected. The model shows the same general relationship, with lower density in the Labrador Sea, and higher density to the south. Note that density units are in kg/m2 rather than kg/m3 because they are integrated over the layer. Right- Changing density in the Labrador Sea over time. Red and black lines show changes in density from two different data sets. In general, the peaks and troughs of these data sets match up. Blue dots are measurements of the rate of circulation from a project that began collecting data in 2004. See the References section for more information about the relevant studies. Source: Karla Panchuk (2018) CC BY-SA 4.0. Modified after Jon Robson (2013) CC BY-SA 4.0 view source

Is Atlantic Circulation Slowing Down More than Usual?

Changes in the Atlantic meridional overturning circulation (AMOC) happen from decade to decade. To know whether circulation is changing compared to what is normal, it is necessary to get information about what circulation looked like in the past. This is difficult to do, because measurements of circulation rates, temperature, and salinity don’t go back as far as we need them to.

In a new study, Thornalley et al. (2018) have used geochemical analyses of microfossils to build a longer-term record of temperatures, then used that record to look for the temperature “fingerprint” of slowing AMOC, an increasing difference in the temperatures of surface waters compared to deeper waters in the downwelling zone (Figure 16.13, top). They observe a longer-term cooling trend beginning at the close of the Little Ice Age, suggesting that less heat is being moved toward the downwelling zone (labelled A on the globe in Figure 16.13).

Figure 16.13 Long-term record of Atlantic Meridional Overturning Circulation (AMOC). Top- Temperature fingerprint of circulation determined using geochemical analyses of marine microfossil shells. Results show the difference between temperatures measured at depth at A on the globe, and temperatures measured near the surface at B. Cooling at A relative to B is indicative of weakening AMOC. Bottom- Changes in silt grain-size used to show changes in the velocity of currents at depth at C on the globe. A smaller average grain size means a slower current, and weaker AMOC. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Thornalley et al. (2018). Locator globe modified after Reisio (n.d.) Public Domain view source

 

They also measured the size of silt grains on the sea floor, above which a southward-moving component of the AMOC system flows (location labelled C on the globe in Figure 16.13). Grain size is used as a substitute for a direct measure of ocean current velocity because the velocity determines what grain size can be carried. They identified a lower average grain size over the past ~150 years compared to the average grain size from earlier (Figure 16.13, bottom, dashed lines). The authors of the study note that the average grain size changes more during cold events in the Northern Hemisphere (the Dark Ages Cold Period and the Little Ice Age).

Thornalley et al (2018) conclude that the AMOC has been weaker on average during the past ~150 years than during the previous ~1,500 years. However, they cannot say for sure how much of that change is from melting that occurred at the close of the Little Ice Age, from melting triggered by warming since the Industrial Revolution, or some combination of the two. Direct measurements of density and current velocity tell us that as of 2017, the AMOC continues to weaken.

Plate Tectonics and Heat Transport

The opening of the Drake Passage is one example of how plate tectonic changes can affect ocean heat transport, and therefore climate. Plate tectonic changes that build or break up continents also play a role. When continents become large, ocean currents warm their margins, but the interiors can be much cooler. Anyone living on the Canadian prairies who has shivered through -40 °C temperatures in the winter, while watching news reports of rain in Vancouver will be familiar with this effect. When the supercontinent Gondwana was over the south pole approximately 300 million years ago (Figure 16.14), this triggered an ice age. The build-up of ice was hastened by the ice albedo feedback effect.

Figure 16.14 Glaciation on the supercontinent Gondwana. Paleogeographic reconstruction for 306 million years ago. Source: C. R. Scotese, PALEOMAP Project (www.scotese.com) view source. Click the image for terms of use.

Short-Term Cycles in Heat Transport: El Niño Southern Oscillation

The El Niño Southern Oscillation (ENSO) operates on a much shorter timescale than climate forcings driven by plate tectonics or orbital cycles, alternating between El Niño and La Niña events on timescales of between two and seven years (Figure 16.15).

Figure 16.15 Variations in the ENSO index from 1950 to 2015. Source: Steven Earle (2015) CC BY 4.0 view source. Modified after Klaus Wolter/ NOAA (n.d.) Public Domain view source

Under normal conditions, strong winds blowing westward across the Pacific cause water to pile up in the western Pacific. This forces deeper colder water to the surface in the eastern Pacific (Figure 16.16, left). During La Niña events, further intensification of winds causes even more cold water to upwell. During an El Niño event, the winds weaken, allowing water to flow back to the east (Figure 16.16, right). The cold water settles deeper once again, meaning that warmer water is present along the eastern margin of the Pacific Ocean.

Figure 16.16 El Niño Southern Oscillation (ENSO) cycles are driven by changes in wind patterns that affect the distribution of warm and cold water in the Pacific Ocean. Source: Karla Panchuk (2018) CC BY 4.0. Modified after Fred the Oyster and NOAA/PMEL/TAO Project Office. View Normal Conditions/ El Niño

ENSO events affect weather on a global scale (Figures 16.17 and 16.18). In western Canada, El Niño years have warmer than average winters, whereas La Niña years have cooler than average winters.

Figure 16.17 El Niño climate impacts. Source: NOAA Climate.gov (n.d.) Public Domain view source

 

Figure 16.18 La Niña climate impacts. Source: NOAA Climate.gov (n.d.) Public Domain view source

 

Climate Forcing by Changes in the Atmosphere’s Energy Budget

Earth’s atmosphere regulates climate by controlling how much energy from Earth’s surface escapes to space, and how much of the sun’s energy reaches Earth’s surface.

Albedo

Albedo is a measure of the reflectivity of a surface. Earth’s various surfaces have widely differing albedos, expressed as the percentage of light that reflects off a given material. This is important because most solar energy that hits a very reflective surface is not absorbed and therefore does little to warm Earth. Water in the oceans or on a lake is one of the darkest surfaces, reflecting less than 10% of the incident light. Clouds and snow or ice are among the brightest surfaces, reflecting 70% to 90% of the incident light (Figure 16.19).

Figure 16.19 Typical albedo values for Earth surfaces. Surfaces with low values reflect less light than surfaces with high values. Source: Steven Earle (2015) CC BY 4.0 view source

Albedo, Feedbacks, and the Acceptance of Milanković Cycles as a Climate Forcing Mechanism

When Milanković published his hypothesis in 1924, it was widely ignored, partly because it was evident to climate scientists that the forcing produced by the orbital variations alone was not strong enough to drive the climate changes of the glacial cycles. Those scientists did not recognize the power of positive feedbacks. It wasn’t until 1973, 15 years after Milanković’s death, that sufficiently high-resolution data were available to show that the Pleistocene glaciations were indeed driven by the orbital cycles, and it became evident that the orbital cycles were just the first step, initiating a range of feedback mechanisms that made the climate change, many of which were related to albedo.

Consider the following:

Exercise: Albedo Impacts of Vegetation Changes

Changes in climate can cause forests to be replaced by grasslands, which have higher albedo than dark forest cover. If deserts expand, vegetated areas can be replaced by higher-albedo sand. Many human activities affect albedo, including adding urban surfaces to an environment, and planting crops. Figure 16.20 shows a forest that has been clear-cut. If a clear-cut has an albedo similar to that of sand, how would clear cutting change the albedo of the area?

Figure 16.20 A clear-cut near Eugene, Oregon. Source: Calibas (2011) CC BY-SA 3.0 view source

Note that trees cool their environment through transpiration, when they release water vapour from their leaves. Changes in local temperatures when trees are clear-cut also include the effects of reduced evaporative cooling. Changes in vegetative cover also affect the rates of CO2 uptake by plants.

Greenhouse Gases (GHGs)

All molecules vibrate at various frequencies and in various ways, and some of those vibrations take place at frequencies within the range of the infrared radiation that is emitted by Earth’s surface. Gases with two atoms, such as O2, can only vibrate by stretching (back and forth; Figure 16.21 top), and those vibrations are much faster than that of IR radiation. Gases with three or more atoms (such as CO2) can vibrate in other ways, such as by bending (Figure 16.21 bottom). Those vibrations are slower and allow the molecules to absorb and release infrared radiation.

Figure 16.21 Molecules with two atoms (top) vibrate differently from molecules with more than two (bottom), and this determines whether a gas will be a greenhouse gas or not. Source: Steven Earle (2016) CC BY 4.0 view source

When infrared radiation interacts with CO2 or with one of the other GHGs, the molecular vibrations are enhanced because there is a match between the wavelength of the light and the vibrational frequency of the molecule. This makes the molecule vibrate more vigorously, heating the surrounding air in the process. These molecules also emit infrared radiation in all directions, some of which reaches Earth’s surface. Heating due to the vibrations of greenhouse gas molecules is called the greenhouse effect. Water molecules (H2O), and methane molecules (CH4) also interact with infrared radiation when they vibrate, so they are greenhouse gases as well.

Ice core records show that over the last 800,000 years, rapid cycles into and out of glacial temperatures are associated with similarly-timed cycles in atmospheric CO2 levels (Figure 16.21).

Figure 16.22 Variations in atmospheric CO2 levels and temperature over the last 800,000 years. Top- CO2 concentration from ice core data in Lüthi et al (2008). The dashed line shows a recent measurement of atmospheric CO2 levels from the Mauna Loa Observatory. Click to view the latest measurements. Bottom: Temperature record derived from oxygen isotope measurements of water in ice cores. Data from Jouzel et al (2008). Upper dashed line- global surface temperature for 2016 from NASA’s Goddard Institute for Space Studies Reference line. Click to view the most recent anomaly. Lower dashed line: average temperature for the past 1000 years. Source: Karla Panchuk (2018) CC BY 4.0, modified after National Research Council (2010) view source. Click the image for terms of use.

Earth-System Response Time

You might have noticed that for most of the 800,000-year record in Figure 16.22, there is a fairly consistent relationship between scale of change in atmospheric CO2 levels and the resulting change in temperature. You might also have noticed that compared to most of the record, the rise in temperature since 1950 is unexpectedly small, given the increase in atmospheric CO2 levels since that time. The reason for the relatively small temperature increase in response to the recent CO2 increase is in large part because the recent rise in CO2 is happening far more rapidly than other parts of the Earth system can respond. The ocean in particular is slowing down the response.

The ocean takes up heat from the atmosphere, and thus helps to determine surface temperatures. A relatively cool ocean can take up more heat from the atmosphere, reducing warming. The fastest way for the ocean as a whole to take up heat is through the “stirring” that happens with ocean circulation, but circulation happens on thousand-year timescales. The slow rate of circulation means that centuries from now, there will still be cool water rising up from the deep ocean that has yet to be exposed to the warmer surface conditions. At other times in the 800,000-year record, changes in CO2 levels happened on timescales much closer to those at which the ocean takes up heat.

Atmospheric Effects of Volcanic Eruptions

Volcanic eruptions don’t just involve lava flows and exploding rock fragments. Eruptions also release particles and gases into the atmosphere. Important volcanic gases include water vapour, CO2, and sulphur dioxide (SO2). Volcanic CO2 emissions can contribute to climate warming if a greater-than-average level of volcanism is sustained over a long time. At the end of the Permian Period, the massive Siberian Traps were produced by eruptions lasting at least a million years. Large quantities of CO2 were released, warming the climate and triggering a cascade of Earth-system responses. The end of the Permian Period at 252 Ma is marked by the greatest mass extinction in Earth history.

Over the shorter term, however, volcanic eruptions can have the opposite effect, cooling the climate. SO2 reacts with water in the atmosphere to make droplets of sulphuric acid. The sulphuric acid droplets scatter sunlight, reducing how much of the sun’s energy can reach Earth’s surface. They also affect cloud formation. The volcanic cooling effect is relatively short-lived, because the particles settle out of the atmosphere within a few years.

Exercise: Climate Change at the end of the Cretaceous Period

The large extraterrestrial impact at the end of the Cretaceous Period 66 Ma ago is thought to have produced a massive amount of dust, which may have remained in the atmosphere for several years. It may also have produced a great deal of CO2. What do you think would have been the short-term and longer-term climate-forcing implications of these two factors?

References

Caesar, I., Rahmstorf, S., Robinson, A., Feulner, G., & Saba, V. (2018). Observed fingerprint of a weakening Atlantic Ocean overturning circulation. Nature 556. 194-196. https://doi.org/10.1038/s41586-018-0006-5 View abstract

Ingleby, B., & Huddleston, M. (2007). Quality control of ocean temperature and salinity profiles—Historical and real-time data. Journal of Marine Systems, 65, 158-175. doi:10.1016/j.jmarsys.2005.11.019 Full text

Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S., Hoffmann, G., Minster, B., Nouet, J., Barnola, J.M., Chappellaz, J.A., Fischer, H., Gallet, J.C., Johnsen, S.J., Leuenberger, M., Loulergue, L., Luethi, D., Oerter, H., Parrenin, F., Raisbeck, G.M., Raynaud, D., Schilt, A., Schwander, J., Selmo, E., Souchez, R., Spahni, R., Stauffer, B., Steffensen, J.P., Stenni, B., Stocker, T.F., Tison, J.L., Werner, M., & Wolff, E.W. (2007). Orbital and Millennial Antarctic Climate Variability over the Past 800,000 Years. Science 317(5839), 793-797. Get data

Lüthi, D., Le Floch, M., Bereiter, B.; Blunier, T., Barnola, J.M., Siegenthaler, U., Raynaud, D., Jouzel, J., Fischer, H., Kawamura, K., & Stocker, T.F. (2008). High-resolution carbon dioxide concentration record 650,000-800,000 years before present. Nature (453), 379-382. Get data

Met Office Hadley Centre (n.d.) EN3: quality controlled subsurface ocean temperature and salinity data. Visit website (data available). Note: EN4 data set (new version) available here.

RAPID-AMOC (n.d.) RAPID: monitoring the Atlantic Meridional Overturning Circulation at 26.5°N since 2004. Visit website (data available)

Robson, J., Hodson, D., Hawkins, E., & Sutton, R. (2014). Atlantic overturning in decline? Nature Geoscience 7(2-3). https://doi.org/10.1038/ngeo2050 Open access version

Shaffrey, L.C., Stevens, I., Norton, W. A., Roberts, M. J., Vidale, P. L., Harle, J. D., Jrrar, A., Stevens, D. P., Woodage, M. J., Demory, M. E., Donners, J., Clark, D. B., Clayton, A., Cole, J. W., Wilson, S. S., Connelley, W. M., Davies, T. M., Iwi, A. M., Johns, T. C., King, J. D., New, A. L., Singlo, J. M., Slingo, A., Steenman-Clark, L., & Martin, G. M. (2009). U.K. HiGEM: The New U.K. High-Resolution Global Environment Model—Model Description and Basic Evaluation. Journal of Climate 22, 1861 – 1896. doi:10.1175/2008JCLI2508.1 Full text

Smith, D. M., & Murphy, J. M. (2007). An objective ocean temperature and salinity analysis using covariances from a global climate model. Journal of Geophysical Research, 112(C02022). doi:10.1029/2005JC003172

Thornalley, D. J. R., Oppo, D. W., Ortega, P., Robson, J. I., Brierley, C. M., Davis, R., Hall, I. R., Moffa-Sanchez, P., Rose, N. L., Spooner, P. T., Yashayaev, I., & Keigwin, L. D. (2018). Anomalously weak Labrador Sea convection and Atlantic overturning during the past 150 years. Nature 556, 227-230. https://doi.org/10.1038/s41586-018-0007-4 View abstract

109

16.3 Methods for Studying Past Climate

Whereas weather refers to day-to-day variations in temperature, precipitation, winds, and so on, climate refers to long-term trends in weather patterns (over decades or more). The term paleoclimate refers to Earth’s climate in the past. The information we have about Earth’s past climates can be classified as direct data or proxy data. Direct data are information derived from first-hand observations of climate. Direct data can be instrumental data, derived from tools designed to quantify observations, or from qualitative descriptions.

Proxy data are information derived from natural materials with characteristics that are affected by climate in a systematic way. This could also be said of some instrumental data: an alcohol thermometer uses the fact that the volume of alcohol changes in a consistent way in response to temperature. Proxy data rely on relationships that are also as systematic and consistent, but there are important differences:

Types of Direct Data

Instrumental records of climate are those derived from tools such as thermometers, rain gauges, or satellite measurements of the extent of ice sheets. Instrumental records are a recent development, as the history of the Earth system goes. The oldest known temperature measurements cover the period from 1654 to 1670, and were made by monks and Jesuit priests who operated stations within a meteorological network supported by the Medici family of Florence.

Non-instrumental historical records of climate also exist, and cover periods of human history prior to the development of the climate-measuring tools we have now. With detective work, these can be used to paint a detailed picture of past climates. Non-instrumental historical records include written records about how long ice and snow were present in a particular year, when harvests occurred, when floods happened, and shipping records that report the extent of sea ice. Paintings of alpine glaciers give information about how far the ice extended, and this can be used to reconstruct temperatures.

The Challenges of Getting Climate Information from Historical Records

In their paper Historical Climate Records in China and Reconstruction of Past Climates, Jiacheng Zhang and Thomas Crowley used official Chinese records extending as far back as 1000 CE to get a detailed picture of climate. This involved transforming descriptions of weather events into a systematic scale. One challenge is defining the scale, but another is deciding what individual accounts actually mean. The authors point out that records of rain or drought can reflect the perceptions and generalizations of the people who wrote about the weather, rather than what actually happened. Consider the following description of rainfall in 1644 from Diary of Qi Zongmin:

“in Wyzhong there is no rain for six months from May”

Does this mean there was no rain at all prior to May, or just very little? Was it actually six months since there had been rain, or is that an approximation? How do we reliably translate different systems of time measurement into durations and dates, like “six months” or “May?” How can we tell for sure where a location is if place names or political boundaries change?

As the authors caution, “a great deal of cross-checking [is required] in order to arrive at a useful descriptive account of climate anomalies.”

Sources of Proxy Data

Tree Rings

The study of tree growth rings for the purpose of understanding past states of the Earth system is called dendroclimatology. Temperatures and a history of drought or wet periods can be reconstructed from the widths of tree rings. Because tree rings form annually, these records can also be well constrained in time. The widths of tree rings reflect how fast the tree grows in a given season. There are factors other than temperature and moisture that affect growth rate, so a sampling strategy must be carefully designed to ensure confidence in climate reconstructions.

Stable Isotopes

Atoms have a nucleus made of protons and neutrons. The number of protons in the nucleus determines what element the atom is, and will always be the same for a given element. In contrast, the number of neutrons can vary for an element. Versions an element having different numbers of neutrons are the isotopes of that element.

Sometimes an atom has a number of neutrons that makes it unstable. Those atoms eventually break apart, releasing energy, and are called radioactive isotopes. The decay rate of radioactive isotopes is known, making it possible to use them to find the ages of natural materials. For example, carbon-14 dating makes use of the radioactive isotope of carbon, 14C, which has eight neutrons instead of the usual number, six (the 14 refers to 6 protons + 8 neutrons).

For investigating Earth’s past climate, stable isotopes, which do not decay, are used instead. Stable isotopes of the same element are measured in natural materials, and their ratios compared. Both isotopes are involved in the same chemical reactions and physical processes, but the slight difference in mass caused by one or two extra neutrons means that those processes are more likely to take up the lighter isotope than the heavier one. Some processes do this in such a particular way that evidence of their occurrence is left behind as a distinctive fingerprint in the stable isotope composition of materials formed in their environment.

The pair of isotopes used to reconstruct past temperatures are the oxygen isotopes 16O and 18O. The ratio of 18O to 16O in water is reflected in the calcium carbonate of shells that form in the water. The shells may remain in the geologic record long after the water is gone, making it possible to know the oxygen isotope compositions, and thus temperatures, of water bodies that existed in the distant past.

Ice Cores

The ice in polar glaciers and mountain glaciers preserves a detailed snapshot of Earth’s atmosphere and climate. A sample of Earth’s atmosphere, including gases and particles, is captured and held within the ice, and buried beneath subsequent ice layers. The annual layers in the ice can be used to determine a timescale for the data. The gases in air bubbles trapped within ice (Figure 16.23) are analyzed to determine the chemical composition of the atmosphere at the time the gasses were trapped.

Figure 16.23 A researcher holds a fragment of ice from Antarctica. The dots in the fragment are air bubbles containing samples of Earth’s past atmosphere. Source: Atmospheric Research, CSIRO (2000) CC BY 3.0 view source

Ice cores (Figure 16.24) are cylinders of ice retrieved using a specialized drill bit. The cores are carefully packaged and stored in specially designed facilities (Figure 16.25) until they are analyzed.

Figure 16.24 A scientist weighs and measures a cylinder of core from the West Antarctic Ice Sheet before she packages it for transport. Source: NASA/Lora Koenig (2010) CC BY 2.0 view source
Figure 16.25 National Science Foundation Ice Core Facility in Lakewood, Colorado. Ice cores are housed in tubes 1 m long. The main storage facility is kept at -36 ºC. Fortunately, scientists can examine the cores under much warmer conditions in a nearby room maintained at -24 ºC. Source: U. S. Geological Survey/ Eric Cravens (n.d.) Public Domain view source

Rock and Fossil Distributions

Earth can be divided into six main climate zones (Figure 16.26). The zones run roughly along lines of latitude, so that the climate zone changes as you move north or south of the equator. When the climate warms, the zones shift away from the equator; an area now in the boreal climate zone might have been in the warm temperate climate zone when Earth’s climate was warmer. When the climate cools, the zones shift toward the equator.

Figure 16.26 Six climate zones of the Köppen-Geiger classification. Source: Karla Panchuk (2018) CC BY-SA 4.0. Modified after LordToran (2007) CC BY-SA 3.0 view source

Some rock types are characteristic of particular climate zones. For example, coal deposits are characteristic of a subtropical climate. Limestone with coral reef fossils is characteristic of a tropical climate. If a rock type is found outside of its climate zone, that might indicate a change in climate. Coal can be found near Estevan, Saskatchewan, now in the warm temperate climate zone. This suggests that at one time, a warmer climate resulted in a northward shift of the subtropical climate zone. Some of the oil in western Canada is present in pore spaces within ancient coral reefs. The warm temperate climate zone cannot have been at its present location when those reefs formed.

Fossils can be used similarly. If an organism lives in a habitat with a particular climate, then evidence that the organism has migrated away from the equator could indicate warming. Migration toward the equator could indicate cooling.

The study of pollen and plant spores, called palynology, is very helpful for determining the distribution of plants when evidence of larger plant parts (e.g., fossil leaves and bark) is absent. Pollen and spores are very tough, and will survive in the environment when other plant materials do not. A detailed record of pollen and spores, and hence of the climate zones in a particular location, can be derived from lake sediments. Lakes in climates with strong seasonality (a distinct difference in temperature as seasons progress) can accumulate distinct annual sediment layers, called varves (Figure 16.27). Each year is represented by a light layer and a dark layer. The light layers consist of sand and silt from spring runoff. The darker layers include organic matter accumulated during the year.

Figure 16.27 Varves in a core from Canoe Brook, Drummerston, Vermont. Each pair of light and dark layers represents one year. The top of the core is to the right. Source: Karla Panchuk (2017) CC BY-NC-SA 4.0 (labels added). Modified after Jack Ridge/ North American Glacial Varve Project (2008) view source. Click the image for terms of use.

Varves can be counted to determine the age of the sediment, and the pollen and spores within the sediment can be extracted to see what types of vegetation were present at different times.

References

Ridge, J.C. (2008) The North American Glacial Varve Project. Retrieved from http://eos.tufts.edu/varves

Zhang, J., & Crowley, T. (1989). Historical Climate Records in China and Reconstruction of Past Climates. Journal of Climate 2(8), 833-849. https://doi.org/10.1175/1520-0442(1989)002<0833:HCRICA>2.0.CO;2 Full text

110

16.4 Computer Models of the Earth System

Earth-system interactions are so complex that it is next to impossible to follow all of the connections and implications without help. So, scientists use Earth-system computer models to assist. Earth-system models incorporate knowledge of the many components of the Earth system in a way that makes it possible to test how important any one change is.

Earth-system models vary in how many aspects of the Earth system they include, and how detailed their representations of those aspects are. Models are designed to answer particular kinds of questions so their performance can be optimized; a study that is concerned only with large-scale global changes might not require a model with a highly detailed representation of Earth’s coastlines. It takes more time to run a complicated model, so this saves on computing resources.

When computer models are discussed, we acknowledge that there is a difference between measurements of the real world, and the output from the model. Modelers are careful to refer to measurements of the real word as data, and output from the model as results. This also helps to avoid confusion when comparing models to real-world measurements to gauge how realistic the model output is.

What Are Computer Models, Exactly?

Computer models describe natural phenomena using mathematical equations. On the most basic level, computer models take some quantity—whether heat, water, or the concentration of a pollutant—and calculate how it moves through a system. Sometimes they look only at how that quantity changes through time. A computer model of the water volume in a bathtub could be limited to looking at how rapidly water flows in through the tap, and how rapidly it flows out through the drain. But sometimes models look at how a quantity changes in space as well as through time.

A study of wind-driven currents in a lake must include information about the shape and depth of the lake to capture how friction at the lake bottom and along the sides affects water flow (Figure 16.28). Data about the lake shape and depth (Figure 16.28, top) is translated to a model grid (Figure 16.28, bottom). Calculations are done to see how wind and friction control how water moves into and out of each cell in the grid.

Figure 16.28 Set-up for a model of wind-driven current flow in Lake Ontario. Top: Map of Lake Ontario showing water depth and the location of current meters. Bottom: Grid used to translate water depth information for model calculations. Source: Karla Panchuk (2002) CC BY 4.0. Based on the exercise described in Chapter 10 of Slingerland & Kump (2011).

The model produces information about wave height (Figure 16.29, left) and shows the direction and speed of water flow across the lake using arrows of different sizes (Figure 16.29, right). If scientists are interested in how a pollutant would move around the lake, they can include the location where the pollutant is added, and how rapidly it is added, and track how it moves.

Figure 16.29 Height of the water surface (left) and current velocity (right) from a model of wind-driven flow in Lake Ontario. The length of current arrows shows the speed of the current, and the arrow points in the direction of flow. Source: Karla Panchuk (2002) CC BY 4.0. Based on the exercise described in Chapter 10 of Slingerland & Kump (2011).

If the model is to be used to track the movement of a pollutant through the lake, it is important to know that it has done a good job of calculating the current velocity. In this case, data from current meters in the lake can be compared to the current velocities that the model calculates (Figure 16.30). The model captures the fact that flow is northward near the margins of the lake, and southward in the middle, but the model current velocities are not exactly the same. This means that the model would do a good job of predicting where the pollution went, but not as good a job at predicting how fast it got there.

Figure 16.30 Comparison of model results with current meter data. Points plotted beneath the blue dashed line indicate northward flow. Points above indicate southward flow. Source: Karla Panchuk (2002) CC BY 4.0. Data from Simons and Schertzer (1989).

The model in this case had a relatively course representation of the lake geography, so a first step would be to make the grid cells smaller to do a better job of simulating the shape of the lake, then look at the flow in finer detail. Another step would be to represent the lake water using a vertical stack of several grid cells to better capture the extent to which bottom friction and wind force affect the lake water at different depths, and to do a better job of representing the water depth.

An Example of Using a Computer Model to Study Past Earth-System Change

The Paleocene-Eocene Thermal Maximum (PETM) was a sudden global warming event that happened approximately 56 million years ago. There was interest in studying this event because its suddenness was thought to be a good analogy for the rapid changes happening in the Earth system today. Cores from ocean-floor rocks show that the oceans became so acidified during the PETM that calcium carbonate sediments dissolved over vast areas of the ocean floor, vanishing entirely from some regions. The same cores also showed a shift in the carbon-isotope composition of calcium carbonate sediments.

Both the acidification and the carbon-isotope shift indicated that a large amount of carbon was added to the Earth system to trigger the PETM. The problem was that there were a number of possible sources for the carbon, and thus a number of possible triggers for the event. Although scientists provided reasoned arguments for their favourite hypotheses, there was no way to know for sure which was the best answer.

To solve this problem, an Earth-system model was used that could test which scenario could best account for the pattern of dissolving calcium carbonate. It took into account the shape of ocean basins, ocean current circulation, and carbonate system chemistry in ocean water and in sediments. It also took into account changes in sediments once they were deposited.

The steps to using this model were the following:

  1. A search was done to locate as many studies of ocean floor sampling sites as possible that had information about changes in the amount of calcium carbonate during the PETM.
  2. The model was set up so that it did a good job of reproducing the distribution of calcium carbonate before the PETM happened. This was important to ensure that model scenarios began with a realistic set of conditions.
  3. Each of the possible scenarios involved carbon coming from different sources in the Earth system, meaning that each scenario could be represented in the model by adding to the atmosphere different amounts of carbon with different carbon isotope compositions. The more carbon a scenario required, the more the calcium carbonate sediments would have dissolved in real life.
  4. The model was run for each different amount of carbon. For each scenario, the pattern of calcium carbonate sediments that the model gave was compared to the actual distribution of calcium carbonate sediments known from the data collected in Step 1.

In the end, the model showed that some of the scenarios did not even come close to matching the observations, either dissolving way too much calcium carbonate, or far too little. The model showed that two scenarios did come close to reproducing the pattern of calcium carbonate, and that one did a better job of matching the observations than the other. When it came time to write a report about the experiments, the scientists learned that newly published measurements from another study supported the scenario that the model suggested was best. It would have been acceptable to write a paper describing the model results, and which scenario worked best. However, also being able to comment about new supporting data meant there was a better chance of convincing other scientists that the model results were meaningful.

Predicting the Future of the Earth System with Models

Using models to investigate the Earth system requires careful consideration of how to build the model and run experiments. But it also requires skillful use of real-life measurements to set up the model, and to interpret and evaluate its results. The PETM model study was an example of how a model can be used to test hypotheses about past behaviour of the Earth system. There were data from before, during, and after the event to help set up the model and gauge its effectiveness.

Using Earth-system models to predict the future is a different kind of modeling challenge, because we don’t already know what the right answer is. The situation being modeled hasn’t happened yet. Scientists who try to predict the future of the Earth system have to do things a bit differently in order to have some confidence in the reliability of their model outcomes:

References

Simons, T. J., & Schertzer, W. M. (1989) The circulation of Lake Ontario during the summer of 1982 and the winter of 1982/83. Burlington, ON: Environment Canada. Full text

Slingerland, R., & Kump, L. (2011). Mathematical Modeling of Earth’s Dynamical Systems: A Primer. Princeton, NJ: Princeton University Press.

111

16.5 Humans in the Earth System

The Start of Human Influence on the Earth System

Anthropogenic change in the Earth system is change caused by humans. Many discussions of anthropogenic climate-change place the start of human impacts on the Earth system at the beginning of the industrial era, in the mid 18th century. The industrial era was when humans began to use fossil fuels—at the time, mostly coal—on a much larger scale than before to do things like run manufacturing machinery and trains.

Some climate scientists place the first anthropogenic impacts much earlier, however. Some suggest that anthropogenic climate change began around 8,000 BCE when humans cleared land for agriculture in Europe and the Middle East. Clearing forests for crops is a type of climate forcing because the CO2 storage capacity of the crops is generally lower than that of the trees they replace. Some climate scientists also point to the creation of wetlands to grow rice in Asia around 5,000 BCE. Creating wetlands is a type of climate forcing because the anaerobic bacterial decay of organic matter within wetlands produces CH4.

Whether anthropogenic climate change began with the Agricultural Revolution or the Industrial Revolution may be a matter for debate for some, but it is clear that Earth-system change accelerated once the Industrial Revolution began. Part of this is due to the fact that agricultural activities had to be scaled up to feed an ever-growing population. When humans first started growing crops, the world population was approximately 5 million (Figure 16.31), fewer people than live in Toronto today. The world population rose to approximately 18 million when wetland rice cultivation began (fewer people than live within the city limits of Beijing today), to over 800 million at the start of the Industrial Revolution. The world population was estimated at 7,600 million in 2018.

Figure 16.31 World population growth over the past 12,000 years. Source: Steven Earle (2015) CC BY 4.0, view source. Data from Roser and Ortiz-Ospina (2018) view source/ view data file

The other reason humans accelerated Earth-system change after the start of the industrial era is that human activities required a source of energy, and fossil fuels such as coal and oil were that source. Fossil fuels are those derived largely from plant material that grew, died, and was partially preserved at various times throughout Earth history. The plants removed CO2 from the atmosphere when they were alive, and stored it in organic compounds in their tissues. The materials accumulated over hundreds of millions of years in settings like swampy forests, shallow seas, and deltas. When fossil fuels are burned, the stored carbon is released back into the atmosphere as CO2.

The Carbon-Isotope Fingerprints of Fossil Fuel

Carbon isotopes provide insights into the extent to which fossil fuels have impacted the Earth system, because fossil fuels have a unique carbon-isotope fingerprint that is detectable in the atmosphere and in geological materials.

Stable Carbon Isotopes (12-Carbon and 13-Carbon)

When plants transform CO2 into tissues, the process imparts a unique carbon-isotope signature to the resulting organic matter. Plants preferentially take in CO2 with the isotope 12C over  CO2 with isotope 13C. They do so in a consistent way, giving plant tissues a distinctive ratio of 13C to 12C. Fossil fuels are derived from plant materials, and they preserve this isotopic ratio.

The ratio of 13C to 12C is commonly expressed relative to a standard to give numbers that are easy to work with and compare. The notation δ13C refers to the ratio of 13C to 12C in a sample compared to the ratio in a standard, and is expressed in parts per thousand (or per mil, ‰). The standard has a δ13C of 0‰. Carbon in plant tissues has a δ13C of -25‰ to -30‰, meaning it has a 13C to 12C ratio that is 25 to 30 parts per thousand lower than the standard. Burning fossil fuel releases CO2 with that ratio into the atmosphere.

For most of the past 1000 years, the atmosphere has had a δ13C of approximately -6.5‰. The carbon-isotope composition of organic matter is much lower than that of the atmosphere, so the mixing in of carbon from fossil fuels causes the over-all carbon-isotope composition of the atmosphere to decrease. An analogy for mixing low δ13C CO2 into the atmosphere is rapidly adding cold water to a hot bathtub. The faster the cold water is added, the faster the bathwater will cool. The colder the water being added, the faster the bathwater will cool. In this analogy, the atmosphere is the bathtub, and fossil fuels are the water being added. The low δ13C value of fossil fuels (-25‰ to -30‰) is like very cold water being added.

As we would expect, the carbon isotope composition of the atmosphere takes a sudden downward turn at the same time that humans undertake the Industrial Revolution, and begin burning large quantities of fossil fuels, adding CO2 to the atmosphere at an accelerating rate (Figure 16.32).

Figure 16.32 A 1000-year record of atmospheric CO2 levels (blue circles) and carbon isotope composition (grey circles) measured in Antarctic ice cores. The Industrial Revolution (grey shading), marking the start of the industrial era and the large-scale use of fossil fuels by humans, coincides with a sudden rise in CO2 levels, and a fall in the carbon-isotope composition of atmospheric CO2. Source: Karla Panchuk (2018) CC BY 4.0. Data from Rubino et al (2013).

Scientists who study past climates on Earth are familiar with carbon-isotope records like this one, because such records are used to reconstruct major changes in the Earth system through their impact on the carbon cycle. In carbon-isotope records from the distant past, a shift of more than 1.5‰ would be enough to catch the attention of a researcher and make them wonder what could have happened.

What is unusual about the 1.5‰ drop today in comparison to those observed in the geological record is how rapidly it is happening. It is more common to see such changes happen over millions of years, not hundreds of years. The rate at which atmospheric CO2 δ13C is dropping is approximately 10 times faster than the carbon-isotope shift at the PETM, which is the fastest event ever documented in the rock record.

Carbon dioxide in the atmosphere mixes into the oceans, where organisms take up carbonate ions to make calcium carbonate shells. The 1.5‰ drop has been imprinted in the calcium carbonate of marine organisms like sponges (Böhm et al, 2002), and will remain in the rock record globally, as evidence of human activity. Because of this, and because of many other such markers that are being left in the rock record by human activities (the presence of plastic, for example), some have suggested that it is time to define a new division of geological time, the Anthropocene Epoch. The start of the Anthropocene Epoch would mark the point at which human activities became evident in the geological record.

Radioactive Carbon (14-Carbon)

Carbon-14 dating relies on the fact that 14C decays to 14N at a known rate. By knowing the rate, and how much 14C and 14N are present, we can work out how long the decay has been happening. Knowledge of the decay rate of 14C also makes it useful to track fossil fuel additions to the atmosphere.

The rate of decay of a radioactive isotope is expressed as a half-life, which in this case is the amount of time it would take half of the 14C atoms in a sample to decay to 14N. The half-life of 14C is 5,730 years. After 10 half-lives, or 57,300 years, there isn’t enough 14C left to do an age measurement. Fossil fuels are millions to hundreds of millions of years old, long enough for there to be none of the 14C originally contained by the plant material.

There is a notation system for 14C similar to the δ13C notation system for the ratio of 13C to 12C, in which the amount of 14C is compared to a standard. Carbon-14 amounts are reported as Δ14C values in units of ‰. In that system, the atmosphere as a whole had a Δ14C of 45‰ in 2010, and fossil fuels have a Δ14C of -1000‰. Effectively, the atmosphere appears to be aging rapidly. In the bathtub analogy for carbon isotopes, adding CO2 from fossil fuels is like dumping ice into the tub.

The effects of fossil fuel CO2 on atmospheric Δ14C levels must account for 14C being made through natural processes in the atmosphere, and decaying away; for the decay of a large pulse of 14C created by nuclear bomb tests; and for other sources of carbon with very low Δ14C values. Fortunately for scientists tracking fossil fuels by their impact on atmospheric Δ14C, the contribution of low Δ14C CO2 from other sources is tiny compared to known rates of fossil fuel emissions, and the other quantities are also well known. Thus, they have been able to determine a decrease in Δ14C of 3‰ for every 1 ppm of CO2 added from fossil fuels.

The Carbon Cycle and Change in Today’s Earth System

Change in the Earth system is strongly driven by Earth’s carbon cycle, the interrelated materials and processes that change carbon from one form to another, and move it from one reservoir to another (Figure 16.33). The CO2 in the atmosphere is just one part of the carbon cycle. Carbon in the atmosphere is taken in by marine and terrestrial plants, and released when they are decomposed. Microbial activities in the soil and respiration by plants release carbon. Carbon also moves into and out of the ocean through exchange processes at the ocean’s surface.

Figure 16.33 Flows of carbon in the Earth system. Numbers are rates in billions of tons of carbon (gigaton, Gt) per year. Yellow numbers are rates unrelated to human activity. Red numbers show the contribution of human activities as of 2012. Source: U.S. Department of Energy (2012) Public Domain view source

In the carbon cycle today, natural processes as a whole comprise far more of the flow in the carbon cycle than human activities do. For comparison, the relative sizes of flows in Figure 16.33 are illustrated by the size of the arrows. As of 2012, human activities were responsible for approximately 9 billion tons (9 Gt) of carbon added to the atmosphere per year. A large part of the 9 Gt comes from burning oil, coal, and gas, and some from changes in how land is used (e.g., clearing forests to plant crops, Figure 16.34). Some comes from changes that humans have made that affect the ability of the Earth system to take up carbon.

Figure 16.34 Flow diagram illustrating the pathways through which human activities produce greenhouse gases. The diagram connects the items in each column with flows that ultimately lead to the type of fuel used, and the greenhouse gasses produced. The width of each band is proportional to the quantity flowing from one column to the next. Note that F-Gas refers to anthropogenic fluorinated gases, which are extremely powerful greenhouse gases. Source: Fischedick et al. (2014), Figure 10.1, based on Bajželj et al. (2013). View source (p. 745). Click the image for terms of use.

The Earth system has accommodated the 9 billion tons by taking up an additional 3 billion tons per year in photosynthesis, and dissolving an additional 2 billion tons per year in the ocean. The remaining 4 billion tons accumulates in the atmosphere each year because the Earth system does not presently have the capacity to remove it.

The fossil fuels added by humans are particularly problematic because burning them means releasing hundreds of millions of years worth of plant-stored carbon that would otherwise not have been an active part of the carbon cycle today. Contrast this with cutting down a tree and burning the wood. Burning the wood also releases CO2 from carbon that was stored in plant tissues, but the difference is in timescale and quantity. If a tree grows for 50 years before it is used as fuel, then over a century there is effectively no change in atmospheric CO2. What carbon the tree took out of the atmosphere decades before, burning and decomposition have returned.

For fossil fuels, on the other hand, the carbon was removed from the atmosphere tens or even hundreds of millions of years ago. Trees draw down CO2 before we burn them, balancing out the equation, but with fossil fuels there is no initial draw-down from our present atmosphere. Releasing the carbon stored in those fuels results in a net addition to the atmosphere. What makes this even worse is that because fossil fuels have been accumulating for so long, there is an enormous quantity that can be burned. Trees can only be burned as fast as they replace themselves, but with fossil fuels it is like accumulating trees for millions of years, then burning them all at once.

Signals of Present-Day Earth-System Change

Rising Temperatures

From studies of Earth’s past climate history, it is clear what to expect as atmospheric CO2 levels rise. Climate warming is one outcome. We know from ice core records that global average temperatures are warmer now than they have been for most of the last 800,000 years (Figure 16.22). Over the shorter term, direct measurements show that the climate has been on a warming trend after the start of the Industrial Revolution (Figure 16.9). Proxy data making up a revised version of the “hockey stick” diagram—so named because the shape reminded some people of a hockey stick laying on its side—take the record back to 1000 years ago, and show global average temperatures falling until the onset of the industrial era (Figure 16.35).

Figure 16.35 Global average temperature change for the last 1000 years. Blue- The original “hockey stick” diagram showing a reconstruction of northern hemisphere temperatures using tree rings as a proxy. Red- Direct temperature measurements. Green dots- Global temperature reconstruction using a wide range of direct measurements, historical records, and proxies (sediments, ice cores, tree rings, corals, stalagmites, pollen). The original hockey stick diagram was the focus of much controversy because it was the first evidence of anthropogenic climate change that could be understood by the general public. The PAGES2K project sought to bring vast quantities of data to establish once and for all whether a global signal of warming could be reliably discerned. The result was very similar to the original hockey stick. Source: Karla Panchuk (2018) CC BY-SA 4.0. Modified after Klaus Bittermann (2013) CC BY-SA 4.0 view source. Learn more about PAGES2K and find data.

Sea Level Change

As of April 2018, global sea level has risen approximately 28 cm since 1800. According to satellite data, the average rate of change since 1993 has been a rise of approximately 3 mm per year. Part of the rise is due to the expansion of seawater as it warms. Another part of the rise is from water added by melting glaciers and other year-round land-based snow and ice. Note that melting of sea ice—ice already floating in the ocean—does not contribute directly to sea-level rise because the ice is already floating in the ocean.

Based on how much melting has occurred thus far, sea levels are projected to rise to between 47 cm and 130 cm above 1880 levels (Figure 16.36). However, there is some uncertainty about how melting rates will respond to changes in the Earth system that result from climate change, such as changes in currents, or seawater beneath the leading edge of melting ice sheets warming the ice from beneath. With that uncertainty factored in, sea level rise could be as low as 33 cm above 1880 levels, or more than 2 m higher.

Figure 16.36 Measured and projected change in global average sea level. Data come from proxy records as well as from direct measurements from tidal gauges and satellite data. Projected sea level rise could be as little as 33 cm over 1800 levels, or as much as 206 cm. Source: Karla Panchuk (2018) modified after Steven Earle (2015) CC BY 4.0 view source and J. Willis, Jet Propulsion Laboratory (2013). View source and more information about this figure. Click the image for terms of use.

Keep in mind that the global average is indeed an average. Where ocean waters experience more warming, and thus more thermal expansion, sea level rise may be greater than elsewhere. Regions that are rebounding as ice melts could experience less sea level rise, or even a fall in sea level, because the elevation of the terrain is actually increasing over time. On the other hand, regions on the peripheral bulge around the margins of ice sheets could experience greater than average sea level rise because the terrain will subside at the same time that the oceans are gaining volume.

Areas that become flooded could experience greater than average sea level rise, because the weight of water causes the land to subside further. In the aftermath of Hurricane Harvey in September of 2017, measurements were reported that showed subsidence of up to 1.5 cm in the region of Houston, Texas. In this case, some of the subsidence could have been from sediments being compressed under the weight of flood waters, however the weight of water, like the weight of ice, does cause the crust to float lower in the mantle.

Melting Ice Sheets

Keeping track of how rapidly ice sheets are melting is important both for being able to predict future sea level change, and for knowing in general how rapidly the Earth system is changing. In a recent study, Bamber et al. (2018) analyzed satellite measurements to determine how much mass had been lost from the Antarctic ice sheets, the Greenland Ice Sheet, and from other glaciers and ice caps around the world since 1992 (Figure 16.37).

Figure 16.37 Large ice sheets of Antarctica and Greenland (blue) and glaciers and ice caps (yellow). Circles are proportional to the area of each region that is covered by glaciers. The green part of the circle indicates the proportion of the ice with margins resting on land, and the blue part indicates margins in the ocean. This difference is important in part because of the potential for faster melting when the base of an ice sheet is in contact with warming seawater. Source: Bamber et al. (2018) CC BY 4.0. View source (see Fig. 1)

The study found that over all, the mass of ice in ice sheets, ice caps, and glaciers has been falling at an increasing rate since 1992, and therefore adding to sea level at an increasing rate (Figure 16.38). The exception is the East Antarctic Ice Sheet, which actually showed an increase in mass during the studied interval. This is because snowfall has increased in the East Antarctic, to the point where more snow is falling now than at any time in the past 2000 years (Medley et al., 2017). The East Antarctic is warming just as the West is, but the difference is that the winds that preferentially bring precipitation to the East rather than the West can carry more moisture because the air is warmer.

Figure 16.38 Results of a study of the change in mass of ice on Earth’s surface. Satellite data show that over all, melting has accelerated since 1992. Source: Karla Panchuk (2018) CC BY 4.0, modified after Bamber et al. (2018) CC BY 4.0. View source (see Fig. 11)

 

References

Bajželj, B., Allwood, J. M., and Cullen, J. M. (2013). Designing Climate Change Mitigation Plans That Add Up. Environmental Science & Technology 47, 8062-8069. doi: 10.1021/es400399h. Retrieved from http://pubs.acs.org/doi/pdf/10.1021/es400399h

Bamber, J. L., Westaway, R. M., Marzeion, B., & Wouters, B. (2018). The land ice contribution to sea level during the satellite era.  Environmental Research Letters 13(2018). https://doi.org/10.1088/1748-9326/aac2f0 Full text

Böhm, F., Haase-Schramm, A., Eisenhauer, A., Dullo, W.-C., Joachimski, M. M., Lehnert, H., & Reitner, J. (2002). Evidence for preindustrial variations in the marine surface water carbonate system from coralline sponges. Geochem. Geophys. Geosyst., 3(3), 10.1029/2001GC000264. Retrieved from http://onlinelibrary.wiley.com/doi/10.1029/2001GC000264/epdf

Earth System Research Laboratory, Global Monitoring Division, NOAA (n.d.). The Data: What 14C Tells Us. Visit website

Fischedick M., Roy, J., Abdel-Aziz, A., Acquaye, A., Allwood, J. M., Ceron, J.-P., Y. Geng, Y., Kheshgi, H., Lanza, A., Perczyk, D., Price, L., Santalla, E., Sheinbaum, C., and Tanaka, K. (2014). Industry. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Edenhofer, O., Pichs-Madruga, R., Sokona, Y., Farahani, E., Kadner, S., Seyboth, K., Adler, A., Baum, I., Brunner, S., Eickemeier, P., Kriemann, B., Savolainen, J., Schlömer, S., von Stechow, C., Zwickel, T., and Minx, J. C. (eds.). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Retrieved from https://www.ipcc.ch/pdf/assessment-report/ar5/wg3/ipcc_wg3_ar5_chapter10.pdf 

Madrigal, A. C. (2017, September 5) The Houston Flooding Pushed the Earth’s Crust Down 2 Centimeters. View at The Atlantic.

Medley, B., McConnell, J. R., Neumann, T. A., Reijmer, C. H., Chellman, N., Sigl, M., & Kipfstuhl, S. (2018). Temperature and Snowfall in Western Queen Maud Land Increasing Faster Than Climate Model Projections. Geophysical Research Letters 45(3), 1472-1480. https://doi.org/10.1002/2017GL075992

Rahmstorf, S. (2013). Most Comprehensive Paleoclimate Reconstruction Confirms Hockey Stick. View at ThinkProgress.org

Roser, M., & Ortiz-Ospina E. (2018). World Population Growth. View at OurWorldInData.com

112

16.6 Welcome to the Anthropocene

Geologists still have not agreed on an official start date for the Anthropocene Epoch, but it is unlikely that they would disagree that we are presently in it. What this means in practical terms is that humans today are experiencing the results of past human influence on the Earth system, and humans in the future will experience the results of decisions made today.

Data on Key Sources of Radiative Forcing

The Intergovernmental Panel on Climate Change (IPCC) was established by the United Nations in 1988 to help with those decisions. Is responsible for reviewing the scientific literature on climate change and issuing periodic reports on several topics, including the scientific basis for understanding climate change, our vulnerability to observed and predicted climate changes, and what we can do to limit climate change and minimize its impacts.

In their Fifth Assessment Report, Summary for Policy Makers, the IPCC identified the main contributions to heating or cooling our atmosphere, and the impact they had in 2011 compared to 1750, before the industrial age (Figure 16.39). The main natural source of radiative forcing is the sun, but it contributes relatively little compared to anthropogenic sources. The greatest contribution over all comes from anthropogenic CO2 (top row). Note that the bar sizes indicate the relative contribution, and correspond to the scale at the bottom of the diagram. Bars to the left of the zero line indicate cooling of the atmosphere, and bars to the right indicate warming. The column on the far right indicates how confident scientists are in their assessment of radiative forcing for each source (VH = very high; H = high; M = medium; L = low).

Figure 16.39 Anthropogenic and natural contributions to radiative forcing in 2011 compared to 1750. Source: IPCC (2013) View source (SPM.5, p. 14). Click the image for terms of use.

The biggest anthropogenic contributor to warming, CO2, accounts for 50% of positive forcing. CH4 and its atmospheric derivatives (CO2, H2O, and O3) account for 29%, and the halocarbon gases (mostly leaked from air-conditioning appliances) and nitrous oxide (N2O, from burning fossils fuels) account for 5% each. Carbon monoxide (CO, also produced by burning fossil fuels) accounts for 7%, and the volatile organic compounds other than methane (NMVOC) account for 3%.

CO2 emissions come mostly from coal- and gas-fired power stations, motorized vehicles (cars, trucks, and aircraft), and industrial operations (e.g., smelting), and indirectly from forestry. CH4 emissions come from production of fossil fuels (escape from coal mining and from gas and oil production), livestock farming (mostly beef), landfills, and wetland rice farming. N2O and CO come mostly from the combustion of fossil fuels. In summary, close to 70% of current anthropogenic GHG emissions come from fossil fuel production and use, while most of the rest comes from agriculture and landfills.

Projected Warming

The IPCC projects future warming by using scenarios that represent different possible paths that take into account global population, economic development, and global co-operation. In their best-case scenario, called RCP2.6 (RCP stands for representative concentration pathway, and 2.6 refers to a radiative forcing of 2.6 W/m2), global CO2 emissions begin to decline after 2020, and go to zero by 2080. Their worst-case scenario is RCP8.5, in which global CO2 emissions continue to rise rapidly. Intermediate scenarios reflect cases where global CO2 emissions eventually peak, but do not go to zero.

Figure 16.40 Model projections to 2100 for surface temperatures (a) and the extent of sea ice in September in the Northern Hemisphere (b). Black line: data; Grey shading: model attempts to simulate conditions from 1950 to 2005; Blue: best-case scenario with peak CO2 emissions in 2020 and zero emissions by 2080; Red: worst-case scenario with no decline in emissions. Numbers indicate the number of models run for each scenario. Shading indicates range of uncertainty. Source: IPCC (2013) View source (Fig. SPM.7, p. 14). Click the image for terms of use.

In the best-case scenario, temperatures stabilize at approximately 1 ºC above 1996 temperatures by the end of the century. As of 2018 there is no global plan in place that will meet the timeline of RCP2.6, and cause emissions to decline after 2020. For sea ice, all but the best-case scenario have the average sea ice extent beneath the dashed line, which effectively means the northern hemisphere would be ice-free.

Extreme Climate-System Events

In 2015 the World Meteorological Organization (WMO) published an analysis of extreme climate-system events, and showed that the number of such events globally was almost 5 times higher in the decade 2001-2010 than in the decade 1971-1980 (Figure 16.41).

Figure 16.41 Numbers of climate-system disasters between 1971 and 2010. Source: WMO (2015) Atlas of Mortality and Economic Losses from Weather, Climate and Water Extremes: 1970-2012. View source (p. 9). Click the image for terms of use.

Analyses of natural disasters and extreme weather events by the European Academies’ Science Advisory Council (EASAC) showed that from 1980 to 2010, extreme weather, floods, and mass movement related to floods increased by more than three-fold in Europe. In the same study, analyses of insurance-industry data showed that globally, Europe and South America are the least affected. The number of events has been more than twice as great in Asia, Australia/Oceania, and North America (EASAC, 2013). The EASAC recently published an updat to their study, to include data up to 2016. They now say that floods and mass movement events have increased in Europe by more than 4 times from 1980 levels (EASAC, 2018).

Extreme Temperatures

The major types of disasters reported by the WMO report are related to climate are floods and storms, but the health implications of extreme temperatures are also becoming a great concern. In the decade 1971 to 1980, extreme temperatures were the fifth most common natural disasters; by 2001 to 2010, they were the third most common.

For several weeks in July and August of 2010, a massive heat wave affected western Russia, especially the area southeast of Moscow. Temperatures soared to over 40°C, as much as 12°C above normal over a wide area, and wildfires raged in many parts of the country. Over 55,000 deaths are attributed to the heat and to respiratory problems associated with the fires (Figure 16.42).

Figure 16.42 Temperature anomalies across Russia and neighbouring regions during July 2010. Source: Earth Observatory (2010) Public Domain view source

A clue to explaining the heatwave was simultaneous flooding in Pakistan: in July, rainfalls were 70% higher than normal, and in August rainfalls were 102% above normal. Flooding from heavy rains combined with a burst dam impacted 18 million people, killing 1,985. These events were linked by a change in atmospheric circulation that caused weather patterns to stall out over Russia and Pakistan. The change itself was due in part to a decrease in the temperature differences between high and low latitudes, caused by warming in the north.

Sea-Level Rise and Tropical Storm Impacts

Projections for sea-level rise to the end of this century vary widely (Figure 16.36), however it is already the case that low-lying coastal regions are experiencing increased flooding during storm events. Flood defenses designed decades ago are or will soon be insufficient in the face of increasingly frequent and more extreme storm events. In a 2008 report, the Organisation for Economic Co-operation and Development (OECD) estimated that by 2070 approximately 150 million people living in coastal areas could be at risk of flooding due to the combined effects of sea-level rise, increased storm intensity, and land subsidence. The assets at risk (buildings, roads, bridges, ports, etc.) are on the order of $35 trillion ($35,000,000,000,000). Countries with the greatest exposure of population to flooding are China, India, Bangladesh, Vietnam, U.S.A., Japan, and Thailand. Some of the major cities at risk include Shanghai, Guangzhou, Mumbai, Kolkata, Dhaka, Ho Chi Minh City, Tokyo, Miami, and New York.

One of the other risks for coastal populations, besides sea-level rise, is that climate warming is also associated with an increase in the intensity of tropical storms (e.g., hurricanes or typhoons), which almost always bring serious flooding from intense rain and storm surges. It is difficult for climatologists to ascribe all of the characteristics of a particular tropical storm to climate change, but it is much easier to make a case for why tropical storms in general would become stronger, and have more of an impact on coastal regions.

When Hurricane Sandy hit New Jersey and New York in 2012 (Figure 16.43), it had a devastating impact in part because it smashed into the coast directly rather than moving parallel to the coast. There was some speculation that the unusual path was the result of changes in atmospheric circulation that were related to climate change. We can’t say for certain whether climate change caused Sandy to move the way it did, but we can say with confidence that climate change worsened the impact of Sandy through sea-level rise. Consider that, over the previous 100 years, sea level had risen a foot at the southern-most tip of Manhattan Island. The island was hit by a 14-foot storm-surge, higher than it would have been without sea-level rise.

Figure 16.43 Damage from Hurricane Sandy on the coast of New Jersey. Source: Master Sgt. Mark C. Olsen, New Jersey National Guard (2010) Public Domain view source

Tropical Storm Intensity

Tropical storms get their energy from the evaporation of warm seawater in tropical regions. In the Atlantic Ocean, this takes place between 8° and 20° N in the summer. A correlation has been observed between variations in the sea-surface temperature (SST) of the tropical Atlantic Ocean (Figure 16.44, red) and the amount of power represented by Atlantic hurricanes between 1950 and 2008 (blue). Not only has the overall intensity of Atlantic hurricanes increased with the warming since 1975, but the correlation between hurricanes and sea-surface temperatures is very strong over that time period.

Figure 16.44 Relationship between Atlantic tropical storm cumulative annual intensity and Atlantic sea-surface temperatures. Source: Steven Earle (2015) CC BY 4.0 view source/ view data source 

Because warm air is able to hold more water than cold air, the general global trend over the past century has been one of increasing precipitation (Figure 16.45). Note that this does not mean that precipitation has increased everywhere in the world. It means that even taking into account regions that have become drier, there is still more precipitation over all.

Figure 16.45 Global precipitation from 1901 to 2000. Source: Karla Panchuk (2018) CC BY 4.0, with data from NASA/GISS. Get data

Thawing Permafrost

Most of northern Canada has a layer of permafrost that ranges from a few centimetres to hundreds of metres in thickness. The same is true in Alaska, Russia, and Scandinavia. One problem that northern communities are currently dealing with is that thawing permafrost is weakening sediments that were previously solid and stable. This can cause structures to settle, or even to sustain serious damage. Another problem with the weakening that occurs when permafrost thaws is that mass wasting can occur (Figure 16.46). This is particularly dangerous for northern communities along coastlines, where homes may be near the shore and the slopes beneath them further weakened by strong wave action during severe storms.

Figure 16.46 A degrading permafrost site on the north coast of Alaska. Source: Alaska Science Center, U. S. Geological Survey (2016) Public Domain view source

A longer-term issue with thawing permafrost is that micro-organisms begin to decompose organic matter within the newly thawed sediments, releasing CO2 and CH4. The amount of these greenhouse gases released by decomposing organic matter could be enough to generate a significant positive feedback, accelerating warming further. In some polar regions, including northern Canada, permafrost also includes methane hydrate, a highly concentrated form of CH4 trapped in solid form. Breakdown of permafrost releases this CH4 as well.

Pests and Disease

The geographical ranges of diseases and pests, especially those caused or transmitted by insects, have been shown to extend toward temperate regions because of climate change. West Nile virus and Lyme disease are two examples that already directly affect Canadians, while dengue fever could be an issue in the future. Canadians are also indirectly affected by the increase in populations of pests such as the mountain pine beetle (Figure 16.47).

Figure 16.47 Mountain pine beetle damage in Manning Park, British Columbia. Source: Jonhall (2010) CC BY 3.0 view source

References

Berwyn, B. (2017, 27 March). Climate Change-Fueled Jet Stream Linked to Brutal Floods and Heatwaves, Says Study. View at Inside Climate News

Charley, S. (2013, 13 June). Hurricane Sandy took highly unusual path, but climate change doesn’t get the blame – yet. View at GeoSpace

European Academies’ Science Advisory Council. (2013). Trends in Extreme Weather Events in Europe: Implications for National and European Union Adaptation Strategies. Halle (Saale), Germany: German National Academy of Sciences. Full text

European Academies’ Science Advisory Council. (2018). Extreme weather events in Europe. Preparing for climate change adaptation: An update on EASAC’s 2013 study. Halle (Saale), Germany: German National Academy of Sciences. Full text

Furphy, D. (2013, 28 September). What on earth is an RCP? A quick guide to carbon dioxide emissions scenarios used by the IPCC Assessment Report 5. View at Medium

Intergovernmental Panel on Climate Change. (2013). Summary for Policymakers. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker,T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Full text

Scott, M. (2011, 6 April). Heavy Rains and Dry Lands Don’t Mix: Reflections on the 2010 Pakistan Flood. View at Earth Observatory

World Meteorological Organization. (2014). Atlas of Mortality and Economic Losses from Weather, Climate, and Water Extremes (1970-2012). Geneva, Switzerland: World Meteorological Association. Full text

 

113

Chapter 16 Summary

The topics covered in this chapter can be summarized as follows:

16.1 What Is the Earth System?

Viewing Earth as a system allows us to take into account the complex ways in which the atmosphere, hydrosphere, biosphere, and lithosphere interact. Positive feedbacks amplify changes in the Earth system, and negative feedbacks reduce them. The stability of the Earth system will depend on what feedbacks are available. The presence of ice sheets makes the Earth system less stable.

16.2 Causes of Climate Change

Weather describes day-to-day conditions, but climate refers to the long-term average conditions over decades or longer. Climate forcings alter climate. They include processes that change the rate and location of solar energy reaching Earth’s surface; processes that alter how ocean currents move heat around Earth’s surface; and processes that affect how heat moves into and out of the atmosphere. Climate forcings operate on a range of timescales, from billions of years to less than a decade. Changes in greenhouse gas concentrations and albedo are two climate forcings affected by human activities.

16.3 Methods for Studying Past Climate

Climate conditions for some of human history can be determined from direct measurements that have been recorded, but for studying paleoclimate it is often necessary to use proxy data. Proxy data come from natural materials that behave in a systematic way in response to climate conditions like temperature or precipitation. Proxies include tree ring data, stable isotopes, measurements of gas bubbles trapped in ice, and the geographic distribution of rocks and fossils.

16.4 Computer Models of the Earth System

Earth-system models use mathematical equations to simulate Earth-system processes. Models are set up and checked using real-life measurements. Model uncertainty is a number that tells us the likelihood that a particular model result falls within a certain range of values. It is a way to evaluate whether results can be used to draw meaningful conclusions.

16.5 Humans in the Earth System

Data show recognizable anthropogenic influence on the Earth system beginning when humans began to use fossil fuels for industrial purposes. CO2 in the atmosphere has the isotopic fingerprints of fossil fuels. The flow of anthropogenic carbon into the Earth system is relatively small compared to some natural flows, but natural processes do not remove all of what humans put in, causing CO2 to accumulate in the atmosphere.

16.6 Welcome to the Anthropocene

Humans today are experiencing the results of past human influence on the Earth system, and humans in the future will experience the results of decisions made today. The main source of radiative forcing is anthropogenic CO2. Humans are already experiencing extreme climate events related to warming. The best-case projected warming scenario will stabilize global average temperatures at ~1ºC above 1996 temperatures by the end of the century, but this will require a peak in anthropogenic CO2 emissions by 2020, and an end to anthropogenic CO2 emissions by 2080.

Review Questions

  1. If you receive positive feedback on a project, this means someone says you did a good job. Does this mean a positive feedback in the climate system is also something good? Explain.
  2. Why does the presence of ice sheets cause the Earth system to be less stable?
  3. Using the orbital information on eccentricity, tilt, and precession, we could calculate variations in insolation for any latitude on Earth and for any month of the year. Why do we focus on a latitude of 65° N?
  4. Explain how the positioning of Gondwana at the South Pole contributed to glaciation during the Paleozoic.
  5. If the major currents in the oceans were to slow down or stop, how would that affect the distribution of heat on Earth, and what effect might that have on glaciation?
  6. Most volcanic eruptions lead to short-term cooling, but long-term sustained volcanism can lead to warming. Describe the mechanisms for these two different consequences.
  7. What property of greenhouse gases allows them to absorb infrared radiation and thus trap heat within the atmosphere?
  8. What are the advantages of proxy data over direct measurements of climate? What are the disadvantages of using proxy data?
  9. What are some steps that scientists take to help ensure that meaningful conclusions can be drawn from Earth-system computer models?
  10. What evidence tells us that rising atmospheric CO2 levels are primarily the result of humans burning fossil fuels?
  11. The flows of carbon through the carbon cycle that are related to human activities are actually much smaller than some natural flows. Why do human activities have such a large impact on the carbon cycle?
  12. Some people argue that climate has always changed, and present-day climate change is not significant compared to the long-term record of Earth’s climate. Is it true that present-day climate signals show that changes are not significant compared to Earth’s long-term climate history? Explain your answer.
  13. What climate mechanism links both extreme heat waves and droughts, and extreme flooding and rainfall events to climate change?
  14. What evidence suggests that increasingly intense tropical storms can be expected as climate warms?
  15. What are the potential impacts of thawing permafrost?

114

Answers to Chapter 16 Review Questions

1. A positive feedback does not mean that something good is happening in the Earth system. Rather, it means that a change is being amplified by other Earth-system processes that were triggered by the change. From a human perspective, having a change amplified could be good or bad.

2. When ice is present, relatively small changes in temperature can trigger albedo feedbacks that amplify the change. If no ice is present, a small increase in temperature might not have much of an impact. But if ice is present, that small increase could be enough to cause the ice to melt, and decrease Earth’s albedo. If ice is created when the climate cools, further cooling can be triggered if the ice increases Earth’s albedo significantly.

3. We use 65° for estimating the glaciation potential of orbital variations because glaciers are most likely to form at high latitudes, where temperatures are cooler, and insolation less direct. We use 65° N rather than 65° S because for more than 50 million years the continents have been concentrated in the northern hemisphere. We use July instead of January, because for ice to form and remain year-round, it is necessary to have cool summers.

4. Gondwana was situated over the south pole for much of the Paleozoic. Not only was the land subject to temperatures cooler over all, and less direct insolation for part of the year, but the large size of the continent limited the extent to which heat from the ocean could reach the interior of the continent. The development of ice on Gondwana further cooled the planet through albedo feedback.

5. If the major currents in the oceans were to slow down or stop, the tropics would get hotter and the high-latitude areas would get colder, leading to expansion of glaciers and sea ice. The various feedbacks (e.g., higher albedo because of increased ice cover) would result in a cooler climate over all.

6. From a climate perspective, the two important volcanic gases are SO2 and CO2. SO2 is converted to aerosols which block sunlight and can lead to short-term cooling (lasting years). CO2 can lead to warming, but only in situations where there is an elevated level of volcanism over the long term.

7. Greenhouse gases vibrate at frequencies that are similar to those of infrared radiation. When infrared radiation hits a greenhouse gas molecule, the molecule’s vibrational energy is enhanced and the radiation energy is converted into heat, which is trapped within the atmosphere.

8. Proxy data can provide information about climate in the distant past, for times during which there were no direct measurements. As new techniques are developed, analyses of proxy data can improve, whereas direct measurements will have experimental errors related to the instruments used to make them at the time. A disadvantage of proxy data is that we don’t always have a detailed timeline to go with the data, and may need to rely on other lines of evidence to understand when the proxy data should apply. Special equipment or techniques may be needed to get proxy data from geological materials. It is also necessary to ensure that geological processes have not altered the materials in some way that makes the results of analyses unreliable.

9. Real-life measurements are used to set up the model, and to test that it can provide realistic results. Many different models may be used to study the same scenario, and the results compared. This helps scientists to compare the results of representing the Earth system in different ways. An estimate of uncertainty is provided so scientists can decide whether or not the range of uncertainty is too big, given the size of the Earth-system change being studied.

10. Data from ice cores and direct measurements show that atmospheric CO2 levels began to rise very rapidly at the same time that humans began using fossil fuels extensively to power activities like manufacturing and transportation. At the same time, the stable carbon-isotope composition of atmospheric CO2 began to decrease at a rate that is consistent with the stable carbon-isotope fingerprint of plant-derived organic matter. Also simultaneously, the radiocarbon age of atmospheric CO2 began to fall, consistent with carbon from very ancient plant sources being added to the atmosphere. Aside from the timing of geochemical changes matching the onset of the industrial era, there are no other sources of carbon that can account for the geochemical fingerprint in the atmosphere.

11. The reason that relatively small anthropogenic flows of carbon into the carbon cycle can have a big impact is that they aren’t being balanced by natural processes that can remove the extra carbon. The result is that carbon from anthropogenic activities is accumulating in the atmosphere and ocean.

12. In the past, Earth’s climate has been both much colder and much hotter than today. However, present day temperatures are significantly different compared to global average temperatures over the past 1000 years, and stand out in the data from the past 800,000 years. Atmospheric CO2 levels are much higher than at any time during the past 800,000 years, and on that timescale, the increase is almost instantaneous. Therefore, the magnitude of present-day change in the Earth-system stands out clearly from background conditions predating the appearance of anatomically modern humans. It is also happening much faster than even the most rapid carbon-release event that we know of in the rock record.

13. Climate warming happens more rapidly at high latitudes more than low latitudes, decreasing the temperature difference between them. The decreased difference allows atmospheric circulation patterns to “stall out” over a location for longer than usual. What would normally be a short heatwave lasts much longer, and a precipitation event that would normally move elsewhere after a short time is now halted in place for weeks.

14. See Figure 16.44. The power of tropical storms depends on the warmth of seawater. Data show that during particularly warm intervals, tropical storms are stronger. Data also show that the strength of storms over all is increasing as sea surface temperatures rise over time.

15. Frozen soils provide a solid foundation for buildings in northern communities. Melting of permafrost means the soil beneath buildings is weakened, and buildings can suffer structural damage. Thawing means that hillsides are more prone to mass wasting, especially terrain along coasts. In terms of the climate system, thawing of permafrost allows micro-organisms to decompose organic matter within previously-frozen sediments, releasing additional CO2 and CH4, and contributing to a positive feedback.

XVII

Chapter 17. Glaciation

Adapted by Joyce McBeth, University of Saskatchewan
from Physical Geology by Steven Earle

Figure 17.1 Glaciers in the Alberta Rockies: Athabasca Glacier (centre left), Dome Glacier (right), and the Columbia Icefield (visible above both glaciers). The Athabasca Glacier has prominent lateral moraines on both sides. Source: Steven Earle (2015) CC BY 4.0 view source

Learning Objectives

After reading this chapter and answering the review questions at the end, you should be able to:

A glacier is a long-lasting (decades or more) body of ice that is large enough to move under its own weight. They are at least tens of metres thick and at least hundreds of metres in extent. About 10% of Earth’s land surface is currently covered with glacial ice, and although the vast majority of this is in Antarctica and Greenland, there are many glaciers in Canada, especially in the mountainous parts of BC, Alberta, and the Yukon, and in the far north (Figure 17.1). At various times during the past million years, glacial ice has been much more extensive, covering at least 30% of the Earth’s land surface at times.

Glaciers currently represent the largest repository of fresh water on Earth (~69% of all fresh water). They are highly sensitive to changes in climate, and in recent decades have been melting rapidly worldwide (Figure 17.2). Although some of the larger glacial masses may still last for several centuries, smaller glaciers, including many in western Canada, may be gone within decades. For mountainous regions, glaciers are an important sources of drinking water. Rapid glacial melting is a troubling issue for western Canadians because glacial ice is an important part of the hydrologic cycle in glaciated regions. Irrigation systems in BC, and across Alberta and Saskatchewan, are replenished by meltwater originating from glaciers in the Coast Range and the Rocky Mountains.

Figure 17.2 Example of rapid melting of a glacier over a 63-year period. Muir Glacier, Alaska. Source: NASA Climate 365 Project (n.d.) Public Domain view source

 

Figure 17.3 Part of the continental ice sheet in Greenland, with some outflow alpine glaciers in the foreground. Source: Steven Earle (2015) CC BY 4.0 view source

115

17.1 Types of Glaciers

There are two main types of glaciers: continental glaciers and alpine glaciers. Latitude, topography, and global and regional climate patterns are important controls on the distribution and size of these glaciers.

Continental Glaciers

Continental glaciers cover vast areas of land. Today, continental glaciers are only present in extreme polar regions: Antarctica and Greenland (Figure 17.3). Historically, continental glaciers also covered large regions of Canada Europe, and Asia, and they are responsible for many distinctive topographic features in these regions (Section 17.2 and 17.3).

Continent glaciers can form and grow when climate conditions in a region cool over extended periods of time. Snow can build up over time in regions that do not warm up seasonally, and if the snow accumulates in vast amounts, it can compact under its own weight and form ice.

Earth’s two current continental glaciers, the Antarctic and Greenland Ice Sheets, comprise about 99% of Earth’s glacial ice, and approximately 68% of Earth’s fresh water. The Antarctic Ice Sheet is vastly larger than the Greenland Ice Sheet (Figure 17.4) and contains about 17 times as much ice. If the entire Antarctic Ice Sheet melted, sea level would rise by about 80 m and most of Earth’s major coastal cities would be submerged.

Figure 17.4 Simplified cross-section profiles of the Antarctic and Greenland continental ice sheets. Both ice sheets are drawn to the same scale (exaggerated in the vertical direction). Source: Steven Earle (2015) CC BY 4.0 view source

Continental glaciers generally cover areas that are flat, but the force of gravity still acts on them and causes them to flow. Continental glacier ice flows from the region where it is thickest toward the edges where it is thinner (Figure 17.5). In the central thickest parts, the ice flows almost vertically down toward the base, while at the edges of the glacier, it flows horizontally out toward the margins. In continental glaciers like the Antarctic and Greenland Ice Sheets, the thickest parts (4,000 m and 3,000 m thick, respectively) are the areas where the rate of snowfall, and therefore of ice accumulation, are greatest. In Antarctica, the ice sheet flows out over the ocean, forming ice shelves. Ice shelves can slow the flow of continental glaciers outward. Conversely, if ice shelves break down continental glacier flow can speed up.

Figure 17.5 Cross-section showing ice-flow in the Antarctic Ice Sheet. Source: Steven Earle (2015) CC BY 4.0 view source

Alpine Glaciers

Alpine glaciers (aka valley glaciers) originate high up in the mountains, mostly in temperate and polar regions (Figure 17.1), but also in tropical regions in high mountains (e.g. in the Andes Mountains of South America).

The flow of alpine glaciers is driven by gravity, and primarily controlled by the slope of the ice surface (Figure 17.6). Alpine glaciers grow due to accumulation of snow over time. In the zone of accumulation, the rate of snowfall is greater than the rate of melting. In other words, not all of the snow that falls each winter melts during the following summer, and the ice surface in the zone of accumulation does not lose its annual accumulation of snow cover over the course of the year. In the zone of ablation, the rate of melting exceeds accumulation. The equilibrium line marks the boundary between the zones of accumulation (above) and ablation (below) (Figure 17.6).

Figure 17.6 Schematic diagram illustrating alpine glacier ice-flow. Source: Steven Earle (2015) CC BY 4.0 view source

Above the equilibrium line of a glacier, winter snow will remain even after summer melting, so snow gradually accumulates on the glacier over time. The snow layer from each year is covered and compacted by subsequent snow, and it is gradually compressed and converted to firn (Figure 17.7).

Figure 17.7 Steps in the process of formation of glacial ice from snow, granules, and firn. Source: Steven Earle (2015) CC BY 4.0 view source

Firn is a form of ice that forms when snowflakes lose their delicate shapes and become granules due to compression. With more compression, the granules are squeezed together, and air is forced out. Eventually the granules are “welded” together to create glacial ice (Figure 17.7). Downward percolation and freezing of water from melting contributes to the process of ice formation.

The equilibrium line of a glacier near Whistler, BC, is shown in Figure 17.8. Below this line is the zone of ablation. In the zone of ablation, bare ice is exposed because the previous winter’s snow has all melted. Above this line the ice is still mostly covered with snow from the previous winter.

Figure 17.8 The approximate location of the equilibrium line (red) in September 2013 on the Overlord Glacier, near Whistler, B.C. Source: Steven Earle (2015) CC BY 4.0, after Isaac Earle (n.d.) CC BY 4.0 view source

The position of the equilibrium line changes from year to year as a function of the balance between snow accumulation in the winter, and snow and ice melt during the summer. If there is more winter snow and less summer melting, this favours the advance of the equilibrium line down the glacier (and ultimately increases the size of the glacier). Between accumulation and melting, the summer melt matters most to a glacier’s ice budget. Cool summers promote an increase in glacier size, and thus lead to advance of the equilibrium line. Warm summers promote melting, and retreat of the equilibrium line.

Alpine glaciers move because they are heavy, and the force of gravity acts on the ice in the glacier to pull it down the slope of the mountains where they form. The movement of the glacier generates stress in the ice, which is proportional to the slope of the glaciers surface features of the underlying rock surface, and to the depth within the glacier.

As shown in Figure 17.9, the stresses are relatively small near the ice surface but much larger at depth. Stresses are greater in areas where the ice surface is relatively steep.

Figure 17.9 Stress within an alpine glacier (red numbers) as determined from the slope of the ice surface and the depth within the ice. The ice will deform and flow where the stress is greater than about 100 kilopascals, and regions with higher rates of deformation are depicted by the red arrows. Any motion of the lower ice will be transmitted to the ice above it, so although the red arrows get shorter toward the top, the ice is still moving (blue arrows in centre of diagram inset illustrate rate of ice motion). The upper ice (above the red dashed line) does not flow plastically, but it is carried along with the lower ice. Source: Joyce McBeth (2018) CC BY 4.0, modified after Steven Earle (2016) CC BY 4.0 view source

Like rock, ice behaves in a brittle fashion under low pressure conditions (shallow depths in the glacier), and plastically at higher pressures (deeper in the glacier). Stress also affects how ice deforms; at high stress ice will either break or deform plastically (ductile deformation) depending on the pressure conditions. Under brittle deformation conditions (low pressures, shallow depths in the glacier), stress is released when the ice cracks, so does not build up to high values. Within the upper 50 – 100 m of ice (above the dashed red line, in Figure 17.9), flow is brittle: the ice is rigid and will crack in response to stress.   Under ductile deformation conditions (higher pressures deeper in the glacier), stress can accumulate, and the ice will flow plastically in response to that stress. Ice deforms plastically if deeper than about 100 m in the glacier, and in this region stress levels can accumulate to high values (100 kilopascals or greater, Figure 17.9).

When the lower ice of a glacier flows, it moves the upper ice along with it. It may seem from the stress patterns (red numbers and arrows in Figure 17.9) that the lower ice moves more or faster than the upper ice, but this is not the case. The lower ice deforms (flows) and the upper part is carried along and deforms through brittle deformation if subjected to sufficient stress. The upper part of the glacier moves faster than the base of the glacier because there is friction between the base of the glacier and the surface beneath it that slows the movement of the ice at the base.

The plastic lower ice of a glacier can flow over irregularities in the rocks under the glacier. However, the upper rigid ice cannot flow in this way, and because it is being carried along by the lower ice, it tends to crack in locations when the lower ice flows over changes in the topography below the glacier. This leads to formation of crevasses in areas where the rate of flow of the deeper, plastic ice is changing. In the area shown in Figure 17.10, for example, the glacier is accelerating over the steep terrain, and the rigid surface ice cracks to release stress that accumulates due to the change in velocity and tension in the ice.

Figure 17.10 Crevasses in a glacier in Mount Cook National Park, New Zealand. Source: Bernard Spragg (2008) CC0 1.0 view source

In addition to deformation, another important aspect of glacier flow is basal sliding, which is sliding movement between the base of the glacier and the underlying material. The base of a glacier can be cold (below the freezing point of water) or warm (above the freezing point). If it is warm, a film of water can form between the ice and the material underneath, and the ice will be able to slide over this surface (Figure 17.11, left). If the base is cold, the ice will be frozen to the material underneath and it will be stuck — unable to slide along its base. In this case, all the movement of the ice will be by internal flow.

Figure 17.11 Differences in glacial ice motion with basal sliding (left) and without basal sliding (right). The dashed red line indicates the upper limit of plastic internal flow. Source: Steven Earle (2016) CC BY 4.0 view source

There are several factors that can influence warming of the ice and basal flow at the base of an alpine glacier. Friction between the base of the glacier and the surface underneath generates heat and can lead to melting of the ice at the base of the glacier. Rainwater and meltwater from upper regions of the glacier can percolate down and transfer heat to warm the base of the glacier and enhance basal sliding, particularly in warmer seasons. Geothermal heat from below also contributes to melting at the base of glaciers in regions with high heat flow due to volcanic activity.

Another factor that controls the temperature at the base of a glacier is the thickness of the ice. The force of gravity acting on thicker ice can enhance friction and melting at the base. Ice is also a good insulator so can prevent accumulated heat from escaping. The leading edge of an alpine glacier is typically relatively thin (see Figure 17.9), so it is common for this part to be frozen to its base while the rest of the glacier is still sliding. Since the leading edge of the glacier is frozen to the ground, and the rest of the glacier behind continues to slide forward, this causes the trailing ice to be pushed (or thrust) over top of the leading edge, forming thrust faults in the ice (Figure 17.12).

Figure 17.12 Thrust faults at the leading edge of the Byron Glacier, Portage Lake, Alaska, USA. The dark stripes are sediments that were entrained in the base of the glacier ice and transported up along the thrust faults. Source: Cindy Zackowitz (2011) CC BY-NC 2.0 view source

Just as the base of a glacier moves slower than the surface, the edges, which are more affected by friction along the channel walls, also move slower. If we were to place a series of markers across an alpine glacier and come back a year later, we would see that the ones in the middle had moved further forward than the ones near the edges (Figure 17.13).

Figure 17.13 Markers on an alpine glacier move forward at different rates over a period of time. Source: Steven Earle (2016) CC BY 4.0 view source

Alpine glacial ice continuously moves down the slope of the ice in response to gravity, but it may not appear to be moving because the front edge of a glacier is also continuously losing volume. It either melts or, if they glacier terminates at a lake or ocean, the front edge will calve into the water (break off pieces of the front edge of the glacier that become icebergs). If the rate of forward motion of the glacier is faster than the rate of ablation (melting), the leading edge of the glacier advances (moves forward). If the rate of forward motion is about the same as the rate of ablation, the leading edge remains stationary, and if the rate of forward motion is slower than the rate of ablation, the leading-edge retreats (moves backward).

Calving of icebergs is an important process for glaciers that terminate in lakes or oceans. An example of such a glacier is the Berg Glacier on Mt. Robson (Figure 17.14), which sheds small icebergs into Berg Lake. The Berg Glacier also lose mass by melting, evaporation, and sublimation.

Figure 17.14 Mt. Robson, the tallest peak in the Canadian Rockies, hosts the Berg Glacier (centre), and Berg Lake. Although there were no icebergs visible when this photo was taken, the Berg Glacier loses mass by shedding icebergs into Berg Lake. Source: Steven Earle (2015) CC BY 4.0 view source

 

Exercise: Ice Advance and Retreat

These diagrams in Figure 17.15 represent a glacier with markers placed on its surface to determine the rate of ice motion over a one-year period. The ice is flowing from left to right.

Figure 17.15 Glacier with markers to show rate of motion. Source: Steven Earle (2015) CC BY 4.0 view source

1. In the middle diagram, the leading edge of the glacier has advanced. Draw in the current position of the markers.

2. In the lower diagram, the leading edge of the glacier has retreated. Draw in the current position of the markers.

 

 

 

116

17.2 Glacial Erosion

Glaciers are effective agents of erosion, especially in situations where the base of the glacier is not frozen to the underlying material and can therefore slide over the bedrock or other sediment. The ice itself is not particularly effective at erosion because it is relatively soft (Mohs hardness 1.5 at 0°C). Glacial erosion is primarily driven by abrasion of the underlying rocks by rock fragments embedded within the ice. These rocks are pushed down onto the underlying surfaces by the ice, and because they are hard they can gouge and grind down the materials beneath the glacier. An analogy for these processes is to compare the effect of a regular piece of paper being rubbed against a wooden surface (“ice eroding rock”) to rubbing a piece of sandpaper over the same surface (“ice with embedded rocks eroding rock”). The results of glacial erosion are different in areas with continental glaciation versus alpine glaciation.

Continental Glacial Erosion Features

Continental glaciation tends to produce relatively flat bedrock surfaces, especially where the rock beneath is uniform in strength. In areas where there are differences in the strength of rocks, a glacier tends to erode the softer and weaker rock more effectively than the harder and stronger rock. Much of central and eastern Canada, which was completely covered by the huge Laurentide Ice Sheet at various times during the Pleistocene Epoch, has been eroded to a relatively flat surface. Glacial deposits have created distinctive topographic features on the landscapes in these regions — such as drumlins, eskers, and moraines (Figure 17.16). These continental glacial features are deposits of glacial materials and are described further in Section 17.3.

Figure 17.16 Landscape features associated with continental glaciation Source: Luis María Benítez (2005) CC BY 4.0 view source

In areas of continental glaciation, the lithosphere is depressed by the weight of glacial ice that is up to 4,000 m thick. Basins formed along the edges of continental glaciers; for example, basins formed around the edges of the Laurentide Ice Sheet that once covered much of Canada (Section 17.4). These basins filled with glacial meltwater, and layers of sediments. Many such lakes, some of them huge, existed at various times along the southern edge of the Laurentide Ice Sheet.

One example of these lakes was Glacial Lake Missoula, which formed within Idaho and Montana, just south of the BC border with the United States. During the latter part of the last glaciation (30 ka to 15 ka), the ice holding back Lake Missoula retreated enough to allow some of the lake water to escape, which escalated into a voluminous and rapid outflow (over days to weeks). During this outflow, most of the lake drained into the Columbia River valley and flowed to the Pacific Ocean. It is estimated that this type of catastrophic outflow happened at least 25 times during this period, and in many cases, the rate of outflow was equivalent to the discharge of all of Earth’s current rivers combined.

Alpine Glacial Erosion Features

Alpine glaciers produce very different topography than continental glaciers. Alpine glaciers produce wide valleys with relatively flat bottoms and steep sides due to the erosion that occurs at the base and edges of the glaciers. These are known as U-shaped valleys (Figure 17.17). In contrast, unglaciated river valleys generally have a V shape.

Figure 17.17 A depiction of a U-shaped valley occupied by a large glacier. Source: Steven Earle (2016) CC BY 4.0 view source

In coastal regions where the bottom of the valley is filled with water, the U-shaped valleys are called fjords. The coastal mountains of BC have many fine examples of U-shaped valleys and fjords. For example, Howe Sound is a fjord that was once occupied by a large glacier. Howe Sound and most of its tributary valleys have pronounced U-shaped profiles due to glaciation (Figure 17. 18).

Figure 17.18 The view down the U-shaped valley of Mill Creek valley toward the U-shaped valley of Howe Sound, with the village of Britannia on the opposite side. Source: Keefer4 (2005) CC BY-SA 2.5 view source

Several other topographic features derived from alpine glacial erosion are found in U-shaped valleys and their tributary valleys (Figure 17.19). Arêtes are sharp ridges formed between U-shaped glacial valleys. Cols are low points (saddles) along arêtes; they form passes (high points) between glacial valleys. Horns are steep peaks that have been eroded by glaciers and freeze-thaw activity on three or more sides. Cirques are bowl-shaped basins that form at the head of a glacial valley, and tarns are lakes that form when cirques are flooded. Hanging valleys form when U-shaped valleys of tributary glaciers connect with a larger U-shaped valley; the tributary valley hangs above the main valley because the larger main-valley glacier is eroded more deeply into the terrain. Truncated spurs (aka “spurs”) are features at the ends of arêtes where the rock is eroded into steep triangle-shaped cliffs by the glacier in the main valley.

Figure 17.19 A diagram of some of the important alpine-glaciation erosion features. Source: Steven Earle (2015) view source. Modified after Luis María Benítez (2005) CC0 1.0 view source

Figure 17.20 shows examples of these features in the Swiss Alps. The area in the image was intensely glaciated during the past glacial maximum and still contains glaciers. The large U-shaped valley in the lower right was occupied by glacial ice historically, and all of the other glaciers shown here were longer and much thicker than they are now. But even at the peak of the Pleistocene glaciation, some of the higher peaks and ridges in this image would have been exposed and not directly affected by glacial erosion. A peak that extends above the surrounding glacier is called a nunatak. In these areas, and in the areas above the glaciers in the image today, most of the erosion is linked to freeze-thaw action.

Figure 17.20 A view of the Swiss Alps from the International Space Station, taken in 2006. The region shown is in the area of the Aletsch Glacier. The prominent peaks labelled “Horn” are the famous mountain peaks the Eiger (left) and Wetterhorn (right). A variety of alpine glacial erosion features are labelled. Source: Steven Earle (2016) view source. Modified after NASA Earth Observatory (n.d.) Public Domain view source

A roche moutonnée is aglacial erosion feature that forms when a glacier moves over an outcrop of bedrock. Roche moutonnées consist of a hill of rock, often with a smooth, often low angle slope on one side, and a steeper and jagged slope on the other side. The side that is smooth and relatively low angle is the side the glacier was flowing from, and the a steep and sometimes jagged side is the direction the ice was moving (Figure 17.21, left).

Figure 17.21 Roche moutonnée near Myot Hill, Scotland. Source: Chris Upson (2006) CC BY-SA 2.0 view source

Glacial grooves (tens of centimetres to metres wide) and glacial striae (millimetres to centimetres wide) are created by the erosion caused by fragments of rock embedded in the ice at the base of a glacier (Figure 17.22, left and right). Glacial striae are very common on rock surfaces eroded by both alpine and continental glaciers. Glacial polish occurs when the abrasion of the rock by the glacier renders the rock so smooth it reflects light.

Figure 17.22 Examples of glacial striae from near Squamish, BC. Ice flow was from right to left in both images. Source: Steven Earle (2016) CC BY 4.0 view source

Lakes are common features in glacial environments. A lake that is confined to a glacial cirque is known as a tarn (Figure 17.23). Tarns are common in areas of alpine glaciation because the ice that forms a cirque typically carves out a depression in bedrock that can then fill with water. Moraines, which are linear deposits of glacial sediments (till) left by the glacier along its edges, can form a dam at the end of a tarn. Often, a series of moraines will form as glaciers recede. These can act as water dams, and result in strings of lakes called rock basin lakes or paternoster lakes.

Figure 17.23 Lower Thornton Lake, a tarn, in the Northern Cascades National Park, Washington. Source: Jeff Pang (2007) CC BY 2.0 view source

A lake that occupies a glacial valley is known as a finger lake. In some cases, a finger lake is confined by a dam formed by an end moraine, in which case it may be called a moraine lake (Figure 17.24). Another type of glacial lake is a kettle lake. These are discussed in section 17.4 in the context of glacial deposits.

Figure 17.24 Peyto Lake in the Alberta Rockies, is both a finger lake and a moraine lake, as it is flooding a glacial valley, and is dammed by an end moraine at right. Source: Jeff Hollet (2016) Public Domain view source

 

Exercise: Identifying Glacial Erosion Features

Figure 17.25 is a photo of Mt. Assiniboine in the BC Rocky Mountains. What are the features at locations a through e? Look for one of each of the following: a horn, an arête, a truncated spur, a cirque, and a col. Try to identify some of the numerous other arêtes in this view, as well as another horn.

Figure 17.25 Glacial erosion features of Mt. Assiniboine. Source: Steven Earle (2015) CC BY 4.0 view source. Modified after Kurt Stegmüller (2008) CC BY 3.0 view source

 

 

 

117

17.3 Glacial Deposits

Sediments transported and deposited during glaciations are abundant throughout Canada. They are important sources of aggregate for construction materials (sand, gravel), and are also important groundwater reservoirs. Because they are almost all unconsolidated, they have significant implications for slope stability and mass wasting.

Figure 17.26 illustrates some of the ways that sediments are transported and deposited by alpine glaciers. The Bering Glacier is the largest glacier in North America, and although most of it is in Alaska, it flows from an icefield that extends into the southwestern Yukon Territory. The surface of the ice is partially, or in some cases completely, covered with rocky debris that has fallen onto the glacier from surrounding steep rock faces. There are muddy rivers issuing from the glacier in several locations, depositing sediment on land, into Vitus Lake, and directly into the ocean. Icebergs are portions of the glacier that have broken off and float away in a lake or ocean. Icebergs are laden with glacial sediments, which are released and deposited as the icebergs melt. Also, not visible in this view, there are sediments being moved along within and beneath the glacier itself.

Figure 17.26 Part of the Bering Glacier in southeast Alaska, the largest glacier in North America. It is about 14 km in width in the centre of this view. Source: Roger Simmon, Landsat 7 Science Team, NASA (2002) Public Domain view source

Sediments are formed and transported in several ways in glacial environments (Figure 17.27). There are many different kinds of glacial sediments, which are generally classified by whether they are transported on, within, or beneath the glacial ice.

Figure 17.27 A depiction of the various types of sediments associated with the Bering Glacier. The glacier is shown in cross-section. Source: Steven Earle (2016) CC BY 4.0 view source

Supraglacial (on top of the ice) and englacial (within the ice) sediments are released from the melting front of a stationary glacier. These sediments can form a ridge of unsorted sediments called an end moraine. The end moraine from the furthest advance of a glacier is called a terminal moraine. The general name for any sediments transported and deposited by glacial ice is till.

Subglacial sediment (e.g., lodgement till) is material that has been eroded from the rock underlying the glacier by the ice and then transported by the ice. It has a wide range of grain sizes, including a relatively high proportion of silt and clay. The larger clasts (pebbles to boulders in size) tend to become partly rounded by abrasion. When a glacier eventually melts, the lodgement till is exposed as a sheet of well-compacted sediment ranging from several centimetres to many metres in thickness. Lodgement till is normally poorly sorted and does not contain bedding features like a lake or stream sediment (Figure 17.28).

Figure 17.28 Examples of glacial till: a: lodgement till from the front of the Athabasca Glacier, Alberta; b: ablation till at the Horstman Glacier, Blackcomb Mountain, BC. Source: Steven Earle (2016) CC BY 4.0 view source

Supraglacial sediments are primarily derived from freeze-thaw eroded material that has fallen onto the ice from rocky slopes above. These sediments form lateral moraines (moraine deposits along the edges of the glacier, see Figure 17.1 for an example). Where two glaciers meet, the sediments form medial moraines (medial moraines are visible in Figure 17.20 and Figure 17.26.) Most of this material is deposited on the ground when the ice melts. This is called ablation till, a mixture of fine and coarse angular rock fragments, with much less sand, silt, and clay than lodgement till (Figure 17.28). When supraglacial sediments become incorporated into the body of the glacier, they are known as englacial sediments (Figure 17.27).

Water flows on the surface, within, and at the base of a glacier, even in cold areas and even when the glacier is advancing. Depending upon its velocity, this water is able to transport sediments of various sizes, and discharges most of these sediments out of the lower end of the glacier, where they are deposited as outwash sediments. These sediments accumulate in a wide range of environments in the proglacial region (the area in front of a glacier). Most of the sediments accumulate in fluvial environments, but some are deposited in lacustrine and marine environments. Glaciofluvial sediments are similar to sediments deposited in normal fluvial environments, but are glacially-derived sediments, and are thus dominated by silt, sand, and gravel. The grains tend to be moderately well rounded and sorted, and the sediments have similar sedimentary structures (e.g., bedding, cross-bedding, clast imbrication [overlapping]) to those formed by non-glacial streams (Figure 17.29).

Figure 17.29 Examples of glaciofluvial sediments: a: glaciofluvial cross-bedded sand of the Quadra Sand Formation at Comox, BC.; b: glaciofluvial gravel and sand, Nanaimo, BC. Source: Steven Earle (2016) CC BY 4.0 view source

A large proglacial plain of sediment is called a sandur (aka outwash plain), and within this area, glaciofluvial deposits can be tens of metres thick. In situations where a glacier is receding, a block of ice might become separated from the main ice sheet and become buried in glaciofluvial sediments. When the ice block eventually melts, a depression forms, known as a kettle, and if this fills with water, it is known as a kettle lake (Figure 17.30, 17.32). Kettle lakes are also known as pothole lakes or prairie potholes.

Figure 17.30 A kettle lake amid vineyards and orchards in the Osoyoos area of BC. Source: Steven Earle (2015) CC BY 4.0 view source

A supraglacial, englacial, or subglacial stream will create its own channel within the ice, and sediments that are being transported and deposited by the stream will build up within that channel. When the ice melts, the sediment will be deposited upon the underlying ground surface to form a long sinuous ridge known as an esker. Eskers are most common in areas of continental glaciation. They can be several metres high, tens of metres wide, and tens of kilometres long (Figure 17.31). Eskers are commonly comprised of well-sorted sands.

Figure 17.31 Part of an esker that formed beneath the Laurentide Ice Sheet in northern Canada. Source: Gord McKenna (1986) CC BY-NC-ND 2.0 view source

Drumlins are elongated, oval shaped ridges of englacial to subglacial sediments that form at the base of continental glaciers. They are often tens of metres high and hundreds of metres long, and often occur in clusters (“fields”) of tens to hundreds of drumlins (Figure 17.32). As the sediments are deposited, the glacier molds the drumlins’ shapes as the glacier moves over and around them. The long axis of a drumlin is aligned with the direction that the ice moved when the drumlin was deposited.

Figure 17.32 Drumlins and kettle lakes viewed from the air near Fort St John, BC. There are numerous drumlins in the image; one is outlined in red. Can you spot the others? Note the alignment of the long axes of the drumlins. Source: Joyce McBeth (2002) CC-BY 4.0. Click the image for a higher resolution version.

Glacial outwash streams commonly flow into proglacial lakes (lakes in front of glaciers) where glaciolacustrine sediments are deposited. These are dominated by silt- and clay-sized particles and are typically laminated (finely layered) on the millimetre scale. In some cases, varves develop. Varves are a series of beds with distinctive summer and winter layers: relatively coarse in the summer when melt discharge is high, and finer in the winter, when discharge is low. Icebergs are common in proglacial lakes, and most of them contain englacial sediments of various sizes. As the icebergs melt, the released clasts sink to the bottom and are incorporated into the glaciolacustrine layers as drop stones (Figure 17.33a). The processes that occur in proglacial lakes can also take place where a glacier terminates in the ocean. The sediments deposited there are called glaciomarine sediments (Figure 17.33b).

Figure 17.33 Examples of glaciolacustrine and glaciomarine sedimentary structures. a: varved glaciolacustrine sediments containing a drop stone, Nanaimo, BC.; and b: a laminated glaciomarine sediment, Englishman River, BC. Source: Steven Earle (2016) CC BY 4.0 view source

 

Exercise: Identifying Glacial Depositional Environments

Refer to the photo of the Bering Glacier in Alaska shown in Figure 17.26. Glacial sediments of many different types are being deposited throughout the region depicted in this photo.

Identify where you would expect to fine the following:
(a) glaciofluvial sand
(b) lodgement till
(c) glaciolacustrine clay with drop stones
(d) ablation till
(e) glaciomarine silt and clay

 

 

118

17.4 Glaciations over Earth's History

We are currently living in the middle of a glacial period, though it is less intense now than it was 20,000 years ago. This is not the only period of glaciation in Earth’s history; there have been many in the distant past (Figure 17.34). In general, however, over the course of Earth’s history the Earth’s surface has been warm and ice-free for longer periods than it has been cold and glaciated.

Figure 17.34 The record of major past glaciations during Earth’s history. Source: Steven Earle (2015) CC BY 4.0 view source

Pre-Cenozoic Glaciations

The oldest known glacial period is the Huronian. Based on evidence of glacial deposits from the area around Lake Huron in Ontario and elsewhere, it is evident that the Huronian Glaciation lasted from approximately 2.4 to 2.1 Ga. Because rocks of that age are rare, we do not know much about the intensity or global extent of this glaciation.

Late in the Proterozoic, for reasons that are not fully understood, the climate cooled dramatically, and Earth had the most intense time of glaciation it has ever experienced. The glaciations of the Cryogenian Period (cryo is Latin for icy cold) are also known as the “Snowball Earth” glaciations. Scientists have hypothesized that the entire planet was frozen at this time — even in equatorial regions — with ice on the oceans up to 1 km thick. A visitor to our planet at that time would not have found it habitable, although life still survived in the oceans.

There were two main glacial periods within the Cryogenian, each lasting for about 20 million years: the Sturtian at around 700 Ma and the Marinoan at 650 Ma. There is also evidence of some shorter glaciations both before and after these longer periods of glaciation. The end of the Cryogenian glaciations coincides with the evolution of relatively large and complex life forms on Earth. This started during the Ediacaran Period, and then continued with the so-called explosion of life forms in the Cambrian. Some geologists think that the changing environmental conditions of the Cryogenian are what triggered the evolution of large and complex life.

There have been three major glaciations during the Phanerozoic (the past 540 Ma). These include the Andean/Saharan (recorded in rocks of South America and Africa), the Karoo (named for rocks in southern Africa), and the Cenozoic glaciations. The Karoo was the longest of the Phanerozoic glaciations, persisting for much of the time that the supercontinent Gondwana was situated over the South Pole (~360 to 260 Ma). Glaciers covered large parts of Africa, South America, Australia, and Antarctica. This widespread glaciation, across continents that are now far apart, was an important component of Alfred Wegener’s evidence for continental drift. Unlike the Cryogenian glaciations, the Andean/Saharan, Karoo, and Cenozoic glaciations only affected parts of Earth. During Karoo times, for example, what is now North America was near the equator and remained unglaciated.

Earth was warm and essentially unglaciated throughout the Mesozoic. Although there may have been some alpine glaciers at this time, there is no evidence for them preserved in the geologic record. The dinosaurs, which dominated terrestrial habitats during the Mesozoic, did not have to endure icy conditions.

Cenozoic Glaciations

A warm climate persisted into much of the Cenozoic; there is evidence that the Paleocene (from about 50 to 60 Ma) was the warmest part of the Phanerozoic since the Cambrian (Figure 17.35).

Figure 17.35 Global temperature trends over the past 65 Ma (the Cenozoic). From the end of the Paleocene to the height of the Pleistocene glaciation, global average temperature dropped by about 14°C. Source: Joyce McBeth (2018) CC BY 4.0. Modified after Steven Earle (2015) CC BY 4.0 (view source) and Makiko Sato & James Hansen (2012), including data from Zachos et al (2008) view original

A number of tectonic events during the Cenozoic have contributed to persistent and significant planetary cooling from 50 Ma to near the present. For example, the collision of the Indian plate with the Eurasian plate and the formation of the Himalayan range and the Tibetan Plateau. Mountain building events such as the formation of the Himalayas are followed by weathering and erosion of the uplifted rocks. Higher than normal global rates of silicate mineral weathering associated with mountain building, especially weathering of feldspar, leads to a decrease in carbon dioxide concentrations in the atmosphere. This contributes to global climate cooling.

Figure 17.36 The Antarctic Circumpolar Current (red arrows) prevents warm water from the rest of Earth’s oceans from reaching Antarctica. Source: Steven Earle (2015) CC BY 4.0 view source

At 40 Ma, ongoing plate motion widened the narrow gap between South America and Antarctica, resulting in the opening of the Drake Passage. This allowed for unrestricted west-to-east flow of water around Antarctica: the Antarctic Circumpolar Current (Figure 17.36), which effectively isolated the Southern Ocean from the warmer waters of the Pacific, Atlantic, and Indian Oceans. The region cooled significantly, and by 35 Ma (Oligocene) glaciers had started to form on Antarctica.

Global temperatures remained relatively steady during the Oligocene and early Miocene, and the Antarctic glaciation waned during that time. At around 15 Ma, subduction-related volcanism between central and South America created the land connection between North and South America, preventing water from flowing between the Pacific and Atlantic Oceans. This further restricted ocean currents that transfer heat from the tropics to the poles, leading to cooling and advance of the Antarctic glaciation.

The expansion of the Antarctic ice sheet increased reflection of solar radiation at the Earth’s surface and promoted a positive feedback loop of further cooling. With more reflective glacial ice, there was more cooling, leading to accumulation of more ice, and so on. By the Pliocene (~5 Ma) ice sheets had started to grow in North America and northern Europe. The most intense part of the current glaciation — and the coldest climate conditions of the current glaciation — has been during the past million years (the last third of the Pleistocene).

The Pleistocene Epoch of the Cenezoic Era (2.58 Ma to 0.126 Ma), is also known as the Ice Age, Pleistocene Glaciation, or Quaternary Glaciation. The Pleistocene has been characterized by temperature fluctuations over a range of almost 10°C on time scales of 40,000 to 100,000 years. These temperature variations have corresponding with expansion and contraction of ice sheets. The temperature variations are attributed to subtle changes in Earth’s orbit, tilt, and wobble. These cyclical changes are called Milankovitch cycles. Over the past million years, the glaciation cycles have cycled over every 100,000 years, approximately (Figure 17.37).

Figure 17.37 Foram oxygen isotope record for the past 5 million years based on O isotope data from sea-floor sediments Source: Steven Earle (2015) CC BY 4.0 view source. Data from Lisiecki and Raymo (2005). Access the data

The Wisconsinian Glaciation

The Wisconsinan Glaciation was the last major continental glaciation in North America (from 150-50 ka). During the Wisconsinan, all of Canada and a small portion of the northern United States was covered with continental glaciers (Figure 17.38). The massive Laurentide Ice Sheet covered most of eastern Canada, as far west as the Rockies, and the smaller Cordilleran Ice Sheet covered most of the western region of present day BC and the Yukon Territory. At various other glacial peaks during the Pleistocene and Pliocene, the ice extent was similarly distributed over North America, and in some cases, was even more extensive. The combined Laurentide and Cordilleran Ice Sheets were comparable in volume to the current Antarctic Ice Sheet.

Figure 17.38 Extent of northern hemisphere ice sheets near the peak of the Wisconsinan Glaciation (grey shading). Interglacial ice is shown in black. Source: Hannes Grobe (2008) CC BY 3.0 view source

Exercise: Pleistocene Glacial and Interglacial Periods

Figure 17.39 shows the past 500,000 years of the data set used in Figure 17.37. The last five glacial periods are marked with snowflakes. The most recent glaciation, which peaked at around 20 ka, is known as the Wisconsinan Glaciation. Describe the nature of temperature change that followed each of these glacial periods.

The current interglacial period (Holocene) is marked with an H. Point out the previous five interglacial periods.

Figure 17.39 Global mean temperatures over the last 500,000 years. Source: Steven Earle (2015) CC BY 4.0 (view source), using data from Lisiecki and Raymo (2005) Access the data

 

References

Hansen, J. E., and Sato, M. (2012). Climate Sensitivity Estimated From Earth’s Climate History. Read the paper

Lisiecki, L. E., and M. E. Raymo (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic d18O records. Paleoceanography, 20, PA1003.  doi:10.1029/2004PA001071. View PDF

Zachos, J. C., Dickens, G. R., and Zeebe, R. E. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 541, 279-283. doi:10.1038/nature06588

119

Chapter 17 Summary

The topics covered in this chapter can be summarized as follows:

17.1 Types of Glaciers

The two main types of glaciers are continental glaciers, which are very large and cover large parts of continents (e.g. the Antarctic Ice Sheet), and alpine glaciers, which occupy mountainous regions. Ice accumulates at higher elevations — above the equilibrium line — where the snow that falls in winter does not all melt in summer. In continental glaciers, ice flows outward from where it is thickest. In alpine glaciers, ice also flows from thicker to thinner regions in the glacier, obeying the law of gravity. At depth in glacier ice, flow occurs through internal deformation, but glaciers that have liquid water at their base can also flow by basal sliding. Crevasses form in the rigid surface ice in places where the lower plastic ice is changing flow rate or shape as it moves over the underlying topography.

17.2 Glacial Erosion

Glaciers are important agents of erosion. Continental glaciers tend to erode land surface into flat plains, while alpine glaciers create a wide variety of different erosional features. The key feature of alpine glacial erosion is the U-shaped valley. Arêtes are sharp ridges that form between two valleys, and horns form where a mountain is glacially eroded on at least three sides. Since tributary glaciers do not erode as deeply as main-valley glaciers, hanging valleys exist where the two meet. On a smaller scale, both types of glaciers form roche moutonnées, d glacial grooves, and striae.

17.3 Glacial Deposits

Glacial deposits form as materials are transported and deposited in a variety of different ways in a glacial environment. Sediments that are moved and deposited directly by ice are known as till. Till deposits left at the edges of the glacier as it recedes are known as moraines. Till can also form features such as drumlins (oval-shaped elongated hills) and kettle lakes. Glaciofluvial sediments are deposited by glacial streams, either forming eskers or large proglacial plains known as sandurs. Glaciolacustrine and glaciomarine sediments originate within glaciers and are deposited in lakes and oceans, respectively.

17.4 Glaciations over Earth’s History

There have been many glaciations in Earth’s past, the oldest known starting about 2.4 Ga. The late Proterozoic “Snowball Earth” glaciations were thought to be sufficiently intense to affect the entire planet. The Pleistocene Glaciation was a series of glacial events over the past 2.85 Ma. The periodicity of glaciations in the Pleistocene is related to subtle changes in Earth’s orbital characteristics (Milankovitch cycles), which are exaggerated by positive climate feedback processes. North America was most recently glaciated during the Wisconsinan Glaciation, from 150-50 ka.

 

Review Questions

1. What feature of a continental glacier causes it to flow?

2. Explain how alpine glacier ice flows.

3. What does the equilibrium line represent in a glacier?

4. Which of the following is more important to the growth of a glacier: very cold winters or relatively cool summers? Why?

5. Describe the relative rates of ice flow, and why the rate distributions are the way they are, within the following parts of a glacier: (a) the bottom versus the top and (b) the edges versus the middle.

6. Figure 17.40 shows thrust faults in the leading edge of the Athabasca glacier in Alberta. Find the thrust faults in the glacier. Note: they are harder to distinguish in this image than in Figure 17.12 because they don’t have sediments along the faults.

Figure 17.40 Edge of the Athabasca glacier in Alberta. Source: Steven Earle (2015) CC BY 4.0 view source

7. What ice conditions are necessary for basal sliding to take place?

8. What sources of heat can lead to melting and/or water accumulation at the base of a glacier?

9. Why do glaciers carve U-shaped valleys, and how does a hanging valley form? How is a river valley different than a glacial valley?

10. A horn is typically surrounded by cirques. What is the minimum number of cirques you would expect to find around a horn?

11. A drumlin and a roche moutonnée are both streamlined glacial erosion features. How do they differ in shape? How can you determine the direction of glacial flow from their shapes?

12. What are drop stones, and under what circumstances are they likely to form?

13. What types of glacial sediments are likely to be sufficiently permeable to make good aquifers?

14. Why are the Cryogenian glaciations called Snowball Earth?

15. Earth cooled dramatically from the end of the Paleocene until the Holocene. Describe some of the geological events that contributed to this cooling.

16. When and where was the first glaciation of the Cenozoic?

17. Describe the extent of the Laurentide Ice Sheet during the height of the last Pleistocene glacial period.

18. Four examples of glacial sediments are shown in Figure 17.41 below. Describe the important characteristics (e.g., sorting, layering, grain-size range, grain shape, sedimentary structures) of each, and give each a name (choose from glaciofluvial, glaciolacustrine, lodgement till, ablation till, and glaciomarine).

Figure 17.41 Examples of glacial sediments. Source: Steven Earle (2015) CC BY 4.0 view source

120

Answers to Chapter 17 Review Questions

1. Continental glaciers flow from the areas where the ice is thickest (and therefore at the highest elevation) toward areas (at the margins) where the ice is thinnest. Ice thickness tends to be related to the rate of ice accumulation.

2.

3. The equilibrium line represents the boundary between the area where ice is accumulating (typically at high elevations), and where it is being depleted (mostly by melting). Above the equilibrium line more snow accumulates in winter than can melt in summer so the glacier is always covered in snow. Below the equilibrium line the snow cover is lost by the end of summer.

4. Relatively cool summers are more important because that controls how much snow will melt in the summer. In many situations very cold winters are associated with less snow accumulation than just cold winters.

5. (a) The ice at the bottom of a glacier flows more slowly than that at the top. In fact if the glacier is frozen to its base the lowermost ice might not be moving at all. (b) The edges also flow more slowly than the middle because there is more friction there between the ice and the valley walls.

6.

7. Basal sliding will take place when the bed of the glacier is warm enough for water to be liquid. The water will act as a lubricant to allow the ice to flow.

8.

9. Glaciers carve U-shaped valleys because they are relatively wide (compared with rivers) and most of the erosion takes place at the base rather than the sides. A hanging valley forms where a tributary glacier joins a larger glacier and where the larger glacier has eroded a deeper valley.

10. There must be at least three cirques to form a horn. In most cases there wouldn’t be room for more than four.

11. A drumlin is relatively steep at the up-ice end and streamlined at the down-ice end. A roche moutonée is streamlined at the up-ice end and jagged at the down-ice end where plucking has taken place.

12. Drop stones are large clasts that are present with lacustrine or marine glacial sediments. They form when coarse material drops from melting icebergs.

13. Glaciofluvial sediments (sand or sand and gravel) are likely to be sufficiently permeable to make good aquifers.

14. The Cryogenian glaciations are called Snowball Earth because it is thought that freezing conditions affected the entire planet and that the oceans were frozen over, even at the equator.

15. The cooling from the end of the Paleocene until the Holocene was related to the formation of mountains including the Himalayas, the Rockies, and the Andes; the opening of the Drake Passage; the development of Antarctic Circumpolar Current; and the closing of the Isthmus of Panama.

16. The first glaciation of the Cenozoic took place in Antarctica during the Oligocene (around 30 Ma).

17. At the height of the last glaciation, the Laurentide Ice Sheet covered almost all of Canada and extended south into the United States as far as Wisconsin.

18. From left to right, the sediments are:

Lodgement till: Poorly sorted, clay to boulders, no layering, large clasts not well rounded, no structures.

Glaciofluvial sand and gravel: Bedded, some fine beds of sand, and some much coarser beds. Clasts are relatively well rounded.

Ablation till: Poorly sorted, pebbles to boulders, no layering, large angular clasts, no apparent structures.

Glaciofluvial sand: Sand- and silt-sized, well-sorted, bedded and cross-bedded.

 

 

 

XVIII

Chapter 18. Geological Resources

Introduction

Learning Objectives

After reading this chapter, completing the exercises within it, and answering the questions at the end, you should be able to:

121

18.1 If You Can't Grow It, You Have to Mine It

It has been said that “if you can’t grow it, you have to mine it,” meaning that anything we can’t grow we have to extract from Earth in one way or another. This includes water, of course, our most important resource, but it also includes all the other materials that we need to construct things like roads, dams, and bridges, or manufacture things like plates, toasters, and telephones. Even most of our energy resources come from Earth, including uranium and fossil fuels, and much of the infrastructure of this electrical age depends on copper (Figure 18.1).
Figure 20.1 The open pit (background) and waste-rock piles (middle) of the Highland Valley Copper Mine at Logan Lake, British Columbia [Photo by Russell Hartlaub, used with permission]
Figure 18.1 The open pit (background) and waste-rock piles (middle) of the Highland Valley Copper Mine at Logan Lake, British Columbia [Photo by Russell Hartlaub, used with permission]

Virtually everything we use every day is made from resources from Earth. For example, let’s look at a tablet computer (Figure 18.2). Most of the case is made of a plastic known as ABS, which is made from either gas or petroleum. Some tablets have a case made from aluminum. The glass of a touch screen is made mostly from quartz combined with smaller amounts of sodium oxide (Na2O), sodium carbonate (Na2CO3), and calcium oxide (CaO). To make it work as a touch screen, the upper surface is coated with indium tin oxide. When you touch the screen you’re actually pushing a thin layer of polycarbonate plastic (made from petroleum) against the coated glass — completing an electrical circuit. The computer is then able to figure out exactly where you touched the screen. Computer processors are made from silica wafers (more quartz) and also include a significant amount of copper and gold. Gold is used because it is a better conductor than copper and doesn’t tarnish the way silver or copper does. Most computers have nickel-metal-hydride (NiMH) batteries, which contain nickel, of course, along with cadmium, cobalt, manganese, aluminum, and the rare-earth elements lanthanum, cerium, neodymium, and praseodymium. The processor and other electronic components are secured to a circuit board, which is a thin layer of fibreglass sandwiched between copper sheets coated with small amounts of tin and lead. Various parts are put together with steel screws that are made of iron and molybdenum.

Figure 20.2 The main components of a tablet computer [SE, base photograph from https://upload.wikimedia.org/wikipedia/commons/8/8d/IPad_Air.png]
Figure 18.2 The main components of a tablet computer [SE, base photograph from https://upload.wikimedia.org/wikipedia/commons/8/8d/IPad_Air.png]

That’s not everything that goes into a tablet computer, but to make just those components we need a pure-silica sand deposit, a salt mine for sodium, a rock quarry for calcium, an oil well, a gas well, an aluminum mine, an iron mine, a manganese mine, a copper-molybdenum-gold mine, a cobalt-nickel mine, a rare-earth element and indium mine, and a source of energy to transport all of the materials, process them, put them together, and finally transport the computer to your house or the store where you bought it.

Exercise 18.1 Where Does It Come From?

Look around you and find at least five objects (other than a computer or a phone) that have been made from materials that had to be mined, quarried, or extracted from an oil or gas well. Try to identify the materials involved, and think about where they might have come from. This pen is just an example.

[https://upload.wikimedia.org/wikipedia/commons/f/fd/Ballpoint-pen-parts.jpg]
[https://upload.wikimedia.org/wikipedia/commons/f/fd/Ballpoint-pen-parts.jpg]

122

18.2 Metal Deposits

Mining has always been a major part of Canada’s economy. Canada has some of the largest mining districts and deposits in the world, and for the past 150 years, we have been one of the most important suppliers of metals. Extraction of Earth’s resources goes back a long way in Canada. For example, the First Nations of British Columbia extracted obsidian from volcanic regions for tools and traded it up and down the coast. In the 1850s, gold was discovered in central British Columbia, and in the 1890s, even more gold was discovered in the Klondike area of Yukon. These two events were critical to the early development of British Columbia, Yukon, and Alaska.

Canada’s mining sector had revenues in the order of $37 billion in 2013. The majority of that was split roughly equally among gold, iron, copper, and potash, with important but lesser amounts from nickel and diamonds (Figure 18.3). Revenues from the petroleum sector are significantly higher, at over $100 billion per year.

Figure 20.3 The value of various Canadian mining sectors in 2013 [SE from data at http://www.nrcan.gc.ca/mining-materials/publications/8772]
Figure 18.3 The value of various Canadian mining sectors in 2013 [SE from data at http://www.nrcan.gc.ca/mining-materials/publications/8772]

 

A metal deposit is a body of rock in which one or more metals have been concentrated to the point of being economically viable for recovery. Some background levels of important metals in average rocks are shown on Table 18.1, along with the typical grades necessary to make a viable deposit, and the corresponding concentration factors. Looking at copper, for example, we can see that while average rock has around 40 ppm (parts per million) of copper, a grade of around 10,000 ppm or 1% is necessary to make a viable copper deposit. In other words, copper ore has about 250 times as much copper as typical rock. For the other elements in the list, the concentration factors are much higher. For gold, it’s 2,000 times and for silver it’s around 10,000 times.

Metal Typical Background Level Typical Economic Grade* Concentration Factor
Copper 40 ppm 10,000 ppm (1%) 250 times
Gold 0.003 ppm 6 ppm (0.006%) 2,000 times
Lead 10 ppm 50,000 ppm (5% 5,000 times
Molybdenum 1 ppm 1,000 ppm (0.1%) 1,000 times
Nickel 25 ppm 20,000 ppm (2%) 800 times
Silver 0.1 ppm 1,000 ppm (0.1%) 10,000 times
Uranium 2 ppm 10,000 ppm (1%) 5,000 times
Zinc 50 ppm 50,000 ppm (5%) 1,000 times
*It’s important to note that the economic viability of any deposit depends on a wide range of factors including its grade, size, shape, depth below the surface, and proximity to infrastructure, the current price of the metal, the labour and environmental regulations in the area, and many other factors.

Table 18.1 Typical background and ore levels of some important metals [SE]

It is clear that some very significant concentration must take place to form a mineable deposit. This concentration may occur during the formation of the host rock, or after the rock forms, through a number of different types of processes. There is a very wide variety of ore-forming processes, and there are hundreds of types of mineral deposits. The origins of a few of them are described below.

Types of Metal Deposits

Magmatic Nickel Deposits

A magmatic deposit is one in which the metal concentration takes place primarily at the same time as the formation and emplacement of the magma. Most of the nickel mined in Canada comes from magmatic deposits such as those in Sudbury (Ontario), Thompson (Manitoba) (Figure 18.4), and Voisey’s Bay (Labrador). The magmas from which these deposits form are of mafic or ultramafic composition (derived from the mantle), and therefore they have relatively high nickel and copper contents to begin with (as much as 100 times more than normal rocks in the case of nickel). These elements may be further concentrated within the magma as a result of the addition of sulphur from partial melting of the surrounding rocks. The heavy nickel and copper sulphide minerals are then concentrated further still by gravity segregation (i.e., crystals settling toward the bottom of the magma chamber). In some cases, there are significant concentrations of platinum-bearing minerals.

Most of these types of deposits around the world are Precambrian in age — probably because the mantle was significantly hotter at that time, and the necessary mafic and ultramafic magmas were more likely to be emplaced in the continental crust.

Figure 20.4 The nickel smelter at Thompson, Manitoba [https://en.wikipedia.org/wiki/Thompson,_Manitoba#/media/File:Vale_Nickel_Mine.JPG]
Figure 18.4 The nickel smelter at Thompson, Manitoba [https://en.wikipedia.org/wiki/Thompson,_Manitoba#/media/File:Vale_Nickel_Mine.JPG]

 

Volcanogenic Massive Sulphide Deposits

Much of the copper, zinc, lead, silver, and gold mined in Canada is mined from volcanichosted massive sulphide (VHMS) deposits associated with submarine volcanism (VMS deposits). Examples are the deposits at Kidd Creek, Ontario, Flin Flon on the Manitoba-Saskatchewan border, Britannia on Howe Sound, and Myra Falls (within Strathcona Park) on Vancouver Island.

VMS deposits are formed from the water discharged at high temperature (250° to 300°C) at ocean-floor hydrothermal vents, primarily in areas of subduction-zone volcanism. The environment is comparable to that of modern-day black smokers (Figure 18.5), which form where hot metal- and sulphide-rich water issues from the sea floor. They are called massive sulphide deposits because the sulphide minerals (including pyrite (FeS2) , sphalerite (ZnS), chalcopyrite (CuFeS2), and galena (PbS)) are generally present in very high concentrations (making up the majority of the rock in some cases). The metals and the sulphur are leached out of the sea-floor rocks by convecting groundwater driven by the volcanic heat, and then quickly precipitated where that hot water enters the cold seawater, causing it to cool suddenly and change chemically. The volcanic rock that hosts the deposits is formed in the same area and at the same general time as the accumulation of the ore minerals.

Figure 20.5 Left: A black smoker on the Juan de Fuca Ridge off the west coast of Vancouver Island. Right: A model of the formation of a volcanogenic massive sulphide deposit on the sea floor. [left: NOAA at: http://oceanexplorer.noaa.gov/okeanos/explorations/10index/background/plumes/media/black_smoker.html, right: SE]
Figure 18.5 Left: A black smoker on the Juan de Fuca Ridge off the west coast of Vancouver Island. Right: A model of the formation of a volcanogenic massive sulphide deposit on the sea floor. [left: NOAA at: http://oceanexplorer.noaa.gov/okeanos/explorations/10index/background/plumes/media/black_smoker.html, right: SE]


Porphyry Deposits

Porphyry deposits are the most important source of copper and molybdenum in British Columbia, the western United States, and Central and South America. Most porphyry deposits also host some gold, which may be, in rare cases, the primary commodity. B.C. examples include several large deposits within the Highland Valley mine (Figure 18.1) and numerous other deposits scattered around the central part of the province.

A porphyry deposit forms around a cooling felsicstock in the upper part of the crust. They are called “porphyry” because upper crustal stocks are typically porphyritic in texture, the result of a two-stage cooling process. Metal enrichment results in part from convection of groundwater related to the heat of the stock, and also from metal-rich hot water expelled by the cooling magma (Figure 18.6). The host rocks, which commonly include the stock itself and the surrounding country rocks, are normally highly fractured and brecciated. During the ore-forming process, some of the original minerals in these rocks are altered to potassium feldspar, biotite, epidote, and various clay minerals. The important ore minerals include chalcopyrite (CuFeS2), bornite (Cu5FeS4), and pyrite in copper porphyry deposits, or molybdenite (MoS2) and pyrite in molybdenum porphyry deposits. Gold is present as minute flakes of native gold.

This type of environment (i.e., around and above an intrusive body) is also favourable for the formation of other types of deposits — particularly vein-type gold deposits (a.k.a. epithermal deposits). Some of the gold deposits of British Columbia (such as in the Eskay Creek area adjacent to the Alaska panhandle), and many of the other gold deposits situated along the western edge of both South and North America are of the vein type shown in Figure 18.6, and are related to nearby magma bodies.

Figure 20.6 A model for the formation of a porphyry deposit around an upper-crustal porphyritic stock and associated vein deposits. [SE]
Figure 18.6 A model for the formation of a porphyry deposit around an upper-crustal porphyritic stock and associated vein deposits. [SE]

 

Banded Iron Formation

Most of the world’s major iron deposits are of the banded iron formation type (classified as a type of chemical sedimentary rock), and most of these formed during the initial oxygenation of Earth’s atmosphere between 2,400 and 1,800 Ma. At that time, iron that was present in dissolved form in the ocean (as Fe2+) became oxidized to its insoluble form (Fe3+) and accumulated on the sea floor, mostly as hematite interbedded with chert (Figure 18.7). Unlike many other metals, which are economically viable at grades of around 1% or even much less, iron deposits are only viable if the grades are in the order of 50% iron and if they are very large.

Figure 18.7 Banded iron formation from an unknown location in North America on display at a museum in Germany. The rock is about 2 m across. The dark grey layers are magnetite and the red layers are hematite. Chert is also present. Source: https://upload.wikimedia.org/wikipedia/commons/5/5f/Black-band_ironstone_%28aka%29.jpg

Unconformity-Type Uranium Deposits

Figure 20.8 Model of the formation of unconformity-type uranium deposits of the Athabasca Basin, Saskatchewan [SE]
Figure 18.8 Model of the formation of unconformity-type uranium deposits of the Athabasca Basin, Saskatchewan [SE]

There are several different types of uranium deposits, but some of the largest and richest are those within the Athabasca Basin of northern Saskatchewan. These are called unconformity-type uranium deposits because they are all situated very close to the unconformity between the Proterozoic Athabasca Group sandstone and the much older Archean sedimentary, volcanic, and intrusive igneous rock (Figure 18.8). The origin of unconformity-type U deposits is not perfectly understood, but it is thought that two particular features are important: (1) the relative permeability of the Athabasca Group sandstone, and (2) the presence of graphitic schist within the underlying Archean rocks. The permeability of the sandstone allowed groundwater to flow through it and leach out small amounts of U, which stayed in solution in the oxidized form U6+. The graphite (C) created a reducing environment (non-oxidizing) that converted the U from U6+ to insoluble U4+, at which point it was precipitated as the mineral uraninite (UO2).

Exercise 18.2 The Importance of Heat and Heat Engines

For a variety of reasons, thermal energy (heat) from within Earth is critical in the formation of many types of ore deposits. Look back through the deposit-type descriptions above and complete the following table, describing which of those deposit types depend on a source of heat for their formation, and why.

Deposit Type Is Heat a Factor? If So, What Is the Role of the Heat?
Magmatic
Volcanogenic massive sulphide
Porphyry
Banded iron formation
Unconformity-type uranium

Mining and Mineral Processing

Metal deposits are mined in a variety of different ways depending on their depth, shape, size, and grade. Relatively large deposits that are quite close to the surface and somewhat regular in shape are mined using open-pit mine methods (Figure 18.1). Creating a giant hole in the ground is generally cheaper than making an underground mine, but it is also less precise, so it is necessary to mine a lot of waste rock along with the ore. Relatively deep deposits or those with elongated or irregular shapes are typically mined from underground with deep vertical shafts, declines (sloped tunnels), and levels (horizontal tunnels) (Figures 18.9 and 18.10). In this way, it is possible to focus the mining on the ore body itself. However, with relatively large ore bodies, it may be necessary to leave some pillars to hold up the roof.

Figure 20.9 Underground at the Myra Falls Mine, Vancouver Island. [SE]
Figure 18.9 Underground at the Myra Falls Mine, Vancouver Island. [SE]

 

Figure 20.10 Schematic cross-section of a typical underground mine. [SE]
Figure 18.10 Schematic cross-section of a typical underground mine. [SE]

 

In many cases, the near-surface part of an ore body is mined with an open pit, while the deeper parts are mined underground (Figures 18.10 and 18.11).

Figure 20.11 Entrance to an exploratory decline (white arrow) for the New Afton Mine situated in the side of the open pit of the old Afton Mine, near Kamloops, B.C. [SE]
Figure 18.11 Entrance to an exploratory decline (white arrow) for the New Afton Mine situated in the side of the open pit of the old Afton Mine, near Kamloops, B.C. [SE]

A typical metal deposit might contain a few percent of ore minerals (e.g., chalcopyrite or sphalerite), mixed with the minerals of the original rock (e.g., quartz or feldspar). Other sulphide minerals are commonly present within the ore, especially pyrite.

When ore is processed (typically very close to the mine), it is ground to a fine powder and the ore minerals are physically separated from the rest of the rock to make a concentrate. At a molybdenum mine, for example, this concentrate may be almost pure molybdenite (MoS2). The rest of the rock is known as tailings. It comes out of the concentrator as a wet slurry and must be stored near the mine, in most cases, in a tailings pond.

The tailings pond at the Myra Falls Mine on Vancouver Island is shown in Figure 18.12, and the settling ponds for waste water from the concentrator are shown in Figure 18.13. The tailings are contained by an embankment. Also visible in the foreground of Figure 18.12 is a pile of waste rock, which is non-ore rock that was mined in order to access the ore. Although this waste rock contains little or no ore minerals, at many mines it contains up to a few percent pyrite. The tailings and the waste rock at most mines are an environmental liability because they contain pyrite plus small amounts of ore minerals. When pyrite is exposed to oxygen and water, it generates sulphuric acid — also known as acid rock drainage (ARD). Acidity itself is a problem to the environment, but because the ore elements, such as copper or lead, are more soluble in acidic water than neutral water, ARD is also typically quite rich in metals, many of which are toxic.

Figure 20.12 The tailings pond at the Myra Falls Mine on Vancouver Island. The dry rock in the middle of the image is waste rock. The structure on the right is the headframe for the mine shaft. Myra Creek flows between the tailings pond and the headframe. [SE]
Figure 18.12 The tailings pond at the Myra Falls Mine on Vancouver Island. The dry rock in the middle of the image is waste rock. The structure on the right is the headframe for the mine shaft. Myra Creek flows between the tailings pond and the headframe. [SE]

 

Figure 20.13 The tailings pond (lower left) at Myra Falls Mine with settling ponds (right) for processing water from the concentrator. [SE]
Figure 18.13 The tailings pond (lower left) at Myra Falls Mine with settling ponds (right) for processing water from the concentrator. [SE]

Tailings ponds and waste-rock storage piles must be carefully maintained to ensure their integrity and monitored to ensure that acidic and metal-rich water is not leaking out. In August 2014, the tailings pond at the Mt. Polley Mine in central B.C. failed and 10 million cubic metres of waste water along with 4.5 million cubic metres of tailings slurry was released into Polley Lake, Hazeltine Creek, and Quesnel Lake (Figure 18.14, a and b). As of July 2015, the environmental implications of this event are still not fully understood.

Figure 20.14a The Mt. Polley Mine area prior to the dam breach of August 2014. The tailings were stored in the area labelled “retention basin.” [https://en.wikipedia.org/wiki/Mount_Polley_mine_disaster]
Figure 18.14a The Mt. Polley Mine area prior to the dam breach of August 2014. The tailings were stored in the area labelled “retention basin.” [https://en.wikipedia.org/wiki/Mount_Polley_mine_disaster]

 

Figure 20.14b The Mt. Polley Mine area after the tailings dam breach of August 2014. The water and tailings released flowed into Hazeltine Creek, and Polley and Quesnel Lakes. [https://en.wikipedia.org/wiki/Mount_Polley_mine_disaster]
Figure 18.14b The Mt. Polley Mine area after the tailings dam breach of August 2014. The water and tailings released flowed into Hazeltine Creek, and Polley and Quesnel Lakes. [https://en.wikipedia.org/wiki/Mount_Polley_mine_disaster]

 

Most mines have concentrators on site because it is relatively simple to separate ore minerals from non-ore minerals and thus significantly reduce the costs and other implications of transportation. But separation of ore minerals is only the preliminary stage of metal refinement, for most metals the second stage involves separating the actual elements within the ore minerals. For example, the most common ore of copper is chalcopyrite (CuFeS2). The copper needs to be separated from the iron and sulphur to make copper metal and that involves complicated and very energy-intensive processes that are done at smelters or other types of refineries. Because of their cost and the economies of scale, there are far fewer refineries than there are mines.

There are several metal refineries (including smelters) in Canada; some examples are the aluminum refinery in Kitimat, B.C. (which uses ore from overseas); the lead-zinc smelter in Trail, B.C.; the nickel smelter at Thompson, Manitoba; numerous steel smelters in Ontario, along with several other refining operations for nickel, copper, zinc, and uranium; aluminum refineries in Quebec; and a lead smelter in New Brunswick.

123

18.3 Industrial Minerals

Metals are critical for our technological age, but there are a lot of other not-so-shiny materials that are needed to facilitate our way of life. For everything made out of concrete or asphalt, we need sand and gravel. To make the cement that holds concrete together, we also need limestone. For the glass in our computer screens and for glass-sided buildings, we need silica sand plus sodium oxide (Na2O), sodium carbonate (Na2CO3), and calcium oxide (CaO). Potassium is an essential nutrient for farming in many areas, and for a wide range of applications (e.g., ceramics and many industrial processes), we also need various types of clay.

The best types of aggregate (sand and gravel) resources are those that have been sorted by streams, and in Canada the most abundant and accessible fluvial deposits are associated with glaciation. That doesn’t include till of course, because it has too much silt and clay, but it does include glaciofluvial outwash, which is present in thick deposits in many parts of the country, similar to the one shown in Figure 18.15. In a typical gravel pit, these materials are graded on-site according to size and then used in a wide range of applications from constructing huge concrete dams to filling children’s sandboxes. Sand is also used to make glass, but for most types of glass, it has to be at least 95% quartz (which the sandy layers shown in Figure 18.15 are definitely not), and for high-purity glass and the silicon wafers used for electronics, the source sand has to be over 98% quartz.

Figure 20.15 Sand and gravel in an aggregate pit near Nanaimo, BC. [SE]
Figure 18.15 Sand and gravel in an aggregate pit near Nanaimo, BC. [SE]

Approximately 80 million tonnes of concrete are used in Canada each year — a little over 2 tonnes per person. The cement used for concrete is made from approximately 80% calcite (CaCO3) and 20% clay. This mixture is heated to 1450°C to produce the required calcium silicate compounds (e.g., Ca2SiO4). The calcite typically comes from limestone quarries like the one on Texada Island, B.C. (Figure 18.16). Limestone is also used as the source material for many other products that require calcium compounds, including steel and glass, pulp and paper, and plaster products for construction.

Figure 20.16 Triassic Quatsino Formation limestone being quarried on Texada Island, B.C. [SE]
Figure 18.16 Triassic Quatsino Formation limestone being quarried on Texada Island, B.C. [SE]

Sodium is required for a wide range of industrial processes, and the most convenient source is sodium chloride (rock salt), which is mined from evaporite beds in various parts of Canada. The largest salt mine in the world is at Goderich, Ontario, where salt is recovered from the 100 m thick Silurian Salina Formation. The same formation is mined in the Windsor area. Rock salt is also used as a source of sodium and chlorine in the chemical industry to melt ice on roads, as part of the process of softening water, and as a seasoning. Under certain conditions, the mineral sylvite (KCl) accumulates in evaporite beds, and this rock is called potash. This happened across the Canadian prairies during the Devonian, creating the Prairie evaporite formation (Figure 6.16). Potassium is used as a crop fertilizer, and Canada is the world’s leading supplier, with most of that production coming from Saskatchewan.

Another evaporite mineral, gypsum (CaSO4.2H20), is the main component of plasterboard (drywall) that is widely used in the construction industry. One of the main mining areas for gypsum in Canada is in the Milford Station area of Nova Scotia, site of the world’s largest gypsum mine.

Rocks are quarried or mined for many different uses, such as building facades (Figure 18.17), countertops, stone floors, and headstones. In most of these cases, the favoured rock types are granitic rocks, slate, and marble. Quarried rock is also used in some applications where rounded gravel isn’t suitable, such as the ballast (road bed) for railways, where crushed angular rock is needed.

Figure 20.17 Slate used as a facing material on a concrete building column in Vancouver [SE]
Figure 18.17 Slate used as a facing material on a concrete building column in Vancouver [SE]

Exercise 18.3 Sources of Important Lighter Minerals

When we think of the manufacture of consumer products, plastics and the heavy metals (copper, iron, lead, zinc) easily come to mind, but we often forget about some of the lighter metals and non-metals that are important. Consider the following elements and determine their sources. Answers for all of these except magnesium are given above. See if you can figure out a likely mineral source of magnesium.

Silicon:

Calcium:

Sodium:

Potassium:

Magnesium:

124

18.4 Fossil Fuels

There are numerous types of fossil fuels, but all of them involve the storage of organic matter in sediments or sedimentary rocks. Fossil fuels are rich in carbon and almost all of that carbon ultimately originates from CO2 taken out of the atmosphere during photosynthesis. That process, driven by solar energy, involves reduction (the opposite of oxidation) of the carbon, resulting in it being combined with hydrogen instead of oxygen. The resulting organic matter is made up of complex and varied carbohydrate molecules.

Most organic matter is oxidized back to CO2 relatively quickly (within weeks or years in most cases), but any of it that gets isolated from the oxygen of the atmosphere (for example, deep in the ocean or in a stagnant bog) may last long enough to be buried by sediments and, if so, may be preserved for tens to hundreds of millions of years. Under natural conditions, that means it will be stored until those rocks are eventually exposed at the surface and weathered.

In this section, we’ll discuss the origins and extraction of the important fossils fuels, including coal, oil, and gas.

Coal

Coal was the first fossil fuel to be widely used. As we learned in section 9.4, coal forms where vigorous growth of vegetation in swampy areas leads to an abundance of organic matter that accumulates within stagnant water. The chemical nature of the water- it being acidic and having little to no oxygen- means the organic matter decays very little. If a thick layer of organic matter is accumulated and then buried, the organic matter begins to change as it is compressed and heated. This situation, where the dead organic matter is submerged in oxygen-poor water, must be maintained for centuries to millennia in order for enough material to accumulate to form a thick layer. Over time, as the organic matter is heated and compressed more and more, the carbon within it becomes concentrated, and it can provide more energy when it is burned. Figure 18.18 shows the classification system for different grades of coal. Increasing pressure and temperature means proceeding clockwise through the diagram, starting with lignite (the lowest grade), then bituminous coal, and finally anthracite, the highest grade.

Coal ranking system used by the United States Geological Survey (USGS). As vegetative organic matter is buried deeper, and experiences higher pressures and temperatures, it progresses clockwise through the diagram, beginning with lignite. Additional heat and pressure result in the coal having a higher concentration of carbon (the vertical axis), and producing more energy (the horizontal axis). [USGS, public domain, https://commons.wikimedia.org/wiki/File:Coal_Rank_USGS.png#mw-jump-to-license]
Figure 18.18 Coal ranking system used by the United States Geological Survey (USGS). As vegetative organic matter is buried deeper, and experiences higher pressures and temperatures, it progresses clockwise through the diagram, beginning with lignite. Additional heat and pressure result in the coal having a higher concentration of carbon (the vertical axis), and producing more energy (the horizontal axis). [USGS, public domain, https://commons.wikimedia.org/wiki/File:Coal_Rank_USGS.png#mw-jump-to-license]

 

There are significant coal deposits in many parts of Canada, including the Maritimes, Ontario, Saskatchewan, Alberta, and British Columbia. In Alberta and Saskatchewan, much of the coal is used for electricity generation. Coal from the Highvale Mine (Figure 18.19), Canada’s largest, is used to feed the Sundance and Keephills power stations west of Edmonton. Almost all of the coal mined in British Columbia is exported for use in manufacturing steel.

Figure 20.19 The Highvale Mine (background) and the Sundance (right) and Keephills (left) generating stations on the southern shore of Wabamun Lake, Alberta [SE]
Figure 18.19 The Highvale Mine (background) and the Sundance (right) and Keephills (left) generating stations on the southern shore of Wabamun Lake, Alberta [SE]

Oil and Gas

While almost all coal forms on land from terrestrial vegetation, most oil and gas is derived primarily from marine micro-organisms that accumulate within sea-floor sediments. In areas where marine productivity is high, dead organic matter is delivered to the sea floor fast enough that some of it escapes oxidation. This material accumulates in the muddy sediments, which become buried to significant depth beneath other sediments.

As the depth of burial increases, so does the temperature — due to the geothermal gradient — and gradually the organic matter within the sediments is converted to hydrocarbons (Figure 18.20). The first stage is the biological production (involving anaerobic bacteria) of methane. Most of this escapes back to the surface, but some is trapped in methane hydrates near the sea floor. At depths beyond about 2 km, and at temperatures ranging from 60° to 120°C, the organic matter is converted by chemical processes to oil. This depth and temperature range is known as the oil window. Beyond 120°C most of the organic matter is chemically converted to methane.

Figure 20.20 The depth and temperature limits for biogenic gas, oil, and thermogenic gas [SE]
Figure 18.20 The depth and temperature limits for biogenic gas, oil, and thermogenic gas [SE]

The organic matter-bearing rock within which the formation of gas and oil takes place is known to petroleum geologists as the source rock. Both liquid oil and gaseous methane are lighter than water, so as liquids and gases form, they tend to move slowly toward the surface, out of the source rock and into reservoir rocks. Reservoir rocks are typically relatively permeable because that allows migration of the fluids from the source rocks, and also facilitates recovery of the oil or gas. In some cases, the liquids and gases make it all the way to the surface, where they are oxidized, and the carbon is returned to the atmosphere. But in other cases, they are contained by overlying impermeable rocks (e.g., mudrock) in situations where anticlines, faults, stratigraphy changes, and reefs or salt domes create traps (Figure 18.21).

Figure 20.21 Migration of oil and gas from source rocks into traps in reservoir rocks [SE]
Figure 18.21 Migration of oil and gas from source rocks into traps in reservoir rocks [SE]

The liquids and gases that are trapped within reservoirs become separated into layers based on their density, with gas rising to the top, oil below it, and water underneath the oil. The proportions of oil and gas depend primarily on the temperature in the source rocks. Some petroleum fields, such as many of those in Alberta, are dominated by oil, while others, notably those in northeastern B.C., are dominated by gas.

Figure 20.22 Seismic section through the East Breaks Field in the Gulf of Mexico. The dashed red line marks the approximate boundary between deformed rocks and younger undeformed rocks. The wiggly arrows are interpreted migration paths. The total thickness of this section is approximately 5 km. [SE after http://wiki.aapg.org/File:Sedimentary-basin-analysis_fig4-55.png]
Figure 18.22 Seismic section through the East Breaks Field in the Gulf of Mexico. The dashed red line marks the approximate boundary between deformed rocks and younger undeformed rocks. The wiggly arrows are interpreted migration paths. The total thickness of this section is approximately 5 km. [SE after http://wiki.aapg.org/File:Sedimentary-basin-analysis_fig4-55.png]

In general, petroleum fields are not visible from the surface, and their discovery involves the search for structures in the subsurface that have the potential to form traps. Seismic surveys are the most commonly used tool for early-stage petroleum exploration, as they can reveal important information about the stratigraphy and structural geology of subsurface sedimentary rocks. An example from the Gulf of Mexico south of Texas is shown in Figure 18.22. In this area, a thick evaporite deposit (“salt”) has formed domes because salt is lighter than other sediments and tends to rise slowly toward the surface; this has created traps. The sequence of deformed rocks is capped with a layer of undeformed rock.

Exercise 18.4 Interpreting a Seismic Profile

The cross-section shown here is from a ship-borne seismic survey in the Bering Sea off the west coast of Alaska. As a petroleum geologist, it’s your job to pick two separate locations to drill for oil or gas. Which locations would you choose?

Interpreting a Seismic Profile

[from USGS at: http://walrus.wr.usgs.gov/infobank/programs/html/definition/seis.html]

What Is Unconventional Oil and Gas?

The type of oil and gas reservoirs illustrated in Figures 18.21 and 18.22 are described as conventional reserves. Some unconventional types of oil and gas include oil sands, shale gas, and coal-bed methane.

Oil Sands

Oil sands are important because the reserves in Alberta are so large (the largest single reserve of oil in the world), but they are very controversial from an environmental and social perspective. They are “unconventional” because the oil is exposed near the surface and is highly viscous because of microbial changes that have taken place at the surface. The hydrocarbons that form this reserve originated in deeply buried Paleozoic rocks adjacent to the Rocky Mountains and migrated up and toward the east (Figure 18.23).

The oil sands are controversial primarily because of the environmental cost of their extraction. Since the oil is so viscous, it requires heat to make it sufficiently liquid to process. This energy comes from gas; approximately 25 m3 of gas is used to produce 0.16 m3 (one barrel) of oil. (That’s not quite as bad as it sounds, as the energy equivalent of the required gas is about 20% of the energy embodied in the produced oil.) The other environmental cost of oil sands production is the devastation of vast areas of land where strip-mining is taking place and tailings ponds are constructed, and the unavoidable release of contaminants into the groundwater and rivers of the region.

At present, most oil recovery from oil sands is achieved by mining the sand and processing it on site. Exploitation of oil sand that is not exposed at the surface depends on in situ processes, an example being the injection of steam into the oil-sand layer to reduce the viscosity of the oil so that it can be pumped to the surface.

Figure 20.23 Schematic cross-section of northern Alberta showing the source rocks and location of the Athabasca Oil Sands [SE]
Figure 18.23 Schematic cross-section of northern Alberta showing the source rocks and location of the Athabasca Oil Sands [SE]

Shale Gas

Shale gas is gas that is trapped within rock that is too impermeable for the gas to escape under normal conditions, and it can only be extracted by fracturing the reservoir rock using water and chemicals under extremely high pressure. This procedure is known as hydraulic fracturing or “fracking.” Fracking is controversial because of the volume of water used, and because, in some jurisdictions, the fracking companies are not required to disclose the nature of the chemicals used. Although fracking is typically done at significant depths, there is always the risk that overlying water-supply aquifers could be contaminated (Figure 18.24). Fracking also induces low-level seismicity (earthquakes).

Figure 20.24 Depiction of the process of directional drilling and fracking to recover gas from impermeable rocks. The light blue arrows represent the potential for release of fracking chemicals to aquifers. [by SE, after https://en.wikipedia.org/wiki/Hydraulic_fracturing#/media/File:HydroFrac2.svg]
Figure 18.24 Depiction of the process of directional drilling and fracking to recover gas from impermeable rocks. The light blue arrows represent the potential for release of fracking chemicals to aquifers. [by SE, after https://en.wikipedia.org/wiki/Hydraulic_fracturing#/media/File:HydroFrac2.svg]

During the process that converts organic matter to coal, some methane is produced, which is stored within the pores of the coal. When coal is mined, methane is released into the mine where it can become a serious explosion hazard. Modern coal-mining machines have methane detectors on them and actually stop operating if the methane levels are dangerous. It is possible to extract the methane from coal beds without mining the coal; gas recovered this way is known as coal-bed methane.

125

18.5 Diamonds

Although Canada’s diamond mining industry didn’t get started until 1998, diamonds are currently the sixth most valuable product mined in the country (Figure 18.3), and Canada ranks sixth in the world in diamond production. Diamonds form deep in the mantle (approximately 200 km to 250 km depth) under very specific pressure and temperature conditions, from carbon that is naturally present in mantle rock (not from coal). The diamond-bearing rock is brought to the surface coincidentally via a type of volcanism that is extremely rare (the most recent kimberlite eruption is thought to have taken place 10,000 years ago and prior to that at around 30 Ma). There is more on the volcanology of kimberlites in section 11.3. All of the world’s kimberlite diamond deposits are situated within ancient shield areas (cratons) in Africa, Australia, Russia, South America, and North America.

It has long been known that diamonds could exist within the Canadian Shield, but up until 1991, exploration efforts had been unsuccessful. In 1980 two geologists, Chuck Fipke and Stu Blusson, started searching in the Northwest Territories by sampling glacial sediments looking for some of the minerals that are normally quite abundant within kimberlites: chromium-bearing garnet, chromium-bearing pyroxene, chromite (Cr2O3), and ilmenite (FeTiO3). These distinctive minerals are used for this type of exploration because they are many times more abundant in kimberlite than diamond is. After more than a decade of exploration, Fipke and Blusson finally focused their search on an area 250 km northeast of Yellowknife, and, in 1991, they announced the discovery of a diamond-bearing kimberlite body at Lac de Gras. That discovery is now the Diavik Mine, and there is another diamond mine — Ekati — 25 km to the northwest (Figure 18.25). There are two separate mines at Diavik accessing three different kimberlite bodies, and there are five at Ekati. See Figure 11.22 for a close-up view of the Ekati Mine. There are six operating diamond mines in Canada, four in the Northwest Territories (including Diavik and Ekati), and one each in Nunavut and Ontario.

Figure 20.25 Diamond mines in the Lac de Gras region, Nunavut. The twin pits of the Diavik Mine are visible in the lower right on an island within Lac de Gras. The five pits of the Ekati mine are also visible, on the left and the upper right. The two main mine centres are 25 km apart. [http://earthobservatory.nasa.gov/IOTD/view.php?id=84085&src=eoa-iotd]
Figure 18.25 Diamond mines in the Lac de Gras region, Nunavut. The twin pits of the Diavik Mine are visible in the lower right on an island within Lac de Gras. The five pits of the Ekati mine are also visible, on the left and the upper right. The two main mine centres are 25 km apart. [http://earthobservatory.nasa.gov/IOTD/view.php?id=84085&src=eoa-iotd]

126

Chapter 18 Summary

The main topics of this chapter can be summarized as follows:

18.1 If You Can’t Grow It, You Have To Mine It

Geological resources are critical to our way of life and important to the Canadian economy. Gold, iron, copper, nickel, and potash are Canada’s most valuable mined commodities.

18.2 Metal Deposits

The proportions of metals in mineral deposits are typically several thousand times higher than those in average rocks, and special processes are required to extract the valuable content. Some deposits form through processes within a magma chamber, others during volcanism or adjacent to a stock, and some are related to sedimentary processes. Mining involves both surface and underground methods, but in either case, rock is brought to surface that can react with water and oxygen to produce acid rock drainage and metal contamination.

18.3 Industrial Materials

Non-metallic materials are very important to infrastructure and agriculture. Some of the major industrial minerals include sand and gravel, limestone for cement and agriculture, salt for a range of applications, potash fertilizer, and decorative stone.

18.4 Fossil Fuels

The main fossil fuels are coal, oil, and gas. Coal forms on land in wet environments where organic matter can remain submerged and isolated from oxygen for millennia before it is buried by more sediments. The depth of that burial influences the grade of coal produced. Oil and gas originate from organisms living in marine environments, and again, fairly rapid burial is required to preserve the organic matter on the sea floor. At moderate burial depth (2 km to 4 km), oil is produced, and at greater depth, gas is produced. Both oil and gas migrate toward the surface and can be trapped beneath impermeable rock layers in structural features, such as anticlines or faults. Some unconventional fossil fuel resources include oil sands, shale gas, and coal-bed methane.

18.5 Diamonds

Diamonds originate in the mantle and are only brought to the surface by the very rare eruption of kimberlitic volcanoes. The relatively recent discovery of diamonds in Canada was based on the exhaustive search for diamond indicator minerals in glacial sediments. There are now six diamond mines in Canada.

Questions for Review

1. What are some of Earth’s resources that are needed to make a compact fluorescent light bulb? The image shows the contents of the ballast.

https://upload.wikimedia.org/wikipedia/commons/3/31/06_Spiral_CFL_Bulb_2010-03-08_%28white_back%29.jpg
https://upload.wikimedia.org/wikipedia/commons/3/31/06_Spiral_CFL_Bulb_2010-03-08_%28white_back%29.jpg
https://en.wikipedia.org/wiki/ Compact_fluorescent_lamp #/ media/File:Elektronstarterp.jpg
https://en.wikipedia.org/wiki/ Compact_fluorescent_lamp
#/ media/File:Elektronstarterp.jpg

2. Explain why nickel deposits are associated only with mafic magma and not with intermediate or felsic magma?

3. What is the composition of the black smoke in a black smoker, and how does that relate to a volcanogenic massive sulphide deposit?

4. How might an epigenetic gold deposit be related to a porphyry deposit?

5. Oxidation and reduction processes are important to both banded iron formation deposits and unconformity-type uranium deposits. Explain the role in each case.

6. A typical kimberlite in northern Canada may look something like the diagram shown. In this case, the diameter at the surface is around 500 m, and the total depth is about 2,500 m. Bearing in mind that an open pit cannot typically be any deeper than it is wide, what mining method(s) might be most applicable to a deposit of this type? kimberlite

7. What mineral is typically responsible for acid rock drainage around mine sites, and why is this mineral so common in this setting?

8. Explain why glaciofluvial gravel is more suitable than till as a source for aggregate.

9. The raw material for making cement is lime (CaO), and this is typically produced by heating limestone (mostly CaCO3) to about 1,000°C. Why is this an environmental issue?

10. Name some important industrial minerals that form in an evaporite setting.

11. If organic matter accumulates at an average rate of 1 mm per year, and if 10 m of organic matter is required to make 1 m of coal, how long must a swampy environment remain stable and wet in order to form a 1.5 m coal seam?

12. What are the ideal characteristics of petroleum source rocks and petroleum reservoir rocks?

13. How deep must the source rocks be buried to produce oil?

14. Why is shale gas an unconventional reserve, and how is it recovered? What are some of the environmental issues associated with that process?

15. If you were sampling glacial deposits to search for kimberlites, why would you be advised to look for kimberlite indicator minerals rather than diamonds?

127

Answers to Chapter 18 Review Questions

1. Some of the components of a compact fluorescent lightbulb (and the resources used to make one) are as follows:

2. Nickel deposits form within mafic and ultramafic igneous bodies because the original magma have relatively high nickel levels to begin with, while intermediate or felsic magma have low levels.

3. The “smoke” in a black smoker is composed of tiny crystals of sulphide minerals. If those include significant quantities of ore minerals like chalcopyrite (CuFeS2), sphalerite (ZnS), and galena (PbS), a VMS deposit could form during this process.

4. A porphyry deposit is situated in the rock around an igneous pluton that has intruded to a relatively high level in the crust (and hence is porphyritic), and they form at least in part from fluids released by the magma. Epigenetic gold deposits may be formed from the same or similar fluids, but are situated at a greater distance from the pluton/

5. Ferrous iron (Fe2+) is soluble in water with a low oxidation potential, and gets converted to insoluble ferric iron (Fe3+) when the water becomes oxidized. The opposite situation happens with uranium. Uranyl uranium (U6+) is soluble under oxidizing conditions, but when the water in which it is dissolved encounters reducing conditions the uranium is converted to the insoluble uranous ion (U4+).

6. It is common for the upper part of a kimberlite to be mined using an open pit (in this case around 500 m wide and up to 500 m deep), and for the lower part to be mined underground.

7. Pyrite (FeS2) is typically responsible for acid rock drainage around mine sites, and it is very common for pyrite to form within the rock at the same time that other metal sulphides (e.g., chalcopyrite) are forming.

8. Glaciofluvial gravels are typically relatively well sorted, and may include clasts ranging in size from coarse sand to pebbles. Till, on the other hand, tends to be poorly sorted and may have clasts ranging from clay to boulders. More processing would be needed to separate the required size ranges, and because till tends to be relatively hard and strong, this would require a lot of effort.

9. During the manufacture of CaO limestone is heated and CO2 is released to the atmosphere, adding to the greenhouse effect. The energy required for this process typically comes from fossil fuels (e.g., natural gas) and the combustion also releases CO2.

10. Some important evaporite minerals include halite (NaCl), sylvite (KCl), and gypsum (CaSO4.2H2O).

11. The 15 m of organic matter required to make 1.5 m of coal, is equivalent to 15,000 mm, and if the organic matter accumulates at 1 mm/y that would require 15,000 years. That organic matter would have to remain submerged in oxygen-poor water for at least that length of time.

12. Petroleum source rocks must have a significant component of organic matter, and then need to be buried to at least 2,500 m depth so that the organic matter can be converted to oil or gas. Reservoir rocks must be both porous and permeable, so that the petroleum liquids can be extracted, and should also take the form of a trap (e.g., an anticline) and capped with impermeable rock.

13. The optimum depth for the generation of oil from buried organic matter is 2,500 to 3,500 m.

14. Shale gas is an unconventional reserve because shale is not permeable enough to allow the gas to be extracted. The rock has to be fractured (fracked) to allow recovery. Fracking involves the use of vast amounts of water, and there is the potential that the fracking fluids can contaminate freshwater aquifers.

15. Kimberlite indicator minerals are much more abundant than diamonds within kimberlites, and so they can typically be detected further away from the kimberlite source, and over a much wider area.

XIX

Chapter 19. Measuring Geological Time

Adapted by Tim Prokopiuk & Karla Panchuk from Physical Geology by Steven Earle

Figure 19.1 Arizona’s Grand Canyon is an icon for geological time; 1,450 million years are represented by this photo. The light-coloured layers of rocks at the top formed at ~ 250 Ma, and the dark ones at the bottom of the canyon at ~ 1,700 Ma. Source: Steven Earle (2015) CC BY 4.0 view source

Learning objectives

After reading this chapter and answering the review questions at the end, you should be able to:

 

Geological Time Is Vast

Time is the dimension that sets geology apart from most other sciences. Geological time is vast, and Earth has changed tremendously during this time. Even though most geological processes are very, very slow, the vast amount of time that has passed has allowed for the formation of extraordinary geological features, as shown in Figure 19.1.

We have numerous ways of measuring geological time. We can tell the relative ages of rocks (e.g., whether one rock is older than another) based on their spatial relationships, we can use fossils to date sedimentary rocks because we have a detailed record of the evolution of life on Earth, and we can use a range of isotopic techniques to determine the absolute ages (in millions of years) of igneous and metamorphic rocks.

But just because we can measure geological time doesn’t mean that we understand it. One of the biggest hurdles faced by geology students—and geologists as well—in understanding geology is to really come to grips with the slow rates at which geological processes happen, and the vast amount of time involved.

128

19.1 The Geological Timescale

James Hutton (1726-1797) was a Scottish geologist, considered by some to be the father of modern Geology. Hutton studied present-day processes and applied his observations to the rock record in order to understand what he saw there. Such a method is now encapsulated in the principle of uniformitarianism, which states that the present is the key to the past. Given that many geological processes that we can see happening around us occur at very slow rates, Hutton concluded that geological time must be very long indeed to account for the large changes apparent in the rock record. But this principle needs to be taken with a grain of salt: there are some processes that have occurred in the past that are no longer occurring (e.g., eruption of ultramafic lavas), as well as some processes that occur so irregularly that we have not yet witnessed such an event in historic time (e.g., impact of a large asteroid with Earth).

William Smith worked as a surveyor in the coal-mining and canal-building industries in south-western England in the late 1700s and early 1800s. While doing his work, he had many opportunities to observe Paleozoic and Mesozoic sedimentary rocks of the region, and he did so in a way that few had done before. Smith noticed the textural similarities and differences between rocks in different locations. More importantly, he discovered that fossils could be used to correlate rocks of the same age. Smith is credited with formulating the principle of faunal succession, the concept that specific types of organisms lived during different time intervals. He used the principle of faunal succession to great effect in his monumental project to create a geological map of England and Wales, published in 1815.

Inset into Smith’s great geological map is a small diagram showing a schematic geological cross-section extending from the Thames estuary of eastern England to the west coast of Wales. Smith showed the sequence of rocks, from the Paleozoic rocks of Wales and western England, through the Mesozoic rocks of central England, to the Cenozoic rocks of the area around London (Figure 19.2).

 

Figure 19.2 William Smith’s “Sketch of the succession of strata and their relative altitudes,” an inset on his geological map of England and Wales (with era names added). Source: Steven Earle (2015) CC BY 4.0 view source, modified after William Smith (1815) Public Domain view map.

Smith did not put any dates on these rocks, because he didn’t know them. But he was aware of the principle of superposition, the idea developed much earlier by the Danish theologian and scientist Nicholas Steno, that young sedimentary rocks form on top of older ones. Therefore, Smith knew that this diagram represented a stratigraphic column. And since almost every period of the Phanerozoic is represented along this section through Wales and England, it is also a primitive geological time scale.

Smith’s work set the stage for the naming and ordering of the geological time periods, which was initiated around 1820, first by British geologists, and later by other European geologists. Many of the periods are named for places where rocks of that age are found in Europe, such as Cambrian for Cambria in Wales, Devonian for Devon in England, Jurassic for the Jura Mountains in France and Switzerland, and Permian for the Perm region of Russia. Some are named for the type of rock that is common during that age, such as Carboniferous for the coal-bearing rocks of England, and Cretaceous for the chalks of England and France.

The early time scales were only relative because 19th century geologists did not know the absolute ages of rocks. This information was not available until the development of isotopic dating techniques early in the 20th century.

The geological timescale is currently maintained by the International Commission on Stratigraphy (ICS), which is part of the International Union of Geological Sciences. The time scale is continuously being updated as we learn more about the timing and nature of past geological events. View the ICS timescale.

Geological time has been divided into four eons: Hadean, Archean, Proterozoic, and Phanerozoic (Figure 19.3). The first three of these eons represent almost 90% of Earth’s history. Rocks from the Phanerozoic (meaning “visible life”) are the most commonly exposed rocks on Earth, and they contain evidence of life forms with which we are familiar.

Figure 19.3 The eons of Earth’s history. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

The Phanerozoic — the past 541 Ma of Earth’s history — is divided into three eras: the Paleozoic (“early life”), the Mesozoic (“middle life”), and the Cenozoic (“new life”), and each era is divided into periods (Figure 19.4). Most of the organisms with which we share Earth evolved into familiar forms at various times during the Phanerozoic.

Figure 19.4 The eras (middle row) and periods (bottom row) of the Phanerozoic. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

The Cenozoic, representing the past 66 Ma, is divided into three periods, the Paleogene, Neogene, and Quaternary; and seven epochs (Figure 19.5). Non-avian dinosaurs became extinct at the start of the Cenozoic, after which birds and mammals radiated to fill the available habitats. Earth was very warm during the early Eocene, and has cooled steadily ever since. Glaciers first appeared on Antarctica in the Oligocene and then on Greenland in the Miocene. By the Pleistocene, glaciers covered much of North America and Europe. The most recent of the Pleistocene glaciations ended ~11,700 years ago. The current epoch is known as the Holocene. Epochs are further divided into ages.

Figure 19.5 The periods and epochs of the Cenozoic Era. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source

Most of the boundaries between the periods and epochs of the geological timescale have been fixed on the basis of significant changes in the fossil record. For example, the boundary between the Cretaceous and the Paleogene coincides exactly with the extinction of the dinosaurs. This is not a coincidence. Many other types of organisms went extinct at this time, and the boundary between the two periods marks the division between sedimentary rocks containing Cretaceous organisms below, and those containing Paleogene organisms above.

References

Smith, W. (1815). A delineation of the strata of England and Wales with part of Scotland [map].

129

19.2 Relative Dating Methods

Relative Dating Principles

The simplest and most intuitive way of dating geological features is to look at the relationships between them. There are a few simple rules for doing this. But caution must be taken, as there may be situations in which the rules are not valid, so local factors must be understood before an interpretation can be made. These situations are generally rare, but they should not be forgotten when unraveling the geological history of an area.

The principle of superposition states that sedimentary layers are deposited in sequence, and the layers at the bottom are older than those at the top. This situation may not be true, though, if the sequence of rocks has been flipped completely over by tectonic processes, or disrupted by faulting.

The principle of original horizontality indicates that sediments are originally deposited as horizontal to nearly horizontal sheets. At a broad scale this is true, but at a smaller scale it may not be. For example, cross-bedding forms at an appreciable angle, where sand is deposited upon the lee face of a ripple. The same holds true of delta foreset beds (Figure 19.6).

Figure 19.6 A cross-section through a river delta forming in a lake. The delta foresets are labeled “Delta deposits” in this figure, and you can quickly see that the front face of the foresets are definitely not deposited horizontally. Source: AntanO (2017) CC BY 4.0 view source

The principle of lateral continuity states that sediments are deposited such that they extend laterally for some distance before thinning and pinching out at the edge of the depositional basin. But sediments can also terminate against faults or erosional features (see unconformities below), so may be cut off by local factors.

The principle of inclusions states that any rock fragments that are included in a rock must be older than the rock in which they are included. For example, a xenolith in an igneous rock, or a clast in sedimentary rock must be older than the rock that includes it (Figure 19.7). A possible situation that would violate this principle is the following: an igneous dyke may intrude through a sequence of rocks, hence is younger than these rocks (see the principle of cross-cutting relationships below). Later deformation may cause the dyke to be pulled apart into small pieces, surrounded by the host rocks. This situation can make the pieces of the dyke appear to be xenoliths, but they are younger than the surrounding rock in this case.

Figure 19.7 Applications of the principle of inclusion. Left- A xenolith of diorite incorporated into a basalt lava flow, Mauna Kea volcano, Hawai’i. The lava flow took place some time after the diorite crystallized (hammer head for scale). Right- Rip-up clasts of shale embedded in Gabriola Formation sandstone, Gabriola Island, BC. The pieces of shale were eroded as the sand was deposited, so the shale is older than the sandstone. Source: Karla Panchuk (2018) CC BY 4.0. Photographs by Steven Earle (2015) CC BY 4.0 view sources left/ right

The principle of cross-cutting relationships states that any geological feature that cuts across or disrupts another feature must be younger than the feature that is disrupted. An example of this is given in Figure 19.8, which shows three different sedimentary layers. The lower sandstone layer is disrupted by two faults, so we can infer that the faults are younger than this layer. But the faults do not appear to continue into the coal seam, and they certainly do not continue into the upper sandstone. So we can infer that coal seam is younger than the faults (because the coal seam cuts across them). The upper sandstone is youngest of all, because it lies on top of the coal seam. An example that violates this principle can be seen with a type of fault called a growth fault. A growth fault is a fault that continues to move as sediments are continuously delivered to the hangingwall block. In this case, the lower portion of the fault that cuts the lower sediments may have originally formed before the uppermost sediments were deposited, despite the fault cutting through all of the sediments, and appearing to be entirely younger than all of the sediments.

Figure 19.8 Superposition and cross-cutting relationships in Cretaceous Nanaimo Group rocks in Nanaimo BC. The coal seam is about 50 cm thick. Source: Steven Earle (2015) CC BY 4.0 view source

The principle of baked contacts states that the heat of an intrusion will bake (metamorphose) the rocks in close proximity to the intrusion. Hence the presence of a baked contact indicates the intrusion is younger than the rocks around it. If an intrusive igneous rock is exposed via erosion, then later buried by sediments, the surrounding rocks will not be baked, as the intrusion was already cold at the time of sediment deposition. But baked contacts may be difficult to discern, or may be minimally developed to absent when the intrusive rocks are low in volume or felsic (relatively cool) in composition.

The principle of chilled margins states that the portion of an intrusion that has cooled and crystallized next to cold surrounding rock will form smaller crystals than the portion of the intrusion that cooled more slowly deeper in the instrusion, which will form larger crystals. Smaller crystals generally appear darker in colour than larger crystals, so a chilled margin appears as a darkening of the intrusive rock towards the surrounding rock. This principle can be used to distinguish between an igneous sill, which will have a chilled margin at top and bottom, and a subaerial lava flow, which will have a chilled margin only at the bottom.

Exercise: Cross-Cutting Relationships

The outcrop in Figure 19.9 has three main rock types:

  1. Buff/pink felsic intrusive igneous rock present as somewhat irregular masses trending from lower right to upper left
  2. Dark grey metamorphosed basalt
  3. A 50 cm wide light-grey felsic intrusive igneous dyke extending from the lower left to the upper right – offset in several places

Using the principle of cross-cutting relationships outlined above, determine the relative ages of these three rock types.

Note: The near-vertical stripes are blasting drill holes. The image is about 7 m across.

Figure 19.9 Outcrop from Horseshoe Bay, BC. Source: Steven Earle CC BY 4.0 view source

 

Unconformities

An unconformity represents an interruption in the process of deposition of sediments. Recognizing unconformities is important for understanding time relationships in sedimentary sequences. An unconformity is visible in the Grand Canyon (Figure 19.10, white dashed line) above Proterozoic rocks that were tilted and then eroded to a flat surface prior to deposition of the younger Paleozoic rocks. The difference in time between the youngest of the Proterozoic rocks and the oldest of the Paleozoic rocks is nearly 300 million years. Tilting and erosion of the older rocks took place during this time, and if there were any deposition occurring in this area during this time, erosion removed those sediments.

19.10 Angular unconformity in the Grand Canyon in Arizona, with a sketch of rock orientations. The tilted rocks at the bottom are part of the Proterozoic Grand Canyon Group (aged 825 to 1,250 Ma). The flat-lying rocks at the top are Paleozoic (540 to 250 Ma). The boundary between the two (dashed white line) represents a time gap of nearly 300 million years. Source: Karla Panchuk (2018) CC BY 4.0. Photograph by Steven Earle (2015) CC BY 4.0 view source

There are four types of unconformities, reflecting different arrangements and types of rocks above and below the surface of non-deposition or erosion (Table 19.1, Figure 19.11).

Nonconformity: A boundary between non-sedimentary rocks below and sedimentary rocks above (Figure 19.11a).// Angular unconformity: A boundary between two sequences of sedimentary rocks where the underlying units have been tilted (or folded) and eroded prior to the deposition of the younger units (Figure 19.10, 19.11b).// Disconformity: A boundary between two sequences of sedimentary rocks where the underlying units have been eroded (but not tilted) prior to the deposition of the younger units (as in Figure 19.8; 19.11c).// Paraconformity: A time gap in a sequence of sedimentary rocks due to non-deposition. The time gap does not show up as an angular unconformity or a disconformity (Figure 19.11d).
Table 19.1 Types of Unconformities. Source: Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0. Click the image for a text version of this table.

 

19.11 The four types of unconformities: (a) a nonconformity between non-sedimentary rock and sedimentary rock, (b) an angular unconformity, (c) a disconformity between layers of sedimentary rock, where the older rock has been eroded but not tilted, and (d) a paraconformity where there is a long period (millions of years) of non-deposition between two parallel layers. Source: Steven Earle (2015) CC BY 4.0 view source

 

Exercise: Relative Dating with Unconformities

  1. The surfaces G and H in Figure 19.12 are unconformities. What kind?
  2. If erosion at the surface stopped and sediments were deposited once again, what kind of unconformity would exist between the layer I and younger rocks?
  3. Provide a list of the events that affected the rocks in Figure 19.12 in order from the oldest event to the most recent event. Note that C and D are faults. The sedimentary rocks labelled A are folded, but the other sedimentary rocks are horizontal.
Figure 19.12 Block diagram showing sedimentary and igneous rocks affected by faults, folds, and erosion. Source: Karla Panchuk (2018) CC BY-SA 4.0, modified after Woudloper (2009) CC BY-SA 1.0 view source

 

 

130

19.3 Dating Rocks Using Fossils

Geologists obtain a wide range of information from fossils. They help us to understand evolution, and life in general; they provide critical information for understanding depositional environments and changes in Earth’s climate; and they can be used to date rocks.

Although the recognition of fossils goes back hundreds of years, the systematic cataloguing and assignment of relative ages to different organisms from the distant past—paleontology—only dates back to the earliest part of the 19th century. The oldest undisputed fossils are from rocks dated ~3.5 Ga, and although fossils this old are typically poorly preserved and are not useful for dating rocks, they can still provide important information about conditions at the time. The oldest well-understood fossils are from rocks dating back to ~600 Ma, and the sedimentary record from this time forward is rich in fossil remains that provide a detailed record of the history of life. However, as anyone who has gone hunting for fossils knows, this does not mean that all sedimentary rocks have visible fossils or that they are easy to find. Fossils alone cannot provide us with numerical ages of rocks, but over the past century geologists have acquired enough isotopic dates from rocks associated with fossiliferous rocks (such as igneous dykes cutting through sedimentary layers) to be able to put specific time limits on most fossils.

A selective history of life on Earth over the past 600 million years is provided in Figure 19.13. The major groups of organisms that we are familiar with appeared between the late Proterozoic and the Cambrian (~600 Ma to ~541 Ma). Plants, which originally evolved in the oceans as green algae, invaded land during the Ordovician (~450 Ma). Insects, which evolved from marine arthropods, invaded land during the Devonian (400 Ma), and amphibians (i.e., vertebrates) invaded land about 50 million years later. By the late Carboniferous, trees had evolved from earlier plants, and reptiles had evolved from amphibians. By the mid-Triassic, dinosaurs and mammals had evolved from reptiles and reptile ancestors, Birds evolved from dinosaurs during the Jurassic. Flowering plants evolved in the late Jurassic or early Cretaceous. The earliest primates evolved from other mammals in the early Paleogene, and the genus Homo evolved during the late Neogene (~2.8 Ma).

Figure 19.13 A selective summary of life on Earth during the late Proterozoic and the Phanerozoic. The top row shows geological eras, and the lower row shows geological periods. Source: Steven Earle (2015) CC BY 4.0 view source

 

If we understand the sequence of evolution on Earth, we can apply this knowledge to determining the relative ages of rocks. This is William Smith’s principle of faunal succession, although in spite of the name, it can apply to fossils of plants and simple organisms as well as to fauna (animals).

The Phanerozoic Eon has witnessed five major extinctions (stars in Figure 19.13). The most significant of these was at the end of the Permian, which saw the extinction of over 80% of all species, and over 90% of all marine species. Most well-known types of organisms that survived were still severely impacted by this event. The second most significant extinction occurred at the Cretaceous-Paleogene boundary (K-Pg, also known the Cretaceous-Tertiary or K-T extinction). At that time, ~75% of marine species disappeared, as well as dinosaurs (but not birds) and pterosaurs. Other species were badly reduced but survived, and then flourished in the Paleogene. The K-Pg extinction may have been caused by the impact of a large asteroid (10 km to 15 km in diameter) and/or volcanic eruptions associated with the formation of the Deccan Traps, but it is generally agreed that the other four Phanerozoic mass extinctions had other causes, although their exact nature is not clearly understood.

It is not a coincidence that the major extinctions all coincide with boundaries of geological periods and/or eras. Paleontologists have placed most of the divisions of the geological time scale at points in the fossil record where there are major changes in the type of fossils observed.

If we can identify a fossil, and we know when the organism lived, we can assign a range of time to the formation of the sediments in which the organism was preserved when it died. This range might be several millions of years, because some organisms survived for a very long time. If the rock we are studying has several types of fossils in it, and we can assign time ranges to all of these fossils, we may be able to narrow the time range for the age of the rock considerably (Figure 19.14).

Figure 19.14 Application of bracketing to constrain the age of a rock based on the presence of several fossils. The yellow bar represents the time range during which each of the four species (A – D) existed on Earth. Although each species lived for several millions of years, we can narrow down the age of the rock to a span of just 1.3 Ma during which all four species coexisted. Source: Steven Earle (2015) CC BY 4.0 view source

Some organisms survived for a very long time, and are not particularly useful for dating rocks. Sharks, for example, have existed for over 400 million years, and the great white shark has survived for 16 million years so far. Organisms that lived for relatively short time periods are particularly useful for dating rocks, especially if they were distributed over a wide geographic area and hence can be used to compare rocks from different regions. These are known as index fossils. There is no specific limit on how short the time span has to be for a fossil to qualify as an index fossil. Some such organisms lived for millions of years, and others for much less than a million years.

Some well-studied groups of organisms qualify as biozone fossils because, although the genera and families lived over a long time, each species lived for a relatively short time and can be easily distinguished from others on the basis of specific features. For example, ammonites have a distinctive feature known as the suture line, where the internal shell layers that separate the individual chambers (septae) meet the outer shell wall (Figure 19.15). These suture lines are sufficiently variable to identify species that can be used to estimate the relative or absolute ages of the rocks in which they are found.

Figure 19.15 The septum of an ammonite (white part, left), and the suture lines where the septae meet the outer shell (right). Source: Steven Earle (2015) CC BY 4.0 view source

Foraminifera—small, calcium carbonate-shelled marine organisms that originated during the Triassic and are still alive today—are also useful biozone fossils. Numerous different foraminifera lived during the Cretaceous Period, some for over 10 million years, but others for less than 1 million years (Figure 19.16). If the foraminifera in a rock can be identified to the species level, the rock’s age can be determined.

Figure 19.16 Time ranges for Cretaceous foraminifera (left), and modern foraminifera from the Ambergris area of Belize (right). Source: Left- Steven Earle (2015) CC BY 4.0, from data in Scott (2014). Right- Steven Earle (2015) CC BY 4.0 view source

Exercise: Dating Rocks Using Index Fossils

Figure 19.17 shows the age ranges for some late Cretaceous inoceramid clams in the genus Mytiloides. Using the bracketing method described above, determine the possible age range of a rock in which all five of these organisms were found.

How would the age range change if M. subhercynius were not present in the rock?

Figure 19.17 Inoceramid ranges. Source: Steven Earle (2015) CC BY 4.0, from data in Harries et al. (1996). View source

Correlation

Geologists employ relative age dating techniques to correlate rocks between regions. Correlation seeks to relate the geological history between regions, by relating the rocks in one region to those in another.

There are different techniques of correlation. The easiest technique is to correlate by rock type, or lithology, called lithostratigraphic correlation. In this method, specific rock types are related between regions. If a sequence of rocks at one site consists of a sandstone unit overlain by a limestone unit, then a unit of shale, and the exact same sequence of rocks—sandstone, limestone, shale—occurs at a nearby site, lithostratigraphic correlation means assuming that the rocks at both sites are in the same rocks. If you could see all of the rock exposed between the two sites, the units would connect with one another. The problem with this type of correlation is that some rocks may only have formed locally, or may pinch out between the two sites, and therefore not be present at the site to which a correlation is being attempted.

Another technique, biostratigraphic correlation, involves correlation based on fossil content. This technique uses fossil assemblages (fossils of different organisms that occur together) to correlate rocks between regions. The best fossils to use are those that are widely spread, abundant, and lived for a relatively short period of time.

Yet another technique, chronostratigraphic correlation, is to correlate rocks that have the same age. This can be the most difficult way to correlate, because rocks are generally diachronous. That is, if we trace a given rock unit across any appreciable lateral distance, the age of that rock actually changes. To give a familiar example, when you go to the beach, you know that the beach itself and the lake bottom in the shallow water is sandy. But if you swim out to deeper water and touch bottom, the bottom feels muddy. The difference in sediment type has to do with the energy of deposition, with the waves at and near the beach keeping any fine sediments away, only depositing them in deeper quieter waters. If you think of this example in time, you realize that the sand at and near the beach is being deposited at the same time as the mud in deeper water. But if lake levels drop, the beach sands will slowly migrate outwards and cover some of the deeper water muds. If lake levels rise, the deeper water muds will slowly migrate landwards and cover some of the shallower water sands. This is an example of Walther’s Law, which states that sedimentary rocks that we see one on top of each other in the rock record actually formed adjacent to one another at the time of deposition. In order to correlate rock units in time, we must target marker beds that formed instantaneously. An example of such would be an ash layer from a nearby volcano that erupted and blanketed an entire region in ash. But such marker beds are usually rare to absent, making such correlation extremely difficult.

References

Harries, P.J., Kauffman, E.G., Crampton, J.S. (Redacteurs), Bengtson, P., Cech, S., Crame, J.A., Dhondt, A.V., Ernst, G., Hilbrecht, H., Lopez, Mortimore, G.R., Tröger, K.-A., Walaszcyk, I., & Wood, C.J. (1996). Mitteilungen aus dem Geologisch – Paläontologischen Museum der Universität Hamburg, 77, 641-671. Full text

Scott, R. (2014). A Cretaceous chronostratigraphic database: construction and applications, Carnets de Géologie, 14(2), 15-37. Full text

131

19.4 Isotopic Dating Methods

Isotope Pairs

Originally, fossils only provided us with relative ages because, although early paleontologists understood biological succession, they did not know the absolute ages of the different organisms. It was only in the early part of the 20th century, when isotopic dating methods were first applied, that it became possible to discover the absolute ages of the rocks containing fossils. In most cases, we cannot use isotopic techniques to directly date fossils or the sedimentary rocks in which they are found, but we can constrain their ages by dating igneous rocks that cut across sedimentary rocks, or volcanic ash layers that lie within sedimentary layers.

Isotopic dating of rocks, or the minerals within them, is based upon the fact that we know the decay rates of certain unstable isotopes of elements, and that these decay rates have been constant throughout geological time. It is also based on the premise that when the atoms of an element decay within a mineral or a rock, they remain trapped in the mineral or rock, and do not escape.

One of the isotope pairs commonly used to date rocks is the decay of 40K to 40Ar  (potassium-40 to argon-40). 40K is a radioactive isotope of potassium that is present in very small amounts in all minerals that contain potassium. It has a half-life of 1.3 billion years, meaning that over a period of 1.3 Ga one-half of the 40K atoms in a mineral or rock will decay to 40Ar, and over the next 1.3 Ga one-half of the remaining atoms will decay, and so forth (Figure 19.18). 40K is called the parent isotope, and 40Ar the daughter isotope, as the parent gives way to the daughter during radioactive decay.

Figure 19.18 The decay of 40K over time. Each half-life is 1.3 billion years, so after 3.9 billion years (three half-lives) 12.5% of the original 40K will remain. The red-blue bars represent 40K and the green-yellow bars represent 40Ar. Source: Steven Earle (2015) CC BY 4.0 view source

In order to use the K-Ar dating technique, we need to have an igneous or metamorphic rock that includes a potassium-bearing mineral. One good example is granite, which contains the mineral potassium feldspar (Figure 19.19). Potassium feldspar does not contain any argon when it forms. Over time, the 40K in the feldspar decays to 40Ar. The atoms of 40Ar remain embedded within the crystal, unless the rock is subjected to high temperatures after it forms. The sample must be analyzed using a very sensitive mass-spectrometer, which can detect the differences between the masses of atoms, and can therefore distinguish between 40K and the much more abundant 39K. The minerals biotite and hornblende are also commonly used for K-Ar dating.

Figure 19.19 Crystals of potassium feldspar (pink) in a granitic rock are candidates for isotopic dating using the K-Ar method because they contained potassium and no argon when they formed. Source: Steven Earle (2015) CC BY 4.0 view source

There are many isotope pairs that can be employed in dating igneous and metamorphic rocks (see Table 19.2), each with its strengths and weaknesses. In the above example, the daughter isotope 40Ar is naturally a gas, and can escape the potassium feldspar quite easily if the feldspar is exposed to heating during metamorphism, or interaction with hydrothermal fluids. Hence we must closely examine the feldspar mineral to determine if there is any evidence of alteration. If some 40Ar has been lost, but the sample is dated anyway, an age that is too young will be calculated.

Table 19.2 Commonly used isotope systems for dating geological materials. Source: Steven Earle (2015) CC BY 4.0. Click the table for a text version.

Each parent isotope has a certain half-life, which ranges from microseconds to billions of years, depending upon the isotope. In dating rocks, we need to select an isotope pair with a parent isotope that has a reasonable half-life. This means that the half-life must not be too short or too long. If the half-life is too short, then most of the parent isotope will have decayed to form the daughter isotope. If we cannot measure the amount of parent isotope very accurately, which will be impossible to do if there is only the tiniest amount of parent isotope left, our calculated age will have huge errors associated with it. The same applies if the half-life is too long. In this case, very little of the daughter isotope will have formed, and our inability to measure the small amount of daughter isotope accurately will again result in huge errors in our calculated age.

Another complicating factor is whether the mineral of interest incorporated any of the daughter isotope into its structure at the time of formation. When we select a mineral and an isotope pair to date that mineral, we make the assumption that all of the daughter isotope we find in the mineral was produced in the mineral by radioactive decay of the parent isotope. But if the mineral formed with some daughter isotope already present in its structure, then the age we calculate will be too old.

A more robust mineral to use to date certain types of igneous and metamorphic rocks is zircon. Zircon is a mineral that incorporates uranium into its structure at the time of formation. One of the isotopes of uranium decays to lead with a long half-life (see Table 19.2). Zircon is a mineral of choice for dating because it takes no lead into its structure when it forms, so any lead present is due entirely to the radioactive decay of the uranium parent. Another reason is because zircon is a very resistant mineral. It can handle exposure to hydrothermal fluids, and all but the highest grades of metamorphism, and not lose any of the parent or daughter isotopes. Hence the age that we calculate tends to be very accurate. One drawback is that zircon tends to form only in felsic igneous rocks. Hence if we are trying to date a mafic igneous rock, we must choose a different mineral.

The Meaning of a Radiometric Date

When we employ isotopic methods on minerals we are measuring an age date. Generally, an age date refers to the time since a mineral crystallized from molten rock (magma or lava). This is when the elements that make up the mineral get locked into the mineral’s structure. But as we have already seen, elevated temperatures can cause elements to escape from a mineral, without the mineral melting. Hence when we date a mineral, we may be dating the time since the mineral crystallized from a melt, or the time since the mineral last experienced a period of heating above its Curie point, which is the temperature beyond which the mineral is able to lose (or gain) elements from its structure, without melting. So we have to know something about the rock before we forge ahead to measure an age. We may choose a mineral and isotope pair that are very resistant to metamorphism, so that we can “see through” the metamorphism, and determine the original age that the mineral crystallized from a melt. Or we may be interested in the age of the metamorphic event itself, so choose a mineral and isotope pair that is susceptible to resetting the isotopic clock during metamorphism (such as by losing all of the daughter isotope).

Absolute age dating is a powerful tool for unraveling the geological history of a region, but we have seen that we must ultimately rely upon igneous rocks (that may have later metamorphosed) for the minerals that we are able to date (see the next section for issues with dating sedimentary rocks directly). Another issue with absolute age dating is that it is expensive, with a single analysis costing several hundreds of dollars. Hence geologists never forget their relative age dating principles, and are always applying them in the field to determine the sequence of events that formed the rocks in a region.

Exercise: Combining Absolute Ages with Relative Dating

The age dates for three igneous rock layers are given. Use relative dating techniques to determine the age ranges for the sets of sedimentary units A, B, and C.

Figure 19.20 A sequence of igneous and sedimentary layers. Age dates  are given for the igneous layers. Source: Karla Panchuk (2018) CC BY-SA 4.0, modified after Jill Curie (2014) CC BY-SA 3.0 view source

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://openpress.usask.ca/physicalgeology/?p=3014

Isotope Dating Techniques and Sedimentary Rocks

A clastic sedimentary rock (e.g., conglomerate, Figure 19.21) is made up of older rock and mineral fragments. These fragments were derived from weathering and erosion of pre-existing rocks. The process of forming a sedimentary rock from sediments generally occurs at low temperatures, so the minerals are not heated beyond their Curie points. Hence the minerals still preserve their original ages (either igneous crystallization age, or a metamorphic age).

Figure 19.21 Conglomerate is a sedimentary rock consisting of large rounded clasts surrounded by finer-grained material. Source: Steven Earle (2015) CC BY 4.0 view source

In almost all cases, the fragments have come from a range of source rocks that all formed at different times. If we dated a number of individual grains in the sedimentary rock, we would likely get a range of different dates, all older than the age of the sedimentary rock. The most that such ages gleaned from a sedimentary rock can tell us is a maximum age of the sedimentary rock. It might be possible to date some chemical sedimentary rocks isotopically, but there are no useful isotopes that can be used on old chemical sedimentary rocks.

Radiocarbon Dating

Radiocarbon dating (using 14C) can be applied to many geological materials, including sediment and sedimentary rocks, but only if the materials in question are younger than ~60 ka, and contain organic material. Beyond this time, there is so little 14C left that it cannot be measured accurately, and the resulting age dates are hence unreliable. Fragments of wood incorporated into young sediment are good candidates for carbon dating, and this technique has been used widely in studies involving late Pleistocene glaciers and glacial sediments. Figure 19.22 shows radiocarbon dates from wood fragments in glacial sediments have been used to estimate the time of the last glacial advance along the Strait of Georgia.

Figure 19.22 Radiocarbon dates on wood fragments in glacial sediments in the Strait of Georgia. Source: Steven Earle (2015) CC BY 4.0 view source, modified after Clague (1976).

Exercise: Radiometric Dating with Potassium-Argon

Assume that a feldspar crystal from the granite shown in Figure 19.19 was analyzed for 40K and 40Ar. The proportion of 40K remaining is 0.91. Using the decay curve shown on this graph, estimate the age of the rock. An example is provided (in blue) for a 40K proportion of 0.95, which is equivalent to an age of approximately 96 Ma. This is determined by drawing a horizontal line from 0.95 to the decay curve line, and then a vertical line from there to the time axis.

Figure 19.23 Decay curve for potassium-argon dating. Source: Steven Earle (2015) CC BY 4.0 view source

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://openpress.usask.ca/physicalgeology/?p=3014

References

Clague, J. (1976). Quadra Sand and its relation to late Wisconsin glaciation of southeast British Columbia. Canadian Journal of Earth Sciences, 13, 803-815.

132

19.5 Other Dating Methods

There are numerous other techniques for dating geological materials, but we will examine just two of them here: dendrochronology—tree-ring dating—and dating based on the record of reversals of Earth’s magnetic field.

Dendrochronology

Dendrochronology can be applied to dating very young geological materials based on reference records of tree-ring growth going back many millennia. The longest such records can take us back over 25 ka, to the height of the last glaciation. One of the advantages of dendrochronology is that, providing reliable reference records are available, the technique can be used to date events to the nearest year.

Dendrochronology has been used to date the last major subduction zone earthquake on the coast of B.C., Washington, and Oregon. When large earthquakes occur in this region, there is a tendency for some coastal areas to subside by one or two metres. Seawater then rushes in, flooding coastal flats and killing trees and other vegetation within a few months. There are at least four locations along the coast of Washington that have such dead trees, and probably many more in other areas. Wood samples from these trees have been studied and the ring patterns have been compared with patterns from old living trees in the region (Figure 19.24).

Figure 19.24 Example of tree-ring dating of dead trees. Source: Steven Earle (2015) CC BY 4.0 view source

At all of the locations studied, the trees were found to have died either in the year 1699, or very shortly thereafter (Figure 19.25). On the basis of these results, it was concluded that a major earthquake took place in this region sometime between the end of growing season in 1699 and the beginning of the growing season in 1700. Evidence from a major tsunami that struck Japan on January 27, 1700, narrowed the timing of the earthquake to sometime in the evening of January 26, 1700. (For more information, see https://web.viu.ca/earle/1700-quake/.)

Figure 19.25 Sites in Washington where dead trees are present in coastal flats. The outermost wood of eight trees was dated using dendrochronology, and of these, seven died during the year 1699, suggesting that the land was inundated by water at this time. Source: Steven Earle (2015) CC BY 4.0 view source, from data in Yamaguchi et al. (1997).

Magnetic Chronology

Changes in Earth’s magnetic field can also be used to date events in geologic history. The magnetic field causes compass needles point toward the north magnetic pole, but this hasn’t always been the case. At various times in the past, Earth’s magnetic field has reversed its polarity, and during such times a compass needle would have pointed to the south magnetic pole. By studying magnetism in volcanic rocks that have been dated isotopically, geologists have been able to establish the chronology of magnetic field reversals going back for ~250 Ma. About 5 Ma of this record is shown in Figure 19.26, where the black bands represent periods of normal magnetism (“normal” meaning a polarity identical to the current magnetic field) and the white bands represent periods of reversed magnetic polarity. These periods of consistent magnetic polarity are given names to make them easier to reference. The current period of normal magnetic polarity, known as Brunhes, has lasted for the past 780,000 years. Prior to that there was a short reversed period and then a short normal period, the latter of which is known as Jaramillo.

Figure 19.26 The last 5 Ma of magnetic field reversals. Source: Steven Earle (2015) CC BY 4.0 view source, modified after U.S. Geological Survey (2007) Public Domain view source

Oceanic crust becomes magnetized by the magnetic field that exists as the crust forms from magma at mid-ocean ridges. As it cools, the magnetic fields of tiny crystals of magnetite that form within the magma become aligned with the existing magnetic field, and remain in this orientation, even if Earth’s magnetic field later changes polarity (Figure 19.27). Oceanic crust that is forming today is being magnetized in a “normal” sense, but crust that formed 780,000 to 900,000 years ago, in the interval between the Brunhes and Jaramillo normal periods, was magnetized in the “reversed” sense.

Figure 19.27 Formation of magnetized oceanic crust at a spreading ridge. Coloured bars represent periods of normal magnetic polarity. Capital letters denote the Brunhes, Jaramillio, Olduvai, and Gauss normal magnetic periods (see Figure 19.26). Source: Steven Earle (2015) CC BY 4.0 view source

Magnetic chronology can be used as a dating technique because we can measure the magnetic field of rocks using a magnetometer, or of entire regions by towing a magnetometer behind a ship or an airplane. For example, the Juan de Fuca Plate, which lies off of the west coast of BC, Washington, and Oregon, is being and has been formed along the Juan de Fuca spreading ridge (Figure 19.28). The parts of the plate that are still close to the ridge exhibit normal magnetic polarity, while parts that are further away (and formed much earlier) have either normal or reversed magnetic polarity, depending upon when the rock formed. By carefully matching the sea-floor magnetic stripes with the known magnetic chronology, we can determine the age at any point on the plate. We can see that the oldest part of the Juan de Fuca Plate that has not yet subducted (off of the coast of Oregon) is just over 8 million years old, while the part that is subducting beneath Vancouver Island is between 0 and ~6 million years old.

Figure 19.28 TThe pattern of magnetism within the area of the Juan de Fuca Plate, off the west coast of North America. Coloured bands represent parts of the sea floor with normal magnetic polarity, and the magnetic time scale is shown using these same colours. Source: Steven Earle (2015) CC BY 4.0 view source

 

Exercise: Magnetic Dating

The fact that magnetic intervals can only be either normal or reversed places significant limits on the applicability of magnetic dating. If we find a rock with normal magnetism, we can’t know which normal magnetic interval it represents, unless we have some other information.

Using Figure 19.26 for reference, determine the age of a rock with normal magnetism that has been found to be between 1.5 and 2.0 Ma based on fossil evidence in nearby sedimentary rocks.

How old is a rock that is limited to 2.6 to 3.2 Ma by fossils, and which has reversed magnetic polarity?

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://openpress.usask.ca/physicalgeology/?p=3026

References

Yamaguchi, D.K., Atwater, B.F., Bunker, D.E., Benson, B.E., & Reid, M. S. (1997). Tree-ring dating the 1700 Cascadia earthquake. Nature389, 922 – 923.

133

19.6 Understanding Geological Time

It is one thing to know the facts about geological time — how long it is, how we measure it, how we divide it into smaller time intervals, and what we call the various periods and epochs — but it is quite another to really understand geological time. The problem is that our lives are short and our memories are even shorter. Our experiences span only a few decades, so we really don’t have a way of knowing what 11,700 years means. What’s more, it is hard for us to understand how 11,700 years differs from 65.5 Ma, or even from 1.8 Ga. It is not that we cannot comprehend what the numbers mean, it is that we cannot really appreciate how much time is involved.

You may wonder why it is so important to understand geological time. There are some very good reasons. One is so that we can fully understand how geological processes that seem impossibly slow can produce anything of consequence. Consider driving from one major city to another, where a journey of several hours might occur at speeds of ~100 km/h. Continents move toward each other at rates of a fraction of a millimetre per day, a speed something on the order of 0.00000001 km/h, and yet, at this impossibly slow rate (try walking at this speed!), they can move thousands of kilometres through geological time. Sediments typically accumulate at even slower rates — less than a millimetre per year — but still they are thick enough to be thrust up to form huge mountains or carved into breathtaking canyons.

Another reason is to understand issues like extinction of endangered species, and human influence on climate. People who do not understand geological time are quick to say that the climate has changed in the past, and that what is happening now is no different. And climate certainly has changed in the past: from the Eocene (50 Ma) to the present day, Earth’s climate cooled by ~12°C on average. This is a huge change that ranks as one of the most important climate changes of Earth’s past, and yet the rate of change over this time was only 0.000024 °C/century. Recent warming has occurred at a rate of ~1.1°C over the past 100 years (NASA GISS), 45,800 times faster than the rate of climate change since the Eocene.

One way to wrap your mind around geological time is to put it into the perspective of single year. At this rate, each hour of the year is equivalent to approximately 500,000 years, and each day is equivalent to 12.5 million years. If all of geological time is compressed down into a single year, Earth formed on January 1, and the first life forms evolved in late March (~3,500 Ma). The first multicellular life forms appeared on November 13 (~600 Ma), plants appeared on land on November 24, and amphibians on December 3. Reptiles evolved from amphibians during the first week of December, and dinosaurs and early mammals evolved by December 13. Non-avian dinosaurs, which survived for 160 million years, went extinct on Boxing Day (December 26). The Pleistocene glaciation began at ~6:30 p.m. on New Year’s Eve, and the last glacial ice melted from southern Canada by 11:59 p.m.

It is worth repeating: on this time scale, the earliest ancestors of the animals and plants with which we are familiar did not appear on Earth until mid-November, the dinosaurs disappeared after Christmas, and most of Canada was periodically locked in ice from 6:30 to 11:59 p.m. on New Year’s Eve. As for people, the first to inhabit Canada arrived about one minute before midnight.

Exercise: What Happened on Your Birthday?

Using the “all of geological time compressed to one year” concept, determine the geological date that is equivalent to your birthday. First, go here to find out which day of the year your birth date is. Divide that number by 365, and multiply that number by 4,570 to determine the time (in millions since the beginning of geological time). Subtract that number from 4,570 to determine the date back from the present.

Example: April Fool’s Day (April 1) is day 91 of the year: 91/365 = 0.2493. 0.2493 x 4,570 = 1,139 million years from the start of time, and 4,570 – 1,193 = 3,377 Ma is the geological date.

Finally, go to the Foundation for Global Community’s Walk through Time website to find out what was happening on your day. The nearest date to 3,377 Ma is 3,400 Ma.

References

NASA Goddard Institute for Space Studies (n.d.). GLOBAL Station Temperature Index in 0.01 degrees Celsius base period: 1951-1980 [data file]. Retrieved from http://data.giss.nasa.gov/gistemp/tabledata_v3/GLB.Ts.txt

134

Chapter 19 Summary

The topics covered in this chapter can be summarized as follows:

19.1 The Geological Time Scale

The work of William Smith was critical to the establishment of the first geological timescale early in the 19th century, but it wasn’t until the 20th century that geologists were able to assign reliable dates to the various time periods. The geological timescale is now maintained by the International Commission on Stratigraphy. Geological time is divided into eons, eras, periods, and epochs.

19.2 Relative Dating Methods

We can determine the relative ages of different rocks by observing and interpreting relationships among them, such as superposition, cross-cutting, and inclusions. Gaps in the geological record are represented by various types of unconformities.

19.3 Dating Rocks Using Fossils

Fossils are useful for dating rocks back to ~600 Ma. If we know the age range of a fossil, we can date the rock in which it is found, but some organisms lived for many millions of years. Index fossils represent shorter geological time spans, and if a rock has several different fossils with known age ranges, we can narrow the time during which the rock formed.

19.4 Isotopic Dating Methods

Radioactive isotopes decay at constant known rates, and can be used to date igneous and metamorphic rocks. Some commonly used isotope systems are potassium-argon, rubidium-strontium, uranium-lead, and carbon-nitrogen. Radiocarbon dating can be applied to sediments and sedimentary rocks, but only if they are younger than 60 ka, and contain organic material, or minerals of calcium carbonate.

19.5 Other Dating Methods

There are many other methods for dating geological materials. Two that are widely used are dendrochronology and magnetic chronology. Dendrochronology, based on studies of tree rings, is widely applied to dating glacial events. Magnetic chronology is based on the known record of Earth’s magnetic field reversals.

19.6 Understanding Geological Time

While understanding geological time is relatively easy, actually comprehending the significance of the vast amounts of geological time is a great challenge. To be able to solve important geological problems and certain societal challenges, we need to really appreciate the vastness of geological time.

Review Questions

  1. A granitic rock contains inclusions (xenoliths) of basalt. What can you say about the relative ages of the granite and the basalt?
  2. Explain the differences between a) disconformity and paraconformity; and b) nonconformity and angular unconformity
  3. What are the features of a useful index fossil?
  4. Figure 19.29 shows a geological cross-section. The granitic rock F at the bottom is the rock that you estimated the age of in the exercise in 19.4, Radiometric Dating with Potassium-Argon. A piece of wood from layer D has been sent for radiocarbon dating and the result was 0.55 14C remaining. How old is layer D?
  5. Based on your answer to question 4, what can you say about the age of layer C in the figure above?
  6. What type of unconformity exists between layer C and rock F?
  7. What type of unconformity exists between layer C and rock B?
  8. We cannot use magnetic chronology to date anything older than ~780,000 years. Why?
  9. How did William Smith apply the principle of faunal succession to determine the relative ages of the sedimentary rocks of England and Wales?
  10. Access a copy of the geological time scale at http://www.stratigraphy.org/index.php/ics-chart-timescale. What are the names of the last age of the Cretaceous and the first age of the Paleogene? Print out the time scale and stick it on the wall above your desk!
Figure 19.29 Geological cross-section (left) and decay curve for 14C ages. Source: Left- Karla Panchuk (2018) CC BY 4.0, modified after Steven Earle (2015) CC BY 4.0 view source. Right- Steven Earle (2015) CC BY 4.0 view source

 

135

Answers to Chapter 19 Review Questions

  1. Xenoliths of basalt within a granite must be older than the granite according to the principle of inclusions.
  2. (a) At both disconformities and paraconformities the beds above and below are parallel, but at a disconformity there is clear evidence of an erosion surface (the lower layers have been eroded). (b) A nonconformity is a boundary between sedimentary rocks above and non-sedimentary rocks below while an angular unconformity is a boundary between sedimentary rocks above and tilted and eroded and sedimentary layers below.
  3. A useful index fossil must have survived for a relatively short period (e.g., around a million years), and also should have a wide distribution so that it can be used to correlate rocks from different regions.
  4. The granitic rock F has been dated to 175 Ma. The wood in layer D is approximately 5,000 years old, so we can assume that layer D is no older than that, although it could be as much as a few hundred years younger if the wood was already old when it got incorporated into the rock.
  5. Layer C must be between 5,000 y and 275 Ma.
  6. The unconformity between layer C and rock F is a nonconformity.
  7. The granite (F) was eroded prior to deposition of C, so it’s likely that layer B was also eroded at the same time. If so, that makes the boundary between C and B a disconformity.
  8. The last magnetic reversal was 780,000 years ago, so all rock formed since that time is normally magnetized and it isn’t possible to distinguish older rock from younger rock within that time period using magnetic data.
  9. William Smith was familiar with the different diagnostic fossils of the rocks of England and Wales and was able to use them to identify rocks of different ages.
  10. The last age of the Cretaceous is the Maastrichtian (71.2 to 66.0 Ma) and the first age of the Paleogene is the Danian (66.0 to 61.6 Ma).

1

Glossary

A

aa  a lava flow that solidifies with a blocky high-relief surface

ablation  melting of ice in the context of glaciation

ablation till  till that is formed when englacial and supraglacial sediments are deposited because the ice that was supporting them melts

abyssal plain  the flat surface of the deep ocean, typically beyond the limits of the continental slopes

abyssal pelagic zone  the deeper parts of the ocean, between 4000 and 6000 m.

accretion (plate tectonics) the process by which continental blocks (terranes) are added to existing continental areas

accretion (planetary) the process by which solid celestial bodies are added to existing bodies during collisions

acid rock drainage (acid mine drainage) the production of acid from oxidation of sulphide minerals (especially pyrite) in either naturally or anthropogenically exposed rock

aeolian  processes related to transportation and deposition of sediments by wind

aerobic processes that take place in the presence of abundant oxygen

aerosol  an aggregate of fine solid particles or a small droplet of liquid suspended in the air

aftershock an earthquake that can be shown to have been caused by another earthquake

aggregate unconsolidated materials (typically sediments) that are used in the construction industry

albedo  the reflectivity of a surface of a planet (expressed as the percentage of light that reflects)

albite  sodium-rich plagioclase feldspar

alpine glacier a glacier formed in a mountainous region and confined to a valley (same as valley glacier)

amphibole  a double-chain ferromagnesian silicate mineral (e.g., hornblende)

amphibolite a foliated metamorphic rock in which the mineral amphibole as an important component

amplification  in the context of seismic shaking the process by which the amplitude of the seismic waves are enhanced

amplitude for any type of wave, the difference in height between a crest and the adjacent trough

anaerobic  processes that take place without oxygen

andesite  a volcanic rock of intermediate composition

anion  a negatively charged ion

angular unconformity  a geological boundary at the base of a sedimentary layer where the sedimentary rock beneath has been tilted or folded and then eroded

anorthite  calcium-rich plagioclase feldspar

Antarctic Bottom Water  water at abyssal depths in the ocean that forms from the sinking of dense cold water adjacent to Antarctica

anticline  an upward fold where the beds are known not to be overturned

anthracite  a high grade of coal (92 to 98% carbon) that is formed from deep burial and weak metamorphism

anthropogenic  resulting from the influence of humans

antiform an upward fold where it is not known if the beds have been overturned

apparent polar wandering path  a path of seeming varying magnetic pole positions defined by paleomagnetic data, which is in fact an artefact of the motion of contients

aphanitic  an igneous texture characterized by crystals that are too small to see with the naked eye

aquifer a body of rock or sediment that has sufficient permeability to allow it to be used as a source of groundwater

aquitard  a body of rock or sediment that has insufficient permeability to allow it to be used as a source of groundwater

arch  a rock weathering remnant in the form of an arch (typically along a coast and resulting from wave erosion)

arenite  a sandstone with less than 15% silt and clay

arête   a sharp ridge that separates adjacent glacially carved valleys

arkose  a sandstone with more than 10% feldspar and more feldspar than lithic fragments

arkosic arenite  an arkose with less than 15% clay/silt matrix

artesian well  a well that is completed in a confined aquifer and in which the water level in the well rises above the top of the aquifer

asteroid  a rocky body orbiting the Sun

asteroid belt  the region between the orbits of Mars and Jupiter that is populated with many asteroids

asthenosphere  the part of the mantle, from about 100 to 200 km below surface, within which the mantle material is close to its melting point, and therefore relatively weak

asymmetrical (folds) where the two sides of the fold make significantly different angles with respect to the axial plane

atoll  a ring-shaped carbonate (or coral) reef or series of islands

atomic mass  the total number of neutrons plus protons in an atom

atomic number  the total number of protons in an atom

attitude  the orientation of a sloping geological feature, such as a bedding plane or fracture

aureole  a zone of metamorphism around a source of heat such as a magma body

axial plane  a plane that can be traced through all of the hinge lines of a fold

B

back reef  the zone of shallow water on the shore-side of a reef

background (geochemistry)  the typical level of an element in average rocks or sediments

backwash  the wash of wave water down the slope of a beach

banded iron formation  an iron-bearing sedimentary rock that is rich in minerals such as hematite and magnetite, and interbedded with chert stained red by hematite

bank-full stage  the water level of stream when it is in flood and just about to flow over its banks

barrier reef  a carbonate (or coral) reef that forms a barrier to waves along a coast

basal sliding  the motion of glacial ice along the base of a glacier that is warm enough to have liquid water

basalt  a volcanic rock of mafic composition

base level (stream) the base level is the lowest level that a stream can erode down to, as defined by the ocean, a lake or another stream that it flows into

batholith an irregular body of intrusive igneous rock that has an exposed surface of at least 100 km2bathypelagic zone  the moderately deep parts of the ocean, between 1000 and 4000 m.

baymouth bar a spit that extends across the mouth of a bay

beach face the part of the beach that is relatively steep and lies between the high and low tide levels

bed  a sedimentary layer

bed load  the fraction of a stream’s sediment load that typically rests on the bottom and is moved by saltation and traction

bedding repeated layering in a sedimentary rock

bentonite  a smectite clay that has strong swelling properties and is effective at absorbing dissolved ions

berm  a flat area of a beach in the backshore area (above the high tide level)

big-bang theory the theory that the universe started by expanding suddenly from a single point approximately 13.77 billion years ago

biochemical sedimentary rock  a rock formed when biological processes cause ions to precipitate (e.g., when organisms make shells of calcite or silica)

biotite  a sheet-silicate mineral (mica) that includes iron and or magnesium, and is therefore a ferromagnesian silicate

biozone  a stratigraphic interval that can be defined on the basis of a specific fossil

bituminous coal  a medium-grade type of coal with 70 to 92% carbon

blueschist  1. (metamorphic rock) a schist with blue colouring due to the presence of the mineral glaucophane. Formed in subduction zones. 2. (metamorphic facies) a facies characterized by relatively low temperatures and high pressures, such as can exist within a subduction zone

body wave  a seismic wave that travels through rock (e.g., a P-wave or an S-wave)

boulder  a sediment clast with a diameter of at least 256 mm

Bowen’s reaction series  the scheme that defines the typical order of crystallization of minerals from magma as the magma cools

braided  a stream pattern which is characterized by abundant sediment and numerous intertwining channels around bars

breakwater a structure built offshore in order to deflect the energy of waves

breccia  a sedimentary or volcanic rock texture characterized by angular clasts

brunisol  a relatively immature forest soil, lacking in well-defined horizons

burial  when a layer of sediment is covered by subsequent sediment accumulation

C

caldera  a volcanic depression that forms when part of the volcano collapses into an empty magma chamber

caliche  a white calcium-carbonate rich layer within soils in arid regions

calving  the loss of ice from the front of a glacier by collapse into water

Canadian Shield  the exposed part of the continent Laurentia

carbonate a mineral for which the anion is CO3-2carbonate compensation depth  the depth in the ocean below which carbonate minerals are soluble

cation  a positively charged ion

cementation  the process by which minerals are precipitated between grains in sediments, locking the grains together

Cenozoic  the most recent of the eras, representing the past 65.5 Ma of geological time

chemical sedimentary rock  a sedimentary rock comprised of material that was transported as ions in solution, then precipitated by inorganic means (e.g., precipitation triggered by evaporation)

chemical weathering  chemical reactions at Earth’s surface which break down rocks and minerals

chernozem  black soil typical of grasslands in cold climates such as the Canadian Prairies

chert  very fine-grained sedimentary rock formed almost entirely of silica

chilled margin  edges of a pluton which cool rapidly through contact with country rock, resulting in finer grain sizes than in the interior of the pluton

chlorite  ferromagnesian sheet silicate mineral, typically present as fine crystals and forming from the low-temperature metamorphism of mafic rock

cinder cone  steep-sided volcano comprised almost entirely of loose rock fragments, and typically formed during a single eruptive event

cirque  a steep-sided semi-circular basin eroded by an alpine glacier at the head of its valley

clast  a sedimentary fragment of mineral or rock

clastic sedimentary rock  a sedimentary rock comprised of material that was transported as clasts or fragments

clay  sediment particle that is less than 1/256 mm in diameter

clay mineral  a hydrous sheet-silicate mineral that typically exists as clay-sized grains

claystone  a sedimentary rock comprised mostly of clay-sized grains

cleavage  tendency for a mineral to break along smooth planes that are predetermined by its lattice structure

climate feedback  a case in which the effects of a climate forcing trigger other changes which either amplify or mute the effects of the initial forcing

climate forcing  a mechanism, such as a change in greenhouse gas levels, that causes the climate to change

coal  an organic sedimentary rock formed by the compression and heating of vegetative organic matter. Types of coal include lignite, bituminous coal, and anthracite.

coal-bed methane  methane that is trapped within the pores of coal within a coal seam

coastal straightening  the tendency for an irregular coast to be straightened over time by coastal erosion processes

cobble  sediment particle that is between 64 and 256 mm in diameter

col  the low point or pass along a ridge between two glacial valleys

columnar jointing  the fractures in volcanic rock forming columns that are typically 6-sided, resulting from cooling and contraction of the rock

composite volcano (or stratovolcano)  a volcano that is constructed of alternating layers of pyroclastic debris and lava flows

concentrate (mining)  a product of ore processing that includes a specific ore mineral, separated from the rest of the rock

concordant  parallel to pre-existing layering or foliation within a rock

cone of depression  the depression of the water table around a well that is heavily pumped

confined aquifer  an aquifer that lies below a confining layer

confining layer  an aquitard that overlies an aquifer and restricts the flow of water down from the surface

confining pressure  pressure resulting from the weight of overlying rocks

conglomerate  a sedimentary rock that is comprised predominantly of rounded grains that are larger than 2 mm

contact metamorphism  metamorphism that takes place adjacent to a source of heat, such as a body of magma

continental drift  the concept that tectonic plates can move across the surface of the Earth

continental glacier  a glacier that covers a significant part of a continent and has an area of at least 50,000 km2continental shelf  the shallow (typically less than 200 m) and flat sub-marine extension of a continent

continental slope  the steeper part of a continental margin, that slopes down from a continental shelf towards the abyssal plain

contractionism  the now discredited theory that mountain ranges formed as a result of the contraction of the Earth

convergent boundary  a plate boundary at which the two plates are moving towards each other

Cordilleran Ice Sheet  the continental glacier that covered part of western North America, including almost all of British Columbia, part of the Yukon, and part of northern Washington, during the Pleistocene glaciations

core  the metallic interior part of the Earth, extending from a depth of 2900 km to Earth’s centre

core-mantle boundary (CMB)  the boundary, at 2900 km depth, between the mantle and the core

Coriolis effect  the tendency for moving bodies (e.g., ocean currents) to rotate on the surface of the Earth, clockwise in the northern hemisphere and counter-clockwise in the southern hemisphere

cosmic microwave background (CMB)  a radiation “fog” left over from the an early stage in the development of the universe, when the universe was too dense to allow photons to travel far without being scattered

country rock  the original rock of a region, into which younger rock (typically igneous) rock has been intruded

covalent bond  a bond between two atoms in which electrons are shared

crater  a volcanic depression that is related to a specific volcanic vent

craton  a region of ancient (typically Precambrian) crystalline rock (equivalent to a shield)

creep  the very slow (mm to cm per year) flow of unconsolidated material on a gentle slope

crest  the highest point on a wave

crevasse  an open fissure on the surface of a glacier

cross bedding  small-scale inclined bedding within larger horizontal beds

crust  the uppermost layer of the Earth, ranging in thickness from about 5 km (in the oceans) to over 50 km (on the continents)

cryptocrystalline  refers to the texture of a rock or mineraloid in which crystals are so small that they are almost undetectable even with magnification

cyanobacteria  photosynthetic bacteria that evolved in the early Archean

D

D” layer (d-double-prime layer)  a low seismic velocity zone within the basal 200 km of the mantle

debris flow  a gravity-driven flow of water and sediment that includes a significant proportion of coarse (cobble to boulder) material

decline (mining)  a sloped tunnel used to access lower parts of a mine with wheeled equipment

decompression melting  melting (or partial melting) of rock resulting from a reduction in pressure without a significant reduction in temperature

dendritic  a pattern of drainage channels that resembles the branches in a tree

density  weight per volume of a substance (e.g., g/cm3)

deposition  when sediments are dropped out of the medium carrying them, and begin to accumulate in layers

deranged (drainage)  a pattern of drainage channels that is chaotic

detrital  referring to fragments of rocks or minerals

diatom  photosynthetic algae that make their tests (shells) from silica

differentiation  the un-mixing of a molten planetary body, resulting in the formation of a metallic core and a silicate mantle

dike  a tabular intrusive igneous body that is discordant to any existing layering in the country rock

diorite  an intermediate intrusive igneous rock

dip  the angle below horizontal at which a sedimentary bed or other feature slopes

directed pressure (also, differential stress, directional pressure) pressure which is greater in one direction than in others (e.g., compression, tension)

discharge  the volume of water flow in a stream expressed in terms of volume per unit time (e.g., m3/s)

discharge area  the part of an aquifer where groundwater discharge takes place

disconformity  a boundary between parallel sedimentary layers where some erosion of the lower layer has taken place

discordant  when a geological feature is not parallel to any existing layering in the country rock

dissolution  when water molecules take a substance apart by capturing its ions and keeping them separated (a type of chemical weathering)

divalent  an ion with a charge or +2 or -2

divergent  a plate boundary at which the two plates are moving away from each other

dodecahedron an object with twelve surfaces, such as a garnet crystal

dolomite a calcium-magnesium carbonate mineral (Ca,Mg)CO3. Also, a rock made out of that mineral (see also dolostone)

dolomitization  the addition of magnesium to limestone during which some or all of the calcium carbonate is converted to dolomite

dolostone  a carbonate rock made up primarily of the mineral dolomite

drainage basin  the catchment area of a stream, including the area where all surface water drains into the stream

drop stone  a fragment of rock within otherwise fine-grained sediment that has been dropped from floating ice on a body of water

drumlin  a streamlined glacial erosional feature comprised of sediments and/or bedrock

dyke see dike

E

eccentricity  (Milankovitch cycles) the degree to which the sun is offset from the geometric centre of the Earth’s orbit

eclogite  a garnet-pyroxene-glaucophane bearing rock that is the product of high-pressure metamorphism of oceanic crustal rock, typically within a subduction zone

effusive  a volcanic eruption dominated by the relatively gentle flow of lava

El Niño  a periodic climatic situation in which warm water extends all or most of the way to the eastern edge of the equatorial Pacific

elastic deformation deformation from which a material can fully recover if the stress is removed

electron  sub-atomic particle with a single negative charge

end moraine sediment deposit that accumulates at the front of a glacier

englacial  within a glacier, referring especially to sediment carried within the glacial ice

epicentre  the location on the surface vertically above the location (i.e., “hypocentre” or “focus”) where an earthquake takes place

epipelagic zone  the upper layer of water (0 to 200 m) in areas of the open ocean

epithermal deposit  a mineral deposit formed near to surface in an area of hydrothermal activity, typically associated with a body of magma

equilibrium line (glacier)  the line between the zone of accumulation and the zone of ablation (in late summer the equilibrium line is the boundary between snow-covered ice and bare ice)

equipotential lines (groundwater)   lines connecting locations with equal hydraulic head or water pressure

erosion  the process of transporting sediments away from their source

esker  a ridge of sediment deposited by a sub-glacial stream

eustatic sea level change  sea level change related to a change in the volume of the oceans, typically because of an increase or decrease in the amount of glacial ice on land

evaporite  a chemical sedimentary rock that forms when evaporation concentrates the ions in a solution to the point where they begin to precipitate out

exfoliation  (weathering) the fracturing of rock that results from a reduction in the pressure when overlying rock is eroded away

exoplanet  a planet that orbits a star other than the sun

extrusive  igneous rock that cooled at Earth’s surface

F

fall  (mass wasting) the vertical or near-vertical downward movement of rock

fault  boundary in rock or sediment along which displacement has taken place

feedback  when one process triggers others which either amplify or mute the original process

feldspar  a very common framework silicate mineral

feldspathic arenite  a sandstone consisting predominantly of sand-sized grains and cement (less than 15% fine-grained matrix), and with more than 10% feldspar grains

felsic  silica rich (>65% SiO2) in the context of magma or igneous rock

ferric  the oxidized form of an ion of iron (Fe3+)

ferromagnesian referring to a silicate mineral that contains iron and or magnesium

ferrous  the reduced (non-oxidized) form of an ion of iron (Fe2+)

fetch  the distance over which wind blows to form waves

finger lake  a lake that occupies a glacial valley

firn  the granular transitional state between snow and ice within a glacier

flood plain  the area that is occupied by water when a stream floods and overtops its banks

flow  a mass-wasting event where material moves which is saturated with water

flow path  the path that groundwater flows along between a recharge area and a discharge area

flowing artesian well  an artesian well in which the water level naturally rises above the surface of the ground

flux melting  melting of rock that is facilitated by the addition of a flux (typically water) which lowers the rock’s melting point

focus (earthquake)  the actual point below surface at which an earthquake takes place (equivalent to hypocentre)

foliation  the alignment of mineralogical or structural features of a rock – especially a metamorphic rock

footwall  the lower surface of a non-vertical fault

foraminifera  single-celled protist with a shell that is typically made of CaCO3fore-reef  the zone on the ocean side of a reef

formation  a unit of sedimentary rock that is lithologically consistent and sufficiently thick and extensive to be shown on a geological map at the scale that is typically used in the area in question

fracking  fracturing rock by injecting water and chemicals down a well at very high pressure (equivalent to hydraulic fracturing)

fractional crystallization  the sequential crystallization of minerals from magma, and the physical separation of early-forming crystals from the magma in the area where they crystallized

fracture  a break within a body of rock in which the rock on either side is not displaced

fringing reef  a reef adjacent to a shoreline where there is either a very narrow back reef area or none at all (in which case the reef is effectively attached to the shore)

frost line (also, snow line)  in the context of newly forming planetary systems, the distance beyond a star at which volatile components (e.g., water, carbon dioxide, methane, ammonia etc.) are frozen

frost wedging physical weathering caused when the expansion of freezing water pries rock apart

G

Ga  (giga annum) billions of years before the present

gabbro  a mafic intrusive igneous rock

Gaia hypothesis  the hypothesis advanced by James Lovelock that the organisms have affected the atmosphere and oceans such that conditions on Earth have been kept habitable, in spite of significantly changing energy received from the Sun

galaxy  gravitationally-bound system of stars and interstellar matter

gas giant  a large planet composed mostly of hydrogen and helium (e.g. Jupiter)

geosyncline  kilometres thick deposit of sediments that has accumulated along the edge of a continent and is sufficient mass to depress the crust beneath it

geothermal gradient  the rate of increase of temperature with depth in the Earth (typically around 30˚C/km within the crust)

giant impact hypothesis  the theory that the Moon formed when a Mars-sized planet (Theia) collided with the Earth at 4.5 Ga

glacial period  a period of Earth’s history during which glacial ice was present over a sufficient extent to have left recognizable evidence

glacial groove  a straight line created on a rock surface by erosion by a rock fragment embedded in the base of glacial ice (larger and deeper than a glacial striation)

glacial striation  a straight line created on a rock surface by erosion by a rock fragment embedded in the base of glacial ice (finer than a glacial groove – typically less than 1 cm wide)

glacier  a long lasting (centuries or more) body of ice on land that moves under its own weight

glaciofluvial  referring to sediments deposited from a stream that is derived from a glacier

glaciolacustrine  referring to sediments deposited within a lake in a glacial environment

glaciomarine   referring to sediments deposited within the ocean in a glacial environment

glaucophane  a blue sodium-magnesium-bearing amphibole mineral that forms during metamorphism at high pressures and relatively low pressures, typically within a subduction zone

gneiss  high-grade foliated metamorphic rock in which the mineral components are separated into bands of different composition

graben a down-dropped fault block, bounded on either side by normal faults

grade  1. (mineral deposit) the amount of a specific metal or mineral expressed as a proportion of the whole rock. 2. (coal) the extent to which carbon has been concentrated within the coal, and the possible energy output on combustion has increased

graded bedding  an individual sedimentary layer that shows a distinctive gradation in grain size (normal graded bedding is finer towards the top, reverse graded bedding is coarser towards the top)

gradient  the slope of a stream bed over a specific distance, typically expressed in m per km

grain size  the diameter of a fragment (clast) of sediment

granite  a felsic intrusive igneous rock

granule  a sedimentary particle ranging in size from 2 to 4 mm in diameter

greenhouse gas  a gaseous molecule with 3 or more atoms that is able to absorb infrared radiation

greenhouse effect (climate) the ability of an atmosphere to absorb infrared radiation due to the presence of greenhouse gases

greenschist  1. (metamorphic rock) a foliated metamorphosed rock (typically derived from basalt) in which the green colouration is derived from either chlorite, epidote, or green amphibole. 2. (metamorphic facies) low-grade metamorphic facies characteristic of regional metamorphism

greenstone  a non-foliated metamorphosed rock (typically derived from basalt) in which the green colouration is derived from either chlorite, epidote or green amphibole. Can be formed by hydrothermal metamorphism on the ocean floor.

greywacke  a sandstone with more than 15% silt and clay, and with a significant proportion of sand-sized rock fragments

groundwater  water that lies beneath the surface of the ground

group  a stratigraphically continuous series of related formations

groyne  a man-made structure extending from the shore built to deflect the energy of waves

gyre  a closed circular ocean current

H

habit  a characteristic crystal form or combination of forms of a mineral

habitable zone  the region around a star that is considered to be suitable for a life-bearing planet

Hadean  the first eon of Earth history, extending from 4.57 to 3.80 Ga

halide  a mineral in which the anion is one of the halide elements (e.g., halite – NaCl or fluorite – CaF2)

halite  NaCl, a halide mineral which consititutes table salt

halogen an element in the second-last column of the periodic table that forms anions with a negative-1 charge

hanging valley  a glacial valley created by a tributary glacier which does not erode as deeply as the main-valley glacier that it joins

hanging wall  the upper surface of a non-vertical fault

headland  a point extending out to sea

horn  a peak that has been eroded on at least three sides by glaciers

hornfels  a fine-grained metamorphic rock that is not foliated. It can have a variety of parent rocks.

horst  an uplifted fault block, bounded on either side by normal faults

hot spot  the surface area of volcanism and high heat flow above a mantle plume

hydrated mineral  a mineral that includes either hydroxyl (OH) or water (H2O) in its chemical formula (e.g., gypsum CaSO4.2H2O)

hydraulic conductivity  an expression of the rate at which a liquid will flow through a porous medium, as determined by the permeability of the medium and the viscosity of the liquid

hydraulic fracturing  fracturing rock by injecting water and chemicals down a well at very high pressure (equivalent to fracking)

hydrolysis  a reaction between a mineral and water in which H+ ions are added to the mineral and a chemically equivalent amount of cations are released into solution

hydrothermal  refers to hot water solutions and processes involving hot water solutions

hydrothermal alteration  chemical alteration of minerals by hot water solutions

hydroxide  the anion OH or an mineral that includes that anion

hypocentre  the actual point below surface at which an earthquake takes place (equivalent to focus)

 

I

ice giant  a planet that is comprised mainly of gases heavier than hydrogen and helium, including oxygen, carbon, nitrogen, and sulfur (e.g., Uranus and Neptune)

igneous  a rock formed from the cooling of magma

illite  a clay mineral with a composition similar to that of muscovite mica

imbricate  aligned and overlapping, like the tiles on a roof

index fossil  a fossil with a distinctive appearance and a wide geographic range but from a relatively restricted time range, thus making it useful for dating a correlating rocks from different regions (the most useful index fossils are from organisms that lived for less than a million years)

index mineral  (metamorphic rocks) a mineral with a stability range of pressures and temperatures sufficiently narrow so as to be useful in indicating the pressures and temperatures at which a metamorphic rock formed.

inert  in chemistry, an element that does not readily react with other elements (e.g., neon)

infiltration  the recharge of groundwater from the downward percolation of surface water

insolation  a measure of the intensity of solar energy at a specific location or time (expressed in W/m2)

intensity  in seismology, a qualitative measure of the amount of shaking at specific location, based on what was felt by observers, or the amount of damage done

Intergovernmental Panel on Climate Change  (IPCC) an international body established in 1988 by the UN’s World Meteorological Organization and the UN Environment Program to prepare periodic reports on the status of global climate change and its mitigation

intrusive  an igneous rock (pluton) that has cooled slowly beneath the surface

ionic bond  a bond in which electrons are transferred from one atom to another, thus forming ions

ion  an atom that has either gained or lost electrons and has thus become charged (or a group of atoms that also has a charge – e.g., HCO3)

isoclinal fold  a tight fold in which the limbs are parallel to each other

isostasy  the equilibrium between a block of crust floating on the underlying plastic mantle

isostatic sea-level change  the effect on relative sea level of a vertical adjustment of the crust resulting from a change in the mass of the crust (e.g., from losing or gaining ice)

isotherm  a surface or line drawn to represent points at the same temperature. (iso = same)

isotope a form of an element that differs from other forms because it has a different number of neutrons (e.g., 16O has 8 protons and 8 neutrons while 18O has 8 protons and 10 neutrons)

J

joint  a fracture in rock where the rock on one side has not moved relative to the other side

jointing  the formation of joints

Jovian planet  a gas giant planet

K

ka (kilo annum) thousands of years before the present

kaolinite  a clay mineral that does not have cations other than Al and Si

karst  the solutional erosion of an area with soluble rock (typically limestone) to form depressions and caves

kettle  a depression formed at the front of a large glacier when a stranded ice block that was surrounded by sediment eventually melts

kettle lake  a lake that forms within a kettle

kimberlite  an ultramafic volcanic rock that originates at significant depth (> 200 m) in the mantle (some kimberlites include diamonds)

Kuiper belt  a region of the Solar System beyond the orbit of Neptune that is populated by small objects and dwarf planets (including Pluto)

L

laccolith  concordant intrusion in which the central part has bulged upward

lahar  a mudflow or debris flow that is either caused by a volcanic eruption, or forms on the flank of a volcano as a result of flooding not related to an eruption

landfill gas  gases produced within a landfill during the microbial breakdown of landfill components (most are dominated by carbon dioxide and methane)

large igneous province (LIP)  a very large area of mafic volcanic rock produced by a massive eruption typically related to a mantle plume

lateral moraine  a deposit of rocky material that forms along the margin of a valley or alpine glacier, mostly from the freeze-thaw release of material from the steep slopes above

lattice  the regular and repeating three-dimensional structure of a mineral

Laurentide Ice Sheet  the continental glacier that extended across central eastern North America during the Pleistocene, covering most of Canada and a significant part of the United States

lava  molten rock on Earth’s surface (cf. magma)

lava levée  a ridge that forms along the edge of a lava flow because the magma at the edge cools faster than that in the middle

lava tube  a tube that forms as mafic lava flows along a channel and lava leveés build up on either side, eventually forming a roof (once a lava tube forms it insulates the flowing magma, allowing it to stay hot a liquid for longer and therefore flow much further)

leachate  in the context of landfills, the liquid (rainwater) that passes through the waste and becomes contaminated with soluble components from the waste

levée  on a stream, the ridge that naturally forms along the edge of the channel during flood events

level  in mining, a horizontal mine opening

light year  the distance that light can travel in one year (9.4607 x 1012 km)

lignite  a low-grade type of coal with less than 70% carbon

limbs  the layers of rock on either side of a fold

limestone  a biochemical sedimentary rock that is comprised mostly of calcite

liquefaction  the tendency for unconsolidated and water saturated sediments to lose strength during seismic shaking

lithic arenite  an arenite in which there is more than 10% lithic clasts and in which there are more lithic clasts than feldspar clasts (see also arenite)

lithic clasts  fragments of another rock which are included in the sand-sized grains in sandstone, or in the larger grains in conglomerate

lithification  the conversion of unconsolidated sediments into rock by compaction and cementation

lithosphere  the rigid outer part of the Earth, including the crust and the mantle down to a depth of about 100 km

lithostatic pressure  pressure due to the weight of overlying rocks

lodgement till  sediment that accumulates at the base of a glacier and typically has a wide range of grain sizes (including clay) and is well compacted

long axis  in a crystal, clast, or grain, the direction in which the length would be the greatest

longshore current  the movement of water along a shoreline produced by the approach of waves at an angle to the shore

longshore drift  the movement of sediment along a shoreline resulting from a longshore current and also from the swash and backwash on a beach face

Love wave  a surface seismic wave, with horizontal motion, that develops in relatively weak (e.g., unconsolidated) materials at surface

luvisol  a cold climate forest soil formed in which clay has been removed from the A horizon and relocated into the B horizon

M

Ma (Mega annum) millions of years before the present

mafic  silica poor (<45% SiO2) in the context of magma or igneous rock, and containing ferromagnesian minerals such as olivine and pyroxene)

magma  molten rock within Earth’s interior (cf. lava)

magnetic chronology  the study of the timing of reversals of the Earth’s magnetic field, and the application of that understanding to dating geological materials

magnitude  a measure of the amount of energy released by an earthquake

mantle  the middle layer of the Earth, dominated by iron and magnesium rich silicate minerals and extending for about 2900 km from the base of the crust to the top of the core

mantle plume  a plume of hot rock (not magma) that rises through the mantle (either from the base or from part way up) and reaches the surface where it spreads out and also leads to hot-spot volcanism

marble  a non-foliated metamorphic rock derived from a limestone or dolostone protolith, in which the calcite or dolomite has been recrystallized into larger crystals

mass wasting  the mass failure, by gravity, of rock or unconsolidated material on a slope

matrix  finer-grained material between larger clasts within a sedimentary rock

maturity  the degree to which a sediment or sedimentary rock exhibits characteristics of prolonged physical and chemical weathering and transpor,t

meander cutoff  the formation of a shorter stream channel across the narrow boundary between two meanders on a stream

meandering  the sinuous path taken by a stream within a wide flat flood plain

mechanical weathering (also, physical weathering)  weathering that occurs when physical processes cause a rock to break into smaller pieces without changing the chemical composition

medial moraine  a lateral moraine that has been shifted towards the centre of a valley glacier at a point where two glaciers meet

member  a subdivision of a formation

mesopelagic zone  the upper middle zone of the open ocean extending from 200 to 1000 m depth

metallic lustre  the lustre of a mineral into which light does not penetrate but reflects off of the surface without being scattered (i.e., shines reflects light like a shiny metal)

metallic bond  a type of bond in which abundant electrons are easily shared amongst cations

metamorphic facies  a group of metamorphic rocks formed under the same range of pressures and temperature conditions, but from different parent rocks

metamorphic grade  refers to the intensity of metamorphism, and increases as pressure and temperature increase

metamorphism  the transformation of a parent rock into a new rock as a result of heat and pressure that leads to the formation of new minerals, or recrystallization of existing minerals, without melting

metasomatism  metamorphism facilitated by ion transfer through water, and which results in a substantial change in the chemical composition (not just the mineral content) of a rock

meteoroid  a small fragment of stony or metallic debris in space

methane hydrate  a combination of water ice and methane in which the methane is trapped inside “cages” in the ice

mica  a sheet-silicate mineral (e.g., biotite, muscovite)

migmatite   rock that is part metamorphic and part igneous, formed at  very high grades of metamorphism when a part of the parent rock starts to melt

Milankovitch cycles  millennial-scale variations in the orbital and rotational parameters of the Earth that have subtle effects on the Earth’s climate

Mohorovičić discontinuity (Moho) the boundary between the crust and the mantle

moment magnitude  a way of estimating earthquake magnitude based on the area of the rupture surface and the amount of displacement

monogenetic  a volcano that forms in a single eruptive event

moraine lake  a finger lake that forms within a glacial valley and is dammed by an end moraine

mud crack  a dessication crack formed when mud shrinks as it dries

mudflow  a mass-wasting event involving the flow of mud (sand, silt and clay) within a channel

mudrock  an inclusive term for mudstone, shale and claystone

mudstone  a fine-grained clastic sedimentary rock with a mixture of silt-sized and clay-sized particles

muscovite  a potassium-bearing non-ferromagnesian mica

N

native element (also, native element mineral) a mineral that consists of only one element (e.g., native gold)

nebula  a large cloud of dust and gas in space, frequently hosting the formation of stars

negative feedback  a process that results in a decrease in that process (in the context of climate change it is a process that reduces the change in climate, such as the enhanced growth of vegetation in response to an increase in atmospheric carbon dioxide)

neutron  a sub-atomic particle with a mass of 1 and a charge of 0

nonconformity  a geological boundary where non-sedimentary rock is overlain by sedimentary rock

non-ferromagnesian mineral  a silicate mineral that does not contain iron or magnesium (e.g., feldsspar)

non-metallic lustre  the lustre of a mineral into which light does penetrate, or which does not produce a bright reflection

normal fault  a non-vertical fault along which the hanging wall (upper surface) has moved down relative to the footwall

normal force  the component of the gravitational force that acts directly into the slope

North Atlantic Deep Water  deep Atlantic Ocean water that has descended in the far north of the basin in the area between Scandinavia and Greenland

nunatuk  a rocky peak that extends above the ice level of a continental glacier

O

obliquity  (Milankovitch cycles) the angle of the tilt of the Earth’s rotational axis with respect to the plane of its orbit around the sun

ocean plain  the extremely flat surface of the deep ocean floor in areas unaffected by plate tectonic processes and volcanism

oil window  the depth range, which is approximately 2000 to 4000 m, within which the temperature is appropriate for the formation of oil from organic matter in sedimentary rock

ooid  a small (approximately 1 mm) sphere of calcite formed in areas of tropical shallow marine water with strong currents

olivine a silicate mineral made up of isolated silica tetrahedra and with either iron or magnesium (or both) as the cations

Oort cloud  a spherical cloud of icy objects extending from between about 5,000 and 500,000 astronomical units (Sun-Earth distances) from the Sun (thought to be a source area of comets)

open-pit mine  a mine that is open to the surface

organic sedimentary rock  a sedimentary rock consisting of materials made of carbon-hydrogen bonds (e.g., animal and plant material)

outcrop  a surface exposure of rock that is part of the crust (bedrock)

outwash plain  an extensive region of sand and gravel deposited by streams flowing out of a glacier (same as sandur)

overturned  a geological feature that has been tilted to the point where it is upside down

oxbow  a part of a stream meander that has become isolated from the rest of the stream as the result of a meander cutoff

oxidation  the reaction between a mineral and oxygen

oxide  a mineral in which the anion is oxygen (e.g., hematite Fe2O3)

P

pahoehoe  a lava flow with a ropy surface texture formed when the surface cools and hardens while the lava beneath is still flowing

paleomagnetic  characterized by past variations in the intensity and polarity of the Earth’s magnetic field

Pangea  che supercontinent that existed between approximately 300 and 180 Ma

paraconformity  an interruption representing a period of non-deposition, without tilting or erosion, in a sequence of sedimentary rocks

parasitic fold  a fold within a fold

parent rock (also, parent material, protolith)  the rock that was already in existence when a process of metamorphism started, or the rock from which sediments were derived

partial melting  the process during which a only specific mineral components of a rock melt

parting  a narrow gap between individual sedimentary layers

passive margin  a boundary between a continent and an ocean at which there is no tectonic activity (e.g., the eastern edge of North America)

paternoster lake  one of a series of rock basin lakes

peat  a product of the first stage of coal formation, where vegetative material undergoes limited decomposition in a low-oxygen, acidic environment

pebble  a sedimentary particle ranging in size from 2 to 64 mm (includes granule)

pelagic  the part of a lake or the ocean that is not close to shore

permafrost  ground that remains frozen for two or more years

permanentism  the now discredited theory that the features on the Earth have not changed significantly over geological time

permeability  an expression of the ease with which liquid will flow through a porous medium

phaneritic  a rock texture in which the individual crystals or grains are visible to the naked eye

Phanerozoic  the most resent eon of geological time, encompassing the Paleozoic, Mesozoic and Cenozoic eras

phenocryst  a relatively large crystal within an igneous rock

phyllosilicate  a silicate mineral in which the silica tetrahedra are made up of sheets

phosphate  a mineral in which the anion is PO43-photic zone  the upper 200 m of the ocean or a lake, where, depending on the turbidity of the water, light can penetrate

phreatic eruption  a steam-drive volcanic eruption that takes place when surface or near-surface water is heated by volcanic activity

phyllite  a metamorphic rock with slaty cleavage and a sheen on the surface produced by aligned micas

physical weathering  (also, mechanical weatheringweathering that occurs when physical processes cause a rock to break into smaller pieces without changing the chemical composition

pillow  a pillow-shaped mass of volcanic rock (typically basalt) formed when magma erupts beneath the surface

pillow lava  a volcanic rock (typically basalt) that is made up primarily of pillows

pipe  a cylindrical body of igneous rock. May feed a volcano or connect plutons

plate  a fragment of lithosphere (crust and upper-most mantle) that is moving across the surface of the Earth as a single unit

plate tectonics  the concept that the Earth’s crust and upper-most mantle (lithosphere) is divided into a number of plates that move independently on the surface and interact with each other at their boundaries

Plinian eruption  a large volcanic eruption in which a column of hot tephra and gases rises many kilometres into the atmosphere

pluton  a body of igneous rock formed by cooling within the Earth (i.e., a body of intrusive igneous rock)

podzol  a soil with well-developed horizons formed in temperate forested regions

podzolization  the process of the formation of podsol

polar wandering path  see: apparent polar wandering path

polymerize  the formation of molecular chains within a fluid (e.g., a magma) that lead to an increase in the fluid’s viscosity

polymorphs  two or more minerals with the same chemical formula but different crystal structures

porosity  the percentage of open pore space within a body of rock or sediment

porphyritic  an igneous texture in which some of the crystals are distinctively larger than the rest

porphyry deposit  a mineral deposit (of copper or molybdenum especially) in which part of the host rock is a porphyritic stock

positive feedback  a process that results in an increase in that process (in the context of climate change it is a process that enhances the change in climate, such as the reduced reflectivity of the Earth’s surface when ice melts)

potassium feldspar  feldspar with the formula KAlSi3O8, and which is a common constituent of felsic igneous rocks

potentiometric surface  the imaginary surface defined by the levels to which water would rise in a series of wells drilled into a confined aquifer

precession (Milankovitch cycles) the variation in the direction at which the Earth’s rotational axis is pointing

pressure-release cracking  cracking of a rock which occurs when overlying rocks are removed by erosion and the outer layer of the rock expands

principle of cross-cutting relationships  the principle that a body of rock that cuts across or through another body of rock is younger than that other body

principle of faunal succession  the principle that life on Earth has evolved in an orderly way, and that we can expect to always find fossils of a specific type in rocks of a specific age

principle of inclusions  the principle that inclusions within a body of rock must be older than the rock

principle of original horizontality  the principle that sedimentary beds are originally deposited in horizontal layers

principle of superposition  the principle that in a sequence of layered rocks that is not overturned or interrupted by faulting, the oldest will be at the bottom and the youngest at the top

proglacial  referring to the area in front of a glacier

protolith  (also, parent rock) the rock which was altered to produce a metamorphic rock

proton  a sub-atomic particle with a mass of 1 and a charge of 1

protoplanetary disk  a rotating cloud of gas and dust surrounding a young star

pumice  a highly vesicular (filled with holes left by gas bubbles) felsic volcanic rock (typically composed mostly of glass)

p-wave  a seismic body wave that is characterized by deformation of the rock in the same direction that the wave is propagating (compressional vibration)

pyroclastic  volcanic material formed during an explosive eruption

pyroclastic density current  a body of hot pyroclastic rock and gases that is flowing rapidly down the flank of a volcano

pyroxene  a single chain silicate mineral

Q

quartz  a silicate mineral with the formula SiO2quartz sandstone (also, quartz arenite) a sandstone in which more than 90% of the grains are quartz

quartzite  a non-foliated metamorphic rock formed from the contact or regional metamorphism of sandstone

R

radial (drainage) a pattern of streams radiating out from a central point, typically an isolated mountain

radioactivity  the natural transformation of unstable isotopes into new elements

radiolaria  microscopic (0.1 to 0.2 mm) marine protozoa that produce silica shells

Rayleigh wave  a surface seismic wave, with vertical motion

recharge  the transfer of surface water into the ground to become groundwater

recharge area  an area of an aquifer where recharge is predominant over discharge

recrystallization  during metamorphism, mineral crystals dissolving and reforming as larger crystals

rectangular drainage a pattern in which tributaries typically flow at right angles to each other and meet at right angles

recumbent fold  a fold that is overturned such that its limbs are close to horizontal

redshift  the increase in wavelength of light resulting from the fact that the source of the light is moving away from the observer

reef  a mound of carbonate formed in shallow tropical marine environments by corals, algae and a wide range of other organisms

regional metamorphism  metamorphism caused by burial of the parent rock to depths greater than 5 km (typically takes place beneath mountain ranges, and extends over areas of hundreds of km2)

remnant magnetism  magnetism of a body of rock that formed at the time the rock formed and is consistent with the magnetic field orientation that existed at that time and place (see also paleomagnetism)

reservoir rock  rock into which petroleum has migrated and is now trapped

residual soil  soil formed by weathering of the underlying rock or sediment

retrograde metamorphism  metamorphism that transforms a higher grade metamorphic rock into a lower grade metamorphic rock

reverse fault  a non-vertical fault along which the hanging wall (upper surface) has moved up relative to the footwall

rhyolite  a felsic volcanic rock

ridge push  the concept that at least part of the mechanism of plate motion is the push of oceanic lithosphere down from a ridge area

rip current  a strong flow of water outward from a beach

ripple  a series of small parallel ridges formed within sediment that has accumulated in moving water or wind

rip-rap  angular rock fragments, typically boulder sized, used to armour slopes and shorelines against erosion

roche moutonée  a product of glaciation in which a bedrock protrusion is eroded into a streamlined shape that has a broken or jagged leading (down-ice) edge

rock avalanche  a rapid turbulent flow of broken bedrock fragments down a steep slope

rock basin lake  a lake situated in a rock basin carved at the upper end of an alpine glacier

rock cleavage  the tendency of a rock to break along planes defined by foliation

rock cycle  the series of processes through which rocks are transformed from one type to another

rock fall   the near-vertical fall or bouncing of rock released from a steep slope

rock slide  the translational motion of an essentially intact body of rock down a slope (rock slides are typically slow, because once they start to move fast the rock body becomes fragmented and then flows as a rock avalanche)

root wedging  a physical weathering process in which roots grow into cracks in rocks and force them open

rounding  describes the extent to which clasts have had their edges and corners smoothed off

runoff  flow of water down a slope, either across the ground surface, or within a series of channels

rupture  breaking of rock subject to stress, typically resulting in an earthquake

rupture surface  the area over which rock rupture takes place during an earthquake

S

sackung  an escarpment or trough at the top of a slow-moving rock slide (sackungen)

saltation  the bouncing of particles along a stream bottom or desert floor

salt wedging  a physical weathering process in which water with dissolved salt flows into a crack, and as the water evaporates, salt crystals grow and push the crack open

sand  a mineral or rock fragment ranging in size from 1/16th to 2 mm

sandstone  a rock that is primarily comprised of sand-sized particles

sandur  an extensive region of sand and gravel deposited by streams flowing out of a glacier (same as outwash plain)

saturated zone  the part of an aquifer, or any body of rock, that is saturated with water

schist  a foliated metamorphic rock with crystals large enough to be visible to the unaided eye

sea cave  a shallow cave formed on a rocky shore by wave erosion

sea cliff  a coastal escarpment that is typically eroding inland as a result of wave action

sea-floor spreading  the formation of new oceanic crust by volcanism at a divergent plate boundary

sector collapse  the sudden collapse of a significant part of the flank of a volcano

sedimentary rock  rock that has formed by the lithification of sediments or by the precipitation of ions from water

sediments  unconsolidated (loose) particles of mineral or rock

seismic  pertaining to earthquakes

seismic moment  a measurement of an earthquake’s energy based on longwave vibrations, or on the product of the fault area and displacement

seismic reflection sounding  measurement of the properties of sediments based on detection of sounds generated at surface and reflected from layers beneath the surface

septae  calcareous partitions between the successive living chambers in a cephalopod

septic system  a system constructed to facilitate the dispersion and detoxification of sewage (typically includes a septic tank and a drainage field)

shaft  a vertical opening at a mine

shale  a silt- and clay-rich rock that has evidence of layering

shatter cone  conical nested fractures that result from extraterrestrial impacts. Cones point toward the impact.

shear force  the component of the gravitational force in the direction parallel to a slope

shear strength  the strength of a body of rock or sediment that counteracts the shear force

shear stress  the stress placed on a body of rock or sediment adjacent to a fault

sheeted dikes  a series of near-vertical dykes formed in the vicinity of a spreading ridge when magma from depth flows into fractures formed by extensional forces

sheet silicate  a silicate mineral in which the silica tetrahedra are combined within sheets

sheetwash  overland flow of water, typically related to a heavy precipitation event

shield  a region of ancient (typically Precambrian) crystalline rock (equivalent to a craton)

shield volcano  a low-profile volcano formed primarily from eruptions of low-viscosity mafic magma

shocked quartz  quartz crystals in which the structure has been deformed by sudden, intense pressure. Deformation is visible as parallel lines within the crystal. with damage along parallel plains

Sial (sialic)  an outdated term referring to rock or magma in which silica and aluminum are the predominant components (generally equivalent to felsic)

silica  a form of the mineral quartz (SiO2)

silica tetrahedron  an ion which is a combination of 1 silicon atom and 4 oxygen atoms that form a tetrahedron shape (SiO44-)

silicate  a mineral that includes silica tetrahedra

silicon  the 14th element

silicone  resin or caulking made from silicon-oxygen chains and various organic molecules

sill  a tabular igneous intrusion (pluton) that is parallel to existing layering in the country rock

silt  sedimentary particles ranging is size from 1/256th to 1/16th of a mm

siltstone  a clastic sedimentary rocks consisting predominately of silt-sized particles

Sima (simatic)  an outdated term referring to rock or magma in which silica, magnesium and iron are the predominant components (generally equivalent to mafic)

skarn  the contact metamorphism (and metasomatism) of limestone

slab pull  the concept that at least part of the mechanism of plate motion is the pull of oceanic lithosphere down into the mantle

slate  a fine-grained metamorphic rock that splits easily into sheets

slaty cleavage  the tendency for slate or phyllite to split into sheets (note that this is the only situation in this textbook where the term “cleavage” is applied to a rock as opposed to a mineral)

slide  the downward movement of rock or sediment on a slope as an intact mass

slump  a slide in which the nature of the motion is rotational (typically only develops in unconsolidated sediments)

smectite  a fine-grained sheet silicate mineral that can accept water molecules into interlayer spaces, resulting is swelling

smelter  a refinery at which minerals are processed to produce pure metals

snow line (frost line)  in the context of newly forming planetary systems, the distance beyond a star at which volatile components (e.g., water, carbon dioxide, methane, ammonia etc.) are frozen

soil horizon  a layer, within a well-developed soil, that is physically or chemically different from layers above or below

solar system  a star and the planets surrounding it. Sometimes used specifically for the sun and its planets, and planetary system used for other stars

solar wind  a stream of ionized (charged) particles away from the sun

solid solution  the substitution of one element for another in a mineral (e.g., in Bowen’s reaction series there exists a continuum of plagioclase feldspar where calcium becomes progressively less common, and sodium more so)

solifluction  the flow of water saturated sediment or soil over a stronger and less permeable substrate

sorting  the extent to which the grain size within a sample of sediment is similar.  Well-sorted sediments have very similar grain sizes, and poorly-sorted sediments have a variety of grain sizes.

source rock  the sedimentary rock from which petroleum originates prior to its migration into a reservoir rock

speleothem  a cave structure formed when calcium carbonate precipitates (see also stalactite, stalagmite)

sphericity  the extent to which a grain is the same diameter in all dimensions (e.g., more like a sphere, but without implying roundness or smoothness)

spit  a sand or coarser deposit extending from shore out into open water

spring  a flow of groundwater onto the surface

stack  a prominent rocky island that is a remnant of the erosion of a headland

stage  the level of water in a stream

stalactite  a cone-shaped speleothem that is suspended from the roof of a cave

stalagmite  a cone-shaped speleothem that forms on the floor of a cave

step-pool  a characteristic of stream flow in which water flows from one pool to another, typically on a stream with a steep gradient

stock  an irregular pluton with n exposed area less than 100 km2stoping  the fracturing and incorporation of fragments of country rock as a magma body moves upward through the crust

strain  the deformation of rock that is subjected to stress

streak  the mark left on a porcelain plate when a mineral sample is ground to a powder by being rubbed across the plate (typically provides a more reliable depiction of the colour than the whole sample)

stream  any body of flowing water

stress a force applied to a rock (specifically, the force per unit area)

stress transfer the change in the pattern of stress on a region of rock as a result of an earthquake (typically stress is reduced in the area of a rupture zone, but is increased elsewhere in the vicinity)

strike  the compass direction of a horizontal line on a sloped surface (e.g., bedding plane, fracture etc.)

strike-slip fault  a fault that is characterized by motion that is close to horizontal and parallel to the strike direction of the fault

subaerial eruption  a volcanic eruption that takes place on land

subaqueous eruption  a volcanic eruption that takes place under water

subducted  when part of a plate is forced beneath another plate along a subduction zone

subduction zone  the sloping region along which a tectonic plate descends into the mantle beneath another plate

subglacial  beneath a glacier

sulphate  a mineral in which the anion is SO42-sulphide  a mineral in which the anion is S2-supergroup   a stratigraphically continuous series of related groups

superterrane  a number of terranes that are contiguous

supraglacial  on the surface of a glacier

surf zone  the near-shore zone where waves are breaking into surf

suture  the line on the surface of a cephalopod that marks the boundary between a septum and the outer shell

swash  the upward motion of a wave on a beach (typically takes place at the same angle that the waves are approaching the shore)

s-wave  a seismic body wave that is characterized by deformation of the rock transverse to the direction that the wave is propagating

symmetrical a fold in which the limbs are at the same angle to the hinge

syncline a downward fold where the beds are known not to be overturned

synform  a downward fold where it is not known if the beds are overturned

T

tabular  referring to a structure that is sheet-like (or like a table top). See also dike, sill

tailings  the fine-grained waste rock from a plant used to concentrate ore minerals

talus slope  a sloped deposit of angular rock fragments at the base of a rocky escarpment

tarn  a lake within a rock basin

tectonic plate  a fragment of the lithosphere that moves across the surface of the Earth as a single unit

tectonic sea level change  relative sea level change related to the vertical motion of a crustal block caused by tectonic processes

tephra  fragments of volcanic rock (including volcanic ash) ejected during an explosive eruption

terminal moraine  and end moraine that marks the farthest forward advance of a glacier

terrane  a block of crust that has geological features which are distinctive from neighbouring regions, and is assumed to have been moved from elsewhere by tectonic processes

terrestrial planet  a planet with a rocky mantle and crust, and metallic core (e.g., Earth)

terrigenous  referring to sedimentary particles that originated on a continent

test  the shell-like hard parts (either silica or carbonate) of small organisms such as radiolarian and foraminifera

thrust fault  a low angle reverse fault

till  unsorted sediment transported and deposited by glacial ice

tiltmeter  a sensitive instrument used to monitor subtle changes in the tilt of the land, particularly in studies of active volcanoes

tombolo  a sand or coarser deposit connecting an island or rocky prominence to a larger body of land

traction  a force that contributes to the movement of particles situated on a stream bed or desert floor

transform fault  a boundary between two plates that are moving horizontally with respect to each other

transportation  refers to moving sediments from one location to another

transported soils  soils which form on sediments that have been moved from their original location.  The soils themselves have not been transported.

travertine  a deposit of calcium carbonate that forms at springs, hot springs or within limestone caves

trellis  a drainage pattern in which tributaries typically flow parallel to one other but meet at right angles

trigger  an event, such as an earthquake or a heavy rainfall, that starts a mass wasting event

trough  the lowest point of a wave

truncated spur  the steep end of a ridge or arête that has been eroded by a main-valley glacier

tsunami  a long-wavelength wave produced by the vertical motion of the floor of the ocean or a large lake, typically related either to an earthquake or a sub-marine mass wasting event

tufa  a form of travertine that is especially porous as it forms around existing vegetative material.

tuya  a flat-topped volcanic hill or mountain that formed when an eruption took place beneath a glacier and the melting led to the formation of a lake that then resulted in the wave-erosion of the top of the volcano

U

unconfined aquifer  an aquifer that is not overlain by a confining layer

unconformity  a geological boundary at the base of a sedimentary layer

unconformity-type uranium deposit  a uranium deposit that has formed at a nonconformity between sandstone and older rock

uncompressed density  the density of planetary material that it would have it was not compressed by the planets gravitational force

underground storage tank (UST) an underground tank for storing liquids, most commonly for liquid fuel

unsaturated zone  the rock or sediment above the water table

U-shaped valley  a relatively straight valley with a flat bottom and steep sides that has been carved by a valley glacier

V

valley glacier  a glacier formed in a mountainous region and confined to a valley (same as alpine glacier)

varve  a recognizable layer within sediments that represents a single year of deposition

vesicular  an igneous texture characterized by holes left by gas bubbles

volcanic glass  lava that has cooled within minutes, not allowing time for the formation of crystals

volcanic-hosted massive sulphide  a mineral deposit hosted by volcanic rocks and including zones where most of the rock is made up of sulphide minerals (including ore minerals and pyrite)

W

wacke  a sandstone with more than 15% clay and silt

water table  the upper surface of the saturated zone in an unconfined aquifer

wave base  the depth of water that is affected by the sub-surface orbital motion of wave action (approximately one-half of the wavelength)

wave-cut platform  a nearly-horizontal bench of rock eroded by waves within the surf zone (equivalent to wave-cut terrace)

wavelength  the distance between the crests of two waves

weathering  a range of processes taking place in the surface environment, through which solid rock is transformed into sediment and ions in solution

wedging  physical (mechanical) weathering processes which involve forcing open cracks in a rock (see also frost wedging, root wedging, salt wedging)

Western Canada Sedimentary Basin  a large basin in the western interior of Canada, east of the Rocky Mountains, extending from the northern United States to the Northwest Territories

Wisconsin Glaciation  the most recent advance of the Pleistocene glaciations, extending from 85 to 11 ka

X

xenolith (zee-know-lith) A fragment of country rock incorporated into igneous rock, commonly as a result of stoping

Y

youthful stream  a stream that is actively down-cutting its valley in an area that has recently been uplifted

Z

zone of ablation  the part of a glacier, below the equilibrium line, where there is net loss of ice mass due to melting and calving

zone of accumulation  the part of a glacier, above the equilibrium line, where there is net gain of ice mass because not all of the snow that falls each winter is able to melt during the following summer

2

Appendix A. List of Geologically Important Elements and the Periodic Table

The following table includes 36 of the geologically important elements, listed alphabetically by their element name, along with their atomic number and the atomic mass of their most stable isotope.

The geologically most important elements are bolded, and the eight main elements of silicate minerals are identified with an asterisk (*).

Symbol Name Atomic No. Atomic Mass
Al* Aluminum 13 27
As Arsenic 33 75
Ba Barium 56 137
Be Beryllium 4 9
B Boron 5 11
Cd Cadmium 48 112
Ca* Calcium 20 40
C Carbon 6 12
Cl Chlorine 17 35
Cr Chromium 24 52
Co Cobalt 27 59
Cu Copper 29 64
F Flourine 9 19
Au Gold 79 197
He Helium 2 4
H Hydrogen 1 1
Fe* Iron 26 56
Pb Lead 82 207
Mg* Magnesium 12 24
Mn Manganese 25 55
Mo Molybdenum 42 96
Ne Neon 10 20
Ni Nickel 28 59
N Nitrogren 7 14
O* Oxygen 8 16
P Phosphorus 15 31
Pt Platinum 78 195
K* Potassium 19 39
Si* Silicon 14 28
Ag Silver 47 108
Na* Sodium 11 23
Sr Strontium 38 88
S Sulfur 16 32
Ti Titanium 22 48
U Uranium 92 238
Zn Zinc 30 65

The periodic table is a list of all of the elements arranged in groups according to their atomic configuration. In this table the elements are colour-coded according to their chemical and physical properties.

periodic table

For an accessible version of the periodic table please see Syngenta Period Table of Elements (http://www.syngentaperiodictable.co.uk/periodic-table.php?keyStage=5)

3

About the Author

Karla Panchuk earned a BSc and MSc from the University of Saskatchewan, and a PhD in Earth-system modeling from the Pennsylvania State University. She teaches geology at the University of Saskatchewan and at St. Peter’s College.