The Story of Earth An Observational Guide 1627402930

Download as pdf or txt
Download as pdf or txt
You are on page 1of 336

The Story of Earth: An Observational Guide

THE STORY OF EARTH: AN OBSERVATIONAL GUIDE

DANIEL HAUPTVOGEL AND JINNY SISSON

CARLOS ANDRADE, JOSHUA FLORES, MELISSA HANSEN, ANA VIELMA, AND


HANNAH ANDERSON
The Story of Earth: An Observational Guide by Daniel Hauptvogel & Virginia Sisson is licensed under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International License, except where otherwise noted.
CONTENTS

Preface 1

About the Authors 1

About this Lab Manual 2

Accessibility 3

Copyright 3

Chapter 0: Geologic Skills 5

0.1 – Let’s Get Started 5

0.2 – Academic Integrity 6

0.3 Skill-building 6

0.4 Reading and Creating Geologic Maps 18

Additional Information 21
Chapter 1: Plate Tectonics 22

1.1 Introduction 22

1.2 Plate Tectonics, Earthquakes, and Volcanoes 26

Additional Information 43
Chapter 2: Earth Materials 44

2.1 Introduction 44

2.2 Minerals 44

2.3 Rocks 54

2.4 Igneous Rocks 56

2.5 Sedimentary Rocks 59

2.6 Metamorphic Rocks 65

2.7 Using Rocks to Interpret Earth’s History 74

Additional Information 84
Chapter 3: Geologic Time 86

3.1 Introduction 86

3.2 Geologic Time 86

3.3 Relative Dating and Correlation 89

3.4 Radiometric Dating 108

3.5 Other Uses for Geologic Dating 110

Additional Information 116


Chapter 4: Sedimentary Structures 117

4.1 Introduction 117

4.2 Types of Sedimentary Structures 119

4.3 Sedimentary Structures and Paleoenvironments 135

4.4 Trace Fossils and Bioturbation 139

Additional Information 145


Chapter 5: Stratigraphy 146

5.1 Introduction 146

5.2 Grain Size 150

5.3 Facies and Lithostratigraphy 154

5.4 Physical Properties 168

5.5 Seismic and Sequence Stratigraphy 172

Additional Information 187


Chapter 6: Fossil Preservation 189

6.1 Introduction 189

6.2 Types of Preservation 190

6.3 Handling of Fossils 196

Additional Information 207


Chapter 7: Fossils 209

7.1 Introduction 209

7.2 Taxonomy 212

7.3 Symmetry 214

7.4 Phylum Cnidaria 216

7.5 Phylum Bryozoa 226

7.6 Phylum Mollusca 229

7.7 Phylum Brachiopoda 236

7.8 Phylum Arthropoda 238

7.9 Phylum Echinodermata 240

7.10 Extinctions 248

Additional Information 250


Chapter 8: Paleoenvironments 252

8.1 Introduction 252

8.2 Marine Paleoenvironments 253

8.3 Continental Paleoenvironments 255

8.4 Fossils and Paleoenvironments in Movies 259

Additional Information 261


Chapter 9: Geologic Structures and Mapping 262

9.1 Introduction 262

9.2 Geologic Structures 268

9.3 Geologic Maps 272

Additional Information 285


Chapter 10: Interpreting Earth's History using Maps 286

10.1 Introduction 286

10.2 How to Locate Fossils 286

10.3 Geologic Provinces 293

Additional Information 303


Chapter 11: Paleoclimate 305

11.1 Introduction 305

11.2 Causes of Natural Climate Change 308

11.3 Oxygen Isotopes as a Climate Proxy 309

11.4 Tree Rings as a Climate Proxy 313

11.5 Fossils as a Climate Proxy 315

11.6 Sediment as a Climate Proxy 322

Additional Information 327


PREFACE

About the Authors


Drs. Daniel Hauptvogel and Virginia Sisson are instructional faculty in the Department of Earth and
Atmospheric Sciences at the University of Houston. We are a team that shares many responsibilities,
including the oversight and training of our department’s ~64 graduate teaching assistants, directors of
our Geoscience Learning Center where our teaching assistants hold their office hours and students come
in for tutoring, coordinators for about 15 sections of physical geology lab per semester, and a few sec-
tions of historical geology lab per year.

I’m an Instructional Associate Professor and have been at UH since 2015. I


currently teach GEOL 1303 Physical Geology and GEOL 1370 Natural Disas-
ters. I was born and raised in New Jersey, where I attended Montclair State
University for my B.S. and M.S. in geology and went to The Graduate
School and University Center of the City University of New York for my
Ph.D. I didn’t start as a geology major though, in fact, I didn’t declare geol-
ogy as my major until my junior year. I sort of floundered around trying to
figure out what I wanted to do, and my grades showed it. However, once I
joined the geology program, I felt at home, and it has been smooth sailing
ever since.

My graduate training in geology was in paleoclimate with a focus on sedi-


ment geochemistry, particularly in Antarctica. I began research on Antarc-
Dr. Daniel Hauptvogel
tica as an undergraduate after my mineralogy professor suggested I get
involved in undergraduate research. I continued to study Antarctica with
different projects throughout my Master’s and Ph.D. Today, I find myself
more interested in geoscience education and student success.

I credit my initial interest in geology to my grandfather, who lived in the Rocky Mountains of northwestern
Montana near Glacier National Park. During my visits, we went to various mountains, lakes, parks, etc.
The most memorable of which was Hidden Lake in Glacier National Park. I was too young to appreciate
the views and the geologic significance of the area, but now I’m longing to return there.

I’m an Instructional Professor and have been at UH since 2008. My teach-


ing includes GEOL 1303 Physical Geology, and I am the co-director of our
summer field geology camp at the Yellowstone Bighorn Research Associa-
tion camp near Red Lodge, Montana. In 1986, I moved to Houston after
finishing my Ph.D. at Princeton University and an A.B. at Bryn Mawr Col-
lege. At first, I did research and teaching at Rice University before moving
across town to UH.

I love to do geology field research and have spent many summers in


Alaska, British Columbia, Guatemala, and Venezuela. You might see some
of these places in the lab exercises as well as other places that I’ve done
field research such as Appalachians (Maine, New Jersey, and Pennsylvania)
and Rocky Mountains (Montana, Wyoming, and Colorado), Grand Canyon,
Dr. Virginia Sisson
Myanmar, Malaysia, Norway, India, Costa Rica, Cascades (Washington), Cal-
ifornia, Utah, and Nevada. In fact, I started majoring in geology when I saw
all the places that my physical geology teachers had traveled.
2 | PREFACE

By trade, I’m a metamorphic petrologist (and sometimes igneous rocks as well) who uses tools such
as geochronology, geochemistry, fluid inclusions, cathodoluminescence, remote sensing, UAVs (drones),
structural geology, and tectonics. Some of my research has touched on global geochemical cycles and
how exhumation may influence local climate. Most recently, I’ve focused on the geology of jade and asso-
ciated rocks. This led to my Andy Warhol 5 minutes of fame when our rediscovery of Olmec blue jade
made the front page of the NY Times.

Other Contributors
Many people contributed to the creation of this lab manual in the form of exercises, images, and editing,
including several graduate students from the University of Houston.

Hannah Anderson – UH M.S. student


Carlos Andrade – UH Ph.D. student (Graduated 2021)
Joshua Flores – UH Ph.D. student
Melissa Hansen – Adjunct faculty at Montclair State University and Kean University
Rosalie Maddox – UH Professor
Carolina Ramon – UH Ph.D. student (Graduated 2020)
Ana Vielma – UH Ph.D. student

About this Lab Manual


The idea for this lab manual came from the lack of available texts and resources that we felt suited the
needs of our teaching style and class. While there are plenty of options available, we felt that none were
truly worth the cost to our students. Hence, we envisioned a completely customized lab manual that
speaks to the nature of our teaching philosophy, giving our students a quality education and encourag-
ing them to observe and think about the world around them. Our Earth has a complicated evolution that
needs stories and observations for students to understand.

Through our own teaching experiences, we have found that students have difficulty making observations
and interpreting them. Perhaps this results from the abundance of standardized tests students must take
throughout their primary school career, but nonetheless, it is a problem that we have observed. Perhaps
Carl Sagan said it best:

“My experience is, you go talk to kindergarten kids, or first-grade kids, you find a class full of science enthu-
siasts. And they ask deep questions! What is a dream? Why do we have toes? Why is the moon round? What
is the birthday of the world? Why is grass green? These are profound, important questions. They just bub-
ble right out of them! You go and talk to twelfth-grade students and there’s none of that. They’ve become
leaden and incurious; something terrible has happened between kindergarten and twelfth grade and it’s not
just puberty.” – Carl Sagan in an interview with Mary Hynes on Canada’s TVO, 1995.

Our goal in creating the material for this lab manual was to focus heavily on students making observa-
tions of geologic data, whether rocks, minerals, fossils, maps, graphs, and other things. We want students
to look at things and wonder why, how, and when. The exercises and examples used in this book are
scattered throughout the world. We wanted to make sure that one region of the world was not the sole
focus of this work.

As you read the chapters, you’ll notice the tone of the text is very casual. This was important to us because
many geology texts are dense and, at times, complicated to read because of the vocabulary. Complicated
text is more than likely not going to be read by students. Therefore, we have tried to break down con-
cepts and terms into simple explanations and discussions to better engage readers.

We were fortunate enough to have our University Library sponsor our participation in the Rebus Text-
PREFACE | 3

book Success Program, a year-long program on creating open-access materials. The experience is a must
for anyone wanting to create quality open-access materials.

If you are a faculty member or teaching assistant using this lab book for the first time, please contact us
for our guide to teaching and possible answers. Many of the answers are open-ended, with several pos-
sible interpretations of the same data. This is part of our philosophy that data is open to interpretation,
and there might not be one correct answer. For students reading this, just try your best and discuss why
you believe your interpretation may be right with your professor or TA.

Accessibility
We have designed this lab manual with accessibility in mind using W3C standards. All text has proper
headings, text and background coloring meet the highest standards with a contrast ratio of at least 7:1,
all images have descriptive alt text, and all hyperlinks open in the same window (you’ll need to right-click
links to open them in a new tab or window). If you find any accessibility errors or have suggestions to
improve accessibility, please contact Daniel Hauptvogel at [email protected].

Copyright
This manual carries a CC BY-NC-SA license, which means you are free to use and adapt this material as
needed provided you give appropriate credit, share it under the same copyright terms, and use it for non-
commercial purposes. Most images used in this lab manual are open-access and have some variation of
a CC license indicated in their captions. Some images have been used with permission from the original
creator, also stated in the captions. For any images and maps that we have created, we can provide the
original Adobe Illustrator files and GMT 5.4.5 coding (for most maps) upon request. Please contact Daniel
Hauptvogel at [email protected]. It is a violation of our creative commons license to upload any of
this material to tutoring sites such as Chegg or Quizlet. Please respect our efforts to create open access
material and do not upload material or answers.

Cover Image Credits


Grand Canyon – National Park Service – Public Domain
Siccar Point – Dave Souza – CC BY-SA
Zion National Park – Dr. Igor Smolyar, NOAA/NESDIS/NODC – CC BY
Plant, Fish, Trilobites, and Ammonite Fossil Images – James St. John – CC BY
Partial Geologic Map – USGS – Public Domain
CHAPTER 0: GEOLOGIC SKILLS

Learning Objectives

The goals of this chapter are to:

• Get to know your classmates


• Summarize the importance of making observations
• Practice sketching for geologic interpretations
• Review the usage of Google Earth
• Create and interpret graphs
• Recognize geologic maps

0.1 – Let’s Get Started


“Chapter 0? I’ve never seen that before; what an odd way to start the book.”

Is it unconventional? Yes, but we like to think this is a more effective way of preparing you for the activities
in this lab manual because, let’s be honest, did you read the Preface? Probably not, and you would have
missed all of this handy information to help make your life easier if we had put this material in there.

This lab manual assumes that most users have already taken an introductory geology course, such as
Physical Geology or Earth Systems. Can you remember anything from your intro course? What about
some of the basic physical processes and vocabulary such as types of plate boundaries, rocks and the
rock cycle, and different classes of minerals such as silicates, carbonates, and oxides? Do you remember
much about the geologic time scale or some of the principles that geologists use to date events in Earth’s
history? If you think you need to review some of these, we recommend browsing through some of those
topics at the following open educational resources (OER):

• Physical Geology 2e – An OER textbook for introductory geology developed by faculty from Earth sci-
ence departments at universities and colleges across British Columbia and elsewhere.
• An Introduction to Geology – An OER textbook for introductory geology developed by a team at Salt
Lake Community College.
• Physical Geology Laboratory: Animations and Interactive Questions – An OER lab book by Elizabeth
A. Johnson that supplements labs and lets you practice skills you have learned.

In any case, the first few chapters are designed to review this material and do more advanced investiga-
tions for these topics. Some of you may be wondering about this, though: Can you complete the exercises
in this lab manual without having a previous geology course? If you’re the type of student who stays on
top of their academic pursuits or easily picks up on new material/concepts, then you most likely won’t
have an issue. If this doesn’t describe you, though, you may find you need to review the material in the
OER links above a little more closely.
6 | CHAPTER 0: GEOLOGIC SKILLS

0.2 – Academic Integrity


As Benjamin Franklin said, “Honesty is the best policy,” and we believe students should follow this adage.
However, in this digital age, cheating has become easier, and catching it has become harder, especially
with large class sizes. Believe it or not, academic integrity (AKA cheating and plagiarism) is a major issue
at most universities. The International Center for Academic Integrity conducted a survey of undergrad-
uate and graduate students between 2002 and 2015. They found that 68% of undergraduate students
and 43% of graduate students admitted to cheating on exams or other graded work at some point. That’s
a staggering yet eye-opening statistic. To put that in perspective, if your lab class has 25 students, that
means about 17 of your classmates have probably cheated at some point in their college career.

One of the common issues we have found with academic honesty cases is that students don’t know they
are cheating or plagiarizing. To help you understand this a bit more, we’ve put together a list of common
offenses:

• Plagiarism – presenting someone else’s work as your own


• Cheating – getting an unfair advantage in a test
• Misrepresentation of facts – distorting the truth or your data
• Encouraging or helping anyone else do any of these things

In this lab, you may benefit from discussing concepts and exercises with your fellow students, teaching
assistants, and professors, which brings us to another gray area: group work. Many of the exercises in
this lab manual are best completed in small groups where each member’s strengths and weaknesses will
shine. This is not a bad thing but rather an opportunity to experience peer-to-peer learning, an effec-
tive learning strategy. Working with others will help you solve the exercises in this lab manual, but the
answers you turn in must be written in your own words. Similarly, sketches and diagrams must also be
your own work, and any data collection, graphs, calculations, or measurements must also be your own.

Learning to write clearly is part of what we hope you learn by answering these questions. So, please be
an ethical student and follow these suggestions for academic integrity. These policies may vary depend-
ing on your university academic code and are general guidelines for how to succeed during these labs
as well as in life. If you are interested in understanding more about plagiarism, consider checking out
www.plagiarism.org and their section titled “What is Plagiarism?“. Posting images and answers from this
lab manual on websites like Chegg and Course Hero violates this work’s copyright, and you can be held
liable.

0.3 Skill-building
There are many skills you will need to be successful in this lab, including how to make observations,
sketching, knowledge of geography and Google Earth, how to compile data, plot it on a graph, and how
to interpret that graph. This information may be overwhelming or sound scary right now, but we have
designed the exercises in this lab manual to guide you through these skills.

Making Observations
Being able to make observations is a critical skill for most of the exercises in this manual. Through our
years of teaching, we’ve come to realize that students have a difficult time making observations. This
is probably a result of the trend in education towards standardized testing, but that’s a discussion for
another time.
CHAPTER 0: GEOLOGIC SKILLS | 7

Exercise 0.1 – Observing and Sketching: Part 1

Before reading any further, let’s see where you stand in terms of your observation skills. Using the space
below, create a sketch of the photo in Figure 0.1 and annotate any observations you can make. Note: there is
no right or wrong answer here; it is merely a means of seeing what you observe.

Sketch Area:

Open your eyes and take note of what is in front of you. It sounds simple, but how do you know what to
look at? That’s the challenge for students making geologic observations; they don’t know what to look for
at first. Take a look at Figure 0.1; what do you see? What’s the first thing you notice? Is it the clouds? Is it
the mountain in the background? Were you able to tell the mountain is a volcano (It’s Mount St. Helens)?
Did you note the landscape around the volcano is barren? What about visible features of an area of land,
its landforms? Or how about that the landscape appears very smooth for being located in a mountainous
region? What about the small valleys carved into the landscape?

Mount St. Helens had a major eruption in 1980 that caused the mountain’s northern slope to collapse,
creating a major landslide that wiped out all of the vegetation north of the mountain. The landslide was
followed by the deposition of volcanic ash and other material, which is why the landscape appears
smooth. Those valleys are being carved out by rainfall that forms small rivers; they easily erode the loose
volcanic material and create the valleys. The region has yet to recover from this disaster, but vegetation
is slowly making its way back in.
8 | CHAPTER 0: GEOLOGIC SKILLS

Figure 0.1 – Photograph of Mount St. Helens in Washington state taken in June 2018. Image credit: Daniel
Hauptvogel, CC BY-NC-SA.

Most people think taking pictures is the best way to record what they see; however, this is not always
true. The lighting may be wrong, or what needs to be observed is under a mineral coating, or there isn’t a
suitable location to take in the entire landscape. Or maybe the opposite; people may need to focus on an
area but can’t get close enough. To overcome these challenges, geologists use sketches to capture their
observations of rocks, fossils, and landscapes.

For a geologist, creating sketches has several distinct advantages over taking photos. When creating
sketches, you can remove unimportant details, use shading and colors to highlight different aspects, and
easily annotate your sketches on-demand in your field notebook. Figure 0.2 is a sketch of a Triceratops
jawbone; the sketch allows us to focus on the important aspects of the fossil, rather than being distracted
by the background or other unimportant details.
CHAPTER 0: GEOLOGIC SKILLS | 9

Figure 0.2 – Sketch of a Triceratops maxilla (jawbone) showing bite marks. This fossil is
from the Hell Creek Formation in central Montana. The crocodilian bite marks are
angular compared to rounded features. When you make sketches, you can consider
adding color or shading to show details. The paleontologist omitted other surrounding
details to focus on the bite marks. Image credit: Jason Poole, CC BY.

Geologists often find themselves doing fieldwork for weeks at a time in remote areas and don’t always
have access to reliable electricity to charge camera batteries or laptops. In those circumstances, sketches
are the preferred method of recording observations. Does that mean geologists never take photos? Of
course not; any geologist has a trove of photos from field locations they visit, but it’s the sketches that
truly help with their interpretations.

The sketch of the north wall of the Grand Canyon in Figure 0.3 was made by John Wesley Powell, who
led the first boat trip through this area in 1869. You can see that he identified three types of rocks and
labeled them A, B, and C. Powell was the first to identify the “Great Unconformity” between the rocks at
the bottom of section A and the tilted rocks in section B. When making sketches, it is important to include
a scale. In this sketch of the Grand Canyon, Powell included a riverboat at the bottom. If possible, use
perspective and shading. Finally, annotate your sketches with notes. Annotations can help with the inter-
pretation of parts of your sketch that are difficult to portray correctly.
10 | CHAPTER 0: GEOLOGIC SKILLS

Figure 0.3 – This is a generalized sketch of the Grand


Canyon, Arizona, that shows several groups of rocks.
Section A is a group of metamorphic rocks (Vishnu
schist) intruded by granite (Zoraster granite). Section
B is a group of tilted sediments of the Grand Canyon
supergroup. Section C is a sequence of Paleozoic
sedimentary units. Powell drew this sketch during a
time when photographs were not readily available. It
gives incredible detail with the horizontal, layered
rocks on top, layers of tilted rocks in the middle, and
crystalline rock at the bottom. There is also a
riverboat drawn to give a sense of scale.
Image credit: Modified from the USGS, Public
Domain.

With all of that said about sketches, you don’t need to be an artist to record what you think is important.
And the more practice you get making sketches, the more your sketching skills are going to improve. It is
better to sketch quickly instead of spending lots of time making them perfect. If you think some aspects
of your sketch are lacking, then make a note of what you are observing.

Exercise 0.2 – Observing and Sketching: Part 2

Search the internet for a geological feature that inspires you. Websites that you can browse for interesting
geology are:

• Pictures that will inspire you to become a geologist


CHAPTER 0: GEOLOGIC SKILLS | 11

• Smithsonian’s list of the ten most spectacular geologic sites


• Awesome geology pictures
• American Southwest geology
• National Geographic’s best geology pictures from ordinary people
• Weird geologic formations

Once you find your geologic muse, complete the following:

• Create a sketch of the feature, and be sure to include some type of scale. Scale is important for the
viewer to get a sense of what you are sketching. For instance, what is the size of the Triceratops max-
illa in Figure 0.2? It could be massive or tiny? You may think that it is massive using your knowledge of
dinosaurs that you’ve seen in museums or elsewhere. But, did you know the smallest dinosaur fossil is
~5 cm (2 inches)? So, add a scale even if there is not one in the picture you are sketching. You can give
your best estimate.
• Add some brief comments about any features you can identify. These can be simple comments such as
the rocks are red and white. Please note whether these are igneous, sedimentary, metamorphic, or a
mix of rock types.

Sketch Area:

Geography
Many students learn some basic geography before college, and almost all seem to forget place names
once they no longer need to know them for a test. Perhaps the best way to learn this information is
to travel and then have memories associated with these places. Oh, you’re not independently wealthy
with enough spare time to travel the world? Not to worry, we have embedded several maps and links
to Google Earth in this text to showcase some geology of the world and hope you will learn enough
to remember some geography after completing this lab manual. You’ll “travel” to the likes of Australia,
British Columbia, Eastern Europe, Chile, and Texas, and that’s just in Chapter 1! We do assume, though,
12 | CHAPTER 0: GEOLOGIC SKILLS

that you can remember the seven continents (Yes, Antarctica is a continent, a landmass larger than the
United States) and the five oceans.

We created many of the maps in this lab manual using Generic Mapping Tools. Most maps are Mercator
projections and therefore distort the sizes of continents and distances as latitude increases. This is why
the United States looks almost as wide as Africa, when, in reality, it is much smaller than Africa (Figure
0.4). In fact, the United States, China, and India can all fit within the area of Africa with room to spare. You
can compare true landmass sizes at The True Size.

Figure 0.4 – A comparison of country sizes between Mercator projection (light blue) and actual country sizes (dark
blue). Mercator projection maps distort the size of continents and distances as latitude increases. Image credit:
Neil R. Kaye, CC-BY

Do you remember anything about latitude and longitude? This is the geographic coordinate system to
locate any point on Earth; think of it as an address. Latitude is the position north or south of the equator
CHAPTER 0: GEOLOGIC SKILLS | 13

and goes from 0° to 90° north and south. Sometimes negative values are used to represent south. Lon-
gitude is the position east or west of the prime meridian, which runs through Greenwich, England, and
goes from 0° to 180°. Positive values represent east of the prime meridian, and negative values represent
west of the prime meridian.

Exercises 0.3 – True Sizes of Landmasses

Since we live on a three-dimensional globe that is often projected in two dimensions, it is difficult to appreciate
the scale of where you live compared to other countries or states on Earth. An easy way to do this is to use The
True Size, a computer visualization tool. Once you are on this website, type in the name of your home state
or country, which will highlight the area on the map. Then, drag it around the world, comparing it to other
countries or states.

a. Find a country or state that is a similar size to your home state or country. ____________________
b. How does the latitude affect the size of your home state or country?

c. What countries are the same size as the United States?

Google Earth (Not Google Maps)


There are many ways to find your location, such as a map app on your digital device. This lab book will use
the web version of Google Earth to show you all of the locations mentioned throughout this lab manual.
Google Earth is a composite of satellite images, aerial photographs, and GIS data on a globe. Coverage
includes about 98% of the Earth. The resolution of the images partly depends on how popular the area
is. For example, remote areas in British Columbia, Canada, have a poor resolution in mountainous areas
except where commercial logging is done. There are many features in the desktop version of Google
Earth, such as historical imagery, measurement tools, three-dimensional imagery, night sky, views of the
Moon and Mars, and bathymetry of the ocean floor. As the web version of Google Earth is updated, we
hope many of these tools will be incorporated in future releases. Please note that all links in this lab man-
ual open in the same browser window according to accessibility standards, you may want to open Google
Earth links in a separate window or tab.

Exercise 0.4 – Using Google Earth

Using Google Earth, find a geologic feature of your choosing (a mountain, fault, desert, specific location, any-
thing really) and answer the following questions about it. You can also use your feature from Exercise 0.2 if
you know its location.

a. Record the latitude and longitude of your feature.

i. Latitude: _______________
14 | CHAPTER 0: GEOLOGIC SKILLS

ii. Longitude: _______________

b. Since one degree of latitude is ~110km, how far from the equator is your geologic feature?
____________________________
c. How about miles? (1km = 0.62 miles) _____________________________
d. Using the ruler tool in Google Earth, how far from your house or dormitory is it to this geologic feature
in km and miles? (The ruler tool is the last icon on the left-hand menu.)
____________________________

Compiling and Plotting Data


In general, data is either quantitative (a measure of how much of something; for example, the tempera-
ture is 10°C) or qualitative (a description of something; for example, the temperature is cold). There are
many ways to display data, and the type of data you collect will determine the type of data plot you will
need. Simple data plots include pie charts, bar charts, timelines, histograms, and scatter plots. Not all
data charts are helpful to interpret trends. Often, we have to try different types of plots to discern what
is important about the data. A lack of a trend is also informative because it means that your assumptions
are incorrect and there is no relationship in your data set.

Let’s look at the scatter plot in Figure 0.5. This graph shows a relationship between the time between two
eruptions and how long (duration) an eruption lasts. This scatter plot shows two types of eruptions: short
eruptions with a short time between them and long eruptions with a long wait time. In this plot, there are
not many data eruption durations between 2.5 and 4.0 minutes. So, if we interpolate between these, we
can estimate the wait time between eruptions.

Generally, scientists look to see if the data is correlated, meaning they are looking for a relationship or
pattern. Suppose there is a positive correlation; as one variable increases, so does the other, just like
Old Faithful eruptions (Figure 0.5). In this plot, as the duration of the eruption increases (x-axis), so does
the time between eruptions (y-axis). Data can also be negatively or inversely correlated; as one variable
increases, the other decreases. For example, as an earthquake’s magnitude increases, its frequency (how
often it occurs) decreases.
CHAPTER 0: GEOLOGIC SKILLS | 15

Figure 0.5 – This scatter plot shows two variables: the duration of
steam eruptions from the Old Faithful geyser in minutes on the
horizontal axis and the time between these eruptions in minutes on
the vertical axis. The blue dotted line shows the positive correlation
because as one variable increases, so does the other. Image credit:
Wikipedia User: Nwstephens, Public Domain.

Most geoscientists use spreadsheet programs such as Microsoft Excel or Google Sheets to analyze their
numerical data. You may not yet be familiar with these programs, but there are many tutorials available
online. Check out this tutorial at Excel Easy or look at an Excel guide to help you get started. Excel is a
powerful program that simplifies many tasks. You will find that many companies and organizations uni-
versally use it. Plus, if you learn one program, you will be able to use other programs because many of
these software packages are very similar.

In geosciences, we often collect data that has three components, such as grain size in sedimentary rocks
and soil, to classify samples using the proportions of sand, silt, and clay-sized particles. These data are
displayed on triangular plots (sometimes known as ternary plots) (Figure 0.6). For most applications, the
three variables (a, b, c) add up to one hundred percent. Since a + b + c = 100 for all components, any one
variable is not independent of the other two. Only two variables are on the graph as c=100−a−b.
16 | CHAPTER 0: GEOLOGIC SKILLS

Figure 0.6 – Triangular plots for sand (red), silt (blue), and clay (green). This version of the plot is typically used to
classify soil types. The four plots are an example of how to plot a soil sample with 50% clay, 26% silt, and 24% sand.
a) blank triangular plot for the relative abundance of clay, silt, and sand; b) First, find the 50 percent line for clay
(green horizontal tie lines); c) Next, find the 24 percent line for silt (blue dotted tie lines); d) Check the result by
plotting the 26 percent line for sand (red dashed tie lines). Image credit: Virginia Sisson, CC BY-NC-SA.

Exercise 0.5 – Plotting Data

Collect some data from your classmates. This is up to you as a class to decide on at least two items. Examples
are: how far from school do they live, what is their major, height, favorite color, how many have dark hair ver-
sus blonde or red hair, how many steps do they take in an average day, how many letters in their names (first,
last, nickname, all three), what kind of pet do they have, what is their resting heart rate, etc. Record your data
in the space below.
CHAPTER 0: GEOLOGIC SKILLS | 17

There are many ways to display data. In general, data is either quantitative related to counts of something
or qualitative about information that can’t be measured. Make a graph of your data using one of the blank
graphs in Figure 0.7. If you have two quantitative (numerical) items, you should make a scatter plot. If you
have only quantitative data, you can create a pie chart, bar chart, or histogram.
18 | CHAPTER 0: GEOLOGIC SKILLS

Figure 0.7 – Blank graphs to be used in Exercise 0.4.

a. What does your graph(s) tell you about your classmates?

0.4 Reading and Creating Geologic Maps


The ability to read a geologic map will be necessary, especially for the second half of this lab manual.
A geologic map contains stories about the region that is covered. Maps contain information about what
CHAPTER 0: GEOLOGIC SKILLS | 19

is on Earth’s surface as well as below. You can use them to make a three-dimensional picture of your
surroundings. Another way to think about it is a geologic map is to a geology major as a wrench is to a
mechanic.

Figure 0.8 shows the first geologic map of the U.S. published by William Maclure in 1809. He subdivided
the rocks into five types based on the Werner classification. This classification, however, is no longer
accepted, and not many people know about this map. Despite this, Maclure made a heroic effort to pro-
duce this map. He supposedly crossed the mountains over fifty times to get the details correct, with walk-
ing and horseback as his only means of transportation.

Figure 0.8 – The first known geological map of the eastern and central parts of the United States by William
Maclure. The colors indicate where primitive (red), transitional (pink), secondary (light blue), alluvial (yellow), Old
Red sandstone (dark blue), and salt and gypsum (green) are located. Even though his understanding of geologic
formations was different from what is used today, the general distributions of different rock types were fairly
accurate. Image credit: David Rumsey, after William Maclure, CC BY-NC.

To an untrained eye, a modern geologic map (Figure 0.9) is a maze of colors in fantastical patterns. We
purposefully did not include a legend for this map because it would be too complicated for most geology
padawans. By the end of this book, though, you will have everything you need to be a geology master.
For now, the colors represent geologic units, which are subdivided by time and rock types. Many geologic
maps use a standard color scheme with colors related to the geologic time scale, such as shades of yel-
low for Quaternary units and blues for Paleozoic units. Still, sometimes the colors used are unique to the
kind of geologic map. The geologic map on this lab book’s cover is for the Grand Canyon region and is
20 | CHAPTER 0: GEOLOGIC SKILLS

mostly blue. So, you can easily guess that most of this area is made of Paleozoic rock units. The standard
color scheme used in the United States is available on the USGS website. Between these colored units
are lines or contacts. The width and type of line designate the type of contact, such as fault (solid line),
intrusive contact (dashed line), or contact that is covered by unconsolidated rocks (dotted line). Typically,
there is a legend that identifies the type of line with different types of contacts.

Figure 0.9 – Modern geologic map of the eastern United States. Note that this map covers about the same area as
William Maclure’s map in Figure 0.8. Image credit: Adapted from the USGS, Public Domain.

Do geologic maps intimidate you? Not to worry, we will break them down for you later on so you can walk
away from this class like a map-reading pro. Now, on to Chapter 1.
CHAPTER 0: GEOLOGIC SKILLS | 21

Additional Information

Exercise Contributions
Virginia Sisson, Daniel Hauptvogel, and Ana Vielma

Google Earth Locations


CHAPTER 1: PLATE TECTONICS

Learning Objectives

The goals of this chapter are to:

• Identify types of plate boundaries and compare their characteristic earthquake and volcanic activities
• Assess the basic lines of evidence supporting plate tectonics
• Explain how ancient plate boundaries affect modern topography

1.1 Introduction
Plate tectonics is the grand unifying theory in geology. It gets that title because many topics in geology can
be explained, in some way, by the movement of tectonic plates. Tectonic plates are composed of Earth’s
crust and the uppermost, rigid portion of the mantle. Together they are called the lithosphere. Earth’s
crust comes in 2 “flavors”: oceanic and continental (Table 1.1).

Table 1.1 – Comparison of oceanic and continental crust

Property Oceanic Crust Continental Crust

Thickness 7-10 km 25-80 km


3 3
Density 3.0g/ cm 2.7g/ cm

Silica (SiO2) Content 50% 60%

Composition Fe, Mg, and Ca silicates K, Na, and Al silicates

Color Dark Light

Lithospheric plates move around the globe in different directions and come in many different shapes
and sizes. Their movement rate is millimeters to a few centimeters per year, similar to the rate that your
fingernails grow. Motion between tectonic plates can be divergent, convergent, or transform. In diver-
gent boundaries, plates are moving away from each other; in convergent boundaries, plates are moving
toward each other; and in transform boundaries, plates are sliding past each other. The type of crust on
each plate determines the geologic behavior of the boundary (Figure 1.1).
CHAPTER 1: PLATE TECTONICS | 23

Figure 1.1 – These models show 6 main types of plate tectonic boundaries. Blue indicates ocean, green indicates
land, brown indicates the lithosphere, and orange is the asthenosphere. The bold arrows on the plates indicate
their relative motion. Also shown are gray volcanoes. Ocean-ocean transform boundaries (not shown) exist on a
small scale associated with spreading at mid-ocean ridges, and continent-ocean transform and divergent
boundaries are rare (former) or don’t exist (latter). Image credit: Adapted from Wikimedia Commons user
Domdomegg, CC BY.

The foundations of plate tectonics began with a German scientist named Alfred Wegener, who proposed
the idea of continental drift in 1915. Think about it, 1915. What kind of evidence could someone possibly
have to propose such a big idea? It turns out that Wegener had 4 pieces of evidence that he claimed
supported his idea: 1) The continents looked like they fit together like pieces of a puzzle; 2) There were
matching fossils on continents that were separated by oceans; 3) There were matching mountain ranges
on continents that were separated by oceans; 4) There was paleoclimate evidence that suggested that in
the past some continents were closer to the polar regions and some were close to the equator. Wegener
took his idea one step further and proposed that all of the continents were together in one giant super-
continent 200 million years ago called Pangea. Like many great ideas in science, Wegener’s idea of conti-
nental drift was not accepted by his peers, in part because he did not have a well-developed hypothesis
to explain what was causing the continents to drift. It wasn’t until the 1960’s that his idea was expanded
upon by scientists like Harry Hess.

Exercise 1.1 – Reconstructing Positions of Continents Using Wegener’s Evidence

When Alfred Wegener came up with his continental drift hypothesis in the early 1900s, he used several lines
of evidence to support his idea. He also proposed that 200 million years ago, all continents were together in a
single supercontinent called Pangea. In this exercise, you will use the fit of the continents and matching fossil
evidence to piece together Pangea. This exercise is adapted from “This Dynamic Planet” by the USGS.

a. Individually or as a group, piece together the supercontinent Pangea.


24 | CHAPTER 1: PLATE TECTONICS

i. Label the landmasses of each continent in Figure 1.2.


ii. Color the fossil areas to match the legend below.
iii. Cut out each of the continents along the edge of the continental shelf (the outermost dark line).
iv. Try to logically piece the continents together so that they form a giant supercontinent.
v. When you are satisfied with the fit of the continents, discuss the evidence with your classmates
and decide if the evidence is compelling or not. Explain your decision and reasoning on the evi-
dence.

b. Pangea began to break apart about 200 Ma resulting in the formation of the Atlantic Ocean. Using the
map in Figure 1.3, calculate the spreading rate of the Mid-Atlantic Ridge in mm/yr. (Hint: measure the
distance from the easternmost tip of South America to the inside curve of western Africa).
____________________

Table 1.2 – Symbol key to Figure 1.2

Symbol Description

The continents are surrounded by the continental shelf (stippled pattern), which extends beyond the
continent until there is a large change in slope.

By about 300 million years ago, a unique community of plants had evolved, known as the European ora.
Fossils of these plants are found in Europe and other areas. Color the areas with these fossils yellow.

Fossils of the fern Glossopteris have been found in these locations. Color the areas with these fossils
green.

Fossil remains of the half-meter-long fresh or brackish water (reptile) Mesosaurus. Mesosaurs flourished
in the early Mesozoic Era, about 240 million years ago. Mesosaurs had limbs for swimming but could also
walk on land. Other fossil evidence found in rocks along with Mesosaurs indicates that they lived in lakes
and coastal bays or estuaries. Color the areas with these fossils blue.

Fossil remains of Cynognathus, a land reptile approximately 3 meters long that lived during the Early
Mesozoic Era about 230 million years ago. It was a weak swimmer. Color the areas with these fossils
orange.

Fossil evidence of the Early Mesozoic, land-dwelling reptile Lystrosaurus. They reproduced by laying eggs
on land. Also, their anatomy suggests that these animals were probably very poor swimmers. Color the
areas with these fossils brown.
CHAPTER 1: PLATE TECTONICS | 25

Figure 1.2 – Continent cut-outs for Exercise 1.1. Image credit: From the USGS, Public Domain.
26 | CHAPTER 1: PLATE TECTONICS

Figure 1.3 – Blank map of the South Atlantic Ocean for Exercise 1.1. Image credit: Daniel Hauptvogel, CC BY-NC-SA.

1.2 Plate Tectonics, Earthquakes, and Volcanoes


Plate tectonic boundaries are often associated with earthquakes and volcanic activity. By looking at maps
for the distribution of earthquakes and volcanoes worldwide (Figures 1.4-1.5), you can interpret the
boundaries between the major tectonic plates. Generally, divergent plate boundaries are characterized
by shallow earthquakes and some volcanism. Convergent boundaries have a range of earthquake depths
from shallow to deep, and many have volcanoes as a result of subduction. Subduction occurs in conver-
gent boundaries where the denser, oceanic plate descends into the mantle beneath the overriding plate.
Convergent boundaries also tend to produce linear and curved mountain belts. Transform boundaries
typically have shallow earthquakes and no volcanoes.
CHAPTER 1: PLATE TECTONICS | 27

Figure 1.4 – This map shows the location of volcanoes that have been active within the past 10,000 years (red
triangles). Map scale reference is 30° latitude. Image credit: Daniel Hauptvogel, CC BY-NC-SA.
28 | CHAPTER 1: PLATE TECTONICS

Figure 1.5 – This map shows the locations of all earthquakes with a magnitude greater than 4.5 for the years 2015
and 2016. The colors indicate earthquake depth; red <35 km, green 35-100 km, and blue >100 km. Please note that
many red dots are overlaid by green dots on this map. Map scale reference is 30° latitude. Image credit: Daniel
Hauptvogel, CC BY-NC-SA.

Exercise 1.2 – Modern Examples of Plate Tectonic Boundaries

Each type of plate boundary has distinct earthquake and volcanic patterns. Using observational and critical
thinking skills, answer the following questions:

a. Observe the patterns amongst the earthquake and volcano location maps (Figures 1.4-1.5). Hypothesize
where you think the major plate boundaries exist and draw those boundaries on the blank map in Figure
1.6 using three different colors to identify the type of motion for each boundary (example: red for diver-
gent boundaries, blue for convergent boundaries, and green for transform boundaries).
b. Which type of boundary (divergent, convergent, or transform) is the most abundant?
______________________________________
c. On the same map where you drew in plate boundaries (Figure 1.6), identify locations where each type of
these boundaries are located:

i. Continent-Continent Convergence (CCC)


ii. Ocean-Ocean Convergence (OOC)
iii. Continent-Ocean Convergence (COC)
iv. Continent-Continent Divergence (CCD)
CHAPTER 1: PLATE TECTONICS | 29

v. Ocean-Ocean Divergence (OOD)


vi. Continent-Continent Transform (CCT)

d. What type of plate boundary is associated with most of the deep earthquakes?______________________
e. Describe the pattern in earthquake depth from the coast to inland at subduction zones.

f. Critical Thinking: The San Andreas Fault in California is a transform fault. Is there any evidence in the
earthquake and volcanic activity that suggests this fault did not always have transform motion? Explain.

Figure 1.6 – This is a blank map of the world to be used in Exercise 1.2. Map scale reference is 30° latitude. Image
credit: Daniel Hauptvogel, CC BY-NC-SA.

Earthquake locations can tell you more about an area than what type of plate boundary is there. For
example, in subduction zones, most earthquakes occur along the boundary between the subducting slab
and the overriding slab. The angle of subduction is not always constant and can vary from one bound-
ary to the next and can even vary along the same boundary. When a plate subducts at a low angle, it is
called flat-slab subduction. The effects of flat-slab subduction are many, including shallower earthquakes,
uplifting of mountains, and the location and activity of volcanoes.
30 | CHAPTER 1: PLATE TECTONICS

Exercise 1.3 – Interpreting Subduction Angle from Earthquake Data

The Western margin of South America is a tectonically active region where the Nazca Plate subducts under the
South American Plate (Figure 1.7), creating the Andes Mountains. Even though the entire coast is part of the
same subduction zone, the subduction process doesn’t look the same throughout. Tables 1.3 and 1.4 contain
earthquake data from two different locations of the subduction zone, one from central Chile and another near
the Chile-Peru border. The location data are the distance each earthquake was from the trench and how deep
within the Earth it was.

a. Using the graph paper provided by your instructor, plot the distance of the earthquake foci (hypocen-
ters) from the trench on the horizontal axis and the depth of the earthquakes on the vertical axis; the
recommended scale is 1 cm = 10 km. Connect the plotted points to create an approximate cross-section
of the subduction zone at the two locations.
b. Look at the graph you made, which region has a steeper subduction angle, the Chile-Peru border or cen-
tral Chile? ____________________
c. Partial melting of the asthenosphere above the subducting slab occurs at a certain depth and leads to
volcanism. In which location do you think volcanoes are closer to the coastline? Why?
CHAPTER 1: PLATE TECTONICS | 31

Figure 1.7 – Plate tectonic map of South America and adjacent plates. Boundaries marked with triangles represent
convergent zones. Two arrows in opposing directions indicate divergent boundaries. Transform boundaries have arrows
showing motion to both right and left. Red and white colors represent high elevations, greens and yellow represent lower
elevations, and purple and blue represent areas below sea level. Image credit: Tectonic boundary map drawn using
GPlates software.
32 | CHAPTER 1: PLATE TECTONICS

Table 1.3 – Earthquake location data from


the Chile-Peru Border (data from
Martinod et al., 2010)

Distance from trench (km) Depth (km)

160 10

200 30

220 50

300 65

370 125

500 190

300 100

250 65

210 40

280 80

450 175

400 140

410 150

Table 1.4 – Earthquake location data from


central Chile (data from Martinod et al.,
2010)

Distance from trench (km) Depth (km)

100 10

170 40

220 65

400 90

200 50

120 20

500 110

350 85

300 75

250 60

280 75

200 55

260 90

1.3 Plate Tectonics and Topography


Geologists can observe most of the processes occurring at plate tectonic boundaries today (earthquakes,
volcanoes, mountain building, etc.). However, understanding the plate tectonic activity of the geologic
past is more difficult because the events have already happened. Hence, geologists use processes that
occur in the present to interpret processes that occurred in the past. This is known as uniformitarian-
CHAPTER 1: PLATE TECTONICS | 33

ism. One way geologists can interpret ancient plate tectonic activity is to look at the topography of an
area. Topography is the study of shapes and features of the land surface. When studying features on the
seafloor, the topography is instead referred to as bathymetry because this data is referencing how deep
a feature is. There are numerous ways of looking at the topography of Earth’s surface, including satellite
imagery, topographic maps, shaded relief maps, and digital elevation models.

Exercise 1.4 – Interpreting Plate Boundaries from Topographic Profiles

Below are five topographic profiles showing different plate boundary configurations. A topographic profile is
a graph that shows elevation changes as you walk from one point on the Earth to another. These are all made
with vertical exaggeration (length/elevation) of 50:1. This overemphasizes the changes in topography. In all of
these profiles, the 0 value on the vertical axis is at sea level.

a. For the topographic profiles in Table 1.4, identify what types of plate boundaries are shown using the
names from Figure 1.1. Pay close attention to the y-axis versus x-axis.
b. On each profile, draw the location of the boundary between the two plates. You can show this as a single
line.
c. For each profile, label features such as oceanic and/or continental crust, mid-ocean ridges, volcanoes,
mountain belts, and trenches.
d. Indicate which direction each tectonic plate is moving (you can use arrows for this).
34 | CHAPTER 1: PLATE TECTONICS

Table 1.5 – Unknown topographic profiles and answer area for Exercise 1.4

Profile Type of Boundary

Geologists can use topography to get a broad sense of the tectonic history of an area. Generally speaking,
plate tectonic activity tends to produce elevation changes at or near the plate boundary, especially in
convergent settings. The collision of two plates leads to suturing; the two plates become one when the
collision ends. Evidence of these ancient boundaries is most commonly in the form of linear mountain
belts that are not currently near a plate tectonic boundary. For example, an eroded, linear mountain belt
CHAPTER 1: PLATE TECTONICS | 35

in the middle of a continent would indicate that area was part of a convergent boundary deep in the
geologic past and likely a continent-continent collision. The Ural Mountains in Russia fit this description
(Figure 1.8). They formed during an orogeny 240 to 300 million years ago and now serve as the boundary
between Europe and Asia.

Figure 1.8 – Shaded relief map of the Ural Mountains in Russia. Red and white colors represent high elevations,
greens and yellow represent lower elevation, and purple and blue represent areas below sea level. The Ural
Mountains are a narrow, linear chain of mountains that trend north-south through Russia. Map scale reference is
60° latitude. Image credit: Daniel Hauptvogel, CC BY-NC-SA.

Exercise 1.5 – Interpreting Ancient Plate Tectonic Boundaries

a. Over time mountain ranges are weathered and eroded, and the topography slowly goes back to base
level. Which do you think is more likely to be older, a mountain belt with higher elevations or a mountain
belt with lower elevations? Explain your reasoning.
36 | CHAPTER 1: PLATE TECTONICS

b. Look at the topographic map for part of North America (Figure 1.9). Mark two areas that you think have
been through major convergence of tectonic plates.
c. Which of the two areas do you think is older? What evidence from the map supports your hypothesis?

d. Look closely at the western portion of the North American continent. You should be able to notice differ-
ences in the patterns that make up the mountains. Each pattern represents a different geologic region.
Draw boundaries on the map that separate these different geologic regions and then describe the pat-
terns you observed to distinguish them (Hint: There are at least three).

e. Tectonic plate activity is often associated with mountain building. Based on Australia’s topography (Fig-
ure 1.10), explain whether or not you think the region is tectonically active today?

f. On the topographic map of Australia (Figure 1.10), mark an area that you think was a plate boundary in
the geologic past but is no longer active today. Explain your reasoning for marking this area.

g. Critical Thinking: Both of these maps contain areas on continents that are below sea level. Come up with
a hypothesis that explains how this can happen.
CHAPTER 1: PLATE TECTONICS | 37

Figure 1.9 – Shaded relief map of the United States. Red and white colors represent high elevations, greens and yellow
represent lower elevation, and purple and blue represent areas below sea level. Map scale reference is 40° latitude.
Image credit: Daniel Hauptvogel, CC BY-NC-SA.
38 | CHAPTER 1: PLATE TECTONICS

Figure 1.10 – Shaded relief map of Australia. Red and white colors represent high elevations, greens and yellow represent
lower elevations, and purple and blue represent areas below sea level. Map scale reference is -20° latitude. Image
credit: Daniel Hauptvogel, CC BY-NC-SA.

When most people think about tectonic plate boundaries, they often visualize parallel, symmetric lines
separating the plates. This is not always the case in the real world as many plate boundaries are curved
or segmented; this is because Earth is a sphere. Think about this: if you had a ball and tried to wrap it with
a flat sheet of paper, would the paper wrap around it perfectly smooth? The answer is no; the paper will
want to fold in some places and tear in other places. The tectonic plates behave similarly to the paper. Of
course, other factors affect the shape of a boundary. Evidence of these plate boundaries is also contained
in the topography of continents because not all mountain belts are straight lines.

Exercise 1.6 – Ancient Plate Boundaries in Texas, Oklahoma, New Mexico, and Northeastern
Mexico

Below is a topographic map of Texas, Oklahoma, New Mexico, and northeastern Mexico (Figure 1.11). This
CHAPTER 1: PLATE TECTONICS | 39

area is not near an active plate tectonic boundary today; the closest boundary is in the Gulf of Mexico. How-
ever, there is evidence in this topography to indicate it was part of a plate tectonic boundary at least twice in
the geologic past.

a. Based on the topography, mark two areas that have been part of a plate tectonic boundary in the geo-
logic past. Topographic changes do not need to be symmetric as some tectonic processes are not sym-
metric.
b. One of these boundaries is older than the other. Label the old and young boundaries.
c. One of these boundaries has highs and lows within the belt. What tectonic process creates low topogra-
phy?

d. One of these boundaries splits into two branches. What is the angle between these branches?
40 | CHAPTER 1: PLATE TECTONICS

Figure 1.11 – Shaded relief map of Texas. Red and white colors represent high elevations, greens and yellow represent
lower elevations, and purple and blue represent areas below sea level. Map scale reference is 30° latitude. Image
credit: Daniel Hauptvogel, CC BY-NC-SA.

As tectonic plates move, they ride over stationary “hotspots” of heat from the mantle. Hotspots are still
a poorly understood geologic phenomenon, but they allow extremely hot mantle material to rise close
to the surface. Hotspots are marked by volcanoes, which come from melting of the mantle and crust
directly above a hotspot. If they occur under oceanic crust, they produce basalts. On the other hand, if
they are under continental crust, they form both basalts and rhyolites, often called bimodal volcanism.
Under North America, there are two hotspots: The Yellowstone hotspot, which is currently under Yel-
lowstone National Park in Wyoming and Montana, and the Anahim hotspot in central British Columbia,
Canada. As the North American Plate moves over these hotspots, calderas form from volcanic activity;
one of the largest volcanic eruptions ever occurred when the Gray’s Landing volcanics erupted 8.72 mil-
CHAPTER 1: PLATE TECTONICS | 41

lion years ago above the Yellowstone hotspot. One of the controversies is whether the hot spot is still
capable of supereruptions or whether the volume of eruptive material is waning.

Figure 1.12. – A topographic map for the northwestern United States and southwestern Canada. Superimposed is
the distribution of volcanic activity (black areas) for both the Anahim (northern chain of volcanoes and plutons) and
Yellowstone (southern chain of volcanoes). These chains of volcanoes are called hotspot tracks. Next to each
volcanic area is the age of the initial volcanism in millions of years. Some of these locations have more than one
caldera; these overlap in space and time. Red and white colors represent high elevations, greens and yellow
represent lower elevation, and purple and blue represent areas below sea level. Map scale reference is 45° latitude.
Image credit: Virginia Sisson and Daniel Hauptvogel, CC BY-NC-SA. Location of volcanic centers adapted from
Wikimedia Commons Sémhur CC BY-SA for the Anahim hotspot and Kelvin Case CC BY for the Yellowstone hotspot.

Exercise 1.7 – Tracking North American Hotspots

Use Figure 1.12 to answer the following questions about North American hotspots.

a. Use a protractor to measure the direction of the plate motion for each hotspot track. This is called the
42 | CHAPTER 1: PLATE TECTONICS

azimuth and is usually measured clockwise from the north.

i. Angle for Yellowstone: ____________________


ii. Angle for Anahim: ____________________

b. How fast is the North American plate moving over these hotspots? Measure the length of the hotspot
eruptions from the youngest to oldest. If you divide the length by the maximum age, this will give you
the rate of plate motion. Convert this to mm/year (km/my) as most plate motions are low-velocity rates.

i. Rate of plate motion for Yellowstone: ____________________


ii. Rate of plate motion for Anahim: ____________________
iii. Are these the same for the two hotspots? ____________________
iv. If not, why are the azimuth and the rates different if North America is a rigid plate tectonic block?

The answer may not be apparent as we don’t often move things around on a sphere. Instead, we
think of motion as a straight line from point a to point b. These hotspots are on the North Amer-
ican plate which means the plate rotates around a point in the middle of northern Quebec. Since
they rotate around a point on a sphere, different locations on the plate move at dissimilar speeds
and directions. Geoscientists call this an Euler pole.

c. Geoscientists have measured instantaneous global plate motions through various techniques such as
global positioning satellites (GPS) and very-long baseline interferometry (VLBI). This data is used to cal-
culate the rates of motion between two plates. There are several online plate motion calculators avail-
able; we will use the one created by UNAVCO. Use the latitude and longitude for the young end of each
hotspot and calculate the rate and direction of plate motion. Fill this out in Table 1.6. This website will
give you other information that is not relevant to this lab exercise.

Table 1.6 – Answer area for Exercise 1.7c

Hotspot Speed (mm/yr) Azimuth Direction

Yellowstone

Anahim

d. Critical Thinking: Are the UNAVCO results similar to your calculated results from b? If not, why might you
get different answers?

e. Has the rate of motion of the North American plate been constant along the hotspot track?

f. Which hotspot track has more volcanism? Yellowstone or Anahim? Be sure to compare the volcanism
over the same time period. Consider the size of the calderas when you answer this question.
CHAPTER 1: PLATE TECTONICS | 43

g. Critical Thinking: Can you explain why one of these has larger volcanic eruptions? You may be able to
use the topography to help you understand each area’s tectonic history to help you answer this ques-
tion.

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson, Carlos Andrade, Melissa Hansen

References
Knott, T.R., Branney, M.J., Reichow, M.K., Finn, D.R., Tapster, S., and Coe, R.S., 2020, Discovery of two new
super-eruptions from the Yellowstone hotspot track (USA) Is the Yellowstone hotspot waning? Geology,
v. 48, p. 934-938. doi.org/10.1130/G47384.1

Martinod, J., Husson, L., Roperch, P., Guillaume, B., and Espurt, N., 2010, Horizontal subduction zones,
convergence velocity and the building of the Andes. Earth and Planetary Science Letters, v. 299, pp.
299-309. DOI:10.1016/j.epsl.2010.09.010.

Google Earth Locations


CHAPTER 2: EARTH MATERIALS

Learning Objectives

The goals of this chapter are to:

• Identify common minerals based on physical properties


• Classify rocks based on texture and mineral composition
• Interpret environments in which igneous, sedimentary, and metamorphic rocks are formed
• Reconstruct the transport and depositional history of sedimentary rocks

2.1 Introduction
Minerals are the basic building blocks of rocks, which means rocks are made up of different combina-
tions of minerals or just one mineral in some cases. Figure 2.1a is an example of a rock called granite,
which is made up of a combination minerals. A mineral is a naturally occurring, usually inorganic, solid
that can be defined by a chemical formula and a crystal structure. Figures 2.1b and 2.1c are two other
types of rocks, sandstone and gneiss, that contain the same minerals as the granite in Figure 2.1a. Even
though they contain the same minerals, all three of these rocks formed in different ways. We’ll get to that
a little later, so for now, let’s focus on minerals.

Figure 2.1 – These are three different types of rocks that contain the same minerals: A) granite; B) arkose
sandstone; C) gneiss. Pink minerals are K-feldspar, clear minerals are quartz, white minerals are plagioclase
feldspar, and black minerals are biotite. Image Credit: James St. John, A, B, and C.

2.2 Minerals
According to the International Mineralogical Association, there are 5,575 known minerals, but most rocks
are composed of just a few common minerals. In fact, only seven minerals make up 84% of Earth’s crust
(Figure 2.2). So when you find an unknown rock, there is a good chance that at least one of those seven
minerals will be present.
CHAPTER 2: EARTH MATERIALS | 45

Figure 2.2 – This pie chart shows the distribution of minerals in continental crust. Silicate
minerals, structures based around the silicon-oxygen tetrahedron, are the most
abundant, making up 92% of all minerals in Earth’s crust. Image Credit: Daniel
Hauptvogel, CC BY-NC-SA.

Geologists can take rocks and minerals back to their laboratories and run in-depth chemical analyses to
determine what minerals they have collected. You don’t have that luxury when you’re working in the field
or the classroom, though. Instead, many minerals can be identified based on a few observable properties
that don’t require high-tech, expensive equipment. So, distinguishing these seven common minerals and
others is easier than you think. The common, observable properties of minerals are:

• Color
• Luster
• Streak
• Hardness
• Cleavage/Fracture

Color
The color of minerals comes from small impurities in the chemical composition. Sometimes elements
substitute for one another in the atomic structure of minerals, leading to significant changes in color. It
is perhaps the most straightforward observable property, but it is rarely a diagnostic property for most
minerals because the same mineral can come in various colors. Quartz, for example, can come in just
about any color (Figure 2.3).
46 | CHAPTER 2: EARTH MATERIALS

Figure 2.3 – Quartz in six color varieties. Color may not be a diagnostic
property for mineral identification. Other quartz colors include green,
red, orange, and brown. Image credit: Karla Panchuck, CC BY-NC-SA.

Luster
The term luster describes how light interacts with the surface of a mineral. There are many different
terms used to describe the luster of a mineral, but for this lab manual, we are only concerned with metal-
lic and non-metallic lusters (does it look like a metal or does it not look like a metal, respectively; Figures
2.4a and 2.4b). Be careful, though, because shiny doesn’t always mean it’s metallic. The technical differ-
ence between metallic and non-metallic is that metallic minerals do not allow light to pass through the
atomic structure, and non-metallic minerals do allow some light to pass through.

Figure 2.4a – Pyrite, a mineral with a metallic luster.


Image credit: Wikimedia Commons user JJ Harrison,
CC BY-SA.
CHAPTER 2: EARTH MATERIALS | 47

Figure 2.4b – Quartz, a mineral with a


non-metallic luster and six-sided crystal
form. Image credit: Wikimedia
Commons user Didier Descouens, CC
BY-SA.

Streak
Streak is the color of a mineral in powdered form, which is observed by scratching the mineral across a
ceramic streak plate (Figure 2.5). The powder color can be different from the mineral color because, when
in powdered form, the mineral’s impurities do not significantly affect the absorption of light. The way light
is absorbed and reflected is what gives things color. Streak is only a diagnostic property for metallic min-
erals because most other minerals have the same streak color as the hand-sample color or the streak is
white.

Figure 2.5 – Four different dark gray metallic


minerals with varying streak colors. The minerals
are from the upper left clockwise: hematite
(red-brown), magnetite (black), sphalerite
(honey-brown), and galena (dark gray). Image
credit: Karla Panchuck, CC BY-NC-SA.
48 | CHAPTER 2: EARTH MATERIALS

Hardness
The hardness of a mineral is a mineral’s ability to resist abrasion or scratching. Hardness is determined
using the Mohs’ scale of mineral hardness (Figure 2.6), developed by Frederich Mohs in 1822. Mohs’ scale
is qualitative, which means there is no quantitative relationship between hardness values; a 10 on the
Mohs scale is not 10 times harder than a 1. The scale is based on the relative relationship between 10
different minerals.

How do you determine mineral hardness? When identifying mineral hardness in the classroom, you can
use common objects with a hardness value defined on the Mohs scale. These common objects and their
hardness are located on the right side of Figure 2.6.

Figure 2.6 – This is the Mohs scale of relative mineral hardness that uses 10 minerals with no quantitative
relationship between the values. Lower numbers mean softer minerals; higher numbers mean harder minerals. On
the right side is a list of common objects and their Mohs hardness used to test unknown minerals. Note:
manufacturing differences can make the hardness of these common objects vary slightly. Image credit: Virginia
Sisson, CC BY-NC-SA.

Cleavage/Fracture
Cleavage is a term used to describe minerals that break in a predictable way. When minerals break along
planes in which chemical bonds are weak, they produce flat surfaces called cleavage planes. Minerals
can have 1, 2, 3, or more planes of cleavage (Figure 2.7). When there is more than one plane, the angle
between the planes can also be described. Cleavage is one of the most challenging concepts for introduc-
tory geology students, but it is a useful property to identify minerals. For example, amphibole and pyrox-
ene have the same physical properties except for cleavage. Knowing whether a rock contains amphibole
CHAPTER 2: EARTH MATERIALS | 49

or pyroxene could reveal a lot about that rock’s geologic history, such as where and how it most likely
formed.

While the concept of mineral cleavage is difficult, identifying cleavage in minerals doesn’t have to be.
When looking at a mineral, the easiest way to identify cleavage is to shine a light on its surface. If that sur-
face brilliantly reflects light, then you most likely have a cleavage plane. Today, just about everyone has
a cell phone with a flashlight that can be used to identify cleavage. In the absence of that, you can move
the mineral around underneath an overhead light or use sunlight.

Minerals don’t always break in predictable ways, and that’s when we use the term fracture. Fracture
describes irregular breakage in minerals, which usually occurs on mineral planes that don’t have any
weak bonds. This sounds easy enough; flat surfaces mean cleavage, and uneven surfaces mean fracture,
but the truth is that it can be messy. Minerals can appear to have flat surfaces that are not actually cleav-
age, and fracturing can occur along planes of cleavage. This is why you need to look closely at the surface
of a mineral using a hand lens.

A hand lens is a magnifying glass to help geologists look closely at rocks and minerals. To use the hand
lens properly, you need to place the lens up against your face and then move the mineral or rock sample
back and forth to get it into focus. A common mistake students make is holding the mineral or rock still
and moving the hand lens; this is incorrect and will not help you see small-scale features in rocks and
minerals.

Another thing that makes identifying cleavage difficult is the entire mineral sample does not have to take
on the shape of the cleavages. The minerals in Figure 2.7 are examples where the samples do take on
the cleavage shape, but this is not the case for most minerals. Instead, you need to use a hand lens to
look at the small details of a single side of the mineral. There will be slight changes in the heights of the
surface that you wouldn’t be able to see without the hand lens. It is along those “elevation” changes that
you should be looking for signs of cleavage. Then repeat your examination on the other sides.

Figure 2.7 – Types of cleavage with mineral examples. Cleavage differs by the number and angle between cleavage
planes. Image credit: cleavage models adapted from M.C. Rygel; muscovite from Daniel Hauptvogel, all other
minerals from James St. John, CC BY-SA.

Exercise 2.1 – Identifying Common Minerals

Your instructor will provide you with a set of unknown minerals. Identify each mineral based on its physical
properties using either Table 2.1 as your guide or the mineral flow charts in Figures 2.8, 2.9, and 2.10. Fill in
50 | CHAPTER 2: EARTH MATERIALS

this information in Table 2.2. There are also many online guides to help you identify minerals, such as Miner-
alID, Smart Geology (Android), or the app Mineral Identifier (Apple).

Table 2.1 – Physical properties of common minerals

Hardness Color Cleavage Other Mineral Name

1-2 Colorless One plane “Chewy” Clays

Can be peeled into


2-2.5 Colorless, light tan, yellow One plane Muscovite
transparent, flexible sheets

Can be peeled into thin,


2-2.5 Brown, black, green One plane Biotite
flexible sheets

2.5 Colorless, white Three directions at 90° Cubic crystals, salty taste Halite

Three directions not at Rhombic crystals, reacts


3 Colorless, white Calcite
90° with HCl, double refraction

Two directions at ~60° Hornblende


5-6 Dark green to black Elongated crystals
and ~120° (Amphibole)

5-6 Dark green to black Two directions at ~90° Elongated crystals Augite (Pyroxene)

Potassium
6 Colorless, pink, gray Two directions at 90° Stubby, prismatic crystals
Feldspar

Can have small grooves on


Colorless, white, gray, Plagioclase
6 Two directions at 90° one of the cleavages called
black Feldspar
striations

No cleavage, can have Stubby crystals or granular


6.5-7 Green Olivine
conchoidal fracture masses

May occur in 12-sided,


7 Mainly red and black Conchoidal fracture Garnet
circular crystals

May occur in six-sided,


7 Varied Conchoidal fracture Quartz
elongated crystals
CHAPTER 2: EARTH MATERIALS | 51

Figure 2.8 – Flowchart for light-colored minerals with nonmetallic luster. Image credit: Virginia Sisson, CC BY-NC-SA.
52 | CHAPTER 2: EARTH MATERIALS

Figure 2.9 – Flowchart for dark-colored minerals with non-metallic luster. Image credit: Virginia Sisson, CC BY-NC-SA.
CHAPTER 2: EARTH MATERIALS | 53

Figure 2.10 – Flowchart for minerals with metallic luster. Image credit: Virginia Sisson, CC BY-NC-SA.
54 | CHAPTER 2: EARTH MATERIALS

Table 2.2 – Worksheet for Exercise 2.1

Sample
Color Hardness Cleavage Mineral Name
ID

2.3 Rocks
Rocks are what most people think of when they hear “geology.” Rocks are made up of different minerals,
and in some cases, just one mineral. Minerals in rocks can be difficult to identify, though. Even upper-level
geology majors and practicing geoscientists can have trouble identifying minerals in rocks. Nonetheless,
it is an essential skill to practice because it serves as one of the primary ways to identify rocks. When geol-
ogists go into the field to study rocks, one of the first things they do is get up close and personal with the
rock, usually with a hand lens, and identify what minerals are present. Let’s practice with a simple exer-
cise on identifying some common minerals in rocks in Exercise 2.2.
CHAPTER 2: EARTH MATERIALS | 55

Exercise 2.2 – Identifying Minerals in Rocks

Your instructor will provide you with a selection of different rock samples. From the list of minerals in Table 2.3
below, indicate which rock sample(s) contain those minerals by filling out the “Rock Sample” column.

Table 2.3 – Worksheet for Exercise 2.2

Mineral Rock Sample

K-feldspar

Plagioclase feldspar

Quartz

Biotite

Hornblende

Based on how rocks are formed, geologists classify them into three basic types: igneous, sedimentary,
and metamorphic. Igneous rocks form when magma or lava cools and solidifies. Minerals start to crys-
tallize as the magma cools, and they interlock with each other in random orientations. Sedimentary rocks
form when broken pieces of other rocks from weathering processes get cemented together. They can
also form when minerals precipitate out of a solution, typically water, and are cemented together through
sedimentary processes. Metamorphic rocks are pre-existing rocks that are altered by heat and pressure.
Metamorphism often results in some type of pattern or orientation to the minerals, called foliation.

Exercise 2.3- Identifying Types of Rocks

Your instructor will provide you with a selection of unknown rocks. Using what you know about how each rock
type forms, determine the rock type for each of your unknown samples and fill in Table 2.4.
56 | CHAPTER 2: EARTH MATERIALS

Table 2.4 – Worksheet for Exercise 2.3

Rock Type Sample

Igneous

Sedimentary

Metamorphic

2.4 Igneous Rocks


The classification of igneous rocks is based on chemistry and texture, which means it’s based on what
minerals they have and how large the mineral grains are (Figures 2.11 and 2.12). The magma’s chemistry
and temperature determine which minerals will form, and the size of the mineral grains is determined
by how quickly the magma cools. Slow cooling rates (thousands to millions of years) make larger mineral
grains. Fast cooling rates (days to hundreds of years) produce smaller grains. Igneous rocks with large
grains are called intrusive or plutonic rocks, which means the magma cooled slowly within the Earth.
Igneous rocks with tiny grains are called extrusive or volcanic because the lava cooled on the surface
of the Earth. The distinction between small and large grains is whether you (or a well-trained geologist)
can identify the minerals without using any magnification. Using this information, if you had a coarse-
grained igneous rock that contained 10% quartz, 10% potassium feldspar, 50% plagioclase feldspar, 20%
pyroxene, and 10% amphibole, it would be called a diorite. If the mineral grains were fine (small), the rock
would be called andesite.

How would you identify minerals in an extrusive igneous rock if the grains are too small to see? There are
a couple of tricks that geologists use to identify which type of igneous rock they have. Do you remember
how color was not a diagnostic property for minerals? Well, color in an igneous rock can be diagnostic!
Iron and magnesium-bearing minerals are darker in color, like olivine and pyroxene, and they are the
primary minerals that makeup mafic and ultramafic igneous rocks. Potassium, aluminum, and silica-rich
minerals, like quartz and potassium feldspar, are lighter in color and make up intermediate and felsic
igneous rocks. Geologists can use the color of igneous rocks and the relative amount of light- vs. dark-col-
ored minerals to help identify them (Figure 2.12).
CHAPTER 2: EARTH MATERIALS | 57

Figure 2.11 – Classification chart of igneous rocks based on the mineral content. The top part is a graphical
comparison of silica content and mineral content for both extrusive and igneous rocks. To use this graph, you will
need to determine the relative abundance of seven common minerals found in igneous rocks. The bottom two
rows compare rocks with the same mineral composition but different names based on whether they formed
intrusively or extrusively. Image credit: Karla Panchuk after Steven Earle, CC BY-SA-NC.
58 | CHAPTER 2: EARTH MATERIALS

Figure 2.12 – Identifying igneous rocks using mineral color. a) The relative percentages of light (felsic) and
dark-colored (mafic – iron and magnesium-rich) minerals, b) Images you can use to estimate percentages of light-
and dark-colored minerals. Note: our eyes tend to overestimate the amount of dark colors, so be careful of this
when looking at the rocks. Also, for low abundances of accessory minerals, their size will influence your results.
Image credit: a) Karla Panchuk after Steven Earle, CC BY; b) Virginia Sisson, CC BY-NC-SA.

Exercise 2.4 – Identifying Igneous Rocks

Your instructor will provide you with a set of unknown igneous rocks to identify. Use Figures 2.11 and 2.12 to
fill out the information in Table 2.5.
CHAPTER 2: EARTH MATERIALS | 59

Table 2.5 – Worksheet for Exercise 2.4

% Light vs. Dark Minerals (for intrusive)


Sample Intrusive or Extrusive or Rock Name
Rock Color (for extrusive)

2.5 Sedimentary Rocks


Sedimentary rocks can form in several ways (Figure 2.13) and are classified as clastic, chemical, and
organic based on how they form. The most common way sedimentary rocks form is when other rocks
weather into small particles and are transported by wind, water, or ice to an area where they are
deposited and form layers. These are called clastic sedimentary rocks and are primarily classified based
on their grain size. Sedimentary rocks can also form through chemical or organic processes. If you have
a body of water containing dissolved salt and then that water evaporates, it will leave behind the salt as a
solid mineral form called a precipitate, which is a chemical sedimentary rock. Organic sedimentary rocks
form from the accumulation of organic debris. This organic material would need to be buried very quickly,
so it doesn’t decay. Organic material in sedimentary rocks is where coal, oil, and natural gas come from.
Then, there are sedimentary rocks that we can classify as biochemical. For example, foraminifera are tiny,
single-celled organisms that build shells out of calcium carbonate. When they die, their shells can sink to
the bottom of the ocean and accumulate to form a sedimentary rock. In this case, biological processes
caused the chemical precipitation of the mineral for the shell, but it doesn’t contain organic compounds,
so we call it bio-chemical.

In all cases, sediments need to be buried so the lithification process can take place. There are two steps
involved in turning sediment into sedimentary rock; compaction and cementation. When grains are
deposited, they typically have empty spaces between them called pores (porosity). Compaction reduces
porosity by bringing the grains closer together. The best analogy to think about is your garbage can at
home. When the garbage can is full, do you change the bag right away, or do you push down on it to
create more room? By pushing down on trash, you reduce the space between the pieces of garbage and
create more room at the top; this is compaction. Sediments work the same way, but what’s causing the
compaction is the accumulation of more sediment on top of the previously deposited material. The more
sediment that gets piled on top, the more compact the sediment beneath it becomes. Cementation is
a process that happens during compaction. As sediments become compacted, water is squeezed out of
the pore spaces. As the water leaves the sediment, it precipitates minerals into those pore spaces acting
as a glue that holds the sediment together. The common minerals that make up the cement are calcite,
quartz, and pyrite.
60 | CHAPTER 2: EARTH MATERIALS

Figure 2.13 – Sedimentary rock classification table. To use this table, you will need to determine whether the rock is
clastic or chemical/organic. Also, you need to determine both the grain size and mineral composition. For clastic
rocks, you also need to know the roundness of the grains. Also shown are the US Geological Survey standard
patterns for these rock types. These will be used in both stratigraphic columns and maps throughout this lab
manual.

The locations where sediments are deposited are called depositional environments. There are a variety
of depositional environments (Figure 2.14) that can be grouped into three broad categories: continental,
marine, and transitional. Continental environments are areas where sediments are deposited on con-
tinents, marine environments are areas where sediments are deposited in the ocean, and transitional
environments are coastal and tidal locations between continental and marine environments. Table 2.6
contains the most common environments, a brief description of them, and the common sedimentary
rocks that can form in those environments.
CHAPTER 2: EARTH MATERIALS | 61

Figure 2.14 – Common depositional environments for sediments ranging from continental to marine settings.
Image credit: Modified from PePeEfe after Mikenorton, CC BY-SA.
62 | CHAPTER 2: EARTH MATERIALS

Table 2.6 – Depositional environments for sedimentary rocks

Common Types of
Environment Description
Sedimentary Rocks

Continental

Sediment deposited by wind; primarily deserts and coastal regions;


Aeolian well-sorted sand; typically red in color; variable energy. Example in Sandstone
Google Earth: Algeria.

Fan-shaped deposits caused by moving water; usually found in arid or


semi-arid regions; contains gravel, sand, silt, and/or clay; poorly sorted; Conglomerate, breccia,
Alluvial
high energy; creates alluvial fans. Example in Google Earth: Death Valley sandstone, shale
National Park, California.

Sediment deposited by moving water, primarily rivers; can contain


gravel, sand, silt, and/or clay depending on how fast the water moves; Conglomerate,
Fluvial
variable energy; commonly red in color from oxidation. Example in sandstone, shale
Google Earth: Upper Mississippi River in Illinois/Missouri.

Lake settings that can contain sand, silt, or clay; generally low energy.
Lacustrine Sandstone, shale
Example in Google Earth: Lake Winnipesaukee, New Hampshire.

Sediment deposited by glaciers; variable grain sizes; poorly sorted. Conglomerate, breccia,
Glacial
Example in Google Earth: Southern Patagonia, Argentina. sandstone, shale

Forms where water evaporates and leaves behind mineral precipitates. Limestone, rock salt,
Evaporitic
Example in Google Earth: Utah. rock gypsum.

Transitional

Along coastlines, sediment transported by wave action; contains


Beach well-sorted gravel and sand; high energy. Example in Google Earth: Sandstone
Island Beach State Park, New Jersey.

Where a river empties into a body of water; contains gravel, sand, and
Deltaic Sandstone, shale
silt; low energy. Example in Google Earth: Yukon River, Alaska.

A shallow body of water separated from a larger body of water by


Limestone, shale, coal
Lagoonal barrier islands or reefs; very low energy; contains silt and clay. Example
(swamps)
in Google Earth: East Matagorda Bay, Texas.

Affected by the tides; mainly silt and clay; can create tidal flats; low
Tidal Shale
energy. Example in Google Earth: Bay of Fundy, Nova Scotia.

Marine

Located on the continental shelf; mainly sand and silt; energy decreases
Sandstone, shale,
Shallow with distance from shore. Example in Google Earth: Eastern Gulf of
limestone
Mexico.

The deep ocean; very low energy; mainly clay; can contain turbidite
Deep deposits of variable sediment sizes. Example in Google Earth: Pacific Shale, chert, limestone
Ocean.

A bar of rock, sand, or coral. Example in Google Earth: Great Barrier


Reef Limestone
Reef, Australia.

You’ll notice that the same type of sedimentary rock can be found in many different settings, making
it difficult to determine the depositional environment. Sometimes, the color of a sedimentary rock can
indicate the depositional environment because the color is largely determined by how much oxygen is
available as the sediment is buried and lithified (Figure 2.15). Reddish colors can indicate oxygen-rich con-
tinental environments, like a flowing river or desert. Green and light gray colors mean low oxygen, which
can be found in shallow marine environments. Black color means no oxygen, which would indicate a deep
marine or swamp environment. Color is not always a good indicator, though, because the sediment only
alters its color after burial. The conditions affecting buried sediment may not always match the environ-
ment on the surface above it.
CHAPTER 2: EARTH MATERIALS | 63

Figure 2.15 – Typical colors of sedimentary rocks based on their


depositional environment and oxygen availability during burial and
lithification. Image credit: Daniel Hauptvogel, CC BY-NC-SA.

The size and shape of grains within sedimentary rocks can also help you interpret that sediment’s history
(Figure 2.16). When all of the grains in the rock are about the same size, it is called well-sorted. Usu-
ally, sediment needs to travel a far distance from its source to be well-sorted. In contrast, poorly sorted
sediment is very close to its source. The rounding of sediment grains is also an indicator of how far the
sediment traveled. Grains that have smooth edges are considered well-rounded and have traveled a far
distance. These grains start out with rough edges and become smoother as they travel further and fur-
ther. In contrast, grains with sharp edges are considered angular and have not traveled a far distance.

Figure 2.16 – Grain characteristics used to infer the relative distance that sediments have traveled from their
source rock. a) Sorting of sediment; b) Sediment roundness, often called sphericity. Image credit: a) Adapted from
Wikimedia user Woudloper, CC BY-SA; b) Adapted from Wikimedia user Woudloper, CC BY-SA.

Exercise 2.5 – Identifying Sedimentary Rocks

Your instructor will provide you with a set of unknown sedimentary rocks to identify. Use Figures 2.13 to 2.17
and Table 2.2 to fill out the information in Table 2.7 below.
64 | CHAPTER 2: EARTH MATERIALS

Table 2.7 – Worksheet for Exercise 2.5

Clastic, (Bio-) Chemical, or


Sample Rock Name Possible Depositional Environments
Organic
CHAPTER 2: EARTH MATERIALS | 65

Figure 2.17 – Flowchart for identifying sedimentary rocks. Image credit: Virginia Sisson, CC BY-NC-SA.

2.6 Metamorphic Rocks


Metamorphic rocks (Figure 2.18) form when any pre-existing rock is altered by heat and/or pressure. The
66 | CHAPTER 2: EARTH MATERIALS

two sources of heat for metamorphism are the heat from a magma chamber and the geothermal gra-
dient, which is the natural increase in temperature when getting deeper into the Earth. Pressure also
increases with depth in the Earth, but intense pressure also occurs during tectonic plate collisions. This is
why most of the world’s metamorphic rocks are found in mountain belts or ancient mountain belts.

Heat and pressure can cause several changes in rocks, such as recrystallizing minerals, creating new min-
erals, and orientating minerals in a direction perpendicular to pressure. The orientation of minerals is the
most noticeable feature when recognizing metamorphic rocks. Heat works to “soften” minerals, allowing
ions to begin migrating in and out of crystal structures. Then pressure forces the minerals to re-orien-
tate. Heat is the most important factor, though; without heat, the rock would just break under intense
pressure. When minerals take on some type of orientation due to metamorphism, it’s called foliation. As
pressure increases, so does the degree of foliation. Not all metamorphic rocks exhibit foliation because
some minerals change very little under metamorphic conditions (quartz, for example); these types of
metamorphic rocks are called non-foliated or massive and can be very difficult to tell apart from igneous
rocks.

The pre-existing rock, before metamorphism occurred, is called the protolith. If you’ve taken an intro-
ductory Physical Geology class, you learned to differentiate metamorphic rocks mostly by their texture
and classified them as slate, phyllite, schist, and gneiss. You may have even been introduced to the term
migmatite for rocks that had both igneous and metamorphic textures as these were starting to undergo
partial melting. These textural terms are associated with changes in metamorphic temperature and pres-
sure, often referred to as the metamorphic grade or just grade.

Confused by metamorphism? Don’t feel alone, as many students and even professional geoscientists
have difficulties describing and interpreting metamorphic rocks. First, you need to understand all factors
affecting igneous and sedimentary rocks. Then, you add the effects of heat and/or pressure, but unlike
other rocks, this can vary dramatically across a region. Often there has been more than one tectonic
event in that region, adding even more complications for you to interpret. It is akin to understanding a
four- or five- dimensional image and difficult to comprehend all the factors at once. You may want to
review your Physical Geology lab and text before tackling metamorphic rocks and these exercises.
CHAPTER 2: EARTH MATERIALS | 67

Figure 2.18 – Metamorphic rock classification by texture, protolith, mineralogy, and type of metamorphism. Map
symbols refer to the standard symbols used by the US Geological Survey for each rock type. However, there are no
standard patterns for amphibolite and eclogite, so we have designed and included patterns for these two rocks
using the first letter of each rock name. The rocks in the sample pictures are from the Wards Classic American rock
collection at the University of Houston, except for gneiss, migmatite, hornfels, and eclogite, which are all
unpublished photos from VB Sisson. Image credit: Virginia Sisson, CC BY-NC-SA.

Exercise 2.6 – Identifying Metamorphic Rocks

Your instructor will provide you with a set of unknown metamorphic rocks to identify. Use Figures 2.18 and
2.19 to fill out the information in Table 2.8.
68 | CHAPTER 2: EARTH MATERIALS

Table 2.8 – Worksheet for Exercise 2.6

Sample Foliated or Non-foliated Rock Name Protolith


CHAPTER 2: EARTH MATERIALS | 69

Figure 2.19 – Flowchart for identifying metamorphic rocks. Begin by deciding if the metamorphic rock exhibits any type of
foliation. Then determine the grain size. Next, look at other features such as mineralogy, hardness, and how the rock
breaks. Image credit: Virginia Sisson, CC BY-NC-SA.

Let’s build on your metamorphic rock foundation and start to interpret metamorphic rocks using mineral
assemblages. To do this, you have to start by knowing the protolith (parent) material for the metamorphic
rock. If you can distinguish mafic (basaltic protoliths) from shale (sedimentary protoliths), this will be easy.
70 | CHAPTER 2: EARTH MATERIALS

During metamorphism, rocks with a shale protolith will always have quartz, feldspar, and muscovite mica.
So, you are on the lookout for new minerals, often seen as porphyroblasts. Most of these are not taught
in Physical Geology, so here are some minerals you will need to know: chlorite (a dark green mica), biotite,
cordierite, garnet, staurolite, sillimanite, kyanite, and andalusite (Figure 2.20). The last three minerals
on this list you may remember are polymorphs: minerals with the same composition but unique inter-
nal structures (Figure 2.21). If you can identify these alumino-silicate minerals in a schist or gneiss, then
you can determine the pressure-temperature conditions for that terrane. In general, rocks with kyanite
and sillimanite are found in areas that have undergone continent-continent collision and are considered
medium-pressure terranes. In contrast, areas with andalusite and sillimanite have an elevated geother-
mal gradient in response to either divergent zones or unusual ocean-continent collision zones called low-
pressure terranes.

Figure 2.20 – Index minerals used to interpret temperature ranges of metamorphic rocks. These index
minerals are for metamorphosed shales (pelites). These rocks all have quartz, feldspar, and muscovite
mica in addition to other index minerals. Green is for low-grade rocks that have chlorite. As temperature
increases, biotite (brown) begins to form. As the temperature continues to increase, medium-grade rocks
will have garnet (red) and kyanite (blue). These are easy to distinguish as these minerals grow as
porphyroblasts and often form bumps in the rocks. Finally, at the highest grade, sillimanite appears, and
then the rocks begin to melt, forming migmatites. Image credit: Virginia Sisson, CC BY-NC-SA.
CHAPTER 2: EARTH MATERIALS | 71

Figure 2.21 – Pressure temperature diagram for aluminosilicate minerals showing the
stability fields for andalusite, kyanite, and sillimanite. The dashed line is defined by the
stability of pyrophyllite and limits the lower temperature of andalusite and kyanite. This
diagram was calculated using the data of Berman (1991). Image credit: Virginia Sisson,
CC BY-NC-SA.

During a continental collision, the sequence of metamorphic rocks that form can be predictable by fol-
lowing the Barrovian sequence (Figure 2.22). Basically, the further away from the collision or the higher
up in the crust the rock is, the lower the grade of metamorphism, while the closer and deeper the rocks
are, the higher the grade. So, how do these deeper, high-grade metamorphic rocks become exposed at
the surface? The answer is erosion; the low-grade rocks above need to be removed to expose the high-
grade rocks below. This is relatively easy for an uplifting mountain belt because the increased surface
area of the mountain allows more weathering and erosion to occur. If you are looking for exposures of
high-grade metamorphic rocks, your best chance of finding them is near the center of the collision, where
much of the uplift and erosion took place. The Barrovian sequence pattern is when the grade of meta-
morphic rocks increases as you approach the collision center, reaching the highest grade near the center
and then decreasing in grade as you move to the other side (like a mirror image). Of course, this per-
fect, mirror-image pattern is not always seen because the rocks may not be exposed, or maybe faults
have moved rocks from their original positions, or there was significant variability in pressure-tempera-
ture conditions.
72 | CHAPTER 2: EARTH MATERIALS

Figure 2.22 – Metamorphic zones in continental convergence. The colored areas are the metamorphic
zones; these use the same scheme as in Figure 2.20. Green is chlorite, brown is for biotite, red is for
garnet, blue is for kyanite, and gray is for sillimanite. This scheme does not include andalusite as a
separate zone; it would overlap the kyanite zone. The dashed line is the boundary between the continental
crust and mantle, called the Moho. It is shortened from the term Mohorovičić discontinuity after the
Croatian seismologist Andrija Mohorovičić (1857-1936) who discovered this. This figure was simplified
from numerical models by Lyubetskaya and Ague (2010). The position of the Moho shows the initial
temperature gradient, and the metamorphic zones reflect the temperature 30 million years after
continental collision. Image credit: Virginia Sisson, CC BY-NC-SA.

Exercise 2.7 – Understanding the Barrovian Sequence

Continental collision and closure of the Iapetus Ocean in the Ordovician and early Devonian resulted in the
Caledonide Orogeny. By the way, Caledonia is Latin for Scotland. One of the hallmarks of continental collision
are suites of metamorphic rocks such as you’ve been looking at in Exercise 2.6.

Now let’s put together information from the rocks you’ve looked at with a map of the metamorphism of the
Scottish Highlands. British geologist George Barrow mapped the area in Figure 2.23 in the 1890s. He realized
that he could distinguish metamorphic zones by using their mineral assemblages. One more clue you need
to solve this exercise is that there is a relationship between pressure and depth in the Earth as 300 MPa cor-
CHAPTER 2: EARTH MATERIALS | 73

responds to approximately 10 km depth. So, you can convert the depths in Figure 2.22 to pressure with this
approximation.

Figure 2.23 – Barrovian metamorphic zones in northern Scotland. The colors used for the different metamorphic
zones are the same as in Figure 2.20. Green is chlorite, brown is for biotite, red is for garnet, blue is for kyanite, and
gray is for sillimanite. Barrow also mapped a staurolite zone which is not shown here. The black circles indicate the
location of major Scottish cities. The area is bound by two faults, the Great Glen strike-slip fault to the northwest and
the Highland Boundary reverse fault to the southeast. Image credit: Virginia Sisson, CC BY-NC-SA; simplified from
Wikimedia user Woudloper, CC BY-SA.

a. Label the colored areas on the map with the mineral present according to the scheme used in Figure
2.20.
b. Where on this map do you think your samples from Exercise 2.6 occur?
c. On the map, mark which areas you think represent the shallowest and deepest during metamorphism.
d. Draw a line on the map that corresponds to only a change in temperature.
e. On the cross-section shown in Figure 2.22, plot an approximate path through the metamorphic zones. It
will be up to you to decide how your metamorphic rocks evolved and your path may be different from a
colleague’s.
f. To help understand the metamorphic history of a region, geologists will construct geodynamic models
of how heat and pressure in the Earth evolve through time. These models start with an initial set of con-
ditions for the area. In Figure 2.22, the heavy black lines show two continental plates colliding; each has
an initial lithosphere thickness of ~ 35 km. The convergence causes deflection of the Moho (Mohorovičić)
discontinuity between the crust and mantle. Over time, the temperature readjusts. The colored areas on
this plot show the metamorphic assemblages about ~35 million years after collision (Lyubetskaya and
Ague, 2010). The first step to using this diagram is to convert metamorphic depths to pressures. A good
rule of thumb is that 3 km depth is ~100 MPa. Use your results from part e to fill in Table 2.9.
74 | CHAPTER 2: EARTH MATERIALS

Table 2.9 – Worksheet for Exercise 2.7f

Metamorphic Zone Depth Pressure Temperature

Chlorite

Biotite

Garnet

Kyanite

Sillimanite

g. Using your calculated pressures from Table 2.9, draw a line on Figure 2.21 that corresponds to your
sequence of metamorphic assemblages from the map. This is called a pressure-temperature (PT) path.
Geologists now like to add geochronology constraints (t), deformation (d) or compositional constraints
(x), so you’ll hear them discuss PTt, PTtd or PTx paths when trying to determine the tectonic history of a
region.

2.7 Using Rocks to Interpret Earth’s History


Igneous, sedimentary, and metamorphic rocks can all be used to interpret Earth’s history. Igneous rocks
can help determine the plate tectonic setting of an area, sedimentary rocks help determine the environ-
ment of an area, and metamorphic rocks provide information about rocks before their deformation and
the tectonic setting. All the following exercises will use your critical thinking skills.

Exercise 2.8 – Interpreting Volcanic History

There are a few places on Earth with significant mountainous relief as well as igneous activity. One such place
is the western margin of South America, which has been the site of ocean-continent convergence since the
Jurassic (~185 Ma). Thus, it has a long history of almost continuous subduction-related volcanism and igneous
activity. In the Central Andes Mountains, igneous rocks formed during the Neogene portion of geologic time
between 23 and 2.5 million years ago; these include both intrusive and extrusive rocks (Figure 2.24). These
volcanic rocks are overlain by more recent volcanoes.
CHAPTER 2: EARTH MATERIALS | 75

Figure 2.24 – Map of the Neogene igneous rocks of the Peruvian


Andes. Intrusive rocks are shown as red areas, and extrusive volcanic
rocks are shown in the orange areas. Black triangles are locations of
active volcanoes. Note that this map is not oriented north-south.
Image credit: Virginia Sisson, CC BY-NC-SA; simplified from Pfiffner
and Gonzalez, 2013.

a. In the suite of igneous rocks provided by your instructor, which rocks are most closely associated with
the Neogene intrusives?

b. Which rocks would be associated with the Neogene extrusive volcanic rocks?
76 | CHAPTER 2: EARTH MATERIALS

c. Would the same rock types be found in the active volcanoes?

d. Describe at least three differences between the map patterns of the recent volcanoes, Neogene extru-
sives, and Neogene intrusives.

e. Critical Thinking: What can explain some of these differences in the map patterns?

Exercise 2.9 – Rocks of the Grand Canyon

The Grand Canyon has amazed tourists and geoscientists alike, not only for its beauty but by the fantastic
stratigraphic section
sections (layers of sedimentary rock) exposed in its walls. The South Rim of the Grand Canyon
is composed of metamorphic, igneous, and various sedimentary rocks. Go here if want to see representative
samples of some of these rocks. Figure 2.25 shows a stratigraphic column of the Grand-Canyon (see Figure 0.3
for a sketch of the Grand Canyon).
CHAPTER 2: EARTH MATERIALS | 77

Figure 2.25 – Cross-section of the Grand Canyon. This view shows the inner gorge, Temple Butte,
and southern rim on the west bank of the south-flowing Colorado River. Temple Butte’s elevation
is 1618 m; the maximum depth of the Grand Canyon is over a mile or 1.829 km. This cross-section
shows simplified stratigraphy of the sedimentary, igneous, and metamorphic rocks exposed in the
cliffs and walls of the canyon. The different formations are shown using patterns for the different
rock types commonly used by the U.S. Geological Survey. In other regions of the Grand Canyon,
the Grand Canyon Supergroup overlies the Vishnu Schist and Zoraster granite. This is not shown in
this figure. The figure is adapted from several publicly available stratigraphic columns, including
the National Park Service. Image credit: Virginia Sisson, CC BY-NC-SA.

Table 2.10 contains descriptions of some of the different rocks found in the Grand Canyon.

a. Which rocks from your kit resemble the sedimentary units marked with a star on the Grand Canyon
78 | CHAPTER 2: EARTH MATERIALS

stratigraphic section? Fill this out in the third column of Table 2.10.
b. The symbols used to describe these rocks are standardized. Based on your findings from part a, com-
plete the legend in Figure 2.25 by identifying the type of rock each symbol stands for.

Table 2.10 – Rock formations and descriptions for the Grand Canyon for Exercise 2.9a

Rock
Description Matching Rock From Your Kit
Formation*

Group of foliated metamorphic rocks that are the result of


Vishnu Group various rocks, mostly igneous, being exposed to high degree
metamorphism.

Igneous rocks composed of coarse-grained intrusions


Zoroaster Group composed mostly of feldspar and quartz, with minor
hornblende and biotite.

Composed mostly of sand-sized quartz grains with medium


Tapeats
roundness. It is mostly massive, but bedding is visible in
Formation
some areas.

Bright Angel Composed of fine-grained shale, and trilobite fossils are


Formation common within this unit.

Chemical and biochemical sedimentary rocks that react with


Muav Formation
dilute acid.

Extremely fine-grained, dark-colored sedimentary rocks


Hermit which have easily observable bedding. Most minerals are
Formation clay minerals but can’t be observed, even with the help of a
hand-lens, due to their extremely fine grain size.

*See this Wikipedia entry for the differences in stratigraphic unit terminology.

Most of the rocks formed at divergent plate boundaries are igneous, with a thin layer of deep marine
sediments on top. Since these sediments accumulate in the deep ocean far from the continents, there
are only chert and marine clays. However, there are significant accumulations of igneous rocks derived
from the decompression melting of Earth’s peridotite mantle. Do you recall what kind of rock forms from
the partial melting of an ultramafic rock? It will be a mafic rock, either extrusive or intrusive. The oceanic
crust is subdivided into five layers, as shown in Figure 2.26. This entire sequence is called an ophiolite.
Ophiolites are typically exposed on Earth’s surface during the end of the subduction process as two con-
tinents move closer together and collide. During these final stages of subduction, the oceanic lithosphere
gets stuck in the middle of the continental collision and is uplifted with the newly forming mountain belt.
This is why most ophiolites are found in mountains.

Exercise 2.10 – Ophiolites

Ophiolites offer clues to where oceanic crust used to exist, and they help geologists understand processes at
mid-ocean ridges.

a. Your instructor has given you a suite of rocks that represent different layers in an ophiolite sequence.
Identify your rocks and complete the legend at the bottom of Figure 2.26 by filling in the compositions
and rock names.
CHAPTER 2: EARTH MATERIALS | 79

Figure 2.26 – Sequence of rocks in an ophiolite complex. Since it is difficult to find a location with
all of these igneous rock types exposed in one outcrop, this is a composite of several different
ophiolite exposures. The patterns used for the different units are adapted from the U.S. Geological
Survey. Since most ophiolite sequences are composed primarily of mafic to ultramafic igneous
rocks underneath deep marine sediment, we used dark green to brown colors for the different
rock types. Image credit: Virginia Sisson, CC BY-NC-SA.

b. Explain why ophiolites show a sequence that moves from intrusive rock to extrusive. Hint: Think about
80 | CHAPTER 2: EARTH MATERIALS

relative cooling rates.

c. What do pillow lavas tell you about where they were formed?

d. Now use your knowledge of ophiolites to answer some questions about the geologic history of the
Himalayan mountains. In Figure 2.27, name the two tectonic plates. Are these oceanic or continental
tectonic plates? What is the thickness of their lithosphere?

e. What type of rocks are shown in the black blobs on this geologic map? ____________________
f. There are five other groups of rocks in the Himalayan Mountains. Each of these groups has a distinct
geologic history.

i. Why do you think these groups of rocks are distributed as long skinny units?

ii. If the rock units are elongated in a particular direction, geologists call this their strike direc-
tion. Typically strike direction is given as a compass direction. What is the strike direction of
these groups? ____________________
iii. Does the strike change from east to west along the Himalayan Mountains?
____________________
iv. How does the strike of the ophiolites compare to the contacts between the units and to the
thrust and normal faults?

v. What can you infer from your observations of the type of faults and the ophiolites? Hint: do
the ophiolite blobs occur next to both the thrust and normal faults?

g. Make some observations about the map patterns of the ophiolites. Are the ophiolite blobs in a continu-
ous or discontinuous belt?

h. Since ophiolites are signatures of previous ocean basins, what happened to all of the oceanic crust that
CHAPTER 2: EARTH MATERIALS | 81

was once between these two continental tectonic plates?

i. Critical Thinking: Some geoscientists suggest using the term suture zone to describe where ophiolites
occur during continent-continent convergence. Looking at the map, do you think this term is appropri-
ate for the Himalayan Mountains. Explain your answer.

Figure 2.27 – Geologic map of the Himalayan Mountain range. The insert is a shaded relief map of Asia, including Nepal,
Bhutan, Pakistan, Afghanistan, China, Russia and India. Most of the high topography is in an area called the Tibetan Plateau. To
the south is the low area of India and to the north is Russia and China. Map scale reference is 30° latitude. Within the insert,
there is a black box that outlines the area shown below. The location for Mt. Everest (Chomolungma in Tibetan) is shown for
reference. The geologic map has six main lithologies. The northernmost unit includes igneous rocks of the Transhimalayan
batholith and Kohistan arc shown in a pink, standard crystalline rock pattern. Scattered throughout the map are occurrences of
ophiolite sequences indicated as black areas. If you are at the eastern (right) edge of the mountains, the next unit is the Indus
suture zone shown in a blue dot and open circle pattern. Next to this is the Tethyan zone in a brown-stippled pattern. The next
unit is the Greater Himalayan sequence in wavy blue lines. The southernmost unit is the Lesser Himalaya, shown as red dots.
Each unit is separated from the next by either a thrust fault (line with barbed teeth) or a normal fault (line with hatchures).
There is a major strike-slip fault (Karakoram) that offsets part of the Himalayan mountains on the northwest side of the range.
If you look carefully at the shaded relief map, you can pick out how far this unit goes into the Tibetan Plateau. Abbreviations:
NP (Nanga Parbat), A (Annapurna), K (Katmandu), E (Everest), and B (Bhutan). Image credit: This map was simplified from MP
Searle and PJ Treloar, 2019.
82 | CHAPTER 2: EARTH MATERIALS

Exercise 2.11 – Southern Alaska

There is a sequence of metamorphic rocks in southern Alaska called the Chugach Metamorphic Complex (Fig-
ure 2.28). In one area along the Tana River, you can walk upstream from phyllite to gneiss. However, to deter-
mine the metamorphic conditions, you need to identify the mineral assemblages. A team of geologists worked
along this river for three summers and mapped the distribution of metamorphic minerals and the deforma-
tion, textural, and geochronological history. The outcrops occur where a river along a glacier eroded away
the glacial sediments and exposed the bedrock. The protoliths for these rocks are graywacke deposited in an
accretionary prism during the Late Cretaceous. Now that you have the geologic history for this region try to
interpret the metamorphic pattern shown in Figure 2.28.
CHAPTER 2: EARTH MATERIALS | 83

Figure 2.28 – Map of metamorphic zones in southern Alaska along the


Tana River. The color scheme is similar to Figure 2.20, except this area has
andalusite instead of kyanite. Green is chlorite, brown is for biotite, red is
for garnet, blue is for andalusite, yellow for staurolite (a yellow-brown
silicate), and gray is for sillimanite. The boundaries between these zones
are isograds marked as solid black lines. There are two dashed lines
indicating boundaries between rocks with different textures (phyllite,
schist, and gneiss). The blue-stippled pattern indicates the position of the
Tana Glacier in 1988. The gray stipple pattern represents glacial deposits.
The insert map shows an arrow pointing to the location of the Tana River
in southern Alaska. Image credit: This map is adapted from Bowman et al.
(2003) with unpublished mapping of Sisson.

a. Do the textural boundaries between phyllite, schist, and gneiss correspond to mineralogical changes?
If not, why do you suppose they are different?
84 | CHAPTER 2: EARTH MATERIALS

b. Is this a medium- or low-pressure terrane (consider Figure 2.21)? Did it form as a contact aureole or
regional metamorphism?

c. Using a graph of metamorphic temperatures in Figure 2.20, what is the difference in temperature from
north to south along the Tana River?

d. What is the geothermal gradient along the Tana River? To determine the horizontal geothermal gradi-
ent, determine the distance between the two points you used for estimating the metamorphic temper-
ature.

e. How does the horizontal geothermal gradient compare with vertical geothermal gradients in stable con-
tinental crust?

f. What was the heat source for the metamorphism?

g. Critical Thinking: This map shows the distribution of another mineral common in metamorphosed
shales, staurolite. Using the temperature scale, determine the temperature that the staurolite formed.

h. Critical Thinking: Assume this area was at ~10 km depth or ~300 MPa. Then, use your temperature esti-
mates from Exercise 2.7f as well as 2.11g to plot the pressure-temperature history for southern Alaska
in Figure 2.21. This will be the path that the highest-grade rocks would have taken through time.

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson, Carlos Andrade
CHAPTER 2: EARTH MATERIALS | 85

References
Berman, R.G., 1991, Thermobarometry using multi-equilibrium calculations; a new technique, with petro-
logical applications, The Canadian Mineralogist, v. 29 (4), p. 833–855

Bowman, J.R., Sisson, V.B., Pavlis, T.L., and Valley, J.W., 2003, Oxygen isotope constraints on fluid infiltra-
tion associated with high temperature–low pressure metamorphism (Chugach Metamorphic Complex)
within the Eocene southern Alaska fore-arc, in Sisson, V. B., Roeske, S. M., Pavlis, T. P., editors, Geo-
logical consequences of ridge-trench interactions in the northern Pacific, Geological Society of America
Special Publication no. 371, p. 237-252

Lyubetskaya, T., and Ague, J.J., 2010, Modeling metamorphism in collisional orogens intruded by magmas:
I. Thermal evolution: American Journal of Science, v. 310, p. 427-458, DOI: 10.2475/06.2010.02

Pfiffner, O. A., and Gonzalez, L., 2013, Mesozoic–Cenozoic Evolution of the Western Margin of South Amer-
ica: Case Study of the Peruvian Andes, Geosciences 3, no. 2: 262-310, DOI: 10.3390/geosciences3020262

Searle, M.P., and Treloar, P.J., 2019, Introduction to Himamalyan tectonics: a modern synthesis, in
Himalayan Tectonics: A modern synthesis ed by P.J. Treloar and M.P. Searle, Geological Society of Lon-
don Special Publications, v. 483, p. 1-17 https://doi.org/10.1144/SP483-2019-20

Google Earth Locations


CHAPTER 3: GEOLOGIC TIME

Learning Objectives

The goals of this chapter are to:

• Illustrate the immense scale of geologic time


• Summarize the geologic time scale
• Explain the principles of relative dating and unconformities
• Apply principles of relative and absolute dating to determine the ages of rocks

3.1 Introduction
Earth is 4.543 billion years old. That’s 4,543,000,000 years, an amount of time so immense that it’s chal-
lenging to grasp just how long it is. To put this into perspective, if the average human lifespan is 80
years, the Earth has been around for 57,000,000 lifetimes. Or if you have a penny for every year the
Earth has been around, you would have $45.4 million! Constantly writing out millions and billions of years
is time-consuming, so when geologists talk about ages, they use a few abbreviations. The symbols ka
(thousands), Ma (millions), and Ga (billions) refer to points in time like a date. For example, the dinosaur
extinction occurred at 66 Ma. Geologists also use other abbreviations for lengths of time, including ky,
kya, kyr, and k.y. for thousands of years; my, mya, myr, and m.y. for millions of years; and by, bya, byr, and
b.y. for billions of years. All four varieties of abbreviations mean the same thing in this case. Here, you
would say the dinosaurs have been extinct for 66 myr. If this sounds confusing, you’re not alone because
even some geologists use all of the abbreviations interchangeably. There is a debate amongst geologists,
and other sciences, over the notation used for geologic time.

Fun fact: The Tyrannosaurus rex was one of the last dinosaurs to evolve about 70.6 Ma (a specific date).
The first dinosaurs evolved about 230 Ma (a specific date), 159 myr (a length of time) before T-rex evolved.
We are closer in time to the T-rex than the T-rex is to its earliest dinosaur ancestor! See the difference in
abbreviations yet?

3.2 Geologic Time


Since 4.54 byr is a large chunk of time, geologists have divided it into more manageable chunks by cre-
ating a time scale. The commonly accepted time scale comes from the International Commission on
Stratigraphy (Figure 3.1). It is continually revised as new research fine-tunes numbers between time scale
divisions. The one in Figure 3.1 is the most up-to-date at the time of this writing and will be referenced
throughout this manual. The divisions on the time scale are often based on significant events that have
taken place tectonically, biologically, or climatically, and the numerical ages are derived from radiometric
dating of rocks, minerals, and fossils.

Geologic time is first divided into eons; these are the Hadean, Archean, Proterozoic, and Phanerozoic.
The first three eons are often referred to as the Precambrian, which we’ll call a “super” eon. The eons
are subdivided into eras, and eras are subdivided into periods, and periods into epochs, and epochs into
CHAPTER 3: GEOLOGIC TIME | 87

ages. For the purposes of this lab manual, we will refer to this nomenclature. You’ll notice there is a sec-
ond way to refer to these divisions on the time scale. The difference between the two is discussed here,
but it is not important for our purposes. Throughout this lab manual, you will mainly see us referencing
periods of geologic time.

Figure 3.1 – The geologic time scale subdivided by eon, era, system, series, epoch, and stage. What do you observe
about each period of time? It is simple; they are color-coded. We will use these standard colors throughout this lab
manual. The golden spikes at age boundaries indicate a specific place used to define the age boundary. Image
credit: International Commission on Stratigraphy.

Exercise 3.1 – Making Your Own Geologic Time Scale

Many depictions of the geologic time scale don’t show the divisions of geologic time on the same scale. Look
at the time scale in Figure 3.1, for example. The far-right column goes from 4.6 Ga to 541 Ma; that’s about 4
billion years of history in one small column! The other three columns make up the remaining 500 myrs. The
reason for this is that geologists know much more about the last 500 myrs of Earth’s history than the first 4
byrs. So, let’s make a geologic time scale where all geologic time is shown at the same scale.

a. Using a 2.5 m long roll of paper, create your own geologic time scale using the following scale: 1 cm = 20
million years. For the purpose of this exercise, round Earth’s age to 4.6 Ga and use a tick mark spacing
of every 100 myrs.
88 | CHAPTER 3: GEOLOGIC TIME

b. Label the Precambrian and its associated eons.


c. Label the Phanerozoic eon and its associated eras and periods.
d. For the Cenozoic era, label the epochs.
e. Table 3.1 is a list of some major events in Earth’s history. On one side of your time scale, label when you
think these events occurred using a color of your choice.
f. Using a cell phone or laptop, look up when these events actually occurred and label them on the other
side of your time scale. Label physical events in one color and biological events in a different color. Some
of the events will have a range of ages.
g. How close were your guesses?

Table 3.1 – A list of major events in Earth’s history

Formation of Earth’s Moon Great oxidation event


Dinosaur extinction Formation of the Himalayan Mountains

First fossil evidence of life First homo sapiens


First fish First mammals
First reptiles First amphibians
First major glaciation, called Snowball
End of the last ice age
Earth
Breakup of Pangea Formation of Rocky Mountains

Earliest trace of life (bacteria) Oldest oceanic crust

h. Come up with a mnemonic device to help you remember the periods of geologic time in the Phanerozoic
Eon. For example, did you know that the word “scuba,” as in scuba divers, is actually a mnemonic device
that stands for Self-Contained Underwater Breathing Apparatus?
CHAPTER 3: GEOLOGIC TIME | 89

Geologists use abbreviations to refer to the different portions


of geologic time (Table 3.2), particularly on geologic maps. A Table 3.2 – Abbreviations for the Geologic
few of these abbreviations may seem puzzling because the Time Scale
abbreviation isn’t always the first letter of the name. Carbonif- Geologic Time Period Abbreviation
erous got the “C”, so Cambrian got a C with a slash through it
“?“. This left Cretaceous with a “K.” Tertiary had originally Quaternary Q
received the “T”, which left Triassic to be symbolized with a T Neogene Ng
that had an R subscript. In 2003, “Tertiary” was stricken from
the geologic timescale and replaced by the Paleogene and Paleogene Pg
Neogene. You will still see “T” on older maps for Tertiary, and Cretaceous K
some geologists still use the term today. So, be careful! There
were already some M’s and P’s; so, Proterozoic eras were Jurassic J
assigned as X, Y, and Z. Triassic TR

Geologists use two methods for dating events in Earth’s his- Permian P
tory. The first is called relative dating, meaning how events
Carboniferous C
relate to each other in time, or more plainly, they figure out
the sequence of events (what came first, second, third, etc.). Devonian D
Relative dating has no regard for numerical ages. The second
Silurian S
method is absolute dating, where geologists use radioactive
isotopes to figure out the numerical age of a rock or mineral. Ordovician O

Cambrian ?

3.3 Relative Dating and Correlation Neoproterozoic Z

Mesoproterozoic Y
There are several principles geologists use for relative dating. Paleoproterozoic X
The first four principles were developed in the 17th century
by an early geologist named Nicolas Steno, three of which Archean A
pertain to sedimentary rocks. The first is the law of superpo-
sition, which states that in layers of horizontal sedimentary
rocks, the oldest rock layer is at the bottom, and the youngest is at the top (Figure 3.2). The second rule
is the principle of original horizontality, which says that layers of sediment are originally deposited hori-
zontally (Figure 3.2). So, any tilting or folding of the rock occurred after it was deposited (Figure 3.3). The
third principle states that layers of sedimentary rock are continuous, and anything that interrupts the
layer (like a river or canyon) happened after the rock formed. This is called the principle of lateral conti-
nuity (Figures 3.4 and 3.5).
90 | CHAPTER 3: GEOLOGIC TIME

Figure 3.2 – Horizontal layers of red sedimentary rocks from the Moenkopi and Chinle Formations from the Triassic
period in the Virgin River Valley near Zion National Park. Layers of sedimentary rock are originally deposited
horizontally, with the oldest layer at the bottom. This shows both the principles of superposition and original
horizontality. Image credit: James St. John, CC-BY.
CHAPTER 3: GEOLOGIC TIME | 91

Figure 3.3 – These are tilted layers of the Moenkopi Formation from the Lower Triassic, just south of Split
Mountain, Dinosaur National Monument in Utah. They were originally deposited horizontally and were titled at
some point in geologic time. Image credit: James St. John, CC BY.

Figure 3.4 – A depiction of the principle of lateral continuity. The layers of rock on one side of this valley were once
connected to the same rocks on the other side. The valley was carved after these rocks were deposited. The
dashed red lines show how the units were connected. Image credit: Wikimedia user Wouldloper, Public Domain.
92 | CHAPTER 3: GEOLOGIC TIME

Figure 3.5 – A panoramic view of meanders in the San Juan River cutting through layers of sedimentary rock in
Goosenecks State Park, southwestern Utah. If you look carefully, you can trace layers of sedimentary rock that
have been cut by the river across the entire image; an example of lateral continuity. Image Credit: Gernot Keller, CC
BY.

Steno’s fourth principle is cross-cutting relationships. This principle is used when other geologic events
cut through sedimentary rocks, like an igneous dike or a fault. This principle basically states that when a
geologic event cuts across another, the event doing the cutting is younger than the one being cut (Figure
3.6). For example, if sedimentary rocks are cut by an igneous dike, the igneous dike is younger than the
sedimentary rocks it’s cutting through. The same can be said of a fault that cuts through any rock; the
fault has to be younger because the rocks had to exist first to be faulted.
CHAPTER 3: GEOLOGIC TIME | 93

Figure 3.6 – A mafic dike (dark rock) cutting through granitic pegmatite
(light rock) in Ruggles Mine, New Hampshire. To determine which is
younger, look to see which unit cross-cuts the other. Since the mafic
dike is cutting through the pegmatite, the dike is younger by the
principle of cross-cutting relationships. Image credit: James St. John,
CC BY.

Some 200 years later, the fifth principle of relative dating was developed by Charles Lyell called the princi-
ple of inclusions. This principle explained that a clast, or a different-looking rock that is contained inside
of another rock, is older than the rock that contains it (Figure 3.7). How can this happen? Originally, a
mafic magma was cooling quickly, producing the finer-grained mafic rock in the middle of Figure 3.7.
Then, something happened to change the chemistry of the magma to felsic and slowed the cooling rate
to produce the surrounding, coarse-grained granite. The mafic rock formed first, and then the felsic rock
formed around it.
94 | CHAPTER 3: GEOLOGIC TIME

Figure 3.7 – A mafic inclusion in granitic rock. Inclusions are older than the surrounding rocks. A penny shows the
scale of the image. Image credit: Marli Miller, CC BY.

Unconformities
Simply speaking, an unconformity is a pattern that you look for in a group of rocks that tells you erosion
has taken place. Rocks exposed on the Earth’s surface are affected by physical and chemical weathering
processes that work to break them into smaller pieces or dissolve them in water. This material is then
transported away by wind, water, or ice, a process known as erosion. Many people use weathering and
erosion interchangeably, but they do mean different things: weathering is the breakdown of rocks, while
erosion removes the broken down material.

There are four types of unconformities, and each forms in a slightly different way (Figure 3.8). They all
involve sedimentary rocks, changes in sea level, and/or uplift from an orogeny. Each unconformity tells a
unique story of the geologic history of the area they’ve been found.
CHAPTER 3: GEOLOGIC TIME | 95

Figure 3.8 – Types of unconformities. (a) Disconformity; (b) Nonconformity; (c) Angular unconformity; (d)
Paraconformity. Image credit: Wikimedia user ‫דקי‬, CC BY-SA.

A disconformity (Figure 3.8a) is an erosional surface where the rocks below the unconformity are much
older than the rocks above. This type of unconformity typically forms when horizontal layers of sedimen-
tary rock are deposited in a shallow marine environment; then sea level lowers to expose these rocks
and allows erosion to occur; then sea level rises again, and new horizontal layers of sedimentary rock are
deposited. Erosion removed some of the original rock, creating a large age gap between the rocks above
and below the erosional surface. This age gap is the disconformity and is located at the contact point
between the older rock and younger rock. Oftentimes the erosion process leaves behind evidence of river
channels or soil development, which provide clues to geologists to locate the unconformity in what looks
like a continuous succession of sedimentary layers.

A nonconformity (Figure 3.8b) forms when igneous or metamorphic bedrock is eroded, and then horizon-
tal layers of sedimentary rock are deposited directly on top of it. The unconformity is where the bedrock
meets the sedimentary rock. For example, when a mountain belt is eroded below sea level, and afterward
sediments are deposited on top of the igneous or metamorphic rock, the contact is a nonconformity.

An angular unconformity (Figure 3.8c) is created when horizontal layers of sedimentary rock lie on top
of tilted layers of sedimentary rock. The most famous angular unconformity is from Siccar Point in Scot-
land. Figure 3.9 shows the process of creating an angular unconformity. For this to occur, sedimentary
rocks deposited in the marine environment are lifted above sea level by an orogeny or similar event. The
orogeny causes the sedimentary rocks to become tilted or folded. Since these rocks are exposed above
sea level, erosion takes place. The rocks can either be eroded below sea level, or sea level can rise, which
96 | CHAPTER 3: GEOLOGIC TIME

would allow new, horizontal layers of sedimentary rock to be deposited on top of the titled ones. This
creates an angle between the younger, horizontal layers on top and the older, tilted layers below.

Figure 3.9 – A depiction of how an angular unconformity forms. 1) Deposition of sedimentary rocks; 2) uplift and
folding of the sedimentary layers; 3) erosion; 4) resumed deposition on top of the folded layers. Image credit: Utah
Geological Survey, Public Domain.

A paraconformity (Figure 3.8d) is very similar to a disconformity, except the evidence for erosion is not
present. Either no evidence of the erosion was left behind, or erosion didn’t happen, and instead, there
was only a pause in sediment deposition. We know these occur because sediment above and below the
paraconformity have been radiometrically dated and reveal a large gap in time.

Exercise 3.2 – Identifying Unconformities

The images in Table 3.3 are different types of unconformities (Figures 3.10-3.12). In the blank spaces provided
next to the images, create a sketch for each unconformity, label where you think the unconformity is located,
and identify the type of unconformity.
CHAPTER 3: GEOLOGIC TIME | 97

Table 3.3 – Worksheet for Exercise 3.2

Figure 3.10 – Unconformity #1.


Image credit: Marli Miller, CC BY.

Figure 3.11 – Unconformity #2.


Image credit: Callan Bentley, used
with permission.

Figure 3.12 – Unconformity #3.


Image credit: Marli Miller, CC BY.

Exercise 3.3 – Practicing Relative Dating Principles

Part I

Now put all of the principles you’ve learned to work. Below are relative dating outcrop diagrams that represent
sections of rock. Each letter represents the deposition of a different layer of sedimentary rock or geologic
event.

a. Use the principles of relative dating and unconformities to determine the sequence of events for each
diagram in Table 3.4 (Figures 3.13-3.16).
b. Identify the type of each unconformity.

The symbols used to represent common types of rocks are standard USGS symbols. You will only see conglom-
erate, limestone, sandstone, shale, granite, and gneiss in this exercise. The subscript letters stand for igneous
dikes (D), faults (F), and unconformities (U). The colors for each unit are from the geologic time scale shown in
Figure 3.1. Hint: it is easier to start with the oldest event and work your way forward through time.
98 | CHAPTER 3: GEOLOGIC TIME

Table 3.4 – Worksheet for Exercise 3.3

youngest

oldest

Figure 3.13 – Diagram #1. Image credit: Daniel Hauptvogel, CC BY-NC-SA.


CHAPTER 3: GEOLOGIC TIME | 99

youngest

oldest

Figure 3.14 – Diagram #2. Image credit: Daniel Hauptvogel, CC BY-NC-SA.

youngest

oldest

Figure 3.15 – Diagram #3. Image credit: Daniel Hauptvogel, CC BY-NC-SA.


100 | CHAPTER 3: GEOLOGIC TIME

youngest

oldest
Figure 3.16 – Diagram #4. Image Credit:
Daniel Hauptvogel, CC BY-NC-SA.

Part II

The web program Visible Geology lets you create block models with a complicated geologic history. An exam-
ple of a diagram made with this website is in Figure 3.17. Answer the following questions using Figure 3.17:

c. Place the following in order from youngest to oldest: Fault, Dike, Folding. ______________________________
d. If the dike is 100 Ma, what is the age of the normal fault? ____________________

Figure 3.17 – Block diagram created using the website Visible Geology. Sedimentary strata are
shown in yellow, blue, red, white, and tan. The dark brown is an igneous dike. The blue line is a
fault. The dashed lines represent contour lines.
CHAPTER 3: GEOLOGIC TIME | 101

Part III

Create your own block model using the web program Visible Geology.

• Your model should include 2 to 6 sedimentary layers. You may want to vary their thickness. Each rock
type is assigned a unique color. You can also include an unconformity (angular, disconformity, or non-
conformity).
• You can include tilting, folds, thrusts, strike-slip faults, and normal faults.
• You may want to include some igneous rocks as sills or dikes. If you include a dike or sill, you may want
to assign it a specific age, such as 100 Ma. This numerical age can be used to give specific dates to events
in this geologic history.
• Include erosion at some point in your model.

e. When you are done with this, rotate the cube and decide which side has the easiest and most difficult
history to assess. Or perhaps you made a model that is easy on all sides? Explain why there is this dif-
ference.

f. Create a geologic cross-section through your model from corner to corner. Draw a sketch of one of these
cross-sections.

Cross-Section Sketch #1:

g. Sometimes geologists only have cliff faces or road cuts to investigate. Do you think you will get the same
geologic history with this cross-section as one of the sides?

h. Now draw a quick sketch of one of the sides. Show it to one of your classmates and see if they can inter-
pret its geologic history. If you included an igneous feature, assign it an age. If your colleague can’t get
the history, give them some clues to help.

Cross-Section Sketch #2:


102 | CHAPTER 3: GEOLOGIC TIME

Exercise 3.4 – Using Relative Dating Principles

One of the ways geologists investigate Earth’s history is by imaging what is below Earth’s surface; this is called
a seismic survey. It is done by sending sound waves into the ground or ocean. As these sound waves move
through different layers of rock or sediment, some of the waves are reflected back toward the surface and
recorded by geophones
geophones. Different types of sediments or rocks change the characteristics of the wave, such as
its velocity. Geologists set up geophones that record these reflected waves and provide an “image” of what
the subsurface looks like. The signals recorded are called the seismic amplitudes, a measure of the difference
in rock properties between two layers. Seismic data is commonly converted to impedance, or hardness (this
is not the same as Mohs hardness). The relative hardness can be positive, negative, or the same. Thus, many
seismic sections will use three colors to better distinguish different strata.

Part I

The continental slope off of the north island of New Zealand is called the Hikurangi margin. Geoscientists
wanted to explore this area to better understand the plate boundary by drilling a sediment core. The Interna-
tional Ocean Discovery Program (IODP) proposed to put a drill core at Site U1519, but first, they conducted
a seismic survey (Figure 3.18) to assess the possibility of active thrust faults in the
area.

a. What type of unconformity is the yellow line?


____________________
b. What type of unconformity is the pink line?
____________________
c. Critical Thinking: Do you think it would be safe
to drill here? Explain your answer.

Figure 3.18 – Seismic profile for IODP Site U1519. There are
two unconformities shown with yellow and pink lines. Also
shown in a thin black line is the bottom simulating reflector
(BSR). The black color is used for positive amplitudes, red for
negative amplitudes, and white for zero amplitude. Image
credit: Wallace et al., 2019 and IODP, CC BY.

Part II

As the Pacific plate traveled over numerous hot spots, the basement of the Gulf of Alaska became riddled with
seamounts. The seamounts have a variety of ages and sizes, as you can see in Figure 3.19. These are partially
covered by a thick blanket of sediments up to 1.5 km thick.
CHAPTER 3: GEOLOGIC TIME | 103

Figure 3.19 – Seismic profile from the Gulf of Alaska. Since this is a preliminary seismic section, it has not been processed
to show depth. Instead, it shows two-way travel time in seconds for the seismic energy. The red, black, and white lines
indicate sedimentary layers, whereas the red, black, and white dots represent the oceanic crust (basement). The
sedimentary layers are more than 1.5 km thick. The same color scheme is used for seismic rock properties as in Figure
3.18. Image credit: USGS, Public Domain.

d. Identify two areas where faulting has taken place by outlining them in Figure 3.19.
e. Why is the contact between the crystalline basement and the sedimentary layers above it not considered
unconformity?

Sedimentary Rock Correlation


The principles of relative dating allow geologists to compare seemingly similar groups of rocks separated
by some distance. On a small scale, it’s as simple as saying that rocks on either side of a canyon are, in
fact, the same and were once connected before the canyon formed, as in Figure 3.5. On a larger scale
of kilometers to hundreds of kilometers, it can be comparing sets of sedimentary rocks that have similar
patterns.
104 | CHAPTER 3: GEOLOGIC TIME

Exercise 3.5 – Correlating Layers of Sedimentary Rock

Geologists can correlate sedimentary rock units over great distances by matching patterns in the outcrops,
called lithostratigraphic correlation. Figure 3.20 below contains two outcrop drawings of sedimentary rock
outcrops that are separated by 100 km.

a. Draw lines that correlate the rock units across these two areas. You can color the units to help you.
b. One of these areas contains an unconformity. Mark the location of the unconformity in Figure 3.20.
c. What type of unconformity is it? _________________________
d. Critical Thinking: Are any other sedimentary layers missing from one of the outcrops? If so, explain why.

Figure 3.20 – Two stratigraphic columns to be used in Exercise 3.5.

Fossil Correlation
Fossils present in sedimentary rocks can also be used for correlation. This usually involves a fossil assem-
blage, which is just the group of fossils found in a rock layer. By comparing fossil assemblages from one
rock outcrop to another, geologists can determine how the outcrops relate to each other in age. This
CHAPTER 3: GEOLOGIC TIME | 105

can tell geologists that layers of rock may have been deposited simultaneously or can be used to identify
unconformities.

Certain fossils, called index fossils, can be indica-


tive of a narrow span of geologic time. These fos-
sils are easily recognizable, abundant, existed for
a short period of time, and had a wide geographic
distribution. When you see an index fossil in a
rock, it immediately gives you an idea of how old
the rock is. For example, Belemnites (Figure 3.21),
a type of cephalopod (squid), only existed during
the Mesozoic era. So if you were to see a belem-
nite fossil in a rock, you know that rock is from the
Mesozoic (between 252 and 66 Ma). We’ll have lots
more on fossils in later chapters.
Figure 3.21 – A sedimentary rock containing multiple
belemnite fossils. These cone-shaped shells come from
an extinct marine animal similar to a modern squid.
Exercise 3.6 – Using Fossils to Correlate Image credit: Wikimedia user PierreSelim, CC BY-SA.
Rocks and Interpret Age

Fossils are very useful for geologic dating and correlation because the type of sediment deposited at a specific
time in a region can vary. Sand particles (sandstone) can be deposited in one area, while clay particles (shale)
can be deposited in another at the same time. This means that two different sedimentary rocks will have the
same age. In this case, correlating the rock units is more difficult because geologists can’t tell they are the
same age. This is where fossils can help because they can confirm the sedimentary rocks were deposited at
the same time, called biostratigraphic correlation. If both rock layers were deposited at the same time, then
they should contain the same fossils.

a. Correlate the sedimentary layers in Figure 3.22 based on the fossils they contain.
b. Label where any unconformities could be interpreted.
106 | CHAPTER 3: GEOLOGIC TIME

Figure 3.22 – Two stratigraphic columns for Exercise 3.6.

c. You can also use the assemblage of fossils in rocks to correlate sedimentary layers and determine age.
Correlate the rock layers in Figure 3.23 based on the groups of fossils that are found.
d. Label where any unconformities could be interpreted.
CHAPTER 3: GEOLOGIC TIME | 107

Figure 3.23 – Image for Exercise 3.6.

e. Suppose the fossils have age ranges as shown in Figure 3.24. Label the geologic periods for each layer
of sedimentary rock in both columns of Figure 3.23.
108 | CHAPTER 3: GEOLOGIC TIME

Figure 3.24 – Geologic time scale from Cambrian to Triassic that shows fossil age ranges for
Exercise 3.6. The age span for each type of fossil is shown as a black bar above the sketch of
the fossil.

f. Which of the fossils is the least useful for dating? __________________________


g. Which of the fossils is the most useful for dating? __________________________

*Note the fossil age ranges in this exercise are fictional and used for educational purposes only.

3.4 Radiometric Dating


Geologists can determine numerical ages for the formation of rocks, minerals, and some fossils using
radioactive isotope systems. Does radioactivity sound scary? It can be, but what we’re talking about is
happening on such a small scale that you don’t have to worry. The next few paragraphs are a bit techni-
cal, but we will try to break it down for you.

An isotope is an atom of an element with a different number of neutrons than protons in its nucleus. For
example, a typical carbon atom on the periodic table of elements should have six protons and six neu-
trons in its nucleus, called carbon-12. Some atoms of carbon can have an extra neutron in the nucleus,
CHAPTER 3: GEOLOGIC TIME | 109

called carbon-13, and some can have two extra neutrons in the nucleus, called carbon-14. If an atom has
too many neutrons in its nucleus, it can become unstable because the nucleus has too much energy; this
is called a radioactive isotope. Carbon-14 is a radioactive isotope. Just as the periodic table of elements
summarizes the most important information of every element, there are periodic tables of isotopes that
include all the possible isotopes for each element and their relative abundance and properties.

Radioactive isotopes decay to more stable atoms by way of alpha, beta, and/or gamma decay. We don’t
need the particulars of these decay processes, but in short, radioactive atoms can gain or lose pro-
tons, neutrons, and electrons or release energy in other forms to become stable. Carbon-14 radioactively
decays to nitrogen-14, a stable product. This ThoughtCo article gives a very simple explanation of why
atoms become radioactive if you want more information.

Radioactive isotopes don’t immediately decay Table 3.5 – Common isotope systems used for radiometric
to a stable form because it’s a spontaneous dating and their half-lives
process, so it’s impossible to predict when Parent Stable Daughter
and which atom will decay. It’s like corn ker- Half-life
Isotope Product
nels popping in a popcorn maker, you don’t
Uranium-238 Lead-206 4.5 billion years
know when a kernel will pop, but you know it
will happen. Scientists have measured the Uranium-235 Lead-207 704 million years
average rate at which radioactive isotopes
Thorium-232 Lead-208 14.0 billion years
decay and describe this rate with the term
half-life. Half-life is the time it takes for half of Rubidium-87 Strontium-86 48.8 billion years
the unstable atoms (parents) to decay to sta- Potassium-40 Argon-40 1.25 billion years
ble forms (daughters). The length of time for
a half-life is different for each isotope system. Samarium-147 Neodymium-143 106 billion years
For example, the half-life for carbon-14 is Carbon-14 Nitrogen-14 5,730 years
5,730 years, but the half-life for potassium-40
is 1.25 byrs. In a sample with 1 million atoms
of carbon-14, 500,000 of these atoms are expected to decay over the course of 5,730 years. Table 3.5
contains common isotope systems and their half-lives.

Using radioactive isotope systems, geologists can determine quantitative, numerical ages for geological
materials and such as rocks, minerals, and fossils. To calculate the age of a geologic material, geologists
analyze the material’s chemistry and figure out the ratio between parent and daughter atoms. The ratio
tells geologists how many half-lives, or what fraction of a half-life, has passed. If you know the ratio
between parent and daughter atoms and the rate of decay for that isotope system, you can calculate how
old geologic material is.

Geologists use a few assumptions with radiometric dating. One assumption is that when the rock or min-
eral first formed, there were no daughter atoms present. This means all of the daughter atoms found
in the rock had to come from radioactive decay. The second assumption is that the rock or mineral was
a closed system. A closed system means there was no addition or removal of elements at any point
because that could alter the parent-daughter ratio and therefore affect the age you calculate. With that
being said, geologists know that these assumptions may be incorrect for some of the material they work
with, and they have methods that can help correct these issues. But that is beyond our scope. That’s it for
the technical stuff.

Radiometric dating tells you different things depending on which geologic material you’re studying. For
minerals in igneous rocks, the radiometric age tells you when the mineral crystallized from magma. In
metamorphic rocks, the radiometric age of some minerals can give you the age metamorphism took
place, but some isotope systems are not affected by metamorphism and can still tell you the age the min-
erals originally crystallized in the protolith. So, you can determine when the protolith formed and when
it became metamorphosed. For minerals in sedimentary rocks, the radiometric age only tells you when
that mineral crystallized or metamorphosed in its original rock; it doesn’t tell you when the sedimentary
rock formed. The only way to determine the age of sedimentary rocks is to combine the principles of rel-
110 | CHAPTER 3: GEOLOGIC TIME

ative dating with radiometric dating of igneous or metamorphic events that cross-cut the sedimentary
rocks, such as a dike. Ages of sedimentary rocks can also be determined from fossils. Note: It is possible
to radiometrically some sedimentary rocks younger than one million years old, but there are no useful
isotopes that can be used on older sedimentary rocks.

Exercise 3.7 – Calculating Radiometric Ages

To determine the age of a mineral or rock, geologists need to know a few things. First, they need to know how
long a half-life is for the isotope system being used. Second, they need to know the ratio of parent to daughter
isotopes in the mineral, which is measured using scientific instruments like a mass spectrometer.

The standard radiometric age equation is: t=ln(1+(D/P))÷λ

• t = time (age of the material)


• ln = natural log (a function on your calculator)
• P = parent ratio (this is measured)
• D = daughter ratio (this is measured)
• λ = lambda is the decay constant (different for each isotope system, a function of the half-life)

Let’s get some practice by calculating the ages below:

a. Complete the following calculations to determine the ages of minerals in Table 3.6 (Hint: most cell
phones have additional calculator functions when you turn it horizontal. There are also free scientific
calculator apps in the Apple and Google stores.)

Table 3.6 – Worksheet for Exercise 3.7

Decay
Isotope System Parent : Daughter Ratio Age
Constant
-4
Carbon-14 1.21×10 1 : 1.3
-10
Potassium-40 5.34×10 1 : 0.25
-10
Uranium-235 9.72×10 1 : 2.34

b. Look back to Figure 3.16. Let’s say a geologist collected a sample of G (gneiss) and MD (dike) to radiomet-
rically date, with the ultimate goal of trying to constrain ages in this region. She used the potassium-40
isotope system to date metamorphism of hornblende grains and measured a parent-daughter ratio of
1:0.89. She used the uranium-235 isotope system on zircon grains for the igneous dike and measured a
parent-daughter ratio of 1:0.28.

i. How old is the gneiss? ____________________


ii. How old is the igneous dike? ____________________
iii. What is age range of the fault PF? ____________________
iv. What is the age range of sedimentary layers S and D? ____________________

3.5 Other Uses for Geologic Dating


Dating in geology has many uses besides determining when a rock or mineral formed. One common use
for dating is determining the source rock for sediment, which is called provenance. Interpreting sediment
CHAPTER 3: GEOLOGIC TIME | 111

provenance can tell a geologist a lot about the history of an area, including where the sediment came
from, where erosion was taking place, and how sediment was transported. Dating of certain sediment
grains can also tell geologists when rocks were uplifting during an orogeny. Geologic dating principles can
even be used on other planets.

Exercise 3.8 – Using Geologic Dating to Interpret Sediment Provenance

Even though radiometric dating on mineral grains in sedimentary rocks doesn’t give you the age the sedimen-
tary rock formed, it still tells you very useful information about the area’s geologic history. One such use is
determining the sediment provenance, which is the source rock of the sediment. Remember that the age you
determine from sediment grains is the age they formed in their original source rock.

Let’s say you collect some biotite grains from the bottom of the lake in Figure 3.25, and you want to figure out
where they came from. There are two potential sources for those grains: a set of metamorphic rocks to the
west and a set of granites to the south. Little is known about the geology in this area, so you need to figure
out the ages of the potential source rocks and determine where the lake sediment came from.

Figure 3.25 – Google Earth image for Exercise 3.8. The yellow star is the
location where the sediment was collected. The two potential sources are
labeled.

a. You collect five biotite grains from Source 1, five biotite grains from Source 2, and measure them using
the potassium-40 system. What are the ages of these source-rock biotite grains? Fill this out in Table 3.7.
112 | CHAPTER 3: GEOLOGIC TIME

Table 3.7 – Worksheet for Exercise 3.8a

Samples from Source 1 Parent : Daughter Age

1a 1 : 0.808

1b 2 : 1.706

1c 1 : 0.825

1d 3 : 2.349

1e 2 : 1.596

Samples from Source 2 Parent : Daughter Age

2a 2 : 0.626

2b 4 : 1.236

2c 1 : 0.323

2d 3 : 0.903

2e 2 : 0.596

b. What is the average age of the rocks at Source 1? _______________


c. What is the average age of the rocks at Source 2? _______________
d. The lake sediment grains you collected have the following ages. Determine which source area they most
likely came from by filling out Table 3.8.

Table 3.8 – Worksheet for Exercise 3.8d

Lake Sediment Age Source Area

1 523 Ma

2 1,157 Ma

3 517 Ma

4 521 Ma

5 486 Ma

6 1,116 Ma

7 513 Ma

8 501 Ma

9 523 Ma

10 494 Ma

e. What percentage of the lake sediment is coming from Source 1? _______________


f. What percentage of the lake sediment is coming from Source 2? _______________
g. Critical Thinking: What could be the reason(s) why you see this sediment distribution in the lake?
CHAPTER 3: GEOLOGIC TIME | 113

Exercise 3.9 – Using Principles of Relative Dating on Other Planets

Cross-cutting relationships is a universal principle. That means it not only applies to our planet but applies to
other planets. Let’s take a look at some pictures taken from the surface of Mars; these show features such as
craters, rivers, and fractures (a type of fault). These types of images are used to decipher the tectonic, volcanic,
and impact history.

a. Look at Figure 3.26.

i. Which is older, Crater B or Crater S? ____________________


ii. Explain why.

iii. Which is older, Fracture F or Crater B? ____________________


iv. Which is older, Fracture F or Crater S? ____________________
v. Critical Thinking: If you wanted to land a rover to investigate Crater B, where would it be safe to
land? Consider whether or not you think Fracture F is an active fault.
114 | CHAPTER 3: GEOLOGIC TIME

Figure 3.26 – Image for Exercise 3.9a. Martian craters near Cerberus Fossae in the Elysium
Planitia region close to the Martian equator. Photo taken January 2018. Image credit:
Adapted from ESA/DLR/FU Berlin, CC BY-SA IGO.

b. Look at Figure 3.27. Put the craters in relative order from oldest to youngest.
______________________________
CHAPTER 3: GEOLOGIC TIME | 115

Figure 3.27 – Image of Hadley crater for Exercise 3.9b. This is a composite of several
images taken during revolution 10572 on 9 April 2012 by ESA’s Mars Express
centered around 19°S and 157°E; the image has a ground resolution of about 19 m
per pixel. The image shows the main 120 km wide crater, with subsequent impacts
within it. Evidence of these subsequent impacts occurring over large time scales is
shown by some of the craters being buried. This image shows one of the
characteristics of Martian craters as they often have fluidized ejecta that can be
seen both in the bottom right and top left craters, the latter crater reaching a depth
of around 2600 m. Image credit: Adapted from ESA/DLR/FU Berlin, CC BY-SA IGO.

c. Look at Figure 3.28.

i. Which is older, Crater C1 or River R1? ____________________


ii. Which is older, Crater C2 or River R2? ____________________
116 | CHAPTER 3: GEOLOGIC TIME

Figure 3.28 – Image from the Arda Valles region of the Martian highlands for Exercise 3.9c. The region was
imaged by Mars Express on 20 July 2015 during orbit 14649. The image is centered on 19°S / 327°E; the
ground resolution is about 14 m per pixel. Image credit: Adapted from ESA/DLR/FU Berlin, CC BY-SA IGO.

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson, Carlos Andrade

Google Earth Locations


CHAPTER 4: SEDIMENTARY STRUCTURES

Learning Objectives

The goals of this chapter are to:

• Identify types of sedimentary structures


• Explain how sedimentary structures form
• Interpret paleoenvironments using sedimentary structures

4.1 Introduction
Sedimentary structures are features that form in sediment as it is being deposited. These structures are
typically an indication of what the sedimentary environment was like. Sedimentary structures can often
be identified by observable patterns in the sedimentary bedding or distinct shapes within the sediment.
Basically, if the sedimentary rock doesn’t look uniform or has a distinctive feature, there’s a good chance
it’s a sedimentary structure.

Exercise 4.1 – Observing Sedimentary Structures

Part A
For this exercise, your instructor will provide you with a set of sedimentary structures. Look at the samples
closely and use a hand lens if necessary. Create a sketch of each sedimentary structure in the blank space
below, focusing on what you think are the most important characteristics of the sample. Then, in full sen-
tences, describe the structure you just sketched. If it helps, pretend you are describing what the structure
looks like to someone who can not see it.
118 | CHAPTER 4: SEDIMENTARY STRUCTURES

Structure #1 Sketch:

Description:

Structure #2 Sketch:

Description:

Structure #3 Sketch:

Description:
CHAPTER 4: SEDIMENTARY STRUCTURES | 119

Structure #4 Sketch:

Description:

Structure #5 Sketch:

Description:

Part B
Now put your observational and descriptive skills to the test. Read your descriptions to your classmates and see if they can
identify which samples you are talking about.

4.2 Types of Sedimentary Structures


The simplest sedimentary structure is stratification, which is layering that can be observed in sedimen-
tary rocks (Figure 4.1). Layers of sediment that are thicker than 1 cm are called beds and layers thinner
than 1 cm are called laminations. Laminations are typically composed of fine-grained silt and clay-sized
sediment. Structures can be more complex like the wavy pattern seen in ripple marks (Figure 4.2) or
chaotic looking patterns in cross-bedding (Figure 4.3).
120 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.1 – Titled sedimentary beds from Morro Solar near Lima, Peru. Thicker layers
are most likely sandstone, and thinner layers are shale. Layering tilts down to the
right. Image credit: Miguel Vera León, CC BY.
CHAPTER 4: SEDIMENTARY STRUCTURES | 121

Figure 4.2 – Ripples caused by waves. This is a view from the top surface and not the side. This rock is
Permian in age from Nomgon, Mongolia. Image credit: Matt Affolter, CC BY-SA.
122 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.3 – Cross-bedding in sandstone seen on a cliff face in Zion National Park, Utah. You can figure out
the scale of the image by looking for trees or other vegetation. Image credit: NOAA, Public Domain.

Sedimentary structures provide a lot of information about the environment in which they formed, includ-
ing processes that were occurring when sediment was deposited, the environment of deposition, the
direction sediment was traveling, and/or the mechanism for transporting the sediment (wind, water, or
ice). Some sedimentary structures also help you determine which side of the rock was originally facing
upwards, called way-up indicators. When outcrops have overturned rocks (rocks that have been tilted
so far they are upside down), sedimentary structures can be used to tell which way was originally facing
up.

Each structure tells a story that geologists use to interpret Earth’s history. For this chapter, only a few
of them are discussed: dunes and ripple marks, cross-bedding, graded bedding, mudcracks, raindrop
imprints, sole marks, and trace fossils and bioturbation.

Dunes and Ripple Marks


As water or wind moves across sediment, it can shape the grains into wavy patterns called dunes (>10 cm)
and ripples (<10 cm). Symmetrical ripple marks, like those seen in Figures 4.2 and 4.4, are formed by
the back-and-forth flow of water over sediment. These types of ripples are formed in the shallow marine
environment where the back-and-forth motion of waves, or even tides, shape the sediment at the bottom
of the ocean. These ripples have symmetrical limbs, meaning that both sides of the ripple dip at about
the same angle. This video on symmetrical ripples can help you see how this process works. Symmetrical
ripple marks are most commonly found in sandstones from shallow marine environments.
CHAPTER 4: SEDIMENTARY STRUCTURES | 123

Figure 4.4 – Examples of symmetrical ripple marks with limbs at the same angle on either side of the structure. A)
Modern symmetrical ripples from the Bahamas. B, C, and D) Symmetrical ripples in Devonian-Missippian age
sandstone from Ohio, USA. These are all views from the top. Image credits: James St. John, CC BY.

Water moving in one direction, like a river, can produce asymmetrical ripple marks. The limbs on these
ripples are not equal, with one side that is more shallow and one side that is steeper. These types of rip-
ple marks can tell you which direction the river was flowing because sediment moves up the shallow side
of the ripple and gets deposited on the steep side (Figures 4.5 and 4.6). The deposition on the steep side
of the ripple allows the ripple to move in the same direction that water is flowing, as shown in this video.
Wind can also create asymmetrical ripple marks at different scales. Ripple marks at smaller scales can
usually be found along a beach. Large-scale ripple marks are called dunes and are common in deserts
and some coastal environments.
124 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.5 – How asymmetrical ripples form. The red circles represent grains of sediment that move up a
shallow side and fall down a steep side to form cross-bedding. Image credit: Wikimedia user Nwhit, CC BY-SA.
CHAPTER 4: SEDIMENTARY STRUCTURES | 125

Figure 4.6 –
Examples of
asymmetrical ripple
marks. The limbs of
the ripples dip at
different angles, one
shallow and one
steep. Three views
(A, C, and D) are
from the top, and B
is from the side. A)
Modern
asymmetrical ripples
from the Bahamas.
B and C)
Asymmetrical
ripples in sandstone
from Colorado, USA.
D) Precambrian
asymmetrical ripples
in quartzite from
Wisconsin. Image
credits: James St.
John, CC BY.

3D Image 4.1 – Asymmetrical Ripples


126 | CHAPTER 4: SEDIMENTARY STRUCTURES

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://uhlibraries.pressbooks.pub/historicalgeologylab/?p=78

Sedimentary Structures – Ripple Marks 2 by nate_siddle on Sketchfab

Cross-Bedding
The top layer of a ripple or dune is not always preserved in the rock record, so it is rare to find ripples like
those seen in Figures 4.3 and 4.7. Dunes and ripples are constantly moving. As one passes and deposits
its sediment, another follows right behind it to deposit more sediment on top. Geologists typically find the
deposited sediment from the steep side of a series of ripples or dunes in the rock record. The deposition
of the steep side of several dunes or ripples creates a sedimentary structure called cross-bedding (Figure
5). One of the most important pieces of information that cross-bedding gives geologists is the direction
that wind or water was moving. The steep side of a ripple always angles downward toward the direction
the water or wind was moving, as shown by the blue lines in Figure 5b. There are many different types of
cross-bedding, and each form in a similar way.
CHAPTER 4: SEDIMENTARY STRUCTURES | 127

Figure 4.7 – Cross-bedding from ancient sand dunes in Coyote Gulch, part of the Canyons of the Escalante, Utah.
The upper image is uninterpreted; the lower image shows interpretations of four dunes as yellow dashed lines and
the cross-beds in blue. Image credit: G. Thomas, Public Domain.

Sedimentary structures are not limited to Earth since similar features have been found on Mars, Venus,
and Titan, Saturn’s largest moon. Figure 4.8 shows cross-bedding from Mars, and it looks very similar to
the wind-blown sand outcrops commonly found in the southwestern U.S. (see Figure 4.7). Do you think
the scale is similar between these two images? The size of the cross-bedding can help to determine if
these formed in water or air (aeolian). Smaller ripples form in water, while larger ones form in terrestrial
dunes.
128 | CHAPTER 4: SEDIMENTARY STRUCTURES

Graded Bedding
Graded bedding is a common sedimentary struc-
ture where a change in grain size can be observed
within a single sedimentary bed (Figure 4.9). At the
bottom of the bed are mainly coarse particles
which get progressively smaller as you move verti- Figure 4.8 – An outcrop of cross-bedded sandstone on
the lower slope of Mars’ Mount Sharp. The sediment
cally up the bed. Graded beds generally represent transport direction is interpreted as sediments carried by
depositional environments in which transport currents moving down the deltas and into deeper lake
energy decreases over time, like the changing water. This photograph was taken by NASA’s Mars Rover
water velocity in a river. However, these beds can Curiosity on August 27, 2015, using its mast camera. This
also form during rapid depositional events, most area is now known as Whale Rock in the Pahrump Hills
commonly from turbidity currents. Turbidity cur- and far from where Curiosity found evidence of delta
deposits where a stream entered a lake. Image credit:
rents are essentially underwater avalanches of
NASA/JPL-Caltech/MSSS.
sediment that move downslope, usually starting at
the edge of the continental shelf and flowing down
the continental slope. The sediment deposited from a turbidity current is called a turbidite, which often
has graded bedding with the coarsest particles at the bottom of the bed and the smallest at the top.

Figure 4.9 – A sketch and an example of graded bedding. The left side of the figure is a sketch of graded bedding
showing larger grains at the bottom and getting finer towards the top. The right side of the figure is a sample of
graded bedding. Image credits: Sketch from Mike Clark, CC BY-SA; example from James St. John, CC BY.

Mudcracks
Mudcracks, also called desiccation cracks, form when wet sediment, typically clay-rich, dries out (Figure
4.10). Clay minerals expand when they get wet and shrink when they dry out. As the sediment shrinks,
cracks can develop, which form polygons on the surface of the mud. Today, you can find plenty of modern
mudcracks along the margins of rivers or in desert valleys that periodically get inundated with floods.
After a mudcrack forms, it can be filled in with new sediment.

Mudcracks are typically wider at the top of the crack and get progressively smaller toward the bottom of
the crack. Because of this pattern, mudcracks can be a good way-up indicator if you can see a cross-sec-
tion view of the crack.
CHAPTER 4: SEDIMENTARY STRUCTURES | 129

Figure 4.10 – A) A modern example of large mudcracks in a dried-up river bed in the Rio San Juan, Argentina. B) An
ancient example of mudcracks with sediment filling in the cracks from Maryland. Image credits: A) Daniel
Hauptvogel, CC BY-NC-SA; B) James St. John, CC BY.

Mars also has mudcracks (Figure 4.11), one of the pieces of evidence that indicates the red planet used
to have liquid water on its surface. These were found in Gale crater in an exposure of Murray Forma-
tion mudstone on lower Mount Sharp. The white material in the cracks may be a form of calcium sulfate,
either anhydrite or gypsum. This is a guess since the Curiosity rover cannot test the mineral hardness.
130 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.11 – This photograph of mudcracks was taken by NASA’s Mars Rover Curiosity on December 31, 2016. The
view spans about 4 feet (1.2 meters) left-to-right and combines three images taken by the Mars Hand Lens Imager
(MAHLI) camera. The cracks may have formed more than 3 billion years ago. Image credit: NASA/JPL-Caltech/MSSS.

Raindrop Impressions
Raindrop impressions are small, concave imprints made by rain when it falls on soft sediment (Figure
4.12). The impressions or small craters are made from the force of raindrops falling onto the sediment,
which makes these structures good way-up indicators. If you were to see only the bottom of the impres-
sion, it would look like a raised bump (convex).

Raindrop impressions tend to be found in fine-grained rocks like siltstones and shale but not in coarser-
grained sandstones. The impressions likely represent the end of a rainstorm as rain is letting up because
any previously formed impressions would be destroyed by subsequent rainfall. That’s why most raindrop
impressions are very scattered rather than occurring all over the surface. Then the impressions need to
be filled in with sediment before the next rainstorm to be preserved.
CHAPTER 4: SEDIMENTARY STRUCTURES | 131

Sole Marks
Sole marks is a broad term that describes several
different sedimentary structures that appear as
impressions or grooves in sediment, including
flute casts, tool marks, groove casts, and load Figure 4.12 – A) Overhead view of modern raindrop
casts. Typically the cast of the marking (the raised impressions (and mudcracks) from Argentina. B and C)
Ancient raindrop impressions from the Lower Permian,
bump) is preserved at the bottom of a sedimen- Upper Pecos Valley, New Mexico. B is an overhead view, C
tary bed, hence the term “sole” mark, and the is the bottom-up view, showing the convex underside of
mold side (the impression) is filled with sediment. raindrop impressions (not the same sample as B). Image
This makes sole marks good way-up indicators credits: A) Daniel Hauptvogel, CC BY-NC-SA; B and C)
since the cast side is facing down. James St. John, CC BY.

Flute casts are common structures created by turbidity currents (Figure 4.13). The movement of these
sediment avalanches underwater can scour the ocean floor, creating an elongated impression. Flute casts
are usually closely spaced and can be stacked on top of one another. Not only can they tell you which
way is up, but they can also tell you which way the current was flowing. The tapered end of the flute cast
points in the direction of flow.

Figure 4.13 – Flute casts from the central Alps, Switzerland. The view is from the
underside of the rock. Image credit: Chris Spencer, CC BY-NC-ND.

Tool marks are made when an object, such as a stick, is dragged across sediment by a current and leaves
behind what looks like scratches in the soft sediment (Figure 4.14). The elongated scratches can be used
as an indicator of the paleocurrent.
132 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.14 – Tool marks from Banff National Park. The view is
from the underside of the rock. Image credit: Callan Bentley,
used with permission.

Groove casts are raised parallel ridges (Figure 4.15). They are spaced closely together, often appearing in
sets of 2 and 3, but do not occur on top of one another like flute casts. Interpreting the paleocurrent from
groove casts can be difficult because the marking is often symmetrical. Without the addition of other
paleocurrent evidence, you may only be able to narrow down the paleocurrent to two directions that are
180° apart.
CHAPTER 4: SEDIMENTARY STRUCTURES | 133

Figure 4.15 – Groove casts. The yellow arrow indicates the direction of transport. This
arrow is double-ended as there is no indication if the water was flowing up or down the
river before it was tilted and exposed. The view is from the underside of the rock. Image
credit: Brian Ricketts, CC BY.

Load casts form when dense, sandy sediment is deposited on less dense, water-saturated sediment, usu-
ally silt or clay (Figure 4.16). The dense sand load pushes into the soft layer below, creating bulb-like
impressions.

Figure 4.16 – Load casts in arkose sandstone from the Aquitaine Basin near Nontron,
France. The view is from the underside of the rock. Image credit: Rudolf Pohl, CC BY-SA.
134 | CHAPTER 4: SEDIMENTARY STRUCTURES

Exercise 4.2 – Identifying Sedimentary Structures

Your instructor will pass around examples of various sedimentary structures. Now that you know what to look
for in these structures create a detailed sketch of each one. It may help to sketch the structures from several
angles. Identify the sedimentary structures and complete any of the relevant information about them.

Within the sketch area for each structure, provide answers to the following questions:

• What type of sedimentary rock is this sedimentary structure in?


• If your structures can provide the paleocurrent, indicate the direction on your sketch.
• If your structures are way-up indicators, indicate which way is up on your sketch.

Structure #1 Sketch:

Structure #2 Sketch:

Structure #3 Sketch:

Structure #4 Sketch:

Structure #5 Sketch:
CHAPTER 4: SEDIMENTARY STRUCTURES | 135

Structure #6 Sketch:

Structure #7 Sketch:

4.3 Sedimentary Structures and Paleoenvironments


As you may have guessed, sedimentary structures are handy for determining what paleoenvironments
were like. By combining sedimentary structures and the surrounding geology, a geologist could describe
a pretty accurate picture of the environment when these sediments were deposited.

Exercise 4.3 – Summarizing Sedimentary Structures and Paleoenvironments

a. Using the information you have learned about sedimentary structures, complete Table 4.1.
136 | CHAPTER 4: SEDIMENTARY STRUCTURES

Table 4.1 – Worksheet for Exercise 4.3

Type of
Structure Rock it Environment Description
Forms in

Symmetrical
Ripples

Asymmetrical
Ripples

Cross-bedding

Graded Bedding

Mudcracks

Raindrop
Imprints

Sole Marks

b. Below are descriptions of two different environments. Which one is a braided river and which is a
swamp/wetlands?

i. Mostly sand deposits with some sediment beds having a layer of gravel on the bottom and cross-
bedding. ____________________
ii. Mostly silt and clay deposits with root structure
structuress, lots of bioturbation, and many coal layers.
____________________

c. Critical Thinking: The global carbon cycle includes the storage of carbon in sedimentary rocks such as
limestone or disseminated organic matter (kerogen) in mudrocks. Which sedimentary structures are
CHAPTER 4: SEDIMENTARY STRUCTURES | 137

associated with the accumulation of carbon in clastic rocks?

d. Critical Thinking: Can you name two factors that might explain enhanced carbon storage?

Exercise 4.4 – Interpreting Paleoenvironments using Sedimentary Structures and


Sedimentary Rocks

Geologists don’t only focus on a single rock outcrop to interpret the paleoenvironment of a region; they look
at many outcrops so they can see how an environment changes across a region. Figure 4.17 below is a map
of an area where sedimentary rocks and structures were described by a geologist. It is divided up into 4 zones
with the following descriptions of the rocks and sedimentary structures:

A. Sandstone with large-scale cross-bedding and very well-rounded sand grains.


B. Sandstone and mudstone with wavy bedding toward the east and mudcracks toward the west.
C. Fine sandstone with symmetrical ripple marks.
D. Shale with lots of plankton fossils and fine laminations.
138 | CHAPTER 4: SEDIMENTARY STRUCTURES

Figure 4.17 – Map for Exercise 4.4.

Answer the following questions:

a. Come up with a brief depositional environmental interpretation for each zone.

i. Zone A:

ii. Zone B:

iii. Zone C:

iv. Zone D:
CHAPTER 4: SEDIMENTARY STRUCTURES | 139

b. Critical Thinking: In what direction was the ocean in this paleoenvironment? ____________________
c. Critical Thinking: In what type of tectonic environment would you find this sequence of sedimentary
rocks. Explain your answer.

4.4 Trace Fossils and Bioturbation


Organisms burrow and move through sediment on the ocean floor and the bottoms of lakes and rivers;
this is called bioturbation. When organisms disturb the sediment by burrowing, their burrows can be pre-
served when the sediment hardens into rock. In most cases, the burrows will fill with new sediment, but
the outline is preserved. It is difficult to assign a specific organism to the creation of a single burrow.
Instead, geologists look at different burrows that tend to occur together in the rock record to classify
them, a branch of study known as ichnofacies. Each ichnofacies is named after the most common trace
fossil in the facies. Determining which ichnofacies the trace fossil fits into can tell you about the environ-
ment in which the organism lived, including water depth, salinity, energy, and turbidity, and what the sub-
strate was like. Generally, vertical burrows were formed in shallow water environments while horizontal
burrows in deeper water environments. Table 4.2 contains a list of ichnofacies, and Figures 4.18-4.25 are
images of them.
140 | CHAPTER 4: SEDIMENTARY STRUCTURES

Table 4.2 – Common ichnofacies

Ichnofacies Substrate Paleoenvironment Description Image*

Horizontal,
Sandstones, curved, and
may be Terrestrial, rope-like burrows.
Scoyenia associated freshwater, low Unbranching, but
with red energy can cross each
beds other. Occasional
vertical burrows.

Figure 4.18 – Scoyenia burrows from the Grand


Canyon. Image credit: National Park Service,
Public Domain.

Vertical burrows
Coastal, barrier
Variable with J, Y, or U
islands, deltas,
Psilonichnus grain size, shapes. Can also
estuaries, lagoons,
sand have vertebrate
bays
footprints.

Figure 4.19 – A single Psilonichnus burrow from


Holocene beach dunes, San Salvador Island,
Bahamas. Cross-bedding is also present. Image
credit: James St. John, CC BY.
CHAPTER 4: SEDIMENTARY STRUCTURES | 141

Ichnofacies Substrate Paleoenvironment Description Image*

Closely-spaced
Hard rock,
Coastal cliffs, reefs, straight, or slightly
Trypanites carbonate,
beachrock curved, vertical
shells
burrows.

Figure 4.20 – Trypanites burrows in Ordovician


limestone from Kentucky. Image credit: James St.
John, CC BY.

Three-dimensional
network of
Firm but Shallow marine, cylindrical
Glossifungites not lithified marginal marine, burrows and
sediment delta, estuary individual, vertical,
teardrop-shaped
burrows.

Figure 4.21 – Glossifungites burrows in


sandstone from Lima, Peru. Image credit: Miguel
Vera León, CC BY.
142 | CHAPTER 4: SEDIMENTARY STRUCTURES

Ichnofacies Substrate Paleoenvironment Description Image*

Straight, vertical,
burrows that do
Beaches, tidal flats,
not branch or
Skolithos Sand shallow marine,
cross. Can be
above wave base
slightly angled and
J-shaped.

Figure 4.22 – Skolithos burrows in sandstone


from western Maryland. Image credit: James St.
John, CC BY.

Mid to distal Horizontal and


continental shelf, vertical burrows
Sand and
Cruziana below wave base, from a wide
silt
but above storm variety of
wave base organisms.

Figure 4.23 – Cruziana burrows from Portugal.


Image credit: Wikimedia user CorreiaPM, Public
Domain.
CHAPTER 4: SEDIMENTARY STRUCTURES | 143

Ichnofacies Substrate Paleoenvironment Description Image*

A series of
Deep water, base
horizontal
Sand and of continental shelf,
Zoophycos burrows curving
silt may be associated
away from a
with turbidites
central point.

Figure 4.24 – Zoophycos burrows from the Swiss


Alps. Image credit: Chris Spencer, CC BY-NC-ND.

Meandering and
Deep water, open spiraling
Nereites Silt and clay
ocean horizontal trails or
burrows.

Figure 4.25 – Nereites trace fossils. Image credit:


Wikimedia user Richdebtomdom, CC BY-SA.

*No OER images are available to summarize each ichnofacies, so these are single examples of an ichno-
fossil that belongs to the ichnofacies. You can find depictions of each facies here and an extensive list of
ichnofossils here. You need to select invertebrates on this webpage.

Exercise 4.5 – Determining Ichnofacies

Your instructor will provide you with some samples of ichnofacies or ichnofossils. Sketch each sample, paying
special attention to the details of each ichnofossil. Identify the ichnofacies of each sample and the type of sed-
imentary rock. Based on those identifications, give a brief description of the environment.

Trace Fossil #1 Sketch:


144 | CHAPTER 4: SEDIMENTARY STRUCTURES

Trace Fossil #2 Sketch:

Trace Fossil #3 Sketch:

Trace Fossil #4 Sketch:

Trace Fossil #5 Sketch:

For one of these ichnofacies samples, answer the following questions:

a. What is the grain size in the surrounding rock compared to the trace fossil?

b. Do you think this trace fossil could be preserved in coarse-grained sediment? Explain.

c. What was the water depth where this fossil lived?


CHAPTER 4: SEDIMENTARY STRUCTURES | 145

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson, Carlos Andrade

Google Earth Locations


CHAPTER 5: STRATIGRAPHY

Learning Objectives

The goals of this chapter are to:

• Evaluate the origin, composition, distribution, and succession of strata to determine past geologic events
related to sedimentary environments and tectonic settings
• Apply Walther’s Law to marine transgression and regression
• Reconstruct sediment thickness to understand processes of deposition
• Create stratigraphic columns and correlations using multiple techniques

5.1 Introduction
Stratigraphy is the area of geology that deals with sedimentary rocks and layers and how they relate to
geologic time; it is a significant part of historical geology. As you learned in Chapters 2 and 4, one of the
primary goals of studying sedimentary rocks is to determine their depositional environment; stratigraphy
is no different.

Stratigraphy is mainly studied through outcrop observations, the collection of sediment cores, and seis-
mic surveys. Sediment cores are mostly collected from the ocean floor by organizations like the Inter-
national Ocean Discovery Program (IODP) using a ship dedicated to drilling (Figure 5.1a). Geologists can
then study the collected sediment (Figure 5.1b) or send instruments into the drill hole to measure the
geologic properties of the surrounding sediment. The cores are then stored at facilities around the world
for scientists to request samples from (Figure 5.1c). Another way to study stratigraphy is to conduct seis-
mic surveys by sending sound waves into the seafloor or ground and monitoring how the waves reflect
back to the surface (Figure 5.2). Previous seismic surveys were only collected in a line, but now they are
typically collected in grids to get three-dimensional data below the earth’s surface. Geologists use the
patterns of reflectivity to help identify the rock types, deformation features, and where there might be
water or petroleum and other characteristics in the subsurface. If a seismic survey is repeated over the
same area, this gives another dimension of time.

Geologists subdivide stratigraphic columns into formations. A formation is a series of sedimentary beds
distinct from other beds above and below and is thick enough to be shown on geological maps. Some-
times several formations are lumped together to form a group. If you want to propose a new formation
or group, there are strict guidelines set up by the International Commission of Stratigraphy.
CHAPTER 5: STRATIGRAPHY | 147

Figure 5.1 – A) The JOIDES Resolution, one of the research vessels used by the IODP. B) Scientists studying
recovered core during Expedition 378 in the South Pacific. C) An IODP core repository at the University of Bremen
in Germany; others are located at Texas A&M University in Texas, Rutgers University in New Jersey, and Kochi
University in Japan. Image credit: A) Thomas Ronge and IODP, CC BY; B) Lindy Newman and IODP, CC BY; C)
MARUM, CC BY-SA.

An important tool for stratigraphy is to observe seismic characteristics below the surface. Some call this
change in seismic response a seismic facies. Just like grain size is important when describing sedimentary
rocks, so are the features in a seismic response. So, when looking at a seismic profile, make observations
and ask yourself: Do you see areas without many reflectors? This may be related to its lithology. Are the
reflectors continuous? If so, they may be well-bedded. Are they thick or thin? Also, what is their shape –
flat or curved?

If you look closely at Figure 5.2, the sedimentary reflectors are relatively flat and show layering in the
Kumano Basin (top left side of the image or NW end; highlighted in green in Figure 5.2b) as well as to
the right edge to the SE. Layering is flat and parallel to the ocean floor. This pattern implies the sediment
was deposited in relatively quiet, low-energy water. Below these regions, and in the middle of the seis-
mic profile, the layering is disrupted and non-continuous; this probably indicates that these sediments
were deformed shortly after deposition and before they were lithified as the sediment was scraped off
the downgoing oceanic crust. Towards the bottom, the seismic reflectors are very discontinuous as the
oceanic crust has no internal structure. The seismic character of the layering lets you know important
tectonic and sedimentary history along this profile.
148 | CHAPTER 5: STRATIGRAPHY

Figure 5.2 – Two seismic images of an accretionary prism near the Nankai trough offshore Japan. Uninterpreted (A)
and interpreted (B) seismic data from IODP Expeditions 314, 315, and 316 for the Nankai Trough Seismogenic Zone
Experiment (NanTroSEIZE) in 2009. The interpretations on B show several packages of marine sediments overlying
oceanic crust. The yellow line is the bottom simulating reflector (BSR). Image Credit: Moore et al., 2009 and IODP,
CC BY.

Exercise 5.1 – Sediment Coring and Stratigraphic Columns

This exercise is an analogy for how geologists recover sediment cores. Instead of sediment, though, we will
use Play-Doh.

1. Using an empty Play-Doh container (or another container), place layers of different colored Play-Doh in
the container; don’t use more than 10 layers. These layers of Play-Doh will act as our layers of sediment.
2. Take a clear, plastic straw and push it down through the layers of your Play-Doh. This represents geolo-
gists drilling into layers of sediment.
3. Pull the plastic straw out. You should see all your layers of “sediment” represented in the straw, which
geologists recover when collecting a sediment core.

Depending on the type of coring that is done, geologists will split the sediment core in half. One-half of the
core will be sampled and studied by geologists, while the other half will be archived and untouched. One of the
first things geologists do on a recovered core is describe the lithology. You are going to create a stratigraphic
column of your Play-Doh core. A stratigraphic column is a graphical representation of layers of sediment and
sedimentary rock and their characteristics. The bottom of a stratigraphic column is always the oldest rock unit
in the area.
CHAPTER 5: STRATIGRAPHY | 149

a. Measure and record the total length of your Play-Doh core in millimeters. ___________________
b. Measure and record the thickness of each layer in your Play-Doh core in Table 5.1. Start with the upper-
most layer.

Table 5.1 – Play-Doh “sediment”


layer characteristics

Layer Color Thickness (mm)

c. Now let’s visualize this data by creating a stratigraphic column using Figure 5.3. Start with your upper-
most layer, use its thickness to fill in the “Lithology” column. For example, if you measured a thickness
of 15 mm, draw a horizontal line at the 15 mm mark. Color your layer so that it matches your Play-Doh.
This will be your first layer.
d. Now add the second layer to the column. If your second layer is 10 mm, start where the first layer ended
and then measure 10 mm down from there. Finish the stratigraphic column with all of your layers.
e. If the sediment at the bottom of your column was deposited 100,000 years ago, what is the sedimenta-
tion rate of your core in mm/yr? ____________________
150 | CHAPTER 5: STRATIGRAPHY

Figure 5.3 – A blank


stratigraphic column for
Exercise 5.1.

Before you begin the rest of the exercises, you may want to review what you know about classifying sed-
imentary rocks (Chapter 2) and sedimentary structures (Chapter 4). Remember that clastic rocks are
subdivided by grain size, rounding, and sorting. Also, how rocks weather is important in these exercises
because their differences in weathering resistance make it easier to see the different stratigraphic units.
For example, a fissile shale weathers easily compared to sandstone and limestone.

5.2 Grain Size


One of the simplest analyses done on sediment is a grain size analysis. Remember, clastic sedimentary
CHAPTER 5: STRATIGRAPHY | 151

rocks are classified based on their grain size. It can tell you a lot about the transport and depositional
history of the sediment. Some geologists use grain size to help determine the geologic history of the sed-
iment; others use it for studying porosity and permeability to locate fluids (water or oil), or in the case
of engineers, how much weight the sedimentary rock can hold before becoming unstable.

Exercise 5.2 – Grain Size

You can try to visually estimate the distribution of grain size in a sample of sediment, but that can lead to
significant error if you over- or underestimate something. Visual estimates are best for field measurements,
but you need to be more accurate in the lab. Geologists use quantitative approaches with various particle size
analyzers (instruments that can accurately measure grain size and shape), but the traditional way to measure
grain size was to separate the sediment into different size ranges using sieves and measure their weights.
Computerized particle sizers are probably not available for you to use in your lab, so let’s do a grain size analy-
sis on a sample of sediment using sieves.

Note: Grain sizes smaller than 63 μm need to be wet-sieved and dried.

a. Measure the starting weight of the entire sediment sample in grams. ____________________
b. Now separate the sediment into different fractions using the set of sieves provided by your instructor.
c. Measure and record the weight of each size fraction in the first two columns of Table 5.2.
d. What weight percentage does each fraction make up? Fill in the third column of Table 5.2. Weight per-
cent is the percentage of the entire weight of the sediment that each size fraction takes up. Take the
weight of each fraction and divide it by the total weight, then multiply by 100.

Table 5.2 – Grain size analysis

Size fraction range (μm) Weight (g) Weight Percent

e. Take your weight percent data and plot it as a bar graph in Figure 5.4.
152 | CHAPTER 5: STRATIGRAPHY

Figure 5.4 – Grain size graph to use in Exercise 5.2.

f. Another way geologists use sediment grain size is to classify and describe the sediment. Do you remem-
ber the triangle (ternary) plot from Chapter 0? Geologists use ternary plots to compare the sand, silt,
and clay content of the sediment. Table 5.3 contains percentages of sand, silt, and clay. Plot these per-
centages on the triangle graph in Figure 5.5 and record the descriptive term for the sediment size in the
last column of Table 5.3.

Table 5.3 – Particle size percentages to plot on Figure 5.5

Sand % Silt % Clay % Sediment Description

55 23 22

18 27 55

40 48 12

23 7 70
CHAPTER 5: STRATIGRAPHY | 153

Figure 5.5 – Triangular plot for sedimentary rocks using proportions of sand, silt, and clay. Instructions for how to
plot on a triangular plot are given in Figure 0.6. This plot is subdivided into 10 different areas used to assign a rock
name to a sample. The abbreviations for each field are given in the figure. Image credit: Adapted from Krumbein,
W.C., and Sloss, 1963 by VB Sisson CC BY-NC-SA.

Exercise 5.3 – Interpreting Outcrop Stratigraphy

The boundaries between different geologic time periods are often easy to find in the stratigraphic record as
a change in sediment type. Sometimes these correspond to major extinction events. One of these occurs at
the end of the Triassic and beginning of the Jurassic time periods (201.3 Ma). At this time, about 25-30% of
all marine species died out. The cause for the extinction is debatable, with some geoscientists relating it to
increased CO2 in the atmosphere from volcanic eruption of a large igneous province (LIP) in the Atlantic ocean.
The Triassic in this region is represented by the Lilstock Formation, deposited in a warm tropical ocean with
abundant marine life. Above this is the Blue Lias Formation, well known for its fossil ammonites.

Below is an image created by stitching together many images with software that renders them in three dimen-
sions. Once the image is loaded, you can use your mouse to rotate, tilt, and zoom in on the figure. This image
has several numbers that you can click on, highlighting some of the features at the Triassic-Jurassic boundary
near Lyme Regis in southern England. There are also two tape measures (inches and feet) draped across the
outcrop that you can use to get a sense of the scale of this sea cliff exposure.
154 | CHAPTER 5: STRATIGRAPHY

An interactive or media element has been excluded from this version of the text. You can view it online here: https://uhli-
braries.pressbooks.pub/historicalgeologylab/?p=80

Figure 5.6 – Triassic – Jurassic Boundary Section by NHM_Imaging on Sketchfab


Sketchfab, CC BY-NC.

a. Sketch a stratigraphic column of Figure 5.6. This is easy to do because the limestones and shales weather dif-
ferently. Limestones tend to be more resistant to weathering than shales and thus make prominent layers.

Sketch Area:

b. Is there an unconformity in this stratigraphic column? Explain.

c. What is the dominant lithology in the Triassic? ____________________


d. What is the dominant lithology in the Jurassic? ____________________
e. Critical Thinking: Can you explain why the lithology changed?

5.3 Facies and Lithostratigraphy


Lithostratigraphy is a sub-discipline of stratigraphy that deals with the type of rock and depositional
environment of sediments and sedimentary rocks. One of the fundamental concepts in stratigraphy is
Walther’s Law (Johannes Walther,1860-1937), which states that depositional environments vary in both
space and time such that “the facies that occur conformably next to one another in a vertical section
CHAPTER 5: STRATIGRAPHY | 155

of rock will be the same as those found in laterally adjacent depositional environments” (Walther, 1894).
That’s a pretty dense statement, so let’s break it down. First, let’s tackle two definitions; facies and con-
formably. A facies is a characteristic rock that represents a certain depositional environment. These char-
acteristics include physical, chemical, and biological aspects. In geology, sedimentary layers are said to
lie conformably when there is no unconformity between them. And remember, sediment is deposited in
continuous, horizontal layers.

Now for the rest of the law. In the marine environment, sand is deposited close to shore, silt and clay
further away, and maybe limestone very far away (Figure 5.7). All three of these sediments are deposited
simultaneously and form a single, continuous layer, as in Time 1 in Figure 5.7. The boundary between
them is gradational, meaning that the transition from sand to silt, or silt to clay, or clay to limestone is a
gradual transition rather than an abrupt boundary (That’s what the zig-zag lines in the figure mean). This
gradual transition keeps the single-layer continuous and conformable.

Now, what if sea level were to rise (called transgression), as in Time 2 in Figure 5.7? As the next layer of
sediment is deposited, all sediment types are shifted toward the coastline following the sea-level rise. If
sea level continues to rise, the next sediment layer will shift even further, as in Time 3 of Figure 5.7. Look-
ing at Time 3 in Figure 5.7, the shale at the bottom is overlain by limestone, and that limestone is the same
facies as the limestone to the left of that original shale. The same could be said about the sandstone and
shale. That’s what Walther’s Law means, vertical changes in the succession of sedimentary rocks reflect
lateral (horizontal) changes in the environment. If sea level were to lower (called regression), the facies
would move in the other direction, as in Figure 5.8 (Note: if sea level drops enough to expose sediment,
erosion will take place).
156 | CHAPTER 5: STRATIGRAPHY

Figure 5.7 – A time-lapse of sea-level rise, also called transgression. At each time interval, the sediment
package that is deposited moves in the direction sea level is rising. The sediments are stacking on top of
each other toward the coastline, and that pattern is called onlapping. The red boxes in Time 3 represent
sediment cores taken at those locations. By looking at the patterns in the sediment cores, you can tell that
sea level was rising, and the coastline was to the right. The blue pattern represents limestone, the gray
pattern represents shale, and the yellow dotted pattern is for sandstone. Image credit: Daniel Hauptvogel,
CC BY-NC-SA.
CHAPTER 5: STRATIGRAPHY | 157

Figure 5.8 – A time-lapse of sea-level fall, also called regression. At each time interval, the sediment
package deposited moves in the direction sea level is falling. The sediments are stacking on top of each
other away from the coastline, and that pattern is called offlapping. The red boxes in Time 3 represent
sediment cores taken at those locations. By looking at the patterns in the sediment cores, you can tell that
sea level was falling and the coastline was to the right. Same patterns as in Figure 5.7. Image credit: Daniel
Hauptvogel, CC BY-NC-SA.
158 | CHAPTER 5: STRATIGRAPHY

Exercise 5.4 – Transgression and Regression

Identifying transgressive (rise) and regressive (fall) cycles in sea level is a key component of lithostratigraphy
and sequence stratigraphy. Figure 5.9 is a basic stratigraphic column that contains evidence of transgression
and regression.

a. Shade in the facies environment box for each layer of sedimentary rock in the column.
b. Identify where transgression and regression are taking place. You can mark this with vertical arrows on
the side of the figure.
c. Can you tell how much sea level has changed? Explain.

d. Can you tell what direction the coastline was in? Explain.
CHAPTER 5: STRATIGRAPHY | 159

Figure 5.9 – Stratigraphic column for part of Exercise 5.4.


The colors and patterns are similar to Figures 5.7 and 5.8.
You will need to fill in the columns on the right side of the
figure. Image credit: Daniel Hauptvogel, CC BY-NC-SA.

e. Sedimentary layers are laterally continuous, and they transition into other sediment types. This means
that a facies can “pinch out” over horizontal distances. Complete the lithologic correlation in Figure 5.10
below of two sediment cores.
f. Which layer pinches out? ____________________
g. Explain why that layer pinched out.

h. The addition of a second core now lets you interpret the direction of the coastline. Was the coastline to
160 | CHAPTER 5: STRATIGRAPHY

the right or the left? Explain your reasoning.

Figure 5.10 – Two stratigraphic columns for part of Exercise 5.4. You
will need to correlate the lithology between these two columns. Note,
there are no unconformities in these columns. Image credit: Daniel
Hauptvogel, CC BY-NC-SA.

i. Imagine a scenario where you’re working as a hydrogeologist


hydrogeologist, a type of geologist that deals with water.
Your company has assigned you to locate the best source of groundwater for drilling a new water well.
Groundwater is contained in rocks with a lot of pore space between the grains. In this area, sandstones
make the best aquifer
aquiferss. You have found stratigraphic columns for the rocks in the area from wells drilled
CHAPTER 5: STRATIGRAPHY | 161

on nearby properties. The site you are investigating is located between cores 2 and 3. Use this data to
create a cross-section to help you understand the thickness of the various layers and potential reservoirs
for groundwater at the new well location. To do this, connect layers with similar lithology, and be careful
of layers pinching out. The first layer of limestone is already completed for you.
j. Critical Thinking: The best aquifers will be laterally continuous sand layers. Label the layer(s) that you
think would be the best source of water.
k. Critical Thinking: What is the minimum depth you need to drill to install a new water well?
____________________

Figure 5.11 – Four stratigraphic columns for part of Exercise 5.3 and the proposed location for a new water well. The
height of the columns changes from core to core, representing a small hill. The horizontal scale is not the same as the
vertical scale. Image credit: Joshua Flores and Daniel Hauptvogel, CC BY-NC-SA.

Determining facies is an essential part of stratigraphy. In general, there are three main types of facies;
marine, coastal, and terrestrial. Many other interpretations of data from sediments and sedimentary
rocks rely on an accurate facies analysis. For example, if you are looking at changes in sediment chem-
162 | CHAPTER 5: STRATIGRAPHY

istry, they could result from the facies changing. So, you need to have a good and accurate understanding
of the facies. You probably discussed rivers and running water in your Physical Geology class. These both
erode and deposit sediments. In a meandering river, they can erode their banks and form cut banks or
deposit on the opposite shore in a point bar. Further downstream, the sediment transported in the river
gets deposited in deltas. Each one of these facies has a distinctive set of sedimentary structures and
depending on the velocity of the water a narrow range of grain sizes. Other features of rivers include
oxbow lakes, levees, terraces, flood plains, and braided channels. These are all different lithofacies. You
may also categorize facies using chemical or biogenic criteria.

Exercise 5.5 – Facies in a Coastal Environment

In Chapter 2, you learned about different depositional environments. One environment is coastal which is
subdivided into beaches and swamps. Well, beach and swamp environments can be further subdivided based
on their proximity to the ocean. These coastal environments can have many sedimentary facies (Figure 5.12),
and each has something distinctive about it. You’ll notice that these facies are mainly sandstone or mudstone,
which is not distinctive. Sedimentary rocks have other characteristics besides their particle size, including the
abundance of fossils, sedimentary structures, and layering patterns of the strata.

a. Use Figure 5.12 below to match the environments to the lithofacies and descriptions for: Backshore
Backshore,
Dunes
Dunes, Foreshore
Foreshore, Bay/Lagoon
Bay/Lagoon, Marsh
Marsh, Nearshore
Nearshore, and Overwash
Overwash.

Figure 5.12 – A model of coastal sedimentary environments and facies. Image credit: Carolina Ramón Dueñas, CC
BY-NC-SA.
CHAPTER 5: STRATIGRAPHY | 163

Table 5.4 – Lithofacies descriptions for some environments in Figure 5.12

Environment Lithofacies Description

Coarse sand with abundant shell


Sandstone
fragments

Medium to fine sand and rare shell


Sandstone
fragments

Medium to fine sand and no shell


Sandstone
fragments

Fine to very-fine sand, root structures,


Sandstone
and cross-bedding

Horizontal bedding, root structures,


Sandstone with mudstone
and shell fragments

Bioturbation, root structures, and


Mudstone
peat

Bioturbation, root structures, and


Mudstone
shell fragments

b. You have collected a 13 m sediment core from the area. Your team has separated and described the
sedimentary layers. Based on your team’s descriptions, identify the environment for each layer (using
environments from Part I) and fill in the last column of Table 5.5.
164 | CHAPTER 5: STRATIGRAPHY

Table 5.5 – Descriptions of layers in a sediment core for which you need to interpret the environment

Scale (m) Lithology Description Environment

0.00 – 0.50 Sandstone Coarse sand with abundant shell fragments

0.50 – 1.20 Sandstone Medium sand with occasional shell fragments

Coarse, well-sorted sand with abundant shell


1.20 – 1.90 Sandstone
fragments

Medium-grained sand with rare shell


1.90 – 2.20 Sandstone
fragments

Medium-grained sand with rare shell


2.20 – 3.80 Sandstone
fragments

3.80 – 4.80 Sandstone Medium to fine sand, no shell fragments

Fine sand, well-sorted, cross-bedded with


4.80 – 5.60 Sandstone
occasional root structures

Medium to fine sand, no root structures or


5.60 – 7.00 Sandstone
shell fragments

7.00 – 7.45 Sandstone Fine sand, well-sorted with cross-beds

Unconformity at the base, alternating poorly


Sandstone
sorted sand and 1 to 10 cm thick beds of silty
7.45 – 8.05 interbedded with
clay with bioturbation. Root structures and
mudstone
shell fragments throughout.

Silty clay, moderate bioturbation, root


8.05 – 8.60 Mudstone
structures, shells, and peat

Sandstone Alternating medium- to coarse-grained sand


8.60 – 9.20 interbedded with and 1 to 10 cm thick beds of silty clay with root
mudstone structures

Silty clay, moderate bioturbation, root


9.20 – 10.80 Mudstone
structures, shell fragments, and peat

Dark gray silty clay, intense bioturbation, and


10.80 – 13.00 Mudstone
shells

c. Use your information to construct your stratigraphic column. If you have computers available, you can
do this digitally through the free program Sedlog (Some geologists also use another free program, PSI-
CAT); otherwise, you can use Figure 5.13 to hand-draw your column. Indicate which facies of trace fossil
is likely the cause of the bioturbation. Note: a stratigraphic column like this is exactly how geologists
present lithologic data from a sedimentary core.
CHAPTER 5: STRATIGRAPHY | 165

Figure 5.13 – A detailed stratigraphic column that you need to fill out as part of Exercise 5.4. vf=very fine, f=fine,
m=medium, c=coarse, vc=very coarse, gran=granule, pebb=pebble, cobb=cobble, boul=boulder.

d. Geologic History: Briefly describe the geologic history of this core, starting from the oldest layer.
166 | CHAPTER 5: STRATIGRAPHY

Exercise 5.6 – Facies and Biostratigraphy

Remember that not all stratigraphic units are laterally continuous, which can make correlation difficult. In
1820, William Smith noticed that he would occasionally find fossils while digging canals in different parts of
England. He noted that the types of fossils changed according to his location. Whenever he found fossils that
matched, he assumed the locations were from the same layer and made maps detailing this. From the maps,
he predicted the types of rocks and fossils found at different locations while digging more canals. This made
it easy to avoid “problem” rock layers and predict the best way to dig a new canal.

The branch of stratigraphy that uses fossils to determine depositional environments and ages of sediment is
called biostratigraphy
biostratigraphy. In this exercise, you will use a combination of lithostratigraphy and biostratigraphy.

Background: Your cousin has a small fossil shop that is very popular among tourists, but the location where
your cousin has been collecting fossil bivalve shells is no longer accessible. He asked you to help find a new
place to collect more fossils and to determine its sedimentary environment. Your cousin told you there are
5 different lithologies in the area. You go to a dozen locations (geologists call these stops) and describe the
geology at each place.

To help your cousin, you need to create a geologic map. On your map, you need to use colors to differentiate
each unit. On geologic maps colors represent the rock age. For example, rocks from the Cretaceous Period
are shown in green, and Quaternary deposits are shown in yellow. Each geologic unit is marked with a label.
Geologists typically use an abbreviation for the age and name of the formation. For example, in the Grand
Canyon, the label Pk refers to the Permian Kaibab Formation and label Mr designates the Mississippian Red-
wall Formation. A geologic map is usually printed over a topographic base map. Many different types of lines
and symbols are found on geologic maps (more on this in Chapter 9). The most prevalent are thin black lines
that define the contacts between two different mappable units.

Use your notes from Table 5.6 and the map of the stops in Figure 5.14 to see if you can find a new location to
collect more fossils.

a. Group the sedimentary rocks from your seven stops into five lithologic units and list these in Table 5.7.
Each stop can only belong to one unit.
b. On the map in Figure 5.14, draw contact lines that separate the different units. Color each unit according
to its geologic age (see Table 5.7).
c. Circle an area on your map that you would recommend looking for additional fossils.
CHAPTER 5: STRATIGRAPHY | 167

Table 5.6 – Descriptions of the stops for Exercise 5.6

Stop Description

This is where your cousin used to collect fossils, specifically a shell that looks like
Stop 1 a clam. Fossilization has made it have a variety of colors that are dazzling to the
eye.

Here there are lots of fossil imprints of leaves and branches. You don’t see any
Stop 2
sedimentary structures, and the rock is very dark as well as fine-grained.

This unit is light colored limestone and has some fossil fish. A local geologist has
Stop 3
told you that this is the oldest unit in the area.

Cross-bedding! No fossils, but you were fascinated by beautiful cross-bedded


Stop 4
sandstone from what look to have been oscillatory waves.

You have found another limestone. There aren’t as many fossil fish here, but you
Stop 5
have seen a few smaller ones.

So many shell pieces. Unfortunately, everything here is broken up. There are
Stop 6
some fragments of the desired fossil.

Stop 7 A well laminated, dark shale with impressions of fern leaves.

You try to scratch this rock with your geologic hammer but it doesn’t scratch. The
Stop 8
grains are well-rounded and well sorted.

A fissile rock that easily breaks when you try to get a sample with your geologic
Stop 9
hammer.

A fine grained rock with molds and casts of bivalve fossils and a few gastropods
Stop 10
(snails).

Stop 11 This tan fine-grained rock fizzes when you put a dilute acid on it!

Another rock that fizzes when you put dilute acid on it. However, this one has
Stop 12
large, angular, poorly sorted fragments.

Table 5.7 – Worksheet for Exercise 5.6

Unit Stop(s) within unit

1
168 | CHAPTER 5: STRATIGRAPHY

Figure 5.14 – Location map for stops in Exercise 5.6

5.4 Physical Properties


When geologists recover sediment cores, they run an assortment of analyses on the core itself and the
hole it came from. After the drilling is completed, geologists will lower an instrument into the hole to
take physical property measurements. These measurements can be continuous or occur at specific inter-
vals. Properties that are commonly measured are natural gamma radiation, resistivity, conductivity,
density, porosity, and others. These properties are used to give geologists as much information about
the rocks or sediment as possible. Physical properties can also be used to correlate sediment cores to
each other. Geologists make these correlations to easily compare a section on one core to a section on
another.

Exercise 5.7 – Physical Property Correlation

Figure 5.16 below shows the natural gamma radiation logs for two cores recovered by the International Ocean
Discovery Program Expedition 369 from the southwestern coast of Australia in the fall of 2017. These two
cores were only 20 meters away from each other. The unit for natural gamma radiation is counts per sec-
ond (CPS). You don’t need the details behind this unit’s meaning, but the higher the number, the more clay
particles there are in the sediment (more shale-like). Geologists use this property to determine and compare
CHAPTER 5: STRATIGRAPHY | 169

lithology; sand versus clay and mud. This property is also used extensively both in oil and gas exploration and
hydrogeology.

a. Correlate the two sediment cores to each other using natural gamma radiation data. Draw dashed lines
across both columns that match patterns to each other.
b. How many layers did you correlate? _______________
c. Compare your answer to b to that of your classmates. Did you have the same number of layers? Explain
any differences.
170 | CHAPTER 5: STRATIGRAPHY

Figure 5.16 – Natural gamma radiation logs for two holes as part of IODP Expedition 369 off the coast of southwestern
Australia. mbsf=meters below seafloor. Higher gamma counts generally indicate clay-sized particles and shale, while
lower counts indicate silt- and sand-sized grains. The data is from IODP.

Side note: You might be wondering why the IODP collects sediment cores near each other. The reason for
this is that the drill rig is only capable of collecting 1.5 m of core at a time, and when there is a section break
every 1.5 m, some of the sediment is actually lost where one section ends, and the next one starts. Addition-
ally, several reasons could cause incomplete recovery of sediment. So, the IODP collects a few cores close to
each other and slightly offset to ensure they can get a complete record of the sediment. For example, the first
section of the first drill site will start at the seafloor and collect sediment every 1.5 m, but the first section at
the second drill site will only go down 1.0 m, and then the rest will be every 1.5 m.
CHAPTER 5: STRATIGRAPHY | 171

Exercise 5.8 – Paleomagnetic Properties to Determine Depositional Age

Geologists also analyze magnetic properties of sediment cores for correlation (magnetostratigraphy
magnetostratigraphy), but they
are also used to help determine the age that the sediment was deposited. Do you remember how magnetic
minerals in igneous rocks align themselves to Earth’s magnetic field as the mineral crystallizes? Well, sedi-
ments can align themselves while they are being deposited, too. So, paleomagnetic properties in sediment
cores can be correlated to known changes in Earth’s magnetic field to determine the sediment’s depositional
age.

We currently have normal polarity, where the magnetic north pole and the geographic north pole are in the
Arctic. During periods of reversed polarity, the magnetic north pole moves to the geographic south pole in
the Antarctic. These magnetic reversals are a natural process on Earth, and geologists have developed a time
scale of these reversals called the Geomagnetic Polarity Time Scale (GPTS).

a. Use the inclination data from an IODP sediment core in Figure 5.17 to determine periods of normal and
reversed polarity. Normal polarity is a positive number, and reversed polarity is a negative number. Fill
in the blank polarity column by coloring in the normal polarity intervals in black. Reversed polarity inter-
vals should remain white.
b. Correlate your polarity column to the GPTS to determine depositional ages for this sediment core. You
can do this by drawing lines that match your polarity intervals of the sediment core to the polarity inter-
vals of the GPTS.
c. What is the age range of the missing sediment in the core gap? _________________________
172 | CHAPTER 5: STRATIGRAPHY

Figure 5.17 – Inclination data (blue dots) from an IODP drill site in the Izu-Bonin-Mariana fore arc drilled during Expedition 352.
There is a core gap between 38 m and 42 m (gray bar), which means the sediment was not recovered during the drilling
process. Gaps in sediment cores are frequent and can happen for a variety of reasons. Data from IODP.

5.5 Seismic and Sequence Stratigraphy


So far, layers on top of one another have been interpreted as being younger than the layers beneath
them. In sequence stratigraphy, age is interpreted from lateral (horizontal) variations of rocks based on
the sequence of sedimentary strata in sea-level rise and fall cycles. As sea level rises or lowers, sediment
deposition migrates landward or seaward. Sea-level regression is interpreted from (clinoforms), which
appear to lay somewhat on top of each other at an angle that dips toward the deeper ocean (Figure 5.18).
These strata form as land migrates seaward and deposits sediment further away from the original coast.
Sea-level rise results in onlapping, where horizontal layers of sediment are deposited against the slop-
CHAPTER 5: STRATIGRAPHY | 173

ing surface of the clinoforms from the previous sea-level drop. Here is a link to an overview of sequence
stratigraphy.

Sequence stratigraphy comes with a slew of new vocabulary words described below and shown in Figure
5.18.

• Base Level – commonly the same as sea level but changes due to basin size and tectonics
• Clinoforms – inclined sedimentary strata tilting toward the deeper ocean
• Downlap – steeply dipping younger strata against a surface or underlying shallowly dipping strata;
indicative of sea-level fall
• High Stand – sea level rise and relative increase of onlapping sediment packages
• Low Stand – sea level falls
• Maximum flooding surface – a surface that marks the transition from a transgression to a regres-
sion
• Onlap – shallowly dipping, younger strata next to more steeply dipping, older strata; indicative of
transgression
• Sequence Boundary – often an unconformity formed by erosion that separates different deposi-
tional cycles
• Unconformity – surface separating older from younger rocks and representing a gap in the geo-
logic record

Figure 5.18 – Seismic patterns for a stratigraphic sequence. The lines are seismic reflections. Their shape and how
they terminate (shown with the arrowheads) at different locations indicate different geological processes. Image
credit: Adapted from Vail et al. (1977) by Virginia Sisson and Carolina Ramon Dueñas, CC BY-NC-SA.

As you look at this figure, there are several things to note. First, clinoforms are not flat. Why you might
ask yourself. Well, sediments that are deposited on flat surfaces are flat-lying, whereas those that are
deposited on shallowly dipping surfaces are shallowly dipping. New sedimentary layers will mimic the
underlying layers until the underlying surface’s angle exceeds that of the maximum angle of repose. The
angle of repose is the angle at which loose sediments will cascade or landslide down the slope, and it
varies depending on the sediment; sandstone is about 30°, and shale is higher than 40°. So, the slope of
the clinoforms reflects the underlying sloping strata. Continental margins are where you can commonly
find sloping surfaces.

Look carefully at the top sequence boundary. Does it look like a clinoform? Sometimes sequence bound-
aries look similar to clinoforms as the transition between depositional cycles can be very gradual without
174 | CHAPTER 5: STRATIGRAPHY

significant erosion. You have to look closely to see if there is onlap of the next strata to distinguish a cli-
noform from a sequence boundary.

Exercise 5.9 – Sequence Stratigraphy from a Seismic Image

In this exercise, you will work on determining whether sea level influenced the deposition of Cretaceous sed-
iments as well as the different features associated with sequence stratigraphy. To do this, you will investigate
a stratigraphic section in the Colville Basin, Arctic Alaska (Figure 5.19).

Figure 5.19. Geologic map of part of the National Petroleum Reserve of Alaska (NPRA) west of the Colville River. You may
wonder why all the exposure is mapped as Quaternary deposits and not Cretaceous. This is because glaciation and formation
of tundra have covered all of the rock exposures. They are only exposed along rivers and seaside cliffs as well as buried
beneath the tundra. This is one reason you need to use a seismic line to do this exercise. Image credit: Adapted from the
geologic map of Payne (1951) with the location of a seismic line from the U.S. Geological Survey NPRA data archive by Virginia
Sisson, CC BY-NC-SA.

Long before Alaska was granted statehood, oil was found seeping from the ground in the Arctic. This oil
helped lead to creating the National Petroleum Reserve of Alaska (NPRA – an area the size of Indiana). This
area is also known for its coal; some have estimated that this region has ~40% of the U.S. reserves. Arctic
Alaska has provided much of the energy resources for the rest of the United States for many years. To help
define where the oil was located and how it formed, the U.S. Geological Survey conducted many seismic sur-
CHAPTER 5: STRATIGRAPHY | 175

veys (often called lines) across the region. We will look at one line for this exercise. Figure 5.20 shows the Cre-
taceous formations in this section: Pebble Shale unit, Torok Formation, and Nanushuk Formation.

Figure 5.20. Simplified stratigraphic section for the Cretaceous sedimentary


sequence in northern Alaska. Image credit: Adapted from Mull et al. (2003) by
Virginia Sisson, CC BY-NC-SA.

Before you analyze the seismic section, you should know how sea level is changing worldwide. If you know if
the sea level was rising or falling during sediment deposition, this will let you know if you are in a highstand
or lowstand environment.

a. We don’t know the exact age of these sediments, so first, you must determine the absolute age of these
176 | CHAPTER 5: STRATIGRAPHY

formations using Figure 3.1. Fill out the ages in Table 5.8.

Table 5.8 – Worksheet for Exercise 5.9a

Rock Formation Absolute Age Range

Nanushuk Formation

Torok Formation

Pebble Shale unit

b. Use the ages you determined in a and the sea-level record for the Phanerozoic Eon in Figure 5.21 to
determine if sea level was rising, falling, or not changing when these three formations were deposited.
You might wonder how to do this since this figure has three different sea-level curves. It’s up to you to
choose one of these to answer this question. Indicate which curve you used to answer this question.
CHAPTER 5: STRATIGRAPHY | 177

Figure 5.21. Three proposed sea-level curves for the Phanerozoic Eon. The gray area is a compilation
of curves proposed by researchers at ExxonMobil. The dashed purple curve considered global ice
volumes without tectonic influences on accommodation space. The red curve used a geodynamic
model combined with sea level data. Image credit: Simplified by Virginia Sisson CC BY-NC-SA from
Vérard et al. (2015).

c. In this exercise, you will identify the age progression of deposition and the different features associated
with sequence stratigraphy. In the seismic image (Figure 5.22), identify as many clinoforms as you can.
Number them in order of depositional age, with “1” being the oldest. You should be able to identify at
least 10 clinoforms. Hint: There is at least one unconformity. Pay attention to the angles at the beginning
and end of your features and the angle at which they touch another reflector; if it is at a high angle, this
is likely to be either onlap or downlap.
178 | CHAPTER 5: STRATIGRAPHY

Figure 5.22 Seismic cross-section (line 16-81b) across Cretaceous sediments of the northeastern National
Petroleum Reserve of Alaska (NPRA). These sediments filled the Colville foreland basin about 6000′ deep. This view
is looking to the north so that the left side is to the west, whereas the right side is to the east. Figure 5.19 shows
the location of this seismic line. Blue is for positive amplitude, black for negative amplitude, and white is for no
amplitude difference. Image credit: SEPM exercises for sequence stratigraphy with data from U.S. Geological
Survey NPRA data archive public domain.

d. Using the seismic features as your guide, which direction would you go to find a marine basin? Explain
your reasoning.

e. Based on the seismic section, do you think that sea level was rising or falling during the deposition of
this sequence? Explain your reasoning.

f. Critical Thinking: Is your answer to e the same or different from your answer to b, which used the strati-
graphic section and the global sea-level curve? If it’s different, explain what might cause the discrepancy.
Or, if your answers are the same, how does this strengthen your interpretation?

All the previous exercises were aimed at understanding changes in stratigraphy due to environmental
conditions and processes. Another component of stratigraphy is the tectonic setting for the sedimentary
basin. So, which two tectonic settings are important to consider? The next two exercises will address both
convergent and divergent plate boundaries in terms of stratigraphic change. Transform margins may
have overlapping faults that create subsidence and small pull-apart basins.
CHAPTER 5: STRATIGRAPHY | 179

The Appalachian/Caledonide Mountains formed during three separate tectonic events over a period
of 220 million years. These events were the Taconic (470 to 440 Ma), Acadian (420 to 380 Ma), and
Alleghenian (350 to 250 Ma) orogenies. Between each of these tectonic events, the mountains eroded
as they were uplifted. One of the concepts you need to know for this exercise is the effect of increasing
the weight on the lithosphere. You may have learned about isostasy in Physical Geology and how the
lithosphere responds like an iceberg melting in the ocean. Well, during continental collision, the added
weight of the continental mass will cause the lithosphere to bend, creating a foreland or foredeep basin.
For example, the Persian Gulf is a basin formed from the collision of the Asian and Arabian plates cre-
ating the Zagros Mountains. These sedimentary basins often form shallow (30-100 m) repositories for
continentally derived sediments. If the area beside the continental collision is wide, it will form a shallow
inland ocean (sometimes called an epeiric or epicontinental sea), such as Baffin Bay in northern Canada.

Exercise 5.10 – Convergent Tectonics in a Sedimentary Record

The Catskill clastic wedge (Figure 5.23) is a Devonian sediment package sourced from the Appalachian Moun-
tains during the Acadian Orogeny. This sediment package extends from New York through Virginia and West
Virginia and is part of the larger Appalachian Basin. In western Virginia and eastern West Virginia, the Catskill
clastic wedge contains evidence for three phases of the Acadian Orogeny:

Phase 1 – A rapid drop of the seafloor caused by the massive weight of the mountains. The load causes the
lithosphere to bend, creating a concave basin in front of the uplifting mountains. It is characterized by a sud-
den change to deep-water sediments such as shales.

Phase 2 – As the mountains erode, coarse sediment is deposited. These can be deposited in fluvial (rivers) or
near-shore marine depositional settings.

Phase 3 – At the end of the mountain building process, there is a return to a stable passive margin sequence
ranging from fine sediment deposition to shallow marine carbonates.
180 | CHAPTER 5: STRATIGRAPHY

Figure 5.23 Continental collision can result in the formation of inland seas, such as the Kaskaskia Sea and foreland basins.
The brown represents the collision of North America with the Avalon terrane. Image credit: Virginia Sisson, CC BY-NC-SA.

Complete the following questions about the Catskill clastic wedge:

a. Figure 5.24 is a roadside outcrop of the Brallier Formation in Pennsylvania, composed of shale and
fine sandstone layers. Provide a sketch of this outcrop and identify the layers of sandstone vs. the
layers of shale.
CHAPTER 5: STRATIGRAPHY | 181

Figure 5.24 – Outcrop of Devonian Brallier Formation on the north side of the Pennsylvania
Turnpike, central Bedford County, near Mile Marker 138 in Pennsylvania. Sandstone beds near the
bottom of the photo are possible turbidites. Image credit: Wikimedia user Jstuby, CC0 Public
Domain, photo taken August 2011.

Sketch Area:

b. Figure 5.25 is a roadside outcrop of the Hampshire Formation in Pennsylvania, composed of sand-
stone and fine-grained siltstone. Provide a sketch of this outcrop.
182 | CHAPTER 5: STRATIGRAPHY

Figure 5.25 – Point bar deposits in the Hampshire Formation near North Bend, Pennsylvania. Image credit:
Wikimedia user Rygel, M. C., CC BY-SA.

Sketch Area:

c. What sedimentary structure is shown in Figure 5.25, and can you tell in which direction sediment
was being transported?
CHAPTER 5: STRATIGRAPHY | 183

d. Figure 5.26 below contains the sequence of sedimentary rocks associated with the Catskill clastic
wedge. Based on the rock descriptions, identify the depositional environment for each strati-
graphic unit and which orogenic phase it represents.

Figure 5.26 – Stratigraphic units of the Catskill clastic wedge and the dominant lithology of each unit. Image credit:
Adapted by Carlos Andrade and Virginia Sisson CC BY-NC-SA.

Another dynamic aspect of plate tectonics is continental rifting, a process in which a continental mass
splits apart and two separate new landmasses form and drift away from each other. After the last
orogeny in the Appalachian mountains, Pangea broke apart into our current continent configuration
through continental rifting processes. The Recôncavo basin in Brazil near the city of Salvador (Figure 5.27)
is the southernmost basin in northeastern Brazil created by continental rifting of South America and
Africa. This rifting process is well recorded in the eastern Brazilian margin and shows different stages of
this process:

1. Before rifting begins, continental processes such as fluvial, eolian, and lacustrine deposition domi-
nate the sediment deposition.
184 | CHAPTER 5: STRATIGRAPHY

2. Then as the rifting begins, igneous activity also begins. In continental rifts, this is usually bimodal
(rhyolite and basalt) volcanism. Rifting also causes normal faulting of the continent forming grabens
and half grabens resulting in lacustrine (lake) deposits. In some areas, very high-energy processes
dominate. As the continent starts to split apart, the ocean water will flood the depression caused by
the faulting creating shallow-water seas, which sometimes host various shallow marine species.
3. As the shallow seas and lakes are evaporated, massive salt and anhydrite (evaporites) deposits are
formed.
4. Rifting completely separates the continent into two new landmasses, which start drifting apart and
are marked by a “passive margin system.” Tectonic activity is no longer present, and the sediments
are increasingly deeper marine.

2
Figure 5.27 – The Recôncavo basin covers 11,500 km and is a prime source of petroleum. Since 1939, 225 wells
3
have been drilled. By 2012, this basin had produced over 1.6 billion barrels of oil and 70 billion km of gas. Image
credit: Simplified by Virginia Sisson CC BY-NC-SA from Magnavita et al., 2005 with insert map of Brazil from
Wikimedia Commons user Addicted04 CC-SA.
CHAPTER 5: STRATIGRAPHY | 185

Exercise 5.11 – Continental Rifting in a Sedimentary Record

Stratigraphic Evidence of Continental Rifting

a. Determine the depositional environments for the different lithologies in the northeastern Brazilian
Recôncavo Basin’s stratigraphic column in Figure 5.28 below.
b. Identify the 4 stages of rifting (described above Figure 5.28) and mark them 1, 2, 3, and 4 in the
right-most column in Figure 5.28. Mark your boundaries between the stages by drawing a thick line
at the top of the formation that ends each stage.

Figure 5.28 – Stratigraphy of the Brazilian Recôncavo Basin with simplified lithology. Image credit: Carlos
Andrade and Virginia Sisson CC BY-NC-SA.

c. Critical Thinking: What will the sedimentary environment be in this part of Brazil after rifting has
stopped?

d. Critical Thinking: Today, the Atlantic Ocean marks the rift boundary between Brazil and Africa. At
what stage did the rifting stop in the Recôncavo Basin? Why?
186 | CHAPTER 5: STRATIGRAPHY

Outcrop Interpretations

Figure 5.29 – Sandstones and fossiliferous shales from the Dom João Formation on the southwestern side of the
Recôncavo basin at Praia da Pedra Mole (translates approximately to soft rock beach) near Barra do Paraguaçu.
These geologists have a rock hammer on the outcrop, and one is about to take a picture. They are standing among
large blocks with distinctive white spots, which are oyster shells. Image credit: Luzia Brito via Google Earth.

The following questions refer to Figure 5.29.

e. What sedimentary environment do you find oysters in?

f. Why is the rock in Figure 5.29 black?

g. What is the sedimentary environment of the overlying buff-colored layers?

h. How does this change in sedimentary environments fit with the tectonic setting of the Recôncavo
Basin?
CHAPTER 5: STRATIGRAPHY | 187

Figure 5.30 – Dom João Formation along BR 101, about 14 km southwest of Algoinhas, Brazil. Two processes
cause the red color; one is intense tropical weathering creating red soil (the left side of the outcrop is almost
completely converted into red soil as is a layer at the top of the outcrop), and two, hematite oxidation during
deposition of these sandstones and shales. A fault offsets the bedding in this outcrop along the yellow line.
Image credit: Google Street View.

i. What type of fault is present in Figure 5.30? It may help you to draw arrows on either side of the
fault showing the sense of motion. ____________________
j. What is the age of the fault (remember to look at the stratigraphic column for the age of the for-
mation)? ____________________
k. Is this fault related to continental rifting? How does this help you determine the depositional envi-
ronment?

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson, Carlos Andrade, Joshua Flores, Melissa Hansen
188 | CHAPTER 5: STRATIGRAPHY

References
Krumbein, W.C., and Sloss, 1963, Stratigraphy and Sedimentation (2nd edition), W.H. Freeman and Co.,
660 p.

Magnavita, L., Da Silva, R.R., and Sanches, C.P., 2005, Field trip guide of the Recôncavo basin, NE Brazil,
April 2005, Boletim de Geociencias – Petrobras, 13(2), 301-333.

Miller, K.G., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman, P.E., Cramer,
B.S., Christie-Blick, N., and Pekar, S.F., 2005, The Phanerozoic Record of Global Sea-Level Change, Sci-
ence, 310, 1293-1298. doi: 10.1126/science.1116412

Moore, G.F., Park, J.-O., Bangs, N.L., Gulick, S.P., Tobin, H.J., Nakamura, Y., Sato, S., Tsuji, T., Yoro, T.,
Tanaka, H., Uraki, S., Kido, Y., Sanada, Y., Kuramoto, S., and Taira, A., 2009. Structural and seismic strati-
graphic framework of the NanTroSEIZE Stage 1 transect. In Kinoshita, M., Tobin, H., Ashi, J., Kimura, G.,
Lallemant, S., Screaton, E.J., Curewitz, D., Masago, H., Moe, K.T., and the Expedition 314/315/316 Scien-
tists, Proc. IODP, 314/315/316: Washington, DC (Integrated Ocean Drilling Program Management Inter-
national, Inc.). doi:10.2204/iodp.proc.314315316.102.2009

Mull, C.G., Houseknecht, D.W., and Bird, K.J., 2003, Revised Cretaceous and Tertiary Stratigraphic Nomen-
clature in the Colville Basin, Northern Alaska: U.S. Geological Survey Professional Paper 1673, 51 p.

Payne, T.G., 1951, Geology of the Arctic slope of Alaska: U.S. Geological Survey Oil and Gas Investigations
Map 126, 3 sheets, scale 1:1,000,000.

Picard, M.D., 1971, Classification of fine-grained sedimentary rocks, Journal of Sedimentary Petrology, 41,
179-195. https://doi.org/10.1306/74D7221B-2B21-11D7-8648000102C1865D

Vail, P.R., Mitchum, R.M., and Thompson, S., 1977, Seismic stratigraphy and global changes of sea level,
part 3: Relative changes of sea level from coastal onlap, in C.E. Clayton, ed., Seismic stratigraphy – appli-
cations to hydrocarbon exploration: Tulsa, Oklahoma, American Association of Petroleum Geologists
Memoir 26, 63–81.

Vérard, C., Hochard, C., Baumgartner, P.O., and Stampfli, G.M., 2015, 3D paleogeographic reconstructions
of the Phanerozoic versus sea-level and Sr-ratio variations, Journal of Palaeogeography, 4, 64-84.
doi:10.3724/SP.J.1261.2015.00068

Google Earth Locations


CHAPTER 6: FOSSIL PRESERVATION

Learning Objectives

The goals of this chapter are to:

• Explain the different ways organisms become fossilized


• Identify the mode of preservation for fossils

6.1 Introduction
Everyone has heard the word fossil! So, how do you define this term? In this lab, we will define it to mean
any evidence for the existence of prehistoric life. What is difficult to explain in this definition is what is
meant by prehistoric. For example, would you consider bodies preserved at Pompeii to be fossils, or how
about the remains of a frozen mastodon from the Pleistocene that was preserved well enough to be
eaten? Some say anything older than 11,000 years is a fossil, but this part of our definition is a matter of
semantics. An excellent place to learn more about fossils and fossilization is the Digital Atlas of Ancient
Life.

Exercise 6.1 – Fossilization Probability

We start this chapter on how organisms become fossilized with a quick exercise. Figure 6.1 contains three dif-
ferent organisms.

Figure 6.1 – a) Worms, b) Giant kelp; c) Mussel shells. Image credit: a) Soil-Net, CC BY-NC-SA; b) NPS, Public Domain; c)
Linnaea Mallette, Public Domain.

a. Which organism in Figure 6.1 do you think has the highest chance to become a fossil and why?
190 | CHAPTER 6: FOSSIL PRESERVATION

b. Which organism do you think has the highest chance to leave behind a trace fossil? ____________________

6.2 Types of Preservation


Taphonomy is the branch of paleontology that focuses on the fossilization process. Fossils are preserved
by three main methods: unaltered soft or hard parts, altered hard parts, and trace fossils. You already
learned about trace fossils in Chapter 4.

Unaltered fossils are rare except as captured in amber, trapped in tar, dried out, or frozen as a preserved
wooly mammoth. Amber is the fossilized tree resin that can trap flowers, worms, insects, and small
amphibians and mammals. The father of one of the authors was part of a gold mining operation that
discovered a wooly mammoth calf (nicknamed Effie) in Alaska. This was the first mummified mammoth
remains found in North America. Even though it was buried about 21,300 years ago, it still consists of
tissue and hair. Sometimes, only organic residue is left behind and is detected by molecular, biochemical
techniques. Earth’s oldest fossils are only preserved as complex organic molecules.

Soft tissue is hard to preserve because it needs to be buried before bacterial decay can occur. This preser-
vation occurs when remains are buried rapidly in an oxygen-free, low-energy sedimentary environment.
Since these conditions are uncommon, the preservation of soft tissue rarely happens. Instead, typical
examples of unaltered fossils are skeletal material that has been preserved with little or no change. Many
marine invertebrate fossils and microfossils were preserved in this manner. However, paleontologists are
now looking closer at fossils and recognizing thin carbon layers in the rock around fossils as soft tissue.
Recently, a team led by Mark Norell, a paleontologist at the American Museum of Natural History in New
York City, identified a layer of carbon around dinosaur embryos formed over 200 million years ago that
they think was a soft eggshell!

Unaltered fossils contain minerals that were biologically produced; these include apatite (in bones and
teeth and rarely in exoskeletons, hardness = 5), calcite (calcium carbonate found in many organisms such
as shells, hardness = 3, fizzes in acid), aragonite (similar to calcite, but an unstable polymorph) and opal
(a type of silica found in marine animals and plants, hardness = 7). In addition, the hard parts (exoskele-
ton) of some insects and arthropods are made of chitin, a polysaccharide related to cellulose. If you can
identify the minerals present in a fossil, you can distinguish if it is original material or altered.

Alteration of hard parts is much more common in fossils and happens when original skeletal material is
either permineralized, recrystallized, replaced, carbonized, or dissolved (Table 6.1).
CHAPTER 6: FOSSIL PRESERVATION | 191

Table 6.1 – Common types of fossil preservation

Type of Preservation Example

Permineralization occurs in porous tissue such


as bone and wood. In this type of preservation,
minerals dissolved in water such as quartz,
calcite, or pyrite permeate the pore space and
crystallize. The addition of these minerals results
in denser and more durable fossils. The original
bone or wood material may be preserved, or it
may be replaced or recrystallized. Figure 6.2 – Petrified wood from the
Petrified Forest National Park, AZ.
Image credit: Jon Sullivan, Public
Domain.

Recrystallization involves a change in the


crystal structure but not in mineral chemistry,
similar to recrystallization in metamorphic rocks.
For example, the mineral aragonite, a common
mineral of many shells, sometimes changes to
calcite, a more geologically stable form of the
same chemical composition, CaCO3 (aka a
polymorph). Typically, the overall size and shape
of a recrystallized fossil do not vary substantially Figure 6.3 – Recrystallized, Silurian-age
from the original unaltered specimen, but fine coral from Ohio. Image credit: James
details may be lost. St. John, CC BY.

Replacement is the substitution of original


skeletal material by a secondary mineral. For
example, the calcite of an oyster shell may be
replaced on a molecule-by-molecule basis by
silica. Remarkably, the replaced fossil may retain
some of the fine cellular detail present in the
original even though its composition changed. In
this type of fossilization, pore space is not filled,
and the fossils are not as dense. The most
common replacement minerals are silica
(quartz), pyrite, dolomite, and hematite.
Replacement by pyrite creates some spectacular
fossils, especially those hosted by black shales!
Figure 6.4 – Pyritized fossil trilobite
with appendages (Triarthrus eatoni)
from the Whetstone Gulf Formation,
Upper Ordovician; New York State,
USA. Image credit: James St. John, CC
BY.
192 | CHAPTER 6: FOSSIL PRESERVATION

Carbonization is a type of fossil preservation in


which the organism is preserved as a residual,
thin film of carbon instead of the original
organic matter. Leaves, fish, and graptolites are
commonly preserved in this way. Compression
of the original organism results in thin layers of
carbon. Carbonization can also result in the
formation of coal.

Figure 6.5 – Carbonization of


Silurian-aged graptolites from Poland.
Image credit: James St. John, CC BY.

Molds and casts form when the original skeletal


material dissolves. The organism leaves behind
an impression in the sediment called a mold. If
that impression fills with new sediment, it
creates a cast. Thus, casts are made from molds.
Figure 6.6 – A mold (left) and cast
(right) of a trilobite fossil. Image credit:
Roger Wellner.
CHAPTER 6: FOSSIL PRESERVATION | 193

Internal molds form when sediment fills the


inside of a shell before it dissolves; this occurs
inside bivalves, snails, or skulls. Often, people
confuse casts and internal molds because both
have positive relief. However, internal molds
preserve a 3-dimensional mold of the inside of
the organism, whereas a cast preserves the
structure of the outermost part of the organism.

Figure 6.7 – Dissolution of a gastropod


that has left an internal mold of the
organism. Fossil from eastern Iowa.
Image credit: James St. John, CC BY.

Trace fossils, which we discussed in Chapter 4, are not really fossils but the evidence that organisms
affected the sediment by burrowing, walking, or even leaving behind excrement or vomit. No kidding,
there is fossil poop; this kind of trace fossil is called a “coprolite,” from the Greek word kopros, meaning
dung. One last rare type of trace fossil is gastroliths, extremely smooth polished stones that aided diges-
tion in dinosaurs and crocodilia. These are more highly polished than stream-worn gravels. Gastroliths
found in Jurassic sediments in Wyoming may have been carried by sauropods over 1600 kilometers from
their source in Wisconsin.
194 | CHAPTER 6: FOSSIL PRESERVATION

Figure 6.8 – Different parts of organisms compared to how they can be preserved as fossils. The green circles are
common types of fossilization, the light green, stippled circles are less common, and the light green circles are
uncommon to rare ways. Image credit. VB Sisson modified from Ritter and Peterson (2015) CC-BY-SA-NC.
CHAPTER 6: FOSSIL PRESERVATION | 195

Figure 6.9 – Types of fossilization including alteration and replacement of the original shell. Follow the arrows from
box to box to see how different processes can result in molds and casts. The brown color is a sedimentary rock.
The random pattern represents recrystallized carbonate, and the stippled pattern represents secondary minerals
such as silica or pyrite. Image credit: VB Sisson with a shell showing growth lines and internal structure adapted
from Casella et al., 2017 and fossilization processes adapted from Ritter and Peterson (2015) CC-BY-NC-SA.
196 | CHAPTER 6: FOSSIL PRESERVATION

3-D Model Box 6.1 – Molds

This model shows a preserved fossil shell on the right (not a cast, original) and an external mold on the left
of the ammonoid cephalopod Gunnarites sp. from the Cretaceous Lopez de Bertodano Formation of Snow
Hill Island, Antarctica. The fossil specimen is from the collections of the Paleontological Research Institution,
Ithaca, New York. The diameter of the specimen (not including surrounding rock) is approximately 9 cm.

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://uhlibraries.pressbooks.pub/historicalgeologylab/?p=82

Cephalopod: Gunnarites sp. (PRI 61543)


by Digital Atlas of Ancient Life
on Sketchfab

This is an example of an internal (1) and external (2) mold of the gastropod Cassidaria mirabilis from the Cre-
taceous of Snow Hill Island, Antarctica. The specimen is from the collections of the Paleontological Research
Institution, Ithaca, New York. It is approximately 6 cm in length (not including surrounding rock).

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://uhlibraries.pressbooks.pub/historicalgeologylab/?p=82

Gastropod: Cassidaria mirabilis (PRI 58468)


by Digital Atlas of Ancient Life
on Sketchfab
More 3-D models on fossil preservation can be found at the Digital Atlas of Ancient Life.

If you ever get asked by a friend to help identify a fossil, watch out for pseudofossils, which are accidents
of diagenesis that look like a fossil but are just weird sedimentary formations. Pseudofossils include sep-
tarian nodules that are mistaken for reptile skin or turtle shells, concretions are mistaken for eggs, and
manganese oxide dendrites mistaken for ferns moss.

6.3 Handling of Fossils


Some fossils are incredibly fragile. Some delicate samples are prepared by air abrasion with talcum pow-
der to remove the matrix. For some trilobite specimens, this takes thousands of hours to expose their
CHAPTER 6: FOSSIL PRESERVATION | 197

delicate features. Some fossils you will use may be easy to replace and others impossible. Others may be
part of a faculty member’s personal collection. Only handle the specimens that your instructor says you
can. If you are taking this lab when teaching is face-to-face in a lab setting, you will handle both real and
replica specimens of fossils. While these may have been around for millions or billions of years and seem
like they are now rocks, they need to be treated with respect. Some of the fossils that you may handle
may be the only specimen of their kind in the collection. Some of the larger specimens may be heavy,
especially those that are molds filled with sediment. Never try to scratch the specimens for hardness.
Also, never use acid as a mineral test. Finally, if you break or steal a specimen, you will be charged for its
replacement.

You are free to make sketches or photograph the specimens. If you do this, you may want to put a scale
in the image, such as a coin or ruler. This will help you remember the size of the object. Some specimens
will have labels or numbers written on them, and others will not because they may be too fragile to even
be written on. You must put each specimen back in its appropriate box or location in a lab tray. Also, do
not move any of the paper labels from the boxes. This will prevent confusion for other lab students.

If you’ve wondered how to start your own fossil collection, you can either find them on your own or buy
them. The price of fossils for sale ranges from cheap to outrageously expensive. For example, in 2020, an
anonymous collector bought a fossil Tyrannosaurus rex, nicknamed Stan, for $31.85 million. This speci-
men only had 188 bones and was one of the most complete of its species. If this is beyond your budget,
you can find inexpensive fossils such as fossilized snails from Morocco for only $0.30 each.

Exercise 6.2 – Identifying Types of Fossil Preservation

Inspect the first set of samples and fill out the table with information about the presence of original biologic
material, positive and negative relief, and mineral composition of the samples. Identify the mode of preserva-
tion of the fossils. Use the flowchart in Figure 6.10 to help.

Figure 6.10 – Flowchart for identifying the type of fossil preservation. Image credit: Virginia Sisson and Daniel Hauptvogel
CC-BY-NC-SA.
198 | CHAPTER 6: FOSSIL PRESERVATION

Table 6.1 – Worksheet for Exercise 6.2

Original Material Mineral


Sample Relief* Type of Preservation
Present? Composition*

*Note that you may not see relief or be able to identify the mineral. Leave these blank if necessary.

Critical Thinking: Why is replacement the most common mode of preservation?

Exercise 6.3 – Thinking about Preservation

The way an organism can become fossilized depends on many things. Below are some examples to think
about.

a. Examine an external mold in your fossil collection. These commonly preserve details such as the veins in
leaves or scales of fish.

i. What is the grain size of the surrounding rock? ____________________


ii. Do you think these impressions could be preserved in coarse-grained sediment?

b. Look at some examples of carbonization. The dark matter is the remnant of organic carbon that was
CHAPTER 6: FOSSIL PRESERVATION | 199

never oxidized (decayed). Under what conditions might this kind of preservation occur?

c. Your fossil collection may have graptolites, an extinct planktonic, colonial organism that secreted an
organic shell of chitin similar to your cellulose. These colonies are usually preserved as two-dimensional
impressions, almost always black (indicating carbonization of the chitin).

i. What type of rocks is best suited to finding graptolites?

ii. What were the burial conditions?

d. Some bones and teeth can be preserved, such as unaltered bones or shark teeth.

i. How would you distinguish these from permineralized fossil bones?

ii. Can permineralized wood scratch glass?

e. Now consider the wide range of sedimentary environments.

i. What sedimentary environments are not suitable for preserving fossils?

ii. Which sedimentary environments are good for preserving fossils?

iii. Which depositional environments within continental and marine environments are best for pre-
serving fossils? Explain.

iv. Can volcanic eruptions preserve fossils? Explain.

v. How can the energy of the sedimentary environment affect the preservation of fossils?
200 | CHAPTER 6: FOSSIL PRESERVATION

vi. Can you find fossils in metamorphic rocks? If so, what factors aid in their preservation?

f. Critical Thinking: There are more invertebrates than vertebrate fossils in this lab exercise. Explain why
this is.

Exercise 6.4 – Modes of Preservation in an Ancient Reef

During the Permian time, 299 to 252 million years ago, an extensive reef system grew in west Texas at the
edge of a small inland marine basin that extended over 26,000 km² (10,000 square miles). Now it is called the
Delaware basin, home to a major oil field (Figure 6.11). This reef is now exposed in three mountain ranges;
Apache, Guadalupe, and Glass Mountains. Elsewhere, the reef is now buried around the entire rim of the
basin. The northernmost exposure of the reef is part of the Guadalupe Mountains National Park. It is home
to Carlsbad Caverns, which has the largest underground chamber in the United States. The reef was home to
many fossils, including ammonoids, bryozoans, algae, sponges, brachiopods, gastropods, pelecypods, echi-
noids, fusulinids, trilobites, corals, and crinoids.
CHAPTER 6: FOSSIL PRESERVATION | 201

Figure 6.11 – Map of exposed and unexposed Permian reef that encircled the Delaware Basin, an inland sea in Texas and
New Mexico. Image credit: Adapted by VB Sisson from National Park Service CC-BY-NC-SA.

Unlike modern coral reefs, such as the Great Barrier reef of Australia or the reefs off the coast of Florida and
Belize, it was constructed from sponge
sponges, algae, and lacy animals called bryozoa
bryozoa. One magnificent exposure
of this reef is El Capitan in the Guadalupe Mountains National Park. The reef is subdivided into three parts:
back reef, reef, and fore reef. Each had its own unique ecosystem as well as lithology and preservation. The
deep part of this basin reached depths of almost 800 meters (½ mile) and is where a lot of organic matter was
deposited, leaving black shales – the source of petroleum.
202 | CHAPTER 6: FOSSIL PRESERVATION

Figure 6.12 – Schematic cross-section across a reef showing the back reef, reef, and fore reef as well as the marine basin.
Image credit: Adapted by VB Sisson from National Park System CC-BY-NC-SA.

The Delaware inland sea had a narrow outlet to the Panthalassan ocean, much like the Black Sea today. How-
ever, after ~30 million years, the entrance got restricted, and the basin started to dry up, forming extensive
evaporite deposits (Castille and Salado Formations). This created supersaturated, acidic brines that started to
dissolve the underlying carbonate reef, forming extensive caves and karst that you can now visit at Carlsbad
Caverns National Park and Lechuguilla Cave, the 8th longest explored cave in the world at ~220 km or 138
miles long. These brines also dissolved the silica-rich sponges that formed the reef and affected the fossil
preservation in parts of this Permian reef system.

Google Street View in Google Maps has recorded many of the trails in Guadalupe Mountains National Park.
Start at either McKittrick Canyon or the trail to Guadalupe Peak and drag the orange person icon onto one of
the trails to see views of the massive limestone reef.

The stratigraphy of this basin is complicated because it varies with time and position in the reef. According to
recent sequence stratigraphic analysis, there were up to six transgressive-regressive sequences in this basin
(Kerans and Kempter, 2002). Figure 6.13 gives a simplified stratigraphy for the basin during the Permian.
CHAPTER 6: FOSSIL PRESERVATION | 203

Figure 6.13 – Simplified stratigraphy for the Delaware Basin. Image credit: Simplified by Virginia Sisson from Kerans and
Kempter (2002) CC-BY-NC-SA.

Fossils in the Capitan Formation of the Glass Mountains are uniquely preserved (see Figure 6.4). Paleontolo-
gists found that it is easy to dissolve away the host carbonate in weak acid and leave behind spectacular spec-
imens.
204 | CHAPTER 6: FOSSIL PRESERVATION

Figure 6.14 – Fossils from the Capitan Formation of the Glass Mountains, Texas. a) silicified fossils including the wavy
outer layer of a bivalve shell and attached to it, are encrusting organisms (sclerobionts). Other fossils include a long tube
of a rugose coral with a ribbed athyrid brachiopod at its base; b) silicified Productid brachiopod ventral valve; c) silicified
Hercosestria cribrosa; d) silicified Anthozoa brachiopoda and bryozoa. Image credits: a) James Wilson, CC0 Public Domain;
b) Wikimedia user Wilson44691, CC0 Public Domain; c) Wikimedia user Mark A Wilson Wilson44691, CC0 Public Domain;
d) Wikimedia user the paleobear CC BY.

a. The mineral in these fossils is harder than glass and does not fizz as it is no
longer a carbonate. Sometimes this mineral is just a coating, and other times
the entire fossil is this new mineral.
i. What is the mineral? ____________________
ii. What is the mode of preservation for these fossils? ____________________
iii. Were fluids involved in their preservation? If so, what was their composition?

iv. Why do you think this type of preservation is found in this one stratigraphic unit.
CHAPTER 6: FOSSIL PRESERVATION | 205

v. Which part of the reef were these fossils found? Back reef, reef, fore reef or basin?
____________________

Figure 6.15 –
Fossilized jaw with
teeth of a
Helicoprion ferri.
The fish associated
with this ~35 cm
jaw is estimated to
be 4 m (~12′) long,
but they could
reach up to 10 m in
length. Also shown
are two possible
reconstructions
envisioned by
paleontologists.
The first image was
common in the
1900s to 1960s. In
2013, CT imaging
revealed a
different
reconstruction.
Their primary food
source was
bottom-dwelling
organisms and
possibly
ammonites. Image
credit: Photo is
from James St John
CC BY 2.0,
Reconstructions
adapted by VB
Sisson from
Tapinila et al. 2013
CC-BY-NC-SA.

c. Elsewhere in the Skinner Ranch Formation of the Glass Mountains, fossils include this amazing saw-
toothed whorl of teeth from an extinct shark-like fish known as Helicoprion.

i. Using Figure 6.15, what type of sediment is this fossil found in? ____________________
ii. What is this mode of preservation for this fossil? ____________________
iii. Where in the reef did Helicoprion live? Back reef, reef, fore reef, or basin? ____________________
206 | CHAPTER 6: FOSSIL PRESERVATION

Figure 6.16 – Four fossils from El Capitan reef in Guadalupe Mountain National Park. a) Sponge (Porifera), b)
Fennestrate bryozoa, c) Polydiexidina fusulinids with foraminifera, d) Fragment of an ammonite. Image
credit: a) Jennifer Trout National Park Service public domain; b) Peter Scholle, used with permission; c) Image
from Verydeadthings CC; d) Richard Langford UT El Paso, CC.

d. In the Guadalupe Mountains, you can find thick carbonate layers with many fossils, such as in Figure
6.16:

i. What is the mineral? ____________________


ii. What is the mode of preservation for these fossils? Note, there could be more than one mode of
preservation. ____________________
iii. Were fluids involved in their preservation? If so, what was their composition?

iv. Why do you think this type of preservation is found in this portion of the reef?

v. Which part of the reef were these fossils found? Back reef, reef, fore reef or basin?
____________________

e. Critical Thinking: Summarize your observations about the modes of preservation in different parts of the
Permian reef system. Can you explain why the preservation is the same or different around the ancient
CHAPTER 6: FOSSIL PRESERVATION | 207

reef?

Additional Information

Exercise Contributions
Virginia Sisson and Daniel Hauptvogel

References
Casella, L.A., Griesshaber, E., Yin, X., Ziegler, A., Mavromatis, V., Müller, D., Ritter, A.-C., Hippler, D.,
HarperE.M/, Dietzel, M., Immenhauser, A., Schöne, B.R., Angiolini, L., and Schmahl, W.W., 2017, Biogeo-
sciences, 14, 1461–1492, doi:10.5194/bg-14-1461-2017.

Cooper, G.A., and Grant, R.E., 1964, New Permian stratigraphic units in Glass Mountains, West
Texas: American Association of Petroleum Geologists Bulletin 48: 1581-1588.

Cooper, G.A., and Grant, R.E. 1966. Permian rock units in the Glass Mountains, West Texas, In: Contribu-
tions to stratigraphy, 1966: U.S. Geological Survey Bulletin 1244-E: E1-E9.

Kerans, C., and Kempter, K., 2002, Hierarchical stratigraphic analysis of a carbonate platform, Permian
of the Guadalupe Mountains: The University of Texas at Austin, Bureau of Economic Geology (American
Association of Petroleum Geologists/Datapages Discovery Series No. 5), CD-ROM.

Norell, M.A., Weimann, J., Fabbri, M., Yu, C., Marsicano, C.A., Moore-Nall, A., Varricchio, D.J., Pol, D., and
Zelinitsky, D.A., 2020, The first dinosaur egg was soft. Nature, 583, 406-410, Published online June 17,
2020. doi: 10.1038/s41586-020-2412-8

Olszewski, T.D. and Erwin, D.H. 2009. Change and stability in Permian brachiopod communities from
western Texas. Palaios 24: 27-40.

Ritter, S., and Peterson, M., 2015, Interpreting Earth History: A Manual in Historical Geology, Eighth Edi-
tion, Waveland Press Inc., 291 pp.

Scholle, P.A., Goldstein, R.H., and Ulmer-Scholle, D.S., 2007, Classic Upper Paleozoic reefs and bioherms
of west Texas and New Mexico: A field guide to the Guadalupe and Sacramento Mountains of west Texas
and New Mexico. New Mexico Bureau of Geology and Mineral Resources Open File Report 504, 178 pp.

Tapanila, L., Pruitt, J., Pradel, A., Wilga, C.D., Ramsay, J.D., Schlader, R., and Didier, D.D., 2013, Jaws for a
spiral-tooth whorl: CT images reveal novel adaptation and phylogeny in fossil Helicoprion. Biology Letters,
9, 20130057. http://dx.doi.org/10.1098/rsbl.2013.0057
208 | CHAPTER 6: FOSSIL PRESERVATION

Google Earth Locations


CHAPTER 7: FOSSILS

Learning Objectives

The goals of this chapter are to:

• Recognize specimens of the most common invertebrate fossils


• Classify the phyla, order, and/or class associated with common fossils
• Compare and contrast symmetry in fossil specimens

Chapter Notes: The main text and much of the imagery from this chapter comes from the Digital Atlas
of Ancient Life (CC BY-NC-SA), which the Paleontological Research Institution manages. Exercises in this
chapter will heavily rely on sketching, so be sure to review the best practices for sketching from Chapter
0.

7.1 Introduction
Life on Earth has been around for a very long time, at least 3.5 byrs, and organisms have continually
evolved and become extinct during this time. Ever since the first fossil was found, paleontologists have
studied them to help decipher Earth’s history. Early Greek philosophers, such as Xenophanes (570-480
BC) and Herodotus (484-425 BC), recognized fossils as marine organisms. Since then, new discoveries are
being found all the time; these change the way we tell Earth’s stories. For several centuries, fossils were
the only tool geoscientists had to date rocks. Now, fossils are used in a variety of ways. Some of these
include determining the origins of life on Earth and possibly other planets, climate change, biodiversity,
changes in evolution, massive and small extinctions, and understanding deep time. Paleontology involves
science, art, and imagination; it inspires curiosity about ancient life and how the modern world evolved.
Paleontology is important to geobiology and is an interdisciplinary bridge between these two fields.

While most paleontologists are paid professionals, there are a large number of citizen scientists who
make important paleontological finds. Take the rock shop owner who showed Jack Horner a coffee can
filled with tiny fossil bones. He realized that these were fossilized baby dinosaurs. When the two of them
went out to the fossil location, Jack Horner discovered the first fossilized dinosaur eggs in the western
hemisphere. So, you never know if you’ll be the one to discover an interesting fossil that changes the way
we think about the Earth.

Recently, paleontologists have embraced the virtual world and computer-aided visualization to look at
the three-dimensional structures in fossils. From this, they can better look at the fossil’s internal struc-
ture. Much like when your doctor needs to look at your organs in detail, paleontologists use computed
tomography (CT) with multiple x-ray scans to look through a fossil.
210 | CHAPTER 7: FOSSILS

Figure 7.1 – Fossil snake CT scanning at Methodist Hospital in Houston, TX. Image credit: Houston Museum of
Natural Science, CC BY-NC-SA.

Using CT images also helps with fossil preparation as paleontologists often spend hours carefully remov-
ing rock matrix from a fossil. For example, if you have seen trilobite specimens, these are often prepared
by micro air blasting with talc to remove the host limestone. This can take from a few to tens of thousands
of hours. CT is much faster at digitally removing the rock matrix (Figure 7.2).
CHAPTER 7: FOSSILS | 211

Figure 7.2 – CT images of Peltosaurus granulosus (AMNH FR 8138), a lizard that lived in the Oligocene (White River
Fm., Wyoming) and was closely related to today’s Texas alligator lizard. The first image is a rendering of the
specimen with matrix in left lateral view, and the second image has the matrix digitally removed. This is an
excellent example of digital fossil preparation via CT, which takes a fraction of the time that physical preparation
takes and can be done without risk of damage to the specimen. The specimen is 56.4 mm long. Image credit:
University of Texas High-Resolution X-ray CT Facility and the American Museum of Natural History (AMNH),
CC-BY-NC.

Even with newer techniques, the basics of looking at fossils have remained constant for centuries. In fact,
some dismiss paleontology as a dead science because simple observations that are the backbone of pale-
ontology are not seen as a credible test you do in a laboratory or model made with advanced machine
learning. Paleontology is like a rosetta stone and can help you decipher the tapestry of Earth’s history
and become a better scientist. If you remember the skills you have been using, such as sketching, data
collection, and map interpretation, you’ll do well studying fossils.
212 | CHAPTER 7: FOSSILS

7.2 Taxonomy
Taxonomy sounds like a big word, but it’s just the term used to describe how scientists classify things,
which we’ve been doing throughout this lab manual (e.g. rock and minerals). The classification of fossils is
one of the meeting points between geology and biology, and we follow the traditional biology taxonomy
when classifying fossils (Figure 7.3). The major ranks are domain, kingdom, phylum, class, order, family,
genus, and species. For the fossils in this lab manual, we will focus primarily on the phylum rank, with
some organisms at the class and order rankings.

Figure 7.3 – The taxonomic ranks of organisms with the red fox as an example. Rankings start broad at the top with
Domain and become more specific toward Species. Image credit: Annina Breen, CC BY-SA.

Exercise 7.1 – Taxonomic Rankings

Using your phones, laptops, or tablets, look up the taxonomic ranking of the species in Table 7.1.
CHAPTER 7: FOSSILS | 213

Table 7.1 – Worksheet for Exercise 7.1

Ranking Human Bornean orangutan Dog Rice

Domain

Kingdom

Phylum

Class

Order

Family

Genus

Species sapiens pygmaeus lupus familiaris sativa

You can take an entire course dedicated to paleontology, so we will only cover the more common fossils
in the following pages. These fossils have detailed anatomies that are used to differentiate class, order,
family, genus, and species. In this chapter, we want you to be able to identify common fossils at a broad
level. A detailed study is reserved for upper-level classes. Therefore, for most of these organisms, you will
not see a detailed review of their anatomy.

Exercise 7.2 – Comparing Common Fossils

Your instructor will provide you with some common fossils. Create sketches of them and describe the differ-
ences that you can observe.

Sketch of first fossil: Sketch of second fossil:

a. What are the similarities and differences between these fossils?

b. Critical Thinking: What evidence from these fossils suggests that they may be organisms that are now
214 | CHAPTER 7: FOSSILS

extinct?

c. How would you respond to someone who argues that all fossils that are thought to be extinct are actu-
ally still alive but haven’t been “rediscovered” yet?

The geologic record is full of fossils, from dinosaurs to plants to fish and everything in between. Inverte-
brate animals from the marine environment are the most common branch of fossils you will find because
of their abundance and higher probability of fossilization versus land-dwelling organisms, and they will
be the focus of this chapter. Table 7.2 contains a list of the major marine invertebrate phyla we will cover,
and in some cases, the class and order.

Table 7.2 – Phyla, classes, and orders of organisms to know for this chapter

Phylum Class Order Common Names

Rugose (horn) coral,


Rugosa,
tabulate coral,
Cnidaria Anthozoa Tabulata,
scleractinian coral
Scleractinia
(stony or hard coral)

Bryozoa Bryozoan

Bivalvia, Cephalopoda,
Mollusca Clam, squid, snail
Gastropoda

Brachiopoda Brachiopod

Arthropoda Trilobita Trilobite

Echinoidea, Crinoidea, Sand dollar, crinoid,


Echinodermata
Blastoidea, blastoid, starfish

7.3 Symmetry
A helpful characteristic in identifying fossils is the symmetry of the organism. Symmetry is an observable
pattern in the external or internal structure of organisms that allows you to divide that organism into
roughly equal parts that are mirror images of each other. For example, take the face of a human being
which has a plane of symmetry down its center, an example of external symmetry. Internal features can
also show symmetry. For example, the tubes in the human body, such as your arteries and veins, are
cylindrical and have several planes of symmetry. Most multicellular organisms have some form of sym-
metry.

There are several types of symmetry in biology, but the main ones you will see are bilateral, radial,
and spherical (Figure 7.4). Bilateral symmetry is a single plane that divides the organism into two equal,
mirror-image halves. Radial symmetry has several subtypes, but they all describe symmetry lines drawn
through a central point. The common types of radial symmetry for fossils in this chapter are 5-fold radial
symmetry (pentamerous) and 6-fold radial symmetry (hexamerous). Spherical symmetry is when you can
CHAPTER 7: FOSSILS | 215

draw a line of symmetry through a central point of the organism from any direction so that no matter
what the orientation of your symmetry line is, you will always have two equal, mirror-image halves.

Figure 7.4 – Common types of symmetry. Image credit: Charl Hutchings, CC BY.

Exercise 7.3 – Identifying Symmetry in Fossils

Let’s get some practice identifying and describing the differences in symmetry.

a. Your instructor will provide you with a selection of fossils that exhibit bilateral, radial, spherical, or no
symmetry. On a separate sheet of paper, create a sketch of each fossil that highlights its symmetry. Be
sure to draw the line(s) of symmetry on your sketch and indicate which type of symmetry each fossil has.
b. Critical Thinking: Recently, Evans et al. (2020) described Ediacaran fossils of Ikaria wariootia (Figure 7.5)
found in South Australia as the first organisms with bilateral symmetry. Ikaria have also been linked to a
Nereites-type trace fossil. What would be the benefit of this type of symmetry for these organisms?
216 | CHAPTER 7: FOSSILS

Figure 7. 5 – Artist’s reconstruction of Ikaria wariootia


forming a trace fossil trail. Ikaria are ~5 mm long. Image
from Evans et al. (2020).

7.4 Phylum Cnidaria


Cnidaria is a phylum of marine organisms that includes over 11,000 species, including jellyfish, sea
anemones, and coral; most have radial symmetry. The soft bodies of jellyfish and sea anemones are
rarely preserved, whereas the hard structures of corals are quite commonly preserved as fossils. In the
geologic record, three orders of coral are the most abundant: rugose, tabulate, and scleractinian corals.
Corals are carnivores, catching zooplankton with their tentacles and pulling them into their mouths.
Some corals also have a partnership with green algae that live within the coral. As the algae perform pho-
tosynthesis, some of the energy they create is transferred to the coral.

Order Rugosa
Rugose corals (Figure 7.6) are an extinct order of coral that originated in the Ordovician and went extinct
at the end of the Permian. Members of Rugosa are sometimes called horn corals because solitary forms
frequently have the shape of a bull’s horn (if you like the Harry Potter movies, some say they look like the
sorting hat). Rugose corals do have a less common, colonial form that does not have this shape.
CHAPTER 7: FOSSILS | 217

Figure 7.6 – Two solitary rugose corals in a slab of Ordovician limestone from near Cincinnati, Ohio. Note the
similarity of the left specimen’s shape to that of a bull’s horn. Image credit: Digital Atlas of Ancient Life, CC
BY-NC-SA.

Rugose corals reached their peak diversity during the Devonian period when colonial forms were impor-
tant reef builders (Figure 7.7). They went extinct during the end-Permian mass extinction event. Figure 7.7
is the first of many diversity curves you will see in this chapter. Diversity curves allow paleontologists to
graph how many different kinds of fossils existed on earth during different geological time intervals. They
are especially important for telling us (a) when a group of animals first evolved on earth, (b) when that
group diversified (increased in diversity) and became common on earth, and (c) when that group went
extinct.
218 | CHAPTER 7: FOSSILS

Figure 7.7 – Diversity of Rugosa genera. Image credit: Paleobiology


Database, CC BY.

Rugose corals may be either solitary or colonial (Figure 7.8). A solitary coral individual is called a corallum,
while an individual within a colony is called a corallite. Rugose corals made their skeletons from calcite;
this is a significant difference relative to other corals that make their skeletons out of aragonite, which is
more resistant to stress.

The outer part of the corallum (or corallite) is called the theca; think of it as the skeletal wall (Figure 7.8).
The theca may, in turn, be covered by an outermost skeletal sheath called the epitheca. Growth lines are
often apparent on the epitheca; these are also called rugae (“ruga” is Latin for wrinkled), which gives this
group of corals their scientific name (thus, rugose corals are the “wrinkled corals”). Finally, the top of the
corallum (or corallite) is called the calice, and it is the portion of the coral occupied by the living organism,
called a polyp.
CHAPTER 7: FOSSILS | 219

Figure 7.8 – Major features of solitary and colonial rugose corals; labeled features include a corallum, corallites,
epitheca, calices, and growth lines. Left: Heliophyllum halli from the Middle Devonian Moscow Fm. of Erie County,
New York (PRI 70755). Right: Acrocyathus floriformis from the Mississippian St. Louis Limestone of Monroe County,
Illinois (PRI 70756). Both specimens are from the collections of the Paleontological Research Institution, Ithaca,
New York. Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.

The surface of the calice is covered with radiating, vertical structures called septa (singular = septum) that
resemble the spokes of a bicycle wheel (Figure 7.9). Septa sometimes extend through the theca, forming
vertical lines called costae (singular = costa). The septa radiate outwards from a central pillar-like struc-
ture called the columella. In rugose corals, the septa are added during growth in a manner that ultimately
divides them into four quadrants. Spaces between the four quadrants are called fossula. However, it is
important to note that fossula are not readily visible in many specimens, especially when they are poorly
preserved.
220 | CHAPTER 7: FOSSILS

Figure 7.9 – Important features of rugose corals, including a fossula, septa, the columella, and costae. Left: solitary
rugose coral Heliophyllum canadense from the Devonian Onondaga Limestone of Mendon, New York (PRI 76809).
Right: solitary rugose coral Lophophyllidium proliferum from the Pennsylvanian Graham Formation of Coleman
County, Texas (PRI 76802). Both specimens are from the collections of the Paleontological Research Institution,
Ithaca, New York. Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.

As a coral polyp grows upwards, it develops horizontal partitions called tabulae (Figure 7.10).

Figure 7.10 – Tabulae observed in hand samples of two solitary rugose corals. Left: Siphonophrentis halli from the
Devonian Ludlowville Fm. of Skaneateles Lake, New York (PRI 76806). Right: Amplexizaphrentis pellansis from the
Mississippian Pella Beds of Pella, Iowa (most epitheca is eroded away) (PRI 76803). Image credit: Digital Atlas of
Ancient Life, CC BY-NC-SA.
CHAPTER 7: FOSSILS | 221

3-D Model 7.1 – Rugose Coral

This is a model of the rugose coral Heliophyllum halli from the Middle Devonian from Moscow. The growth
lines and costae are clearly visible on the epitheca. Some septa are the only visible features on the calice.

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://uhlibraries.pressbooks.pub/historicalgeologylab/?p=84

Heliophyllum halli (PRI 70755) by Digital Atlas of Ancient Life on Sketchfab


More 3-D models of rugose corals can be found at the Digital Atlas of Ancient Life.

Order Tabulata
Tabulate corals originated in the Early Ordovician period and went extinct at the end of the Permian
period. All tabulate corals were colonial, and some species were important reef makers during the Sil-
urian and Devonian periods. Tabulate corals constructed their skeletons primarily of calcite. Some tab-
ulate corals look superficially like honeycombs (e.g., Favosites; Figure 7.11a), while others look like chain
links (e.g., Halysites; Figure 7.11b) or collections of narrow tubes (e.g., Syringopora; Figure 7.11c). Others can
be encrusted upon other marine invertebrates (including other corals).

Figure 7.11 – Examples of three genera of tabulate coral. a) Favosites in dolostone from the Silurian of Michigan.
Favosites have a distinctive honeycomb shape. b) Halysites corals look like chain links. c) Syringopora corals look like
narrow tubes, from the Lower Carboniferous Boone Limestone near Hiwasse, Arkansas. Image credit: a) James St.
John, CC BY; b) Wikimedia user Wilson44691, Public domain; c) Wikimedia user Wilson44691, Public domain.

The following video from the Paleontological Research Institute gives a great introduction to tabulate
corals.
222 | CHAPTER 7: FOSSILS

A YouTube element has been excluded from this version of the text. You can view it online here: https://uhlibraries.press-
books.pub/historicalgeologylab/?p=84

According to the Paleobiology Database, there are 58 families of tabulate corals, 376 genera, and 511
species. As demonstrated by the genus-level plot of their diversity through time (Figure 7.12), tabulate
corals suffered badly during the Late Devonian extinction event and never again reclaimed their former
levels of diversity. The group went extinct at the end of the Permian period.
CHAPTER 7: FOSSILS | 223

Figure 7.12 – Diversity of tabulate corals. Image credit: Paleobiology


Database, CC BY.

A defining feature of most tabulate corals are structures called tabulae, which give them their name.
Tabulae (singular, tabula; from the Latin for board or tablet) are horizontal plates that span across indi-
vidual corallites (Figure 7.13). They are deposited by polyps as they grow, separating the living animal
from the space(s) occupied earlier in life. Remarkably, we know something about the soft polyps of tab-
ulate corals due to the discovery of calcified polyps in specimens of Silurian Favosites from the Jupiter
Formation of Quebec, Canada. This discovery was reported by Copper (1985), who reported that individ-
ual corallites typically had 12 tentacles, though some had 11 or 13. The discovery of these polyps also
confirmed that Tabulata are indeed Cnidarians, rejecting the hypothesis of some earlier workers that this
group belonged with the sponges.

Unlike rugose and scleractinian corals, most tabulate corals did not have septa. Therefore, as a general
rule, identifying whether or not a specimen of colonial Paleozoic coral has septa is a good indication of
whether it is a rugose coral (septa always present) or a tabulate coral (septa usually absent).
224 | CHAPTER 7: FOSSILS

Figure 7.13 – Corallites and tabulae identified on a specimen of Favosites favosus. Note the absence of septa. The
specimen is from the collections of the Paleontological Research Institution, Ithaca, New York. Image credit: Digital
Atlas of Ancient Life, CC BY-NC-SA.

3-D Model 7.2 – Tabulate Corals

This is a model of the tabulate coral Favosites favosus from the Silurian of Delaware County, Iowa (PRI 76737).
It is the same one pictured in Figure 7.12. The length of the specimen is approximately 10 cm. The specimen
is from the collections of the Paleontological Research Institution, Ithaca, New York. The side view shows the
tabulae, and the top view shows the corallites.

An interactive or media element has been excluded from this version of the text. You can view it online here:
https://uhlibraries.pressbooks.pub/historicalgeologylab/?p=84

Tabulate coral: Favosites favosus (PRI 76737) by Digital Atlas of Ancient Life on Sketchfab
More 3-D models of tabulate corals can be found at the Digital Atlas of Ancient Life.

Order Scleractinia
The order Scleractinia includes the “true corals” or “stony corals,” with about 1,500 living species today.
CHAPTER 7: FOSSILS | 225

Scleractinians first appeared in the early Middle Triassic (Figure 7.14) and have been the dominant reef-
building organisms over the past 240 million years.

Figure 7.14 – Diversity of Scleractinia genera. Image credit:


Paleobiology Database, CC BY.

Scleractinian corals may be either solitary or colonial in form and always have skeletons composed of
aragonite. We are going to focus on the colonial forms of Scleractinia corals. An entire colonial coral
colony is a corallum; the space occupied by a single polyp within the colony is called a corallite (Figure
7.15). Individual corallites have their own radial septa, which often attach to a centralized pillar called the
columella; all familiar terms from rugose and tabulate corals.
226 | CHAPTER 7: FOSSILS

Figure 7.15 – Fossil specimen of the colonial scleractinian coral Astrangia sp. showing the corallites, septa, and
columella. The specimen is from the Late Pleistocene Ft. Thompson Fm. of Hillsborough County, Florida (PRI
76863). The specimen is from the research collections of the Paleontological Research Institution, Ithaca, New York.
Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.

The skeletons of colonial corals exhibit a much greater diversity of forms than those of solitary species.
Often, their growth forms are dependent upon the habitats in which they live. For example, robust,
rounded colonies are often favored in high-energy habitats with lots of wave action (Figure 7.16a), while
delicate branching forms are usually associated with quieter environments (Figure 7.16b).

Figure 7.16 – Two types of Scleractinia corals. a) Stacked fossil specimens of the colonial scleractinian coral
Siderastrea pliocenica from the Plio-Pleistocene of Lee County, Florida (PRI 76859). The specimen is from the
research collections of the Paleontological Research Institution, Ithaca, New York. b) Fossil specimen of the colonial
scleractinian coral Oculina sarasotana from the Plio-Pleistocene of Lee County, Florida (PRI 76894). The specimen is
from the research collections of the Paleontological Research Institution, Ithaca, New York. Image credit: Digital
Atlas of Ancient Life, CC BY-NC-SA.

7.5 Phylum Bryozoa


Bryozoans are filter-feeding invertebrates found in freshwater and marine habitats, where they are often
easy to miss because of their small size (Figure 7.17). In almost all species, tiny (<1 mm diameter) bry-
ozoan individuals, called zooids, live together as a colony that often encrusts surfaces, grows branch-
ing structures, or, in freshwater species, forms a gelatinous blob. They have bilateral symmetry. Despite
CHAPTER 7: FOSSILS | 227

their small size, marine bryozoans are abundant in many fossil assemblages thanks to the preservation
of their hard calcium carbonate skeletons. Bryozoans first appeared in the Ordovician and are still pre-
sent in oceans today (Figure 7.18). They were greatly impacted by several mass extinctions with their
lowest diversity occurring during the Mesozoic. More than 17,800 species of fossil bryozoans have been
described, and more than 6,000 living species are known.

Figure 7.17 – A variety of bryozoan fossils from an Ordovician oil shale from Estonia. Image credit: Mark A. Wilson,
Public Domain.
228 | CHAPTER 7: FOSSILS

Figure 7.18 – Diversity of Bryozoa genera. Image credit: Paleobiology


Database, CC BY.

Bryozoan colonies have highly variable forms (Figure 7.19), from gelatinous blobs to upright branching
structures and sheet-like encrusters, but the general morphology of a zooid is similar across the classes.
Zooids can take several forms, but the most common forms in each class are autozooids, which function
in feeding the colony and excreting waste. Branching bryozoans may look similar to branching corals, but
the zooids in bryozoa do not have septa or a columella as corallites do in corals.
CHAPTER 7: FOSSILS | 229

Figure 7.19 – Common bryozoan fossils. a) fenestrate; b) branching; c) trepostome; d) archimedes. Image credit:
compiled from James St. John, CC BY.

7.6 Phylum Mollusca


Mollusks are a phylum of marine invertebrates that comprise about 23% of all marine organisms. There
are several classes of mollusks, but we will focus on gastropods, cephalopods, and bivalves. All have bilat-
eral symmetry, but it’s more complicated for gastropods.

Class Gastropoda
Gastropods are one of the most evolutionarily successful groups of animals; you may commonly know
them as snails and slugs. They occupy the world’s oceans, freshwater lakes and streams, and terrestrial
ecosystems, including many backyards. Some are algae-eating herbivores, while others are venomous
hunters of fish. Their strong, single-valved shells (Figure 7.20) have left behind a rich Cambrian to Recent
230 | CHAPTER 7: FOSSILS

fossil record and are the focus of many paleobiological studies (Figure 7.21). These shells are primarily
made of calcium carbonate. Gastropods have many different shell shapes, and some with no shell, but
one of the defining characteristics of gastropods is the rotation of their body during their development,
a process called torsion. Torsion does not refer to the coiling shells you are familiar with and only relates
to the gastropod’s coiled body. Before rotation, gastropods exhibit bilateral symmetry, but they lose that
symmetry as they mature.

Figure 7.20 – A variety of gastropod shell shapes. Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.

Figure 7.21 – Diversity of Gastropoda genera. Image credit:


Paleobiology Database, CC BY.

Class Cephalopoda
Cephalopods are marine organisms with bilateral body symmetry, a prominent head, and a set of tenta-
CHAPTER 7: FOSSILS | 231

cles. You may be familiar with some of them, like squids and octopuses. While most modern cephalopods
are completely soft-bodied or have only thin internal shells, their ancient shelled cousins left behind a
detailed fossil record with a remarkable diversity of shell shapes. Cephalopods became abundant during
the Ordovician Period, and many went extinct by the end of the Mesozoic Era with the dinosaurs (Figure
7.22).

Figure 7.22 – Diversity of Cephalopoda orders. Image credit:


Paleobiology Database, CC BY.

One of the most well-known index fossils for cephalopods is the belemnite (Figure 7.23). The belemnite
is an extinct order of mollusk that only lived during the Mesozoic. It had a characteristic elongated, cone-
shaped shell called a guard (Figure 7.23b). The guard is thought to have been a counterbalance that
allowed the belemnite to move horizontally through the water.

Figure 7.23 – Belemnite fossils. a) The entire organism, including the guard and carbonized and preserved soft
parts. The cone-shape on the left side is the guard. Please note it is extremely rare to find any fossilization of the
soft parts. b) Belemnite guard. Image credit: a) Wikimedia user Ghedoghedo, CC BY-SA; b) California Academy of
Sciences, CC BY-NC-ND.

Other well-known cephalopod index fossils are from the subclass Ammonoidea, commonly referred to
232 | CHAPTER 7: FOSSILS

as ammonites (Figure 7.24). They are extinct mollusks that lived from the Devonian to the end of the Cre-
taceous (Figure 7.25). These are ideal for biostratigraphy as there are over 8,000 species of ammonoids,
and they occur over a large geographic range.

Figure 7.24 – Examples of different shapes of ammonoid shells. Image credit: Digital Atlas of Ancient Life, Jonathan
R. Hendricks, CC BY-NC-SA.

Figure 7.25 – Diversity of Ammonoidea genera. Image credit:


Paleobiology Database, CC BY.

One of the ways to distinguish orders of ammonoids is the suture pattern of their shells (Figure 7.26).
Ammonoid sutures fall into three main groups: goniatites, ceratites, and ammonites. As you go from goni-
atites to ceratites to ammonites, the suture patterns become more complex. Goniatitic sutures do not
have subdivided saddles or lobes. Saddles are convex toward the opening of the shell, and lobes are con-
CHAPTER 7: FOSSILS | 233

vex in the other direction. Ceratitic sutures have subdivided lobes but undivided saddles. Finally, both the
saddles and lobes of ammonitic sutures are subdivided, sometimes in an amazingly complex fashion.

One of the oft-quoted principles of evolutionary theory is, “ontogeny recapitulates phylogeny.” This a his-
torical hypothesis from Ernst Haeckel that states the development of the embryo of an animal, from fer-
tilization to gestation or hatching (ontogeny), goes through stages resembling or representing successive
adult stages (recapitulates) in the evolution (phylogeny) of the animal’s remote ancestors. It means that
as an animal develops, different stages of growth may give you hints of the evolution that their ancestors
went through. Modern biology is full of examples of this theory.

Well, for the fossil record, this theory is hard to prove because of incomplete preservation. One of the
best examples in the fossil record that supports this quote is the suture patterns in ammonite shells.
These sutures form at the intersection of the shell wall and the interior septa (walls dividing the interior
cavity). These patterns started as mere wavy lines in the Cambrian and became more complex through
time. Through time, the sutures acquired additional saddles and lobes. Later, paleontologists recognized
that within some species, such as a Permian ammonoid (Perrinites), the sutures are simple patterns in
young animals and get more complex as the animal gets older. Thus, the suture patterns in the young
animals resemble their ancestors.

Figure 7.26 – Examples of ammonoid order sutures: goniatites, ceratites, and ammonites. Image credit: Digital
Atlas of Ancient Life, Jonathan R. Hendricks, CC BY-NC-SA.

Exercise 7.4 – Stratigraphic Analysis and Fossil Evolution

On a class field trip to the Guadalupe Mountains in west Texas (Exercise 6.4), you are taken to a stratigraphic
section that goes through the Carboniferous to Permian up to the Capitanian extinction, which was an extinc-
234 | CHAPTER 7: FOSSILS

tion event that occurred in the late Permian about 260 Ma. Your field trip leader has given you a guide to some
of the important ammonite species used for relative age dating in this area. In addition, she has given you the
patterns for several ammonite species. You prefer looking at fossils compared to stratigraphic sections and
have found two fossils just randomly on the ground. Can you place these on the stratigraphic column?

Figure 7.27 – Suture patterns for Carboniferous to Permian ammonoids,


specifically for Goniatitida (goniatites). The arrow bisects the umbilicus and
points towards the suture of the ammonoid. Image Credit: VB Sisson adapted
the stratigraphic column from Leonova (2018) with unknown ammonites from
Miller and Furnish (1940) CC.

a. Describe how the sutures of species in this stratigraphic column evolve through time.

b. Based on the evolution of the suture patterns, place these two unknown fossils in the column. You can
sketch the sutures in the blank boxes.
c. Explain your answer.

Class Bivalvia
Bivalves refer to a diverse Mollusca class that typically has bilateral symmetry where the upper shell and
inner parts of the organism are a mirror image of the bottom half. Clams, oysters, and mussels belong
CHAPTER 7: FOSSILS | 235

to this class. Their shells are made of calcium carbonate and typically have growth lines (Figure 7.28).
Bivalves first appeared in the fossil record in the Cambrian and are still around today (Figure 7.29).

Figure 7.28 – Examples of bivalves. a) Placopecten clintonius (a type of


scallop) from the Pliocene of North Carolina; b) Exogyra costata (a type of
oyster) from the Cretaceous of South Carolina; c and d) two views of
Idonearca vulgaris (a type of clam) from the Cretaceous of Tennessee; e)
Rastellum carinatum (a type of oyster) from the Cretaceous of Texas.
Image credit: all images from James St. John, CC BY.
236 | CHAPTER 7: FOSSILS

Figure 7.29 – Diversity of Bivalvia genera. Image credit: Paleobiology


Database, CC BY.

7.7 Phylum Brachiopoda


Brachiopods are shelled, filter-feeding marine organisms (Figure 7.30) that inhabit the seafloor and come
in various shapes and sizes. They have been around since the Cambrian with incredible diversity during
the Paleozoic Era (Figure 7.31). Brachiopods are still around today, but their diversity is greatly dimin-
ished.

Figure 7.30 – Examples of brachiopods. Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.
CHAPTER 7: FOSSILS | 237

Figure 7.31 – Diversity of Brachiopoda genera. Image credit:


Paleobiology Database, CC BY.

Superficially, brachiopods may look like bivalves, but the two are not related. One of the biggest differ-
ences between brachiopods and bivalves lies in their symmetry. Both have bilateral symmetry, but the
plane of symmetry in brachiopods is vertical rather than horizontal (Figure 7.32). This means that the left
half of a brachiopod is a mirror image to the right half. In bivalves, the plane of symmetry is horizontal,
and the upper half is a mirror image of the lower half.

Figure 7.32 – Differences in symmetries between a brachiopod (left) and a bivalve (right). Image credit: Jaleigh Q.
Pier, CC BY-SA.

Brachiopod shells have two valves that are distinct in both size and shape (Figure 7.33). The brachial valve
238 | CHAPTER 7: FOSSILS

is usually the smaller of the two valves and has a ridge, or fold, down the middle of the valve. The pedicle
valve is usually larger than the brachial valve and has a valley down the middle. It also has a hole through
which the pedicle passes. The pedicle is a fleshy, stalk-like feature which some groups use to attach them-
selves to hard rocky seafloor. It is rarely preserved, but brachiopod shells will often have a pedicle open-
ing preserved along the hinge-line that varies in shape from rounded to triangular.

Figure 7.33 – Brachiopod morphology. Image credit: Jaleigh Q. Pier, CC BY-SA.

7.8 Phylum Arthropoda


Arthropods are an extremely diverse and abundant phylum of animals. They include organisms such
as spiders, scorpions, crabs, barnacles, and insects. Unfortunately, most of these are not abundant in
the fossil record because of poor preservation, but one class of arthropods does stand out in the fossil
record, the trilobite.

Class Trilobita
Trilobites are among the most well-known fossils, thanks largely to their abundance, diversity, and broad
distribution during the Paleozoic (Figure 7.34). These emblematic arthropods first appeared in the fos-
sil record 521 million years ago and survived until the end-Permian extinction approximately 250 million
years ago (Figure 7.35). Even when they first appeared, trilobites were diverse and widespread, suggest-
ing a long history before the first fossils, likely extending as far back as 700 Ma. During the Cambrian,
trilobites were a dominant marine ecosystem component, but many of the Cambrian trilobite species
died out during the extinction at the end of the Ordovician. Though the class did survive this extinction,
their diversity never matched their earlier history. Because they evolved rapidly and were relatively wide-
spread, trilobite species are often used to provide relative ages of the rocks in which they are found.
CHAPTER 7: FOSSILS | 239

Figure 7.34 – Trilobite fossils. Image credit: Digital Atlas of Ancient Life, CC BY-NC-SA.

Figure 7.35 – Diversity of Trilobite families. Image credit: Paleobiology


Database, CC BY.

Trilobites get their name because their bodies exhibit three distinct, longitudinal sections, and they also
have three segments from “head to toe” (Figure 7.36) and have bilateral symmetry. They fed in several
240 | CHAPTER 7: FOSSILS

ways, including moving across the ocean floor as predators, scavengers, or filter feeders, and some swam
to feed on plankton. They ranged in size up to 45 centimeters long (about 1.5 feet).

Figure 7.36 – Trilobites have three major body sections, from anterior to posterior, called
the cephalon (1), thorax (2), and pygidium (3). The body can also be divided into three
longitudinal lobes, called the right pleural lobe (4), axial lobe (5), and left pleural lobe (6).
Image credit: Sam Gon III, CC0 Public Domain.

7.9 Phylum Echinodermata


Echinoderms represent the largest marine animal phylum and include starfish, sea cucumbers, sand dol-
lars, blastoids, and crinoids. They first appeared in the fossil record in the Cambrian and are still around
today. One of the key features of this group is five-part, radial symmetry (Figure 7.37). There are many
classes of echinoderms, and we will focus on Echinoidea, Crinoidea, and Blastoidea.
CHAPTER 7: FOSSILS | 241

Figure 7.37 – Five-part, radial symmetry of two echinoderms. Image credit: Jaleigh Q. Pier, CC BY-SA.

Class Echinoidea
Echinoids are the most diverse echinoderm class (Figure 7.38). They include spiny-looking sea urchins
(regular echinoids) and organisms commonly known as sand dollars (irregular echinoids). Regular echi-
noids live on the ocean floor and can be herbivores (eating kelp and algae) or carnivores (eating bry-
ozoans). Irregular echinoids live in the sediment on the ocean floor and eat tiny food particles in that
sediment. All modern echinoids have a hard, calcareous, internal skeleton.
242 | CHAPTER 7: FOSSILS

Figure 7.38 – Examples of Echinoidea. a) a living sea urchin; b) sea urchin fossil; c) sea urchin fossil from the
Caribbean; d) sand dollars. Image credit: a) S. Rae, CC BY; b) Marian Garcia, CC BY-NC-ND; c) Tim Sackton, CC BY-SA;
d) Rachel Hathaway, CC BY.

Echinoids first appear in the fossil record in the Upper Ordovician, approximately 460 million years ago,
but they didn’t really diversify until the Jurassic (Figure 7.39).
CHAPTER 7: FOSSILS | 243

Figure 7.39 – Diversity of Echinoidea genera. Image credit:


Paleobiology Database, CC BY.

Class Crinoidea
Crinoids, often referred to as “sea lilies,” may resemble plants (Figure 7.40), but they are actually suspen-
sion-feeding animals that have been around since the Ordovician (Figure 7.41). They use their arms to
catch floating food particles and transfers them to the base of their crown. The crinoid “stem” contains
numerous ring-like elements made of magnesium-rich calcite and is held together by a combination of
ligaments and muscles. The stem of crinoids is most often found in the geologic record (Figure 7.42). The
crown resembles a flower, and this soft tissue is rarely fossilized.

Figure 7.40 – a) Basic anatomy of a crinoid; b) A crinoid fossil from the Permian; c) A living crinoid from Sumilon
Island, Philippines. Image credit: a) William I. Ausich, CC BY; b and c) Klaus Stiefel, CC BY-NC.
244 | CHAPTER 7: FOSSILS

Figure 7.41 – Diversity of Crinoidea families. Image credit: Paleobiology


Database, CC BY.

Figure 7.42 – Crinoid stems in limestone from the Ordovician of Kentucky. Image credit: James St. John, CC BY.
CHAPTER 7: FOSSILS | 245

Class Blastoidea
Blastoids are an extinct class of echinoderm. Similar to crinoids, blastoids had a stalk and were suspen-
sion feeders. The main body is the theca, which was protected by interlocking plates made of calcium
carbonate (Figure 7.43). Blastoids evolved in the Ordovician and went extinct during the End Permian
extinction event (Figure 7.44).

Figure 7.43 – Blastoid from the Lower Carboniferous


of Illinois. Image credit: Wikimedia user
Wilson44691, CC0 Public Domain.

Figure 7.44 – Diversity of Blastoidea genera. Image credit:


Paleobiology Database, CC BY.
246 | CHAPTER 7: FOSSILS

Exercise 7.5 – Fossil Identification

To help you identify these fossils, here is a flow chart with some of the observations that may help you decide
the phylum and class for the specimens.

a. The flow chart has a few blanks. Fill in the blanks with the correct word.
CHAPTER 7: FOSSILS | 247

Figure 7.45 – Flow Chart for Fossil Identification. Each color represents different types of observations you should
make, with the yellow boxes listing the phylum and class for each type of fossil. You need to fill in a few blanks to
be able to use this flow chart. Image credit: VB Sisson.

b. Your instructor will provide you with a selection of fossils for you to identify. On a separate sheet of
paper, make detailed sketches for each fossil from at least two different viewpoints. Be sure to label any
identifiable parts and then identify the fossils.
248 | CHAPTER 7: FOSSILS

7.10 Extinctions
Over Earth’s history, many species have gone extinct. Extinction is defined as the death of the last individ-
ual of a species. You probably already know about the famous extinction event at the end of the Creta-
ceous that killed all the dinosaurs and many other species. But, have you ever wondered how and when
such enormously successful animals went extinct? Are we undergoing a mass extinction right now? If you
want to know more, Kolbert (2014) offers a good introduction to extinctions and speculates that the Earth
is undergoing a sixth massive extinction, the Holocene extinction event caused by human activity.

Some estimate that 99% of all species (~5 billion species) that ever lived on Earth are now extinct. Massive
extinction events are rare, whereas local extinctions are quite common.

Figure 7.46 – Plot of extinction intensity (percentage of marine genera that are present in each interval of time but
do not exist in the following interval) versus time. Geological periods are annotated by abbreviations above. This
shows five major and two of the 19 minor extinctions since the beginning of the Cambrian period, including
Ordovician-Silurian (O-S), Devonian (D), Permian-Triassic (P-Tr), Triassic-Jurassic (Tr-J), and Cretaceous-Paleogene
(K-Pg) with two minor extinctions Cambrian-Ordovician (Cm-O) and Capitanian event (Cap). Image credit: VB Sisson
adapted from Rohde and Muller (2005).

Paleontologists postulate that there have been five major mass extinctions in Earth’s history (Figure 7.46);
these have had an enormous impact on life by quickly killing large numbers of species from land and
sea environments. All of these mass extinctions were caused by major environmental catastrophes. The
most famous example was the end of the Cretaceous mass extinction caused by a meteorite that crashed
to Earth 65 million years ago. It threw so much dust into the atmosphere that the Earth felt like night-
time for many years, preventing algae, plants, and animals from being able to get enough food to survive.
The Late Permian mass extinction was the most devastating of all extinctions. Warming of the Earth’s cli-
mate and changes in ocean chemistry caused by the outpouring of basaltic lava in Siberia may have been
responsible for this extinction. These two mass extinctions so dramatically affected life on earth that they
are used to define the boundaries surrounding the Paleozoic, Mesozoic, and Cenozoic eras.
CHAPTER 7: FOSSILS | 249

Exercise 7.6 – Extinction Exercise

The Paleobiology Database (PBDB), an online resource for fossil research, will help you investigate your
species extinction. Open the PBDB Navigator. The Navigator page has four parts: 1) The map (in the center)
shows continents with dots representing fossil data. The color of these dots represents their geologic age. If
you click on the dots, you can see all of the information on each fossil locality and the plants, animals, and
other fossil life that occur there. 2) The geologic time scale (on the bottom) shows the major eras and peri-
ods (and their subdivisions called stages). If you click on the timescale, the map will show you the location of
all known fossil localities from that time interval. 3) The most prevalent taxa (on the right) is shown, such as
Mollusca, Brachiopoda, etc. 4) The toolbar (on the left) showing the tools you can use to explore the database.
Hover over each tool to see its use. This exercise is inspired by Phil Novack-Gottshall, Benedictine University
Evolution of Extinct Animals.

a. Select a group of extinct animals you want to study for this exercise. Your instructor may provide this for
you. Which group of extinct animals are you studying? ____________________
b. On the upper-right corner, type in the name of your extinct group. Next, make a diversity curve for your
group by clicking the tool on the left side. You will see a new window show up. The default settings are
fine to use. Draw a copy of this graph in the space below.

Sketch Area:

c. The y-axis plots genus diversity, the number of different genera of animals in the fossil record during
each time interval. Describe features of your diversity curve through time. Include features such as when
the largest diversity was observed, when your animal group first evolved, and what time period your
animal group went extinct? (If your group is not extinct yet, then write that it is not yet extinct.) Did your
taxa simply increase in diversity and then suddenly extinct? Or was their evolutionary trends have sev-
eral periods of increased diversity?
250 | CHAPTER 7: FOSSILS

d. Some people think extinct animals were evolutionary failures because they are extinct today, or nearly
so. Some also claim they were never successful or common animals on earth. Based on the diversity
curve, is it fair to claim that your extinct group was a failure? Why or why not? (For comparison, primates
have existed for just 16 million years and our species for only 250,000 years.)

e. How were your taxa affected by one of the five major mass extinction events in Earth’s history? For each
mass extinction, explain whether it (a) had not yet originated, (b) was not affected by the extinction
(meaning it did not have a major drop in diversity at this time), (c) was heavily affected by the extinction
(meaning diversity drops significantly at this time), or (d) it went extinct before the mass extinction.

f. Which mass extinction was the most catastrophic for your group? Explain your answer.

g. If your animal group is extinct today, did its extinction occur during a mass extinction? Explain your
answer.

Additional Information

Exercise Contributions
Virginia Sisson and Daniel Hauptvogel

References
Evans, S.D., Hughes, I.V., Gehling, J.G., and Droser, M.L., 2020, Discovery of the oldest bilaterian from
the Ediacaran of South Australia, Proceedings of the National Academy of Sciences, 117 (14) 7845-7850
doi.org/10.1073/pnas.2001045117

Kolbert, E., 2014, The Sixth Extinction: An Unnatural History, Henry Holt and Co., New York City, 319 p.
CHAPTER 7: FOSSILS | 251

Leonova, T.B., 2018, Paleozoic Ammonoids: Historical Pathways of the Development of Morphological
Diversity, Paleontological Journal, 52, 1746-1755, DOI: 10.1134/S0031030118140113

Miller, A.K., and Furnish, W.M., 1940, Permian ammonoids of the Guadalupe Mountains region and adja-
cent areas, Geological Society of America Special Paper, 26, 242 pp.

Rohde, R.A. and Muller, R.A., 2005, Cycles in fossil diversity, Nature, 434, 209-210. doi:10.1038/
nature03339

Google Earth Locations


CHAPTER 8: PALEOENVIRONMENTS

Learning Objectives

The goals of this chapter are to:

• Categorize fossil assemblages to determine their habitats


• Develop connections between modern and ancient paleoenvironments
• Evaluate paleoenvironments through time and space

8.1 Introduction
Some geologists use fossils to determine stratigraphy and correlate stratigraphic sequences. Others pre-
fer to study paleoecology to determine ancient habitats where both plants and animals co-existed. This
avenue leads to discoveries about past climate, predator-prey relationships, and even ocean depth. Using
this information, artists create reconstructed paleoenvironments (Figure 8.1) to display these ancient
habitats.

Figure 8.1 – Using only fossil pollen, here is a series of reconstructed paleoenvironments for Cretaceous through
Paleogene time in Patagonia. Before the Cretaceous-Paleogene mass extinction, there was diverse plant flora (left
side). After the meteorite impact near Chicxulub, Mexico, most of the species died (middle). Then as plants started
to rebound, the fauna changed (right side). Image credit: Barreda et al. (2012) CC BY.

Remember, fossils are not just exciting curiosities but were once living organisms that reproduced, ate
each other, and interacted with their environment just like we do. A fossil’s paleoecology depends on
many factors of their paleoenvironment. These include physical conditions, such as temperature, light,
water depth, and energy (calm vs. turbulent), and chemical conditions, such as salinity, pH, oxygen, and
toxic chemicals. Some animals will thrive if these variables change, and others will die off. Have you ever
been to the beach and seen large piles of seaweed? They came from the middle of the ocean and were
brought to shore as ocean currents and winds changed. Some organisms can survive environmental
change, such as oysters that can live in both salty and freshwater. Fossils hold the key to determining
what the paleoenvironment was like. This ability to reconstruct a fossil’s paleoenvironment is one of the
most valuable tools that a paleontologist has to offer to broader fields of geology.

Understanding every detail of how a fossil lived is not necessary to help you use fossils to interpret an
ancient environment in the rock record. For example, suppose you know that a phylum is exclusively
CHAPTER 8: PALEOENVIRONMENTS | 253

marine or exclusively terrestrial. In that case, this can easily change your interpretation of an otherwise
dull sedimentary layer. This is just the beginning of using fossils to determine the paleoenvironment; if
you know how it lived, you can determine where it lived.

For any fossil group, ask yourself some basic questions such as the type of the skeleton, its body plan
(symmetry), its habitat, and how it fed and behaved. You might find these are hard to determine with
fossils but give it a try. Table 8.1 is a list of characteristics that you can use for determining its paleoenvi-
ronment:

Where is the skeleton relative to the tissue, as this can be internal, external, or none? Also, look at the
number of parts in the skeleton and its construction. This can be either a single element or multiple ele-
ments like bivalves (two-part), multivalve (range of 3 to ~15 parts), plates, ossicles or bones (many parts
from 10’s to 100’s), or spicules (very many small parts numbering up to 1,000’s).

Symmetry is similar to what you described in the last chapter but now consider how it relates to how
the organism lived. Bilateral symmetry means that both locomotory and sensory organs can be placed
efficiently. In contrast, organisms with radial symmetry tend to be slow or immobile (sessile). They also
respond to stimuli from all directions. Organisms with spherical symmetry will live in open water. Finally,
those with no symmetry are typically immobile (sessile).

Table 8.1 – Comparison of skeleton types and symmetry between marine fossils. *In colonial forms,
symmetry is difficult to identify because of how close together the organisms grow.

Skeleton Type Number of Parts in Skeleton Symmetry

Plates
Single
Organism Internal External Multi-Element or Bilateral Radial
Element
bones

Ammonite X X X

Belemnite X X X

Bivalve X X X

Blastoid X X X

Brachiopod X X X

Bryozoa X X X

Crinoid X X X

Echinoid X X X

Gastropod X X X

Graptolite X X X

Rugosa X X X*

Scleractinia X X X*

Tabulata X X X*

Trilobite X X X

8.2 Marine Paleoenvironments


Most of the fossils you have studied are marine, but remember there are many other environments in the
terrestrial, transitional, and marine realms (Table 8.2). Within these environments, organisms can have
different modes of life, including not moving (sessile), attached by cement to the bottom, or unattached
254 | CHAPTER 8: PALEOENVIRONMENTS

and able to move (motile) (Table 8.3). Some also categorize organisms as benthic and planktonic. Benthic
organisms live at the bottom of the ocean on or in the sediment, and planktonic organisms float or swim.

Table 8.2 – Comparison of habitat between


marine organisms

Habitat

Organism Terrestrial Transitional Marine

Ammonite X

Belemnite X

Bivalve X X

Blastoid X

Brachiopod X

Bryozoa X

Crinoid X

Echinoid X

Gastropod X X X

Graptolite X

Rugosa X

Scleractinia X

Tabulata X

Trilobite X

The way an organism feeds is important for determining its role in the paleoenvironment (Table 8.3).
Does it eat by suspension-feeding as it collects small particles from the water column? Or perhaps it eats
by deposit-feeding and ingesting sediment to get nutrients. Is it an herbivore that only eats plant material,
a carnivore that captures and kills its prey, an omnivore that eats both plants and animals, or a scavenger
that only eats dead material? One fascinating debate among paleontologists is whether Tyrannosaurus
rex was a carnivore or scavenger. Ask yourself: if it was a carnivore, how could it catch its prey with such
tiny forearms? Whereas if it was a scavenger, having small arms might be an advantage for just using its
large teeth to tear apart the dead meat. The final question is to determine if the organism lived a solitary
life or in a colony with others of its own kind. If you can answer some or all of these questions from your
observations, you’ll be able to address more fascinating aspects of ancient life.
CHAPTER 8: PALEOENVIRONMENTS | 255

Table 8.3 – Comparison of mobility and feeding characteristics between marine organisms

Mobility Feeding

Deposit
Organism Attached Mobile Suspension Herbivore Carnivore Scavenger
(Sediment)

Ammonite X

Belemnite X X

Bivalve X X

Blastoid X X

Brachiopod X

Bryozoa X X

Crinoid X X

Echinoid X X X

Gastropod X X X X X

Graptolite X X

Rugosa X X

Scleractinia X X

Tabulata X X

Trilobite X X

8.3 Continental Paleoenvironments


Depending on the continental paleoenvironment, these can be just as fossiliferous as marine environ-
ments. For example, swamps and lagoons can have a wide variety of fish, plant, and insect fossils. In
continental paleoenvironments, you commonly do not find index fossils as these are not as widespread.
Think how quickly the environment can change during a drive through the countryside. Instead these
paleoenvironments can record biodiversity and can be used for paleobehavioral analysis. Often these are
the fossils more commonly known to the general public as they catch our imagination. What child doesn’t
know the name of common dinosaurs or ice age mammals?

Exercise 8.1 – Marine Paleoenvironments

First, review these summaries of the characteristics of Bryozoans and Echinoidea that can be used to deter-
mine their paleoenvironments:

Bryozoans: External skeleton. Individuals (zooids) are bilaterally symmetric, but colonies are typically asym-
metric. Marine immobile (sessile), typically attached to the substrate, benthic, suspension feeders, and lives in
colonies.

Echinoidea: Internal skeleton. Radial (5-fold or pentamerous) symmetry. Marine benthic scavenger or deposit
feeder, mobile, and solitary lifestyle.
256 | CHAPTER 8: PALEOENVIRONMENTS

Using these descriptions as a guide along with your observations, Table 8.1, and material from Chapter 7,
characterize a coral (Rugosa) and brachiopod.

a. Summarize the characteristics that will be useful to determine the paleoenvironment of a Rugosa coral.

b. Summarize the characteristics that will be useful to determine the paleoenvironment of a brachiopod.

c. Next, speculate on the paleoenvironment for all four (bryozoans, echinoids, rugosa, and brachiopod) of
these organisms. Would all of these four organisms live in the same paleoenvironment?

d. Both bryozoans and echinoids lived throughout the fossil record and thus are not good index fossils. In
contrast, rugosa and brachiopods are both more limited in their age range. Using the diversity plots in
Chapter 7, what age could a paleoenvironment be if it had both of these fossils?

Exercise 8.2 – Devonian Reef

Head to the PaleoDB Navigator. A large map of the world should open with many colored circles. The circles
represent locations where fossils have been discovered, and the color of the circle corresponds to the geologic
timescale at the bottom of the page. You can click through the different time periods to see where fossils of
that age have been discovered. On the right side, there is a list of the most abundant fossil groups based on
the time period you have selected on your map view. For example, you could zoom in on fossils found in South
America from the Ordovician Period. The menu on the right will update with what fossils are found on your
map for that period. Depending on your zoom level, you may see that the Trilobita Class is the most abundant,
or you may see it broken down into several orders of Trilobita with the Asaphida Order as the most abundant.
If there is lower diversity of fossils in your map view and time period, you will see the classification broken
down to lower levels. You may find you need to Google some of the names to figure out what phylum, class,
or order they belong to. You can clear whatever filters you select on the lower left side of the page.
CHAPTER 8: PALEOENVIRONMENTS | 257

Figure 8.2 – Distribution of organisms in a typical Devonian reef. The thickness of the organism curves represents the
relative abundance of the organisms. Image credit: Virginia Sisson, CC BY-NC-SA adapted from Corrandini et al. (2002).

a. In the PaleobioDB Navigator, set your time period to Devonian and zoom in on New York State. What is
the most common group of Devonian fossils found in New York? Remember, you may need to search
more technical taxonomic names to determine which fossils you’re looking at (I’m talking about you,
Rhynchonellata and Strophomenata).

b. What are some other organisms that are present?

c. What type of environments are most of these organisms found in?

d. Compare the locations of gastropods and cephalopods. You can do this by clicking on the classes on
the right-hand menu. Look at the two locations of these fossils compared to each other. You may find
it easier to have two browser windows open for this. Based on the locations from the navigator and the
typical locations of fossils around a Devonian reef, what direction was the coastline located?

e. Remove any organism filter and now look at the distribution of all fossils throughout the Devonian in
258 | CHAPTER 8: PALEOENVIRONMENTS

New York. Click on “E”, “M”, and “L” under Devonian to look at the distribution from early to middle to
late Devonian. Was sea level rising or falling in this region during the Devonian (you already know which
way the coastline was based on your previous answer)?

Exercise 8.3 – Himalayan Mountains

Figure 8.3 – Mount Everest showing geological relationships with the unmetamorphosed Qomolangma
Formation overlying the Everest Series metamorphic rocks and the Rongbuk Granite. Image credit: photo of
Everest by Pavel Novak CC SA with geology superimposed.

a. Figure 8.3 shows the tallest peak on Earth, Mount Everest, at 8,848 m (29,029 ft) (Tibetan name
Qomolangma). It has a layer of unaltered limestone from the Ordovician at its summit. What does this
tell you about the environment of Mt. Everest during that time?
CHAPTER 8: PALEOENVIRONMENTS | 259

b. Using the PaleoDB navigator, look at the distribution of fossils for the Himalaya Mountains, particularly
near the Nepal/China border. During which two geologic time periods were fossils most abundant?

c. What were the most abundant organisms for each of those periods? The pane on the right side tells you
the percentage of each organism found in your current map view. You may need to search the technical
taxonomic name to learn which organism it is.

d. Based on the changes in fossils over the time periods, determine whether sea level was rising, falling, or
constant. Be sure to explain your answer.

e. Critical thinking: Does your sea-level interpretation match the global sea level record from Figure 5.21?
Explain your answer.

8.4 Fossils and Paleoenvironments in Movies


We all know that Hollywood stretches the truth (or outright makes it up) when depicting science in
movies. Perhaps the most famous movie depicting fossils and paleoenvironments is the Jurassic Park
series. For example, in the original movie from 1993, there is a scene where a Tyrannosaurus rex chases
a Jeep at over 60 mph. At the time of production, some paleontologists thought T. rex could only run at
a maximum speed of about 45 mph. However, we now know that T. rex didn’t have the hip structure to
support running and was more of a brisk walker. So, you could probably outrun a T. rex.

Exercise 8.4 – Jurassic Park

Using the PaleobioDB Navigator, look up “Dinosauria” (an unranked clade


clade). A clade is a group of organisms
that share a common ancestor, and cladistics is a way to classify organisms based on common characteristics.
You’ll notice that one of the groups in the Dinosauria clad is Aves, which is the class comprising birds. Yes,
according to cladistics and taxonomic rankings, dinosaurs and birds are related.
260 | CHAPTER 8: PALEOENVIRONMENTS

Figure 8.4 – Jurassic Park sign at Universal Studios


Hollywood. Image credit: HarshLight CC SA.

a. During which geologic time periods were Dinosauria most abundant (excluding Aves)?

b. Which one of those time periods were Dinosaurs (not Aves) most abundant?

c. The movie franchise Jurassic Park showcases several different dinosaurs. The table below contains the
seven dinosaur genera shown in the original film from 1993. Complete the table below with what geo-
logic time periods these dinosaurs actually lived and what their preferred habitat was (use Google to
search for the “[genera] habitat.” Remember, the movie depicted all of these dinosaurs living in a tropi-
cal climate.

Table 8.4 – Worksheet for Exercise 8.4.

Genera Geologic Time Period Preferred Habitat

Tyrannosaurus

Brachiosaurus

Triceratops

Velociraptor*

Dilophosaurus

Gallimimus

Parasaurolophus

*In the movies and book Velociraptor was based on a similar dinosaur, called Deinonychus, in almost
every detail, and that only the name had been changed to be more dramatic.

d. Does Jurassic Park seem like a fitting name based on when these dinosaurs lived? Why or why not?
CHAPTER 8: PALEOENVIRONMENTS | 261

Additional Information

Exercise Contributions
Daniel Hauptvogel and Virginia Sisson

Exercise 8.1 was inspired by an exercise on Life Mode Characteristics by Steven J. Hageman at
Appalachian State University

References
Barreda V.D., Cúneo N.R., Wilf P., Currano E.D., Scasso R.A., Brinkhuis H., 2012, Cretaceous/Paleogene
Floral Turnover in Patagonia: Drop in Diversity, Low Extinction, and a Classopollis Spike. PLoS ONE 7(12):
e52455. https://doi.org/10.1371/journal.pone.0052455

Corradini, C., Pondrelli, M., Corriga, M.G., Simonetto, L., Kido, E., Suttner, T.J., Spalletta, C., and Carta, N.,
(2012). Geology and stratigraphy of the Cason di Lanza area (Mount Zermula, Carnic Alps, Italy). Berichte
des Institutes für Erdwissenschaften, Karl-Franzens-Universität Graz. 17. 83-103.

Google Earth Locations


CHAPTER 9: GEOLOGIC STRUCTURES AND
MAPPING

Learning Objectives

The goals of this chapter are to:

• Visualize geologic structures in maps and cross-sections


• Identify geologic structures on a geologic map
• Learn how to read a geologic map

Some of the maps in this chapter can be printed on poster paper from large PDF files found here (opens in a
new tab).

9.1 Introduction
Do you remember in Chapter 0 how we said you would become an expert at reading and interpreting
geologic maps? Now is the time to sharpen this skill. In the past, you probably encountered a variety of
types of maps. If you have used a mapping app to get directions somewhere, you have used a map. Maps
are a scaled, 2-dimensional representation of the surface of an area. Some maps even attempt to por-
tray 3-dimensional landscape features, such as mountains or canyons. This is one of the great difficulties
in geology because geologic maps represent 3-dimensional features in a 2-dimensional format. It can be
challenging for someone unfamiliar with these features to look at a geologic map and visualize reality.
Maps may represent the surface of the land or areas in and around lakes and oceans, the seafloor, or
features known or thought to occur underground. Mapping is also used in astronomy (planet and moons,
regions of space, etc.). Anything that has a visual distribution can be mapped.

In previous chapters, you have seen how colors and patterns can be used on maps to distinguish different
types of rocks, such as volcanic (orangish) versus plutonic (red) in Figure 2.24. Also, the lines between the
different colors represent contacts between different rock types.

Before interpreting or preparing a map, it is important to understand some of the background informa-
tion about the map. Below is a list of information typically associated with maps.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 263

Table 9.1 – Common map information

Information What it tells you

Title Concise information related to geographic information and map contents

The ratio of distance on the map and distance on the ground. The scale should be relative
Scale to both kilometers and miles. Most maps also have a visual scale bar of relative distance on
the map.

Orientation An arrow pointing north and, if possible, latitude and longitude

Explanation or
Explanation of any symbols used on the map, including colors, lines, or icons.
Legend

Reference Selected reference locations or features such as a town, river, state boundary, mountain,
features etc.

A list of authors, compilers, editors, publishers, associated publications (text or guidebook),


Source
URL, and any other information so it can be accurately cited in reference list or
information
bibliography.

Base map What is the source of the geographic information? For example, is it based on a USGS
information topographic map, or is it a satellite image? What is the map projection?

Date published The year the map was released or revised.

Written text Is there a publication or guidebook associated with this map?

What do you need to know for geologic maps? First, there are lots of standard symbols used to help geol-
ogists worldwide use the same terminology. You also need to be able to read strike and dip symbols. If
you didn’t learn about strike and dip in Physical Geology, now is the time to do so. Most geologic maps will
have a stratigraphic column with either colors or patterns to distinguish various rock units. Again, there
are often standard colors for different ages of rocks and patterns for the rock types. Look at the tables in
Chapter 2 Earth Materials to see some common patterns for different rock lithologies. Figure 9.1 shows
the common symbols used for geologic structures.

Figure 9.1 – Common map symbols used on geologic


maps. These are usually included in the map legend.
Image credit: VB Sisson CC-BY.

Exercise 9.1 – Mississippi River Maps

The Mississippi River has evolved throughout its history as the river channel has meandered back and forth
264 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

across the landscape of the central U.S. Different types of maps can provide some insight about previous river
channels and how we can distinguish them. What you learn from the Mississippi River will help you under-
stand river systems throughout the world, regardless of their size.

Part 1: Aerial Imagery

Figure 9.1 is a satellite photograph from a section of the Mississippi River from Google Earth. Think of this
image as what the ground looks like from an airplane. You can see vegetation that has responded to the river
evolving and patterns associated with the river.

a. Trace three previous paths the Mississippi River has taken. It is ok if these channels cross where the river
is currently. Label the different paths a, b, and c, and use color to differentiate them.
b. Put the paths that you identified in order from the most recent to the oldest. ____________________

i. What is your evidence for this order?

c. Locate and mark at least five features associated with a river system. Examples could include cut bank,
point bar, levee, oxbow lake, thalweg, flood plain, terrace, etc.
d. What are some of the difficulties you experienced with completing these two questions?

e. How do you think we could solve some of these problems? What data would you need to accurately
assess the true paths and their ages?
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 265

Figure 9.2 – Google Earth image of a section of the Mississippi River.

Part 2: LiDAR Imagery

Look at the second image (Figure 9.2). This image was made by using LiDAR
LiDAR. This method uses wavelength
reflections to produce detailed images of topography. Different wavelengths are susceptible to changes in
interference and reflection caused by vegetation and sediment changes, which can help identify previous
channels.

f. See if you can locate the channels you found in Part 1. Explain why you agree or disagree with the loca-
tions you placed previous river channels?

g. In Figure 9.2, locate and mark three previous paths the Mississippi has taken that are different from
those you identified from Figure 9.1. Label them with d, e, and f.
h. Using your three paths from Part 1, and your three paths from this part, try to place the six channels in
order from the most recent to the oldest. ________________________

i. What are some of the problems you are encountering doing this?

i. Explain how this image helped or didn’t help you identify the proper paths the Mississippi River has
taken.
266 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

j. Do you think it is necessary to obtain more data to solve our problem of finding previous paths and
ages?

Figure 9.3 – LiDAR image from a section of the Mississippi River. This is the same area as Figure 9.1. Image credit: Daniel
Coe, CC BY-NC-ND.

Part 3: Geologic Map

Figure 9.3 is a map of previous river paths the Mississippi has taken. It was made by a geologist before the
ability to use Google Earth or LiDAR images.

k. Locate as many paths as you can that you have identified from the previous images. Do you still agree
with where you placed the channels you found? Why/why not? Keep in mind that just because a geolo-
gist made this map, it may not be 100% correct.

l. See if you can put the paths you’ve identified from the previous images in order from most recent to the
oldest. Explain if you were able to identify a different order, add more paths you hadn’t seen before, or
if you notice anything else that you think is important.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 267

m. How has this image helped in understanding the proper paths the Mississippi River has taken? Do you
think it is necessary to obtain more data, such as drill cores, to solve our problem of finding previous
paths and ages for analysis? Explain.

k. Critical Thinking: Where would you put a drill core? Explain.

l. Critical Thinking: Could you describe how and where the river moved from its creation to now in such a
manner that someone could draw snapshots in time? Try it.

Figure 9.4 – Meander map from a section of the Mississippi River drafted in 1944 by Harold Fisk. Each color and pattern
represent a previous path of the Mississippi River. This is the same area as Figures 9.2 and 9.3. Image Credit: Harold Fisk
and Army Core of Engineers, Public Domain.
268 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

9.2 Geologic Structures


You already know that the Earth is a dynamic place and that plate tectonic forces bend and break rocks.
You’ve already studied this subject in both Physical Geology text and lab books. Here is a short review
of some of the concepts you will need for this chapter and the next. Some students have an easy time
visualizing in three dimensions. If you have trouble with this, practice with block diagrams.

Remember those lab exercises on geologic time and stratigraphy? While doing these, you looked at ver-
tical sections and views of cliffs, but now you need to be able to decipher geologic patterns on maps. If
you need a refresher or perhaps a glimpse of geologic maps with different map patterns, check out these
links to flat stratigraphy in Nebraska, three types of unconformities in Saskatchewan, thrust fault in the
Olympic Peninsula, plunging anticline in Wyoming, plunging anticlines and synclines, normal faults (core
complex in Arizona), and a strike-slip fault in New Zealand.

Strike and Dip


When deformation occurs, sedimentary layers are no longer flat-lying and instead lie at an angle. This
angle is called the dip (Figure 9.5). The dip is a measure of the steepest possible slope, sometimes
referred to as inclination. You can imagine water running down the surface; it will take the steepest path
in the dip direction. It is common for geologists to pour a little water on a rock surface to see how the
water runs, telling them the dip direction. The strike of a plane is a bit more complicated. Strike is the
horizontal line that forms as a dipping bed intersects with the surface (Figure 9.5) and is always 90° from
the dip direction. Geologists measure the orientations of geologic features using a Brunton Compass, a
special compass that has a set of levels to take accurate measurements of features. To determine the
orientation of these dipping bedding planes, you need to measure both strike and dip.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 269

o o
Figure 9.5 – Block diagram showing an inclined plane that strikes ~40 and dips ~30 to the east. Also shows the
dip direction. Image credit: VB Sisson CC-BY-NC-SA..

There are two directions you can measure the strike that are 180° apart. The dip direction is clockwise
from one and counterclockwise from the other. In most geological fieldwork, the right-hand rule is used.
If you place your right hand on the dipping layer with your fingers pointed toward the dip, then your
thumb is pointing in the strike direction. In other words, the strike is always 90° counterclockwise from
the dip direction.

When writing out the strike and dip, it is a good idea to add the compass direction to the dip measure-
ment, just as a check that you’ve done the right-hand rule measurement correctly. For example, in Figure
9.5, the strike is 40°, and the dip of the plane is 30°, so this would be written as 040/30 SE. This is a plane
that dips at 30° with strike roughly NE. The dip direction is clockwise from the strike, so the dip direction
is SE. You can check this since 225/30 would dip to the NW to follow the right-hand rule, not the NE.

Anticlines and Synclines


Folds are some of the most dramatic geologic structures. These form when the Earth’s crust is shortened
during compression associated with plate tectonic convergence. Folds are important economically as
many resources accumulate in the folds, such as petroleum or gold in anticlines. The points of the tightest
curvature are called the hinge. In contrast, the limbs are the areas of least curvature. If the limbs dip away
from the hinge, this is called an antiform. If the limbs dip toward the hinge, this is called a synform. Note
that if the fold closes on its side, it is impossible to tell if it is an antiform or synform.
270 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

Figure 9.6 – Folded layers with an anticline and syncline pair showing the fold limbs, axial plane, and hinge line.
The layers are from oldest red, tan, dark gray, light gray, yellowish, and youngest is green. The top of the folds is
eroded, showing the relative ages of the layers. Image credit: VB Sisson.

In a stratigraphic sequence of folded sedimentary rocks, the rocks get younger towards the inside of a
syncline and older toward the inside of an anticline, then the fold is a syncline.

The hinges of folds can be connected to create an axial line. If this line is horizontal, the fold is non-plung-
ing. However, if the axial line is tilted, then this is a plunging fold.

If you draw a plane connecting all of the hinges or axial lines of each sedimentary layer in a fold, this
plane defines an axial plane. This can be oriented vertically, be inclined, or even flat-lying or horizontal.
If the axial plane is horizontal, the fold is classified as recumbent. The dip of the axial plane can be used
to classify the type of fold (Table 9.2).

Table 9.2 – Fold information

Dip Fold Type

0-10° Recumbent

10-30° Gently inclined

30-60° Moderately inclined

60-80° Steeply inclined

80-90° Upright
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 271

Faults
Faults are fractures through rocks that have had a significant amount of movement on them. Strictly
speaking, a fault is a single brittle fracture, whereas a fault zone has multiple brittle fractures, and a shear
zone is a ductile fault (Figure 9.7).

Figure 9.7 – An undeformed block, fault, fault zone, and shear zone. If this is a bird’s eye view from above, this is a
strike-slip fault with right-lateral motion. Instead, if this is a view of a vertical outcrop, this would be a reverse fault.
Image credit: VB Sisson adopted from Weldon and Synder CC BY-NC.

The blocks on either side of a fault are called the walls. If the fault has a vertical orientation, this is a strike-
slip fault with horizontal displacement. If the fault has a dip that is not vertical, then one wall is over the
other with dip-slip motion. The blocks are defined as either the footwall or hanging wall. There are two
types of faults with dip-slip motion: normal faults formed during extension and reverse faults formed dur-
ing compression and shortening of the Earth. In normal faults, the hanging wall moves down. In reverse
faults, the hanging wall moves upwards. There is a special type of reverse fault with a gentle fault that
dips less than 30°, called a thrust fault, which is more common than reverse faults (Figure 9.8). For thrust
faults, the block that moves is called the allochthonous block that is thrust over the autochthonous
block. Erosion may isolate some of the allochthonous block as a klippe. It may also expose the autochtho-
nous material in a window.
272 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

Figure 9.8 – Schematic block diagram of a thrust fault. Also shown are the allochthonous and autochthonous
blocks, a window, and a klippe. If you remember the stratigraphy shown in Figure 9.6, you can see that the older
tan rocks are thrust over younger gray rocks. Image credit: VB Sisson CC-BY-NC-SA.

9.3 Geologic Maps


To make a geologic map, you need to use your geological skills to fill in the areas between data points,
such as outcrops, stratigraphic sections, wells, soil profiles, etc. Often the data is imperfect and doesn’t
cover 100% of the area in question. For example, figure 9.9 shows an outcrop map and two possible
attempts at geological maps. One is made without much understanding of geological processes. Although
it satisfies the observations in a simplistic way, it is unlikely to be correct. The other attempt is a more
likely interpretation made with some understanding of geological processes. In this case, there are three
sedimentary units; these probably have parallel boundaries. Also, there is a basalt which you can see
from some of the outcrops that is probably a dike that cross-cuts the sedimentary layers. So, to interpret
this outcrop pattern, you have to use your knowledge of sedimentary contacts (Chapter 4) and cross-cut-
ting relationships of basalt dikes since you know that they are often straight (Chapter 3).

Figure 9.9 – Schematic maps with three sedimentary units and basalt (see Legend) shown in different colors. A. is
an outcrop map with different units as well as contacts between them. B. One possible interpretation of the
outcrop map made without any geological processes. C. Another interpretation of the same outcrop data made
using knowledge of sedimentary and igneous rock types as well as insights about the relative age of the units.
Image credit: VB Sisson CC-BY-NC-SA.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 273

Do you remember the rule of V’s for maps? There are two different versions of this rule; one for topo-
graphic maps and the other for geological maps. If you see a V shape in a topographic map, then the
point of the V points upstream toward the source. In geological maps, this rule governs the intersection
of the geologic features, such as dikes, sills, faults, bedding planes, fold hinges, etc. with topography. The
easiest two features to remember this is that horizontal features such as sedimentary layers and sills will
be parallel to topography, whereas vertical features such as dikes will cross-cut topography (Figure 9.10).
You can use the angle of V formed by the intersection of a linear feature and the topography can help you
determine the dip of the feature. If you need help seeing these relationships in three dimensions, look at
these models.

Cross-sections
You learned how to make topographic profiles and cross-sections in Physical Geology. You may want to
review how to do this before starting the exercises. Remember that vertical scale is one of the features
in a cross-section, and it may not be the same as the horizontal scale. If these are the same, there is no
vertical exaggeration. In many cross-sections, the horizontal scale is longer than the vertical scale. This
is common in many seismic reflection profiles. This may exaggerate sedimentary layer thickness as well
as the dip of some features. In general, the apparent angle of dip increases with increased vertical exag-
o o
geration. For example, for a thrust fault dipping at ~25 , it will appear to dip 60 with ~3x vertical exag-
geration. This will make it difficult to differentiate a thrust fault from a reverse fault in your cross-section.
So, it is important to include the vertical exaggeration for any of the cross-sections you create while inter-
preting geologic maps.

An important feature on cross-sections is the dip of any contacts, which can range from flat to vertical. If
the cross-section is constructed perpendicular to the strike of dipping layers or contacts, then the cross-
section will show the true dip of the features. If the cross-section is not perpendicular to the features, the
apparent dip of the layers will be less than the true dip.

Another way to visualize cross-sections through an area is to construct slices through a three-dimensional
(3D) seismic survey. You may think this is a new technology, but the first 3D survey was made by Exxon
across the Friendswood oil field near Houston, Texas, in 1967. For a 3D seismic survey, seismic data is
acquired along closely spaced parallel lines that are processed and interpreted as a volume of data. Mod-
ern 3D surveys have seismic traces spaced at ~12.5 meters. These seismic volumes are also called seismic
cubes.

Figure 9.11 shows a 3D seismic cube through a folded sequence of sedimentary rocks and several two-
dimensional cuts through this cube. Cross-sections through the 3D seismic cube are identified relative to
the direction that the seismic survey was shot. A cross-section parallel to the direction of the survey is
an inline view. A cross-section that is perpendicular to the survey is a crossline. You can also slice hori-
zontally through the seismic cube and create a time slice. Finally, you can make an arbitrary traverse to
better show structural features.
274 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

Figure 9.11 – A three-dimensional (3D) seismic cube of folded sedimentary strata that includes both anticlines and
synclines. The blue arrow across the top of the cube shows the direction of the seismic survey. The
two-dimensional views include (B) a cross-line at 3000, (C) time-slice at 1000 s, (D) an inline section done at 1500,
and (E) an arbitrary traverse along line A. Scale bar for B-D is 5 kilometers. Blue is for positive amplitude, black for
negative amplitude, and white is for no amplitude difference. Image credit: adapted from FW Schroeder,
Incorporated Research Institutions for Seismology (IRIS) Lesson on 3D Seismic Data.

Exercise 9.2 – Create a Simple Geologic Map and Cross-Section

Your instructor has placed model stations around the classroom. You need to note lithology and collect struc-
tural data from each model station. Go around the lab room using the Clino app (free from the Apple app store
and Google play store) to measure the strike and dip of the model outcrops and make color-coded notes of
their lithology.

a. Make a geologic map using the data points collected from the outcrops you visited in this exercise. Make
sure to use appropriate colors for the rock units and appropriate structural notation. Remember that
your map should have a scale and a north arrow. Draw contacts between the different rock types. At
each station, you should put a strike and dip symbol using the right-hand rule.

Geologic Map:
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 275

b. Make a straight line that crosses the contact lines in your map. If possible, the cross-section should be
perpendicular to your contacts. This line will be where you construct a cross-section of your map. Label
the two ends of the cross-section. Typically cross-sections are labeled as A-A’ or A-B. Show the location
of your cross-section on your geologic map. Remember that since the desks are flat in the lab, you don’t
need to worry about the topographic profile for this cross-section. Did you use any vertical exaggeration
for your cross-section?

Cross-section:

c. Critical Thinking: Imagine you found some fossils in these sedimentary units. Remember your laws of
superposition and add into your cross-section which layer would have a Paradoxides trilobite and which
would have a Mucrospirifer brachiopod.

Exercise 9.3 – Using Google Earth to Interpret Geologic Structures

Head to this Google Earth tour to look at geologic structures in Morocco, a country in northwestern Africa
that borders the Atlantic Ocean and the Mediterranean Sea. Morocco is currently part of a passive margin
but historically was tectonically active. The geologic history of Morocco started about two billion years ago,
in the Paleoproterozoic with the Pan-African orogeny. Then, another tectonic event involved the collision of
Euramerica (or Laurussia) with Gondwana during the Hercynian orogeny (in eastern North America, this is
the Appalachian orogeny). During the Paleozoic, extensive sedimentary deposits preserved marine fossils.
Throughout the Mesozoic, rifting broke up Pangaea forming the Atlantic Ocean. These basins and fault blocks
were blanketed in terrestrial and marine sediments. In the Cenozoic, a microcontinent covered in sedimentary
rocks from the Triassic and Cretaceous collided with northern Morocco, forming the Rif region. Morocco has
extensive phosphate and salt reserves, as well as resources such as lead, zinc, copper, and silver. Morocco is
also a fossil hunter’s paradise with exquisite trilobites and ammonites.

a. What geologic evidence would you look for to see if an area was tectonically active in the past?

Overview

Start the tour and look at the Overview stop. This is the first geologic structure we will look at

b. From this aerial view, how would you describe the appearance of this structure?

c. Do you think this structure is made of igneous, sedimentary, or metamorphic rocks? ____________________
276 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

d. Trace three layers as far as you can on the tour screenshot in Figure 9.12

Figure 9.12 – Screenshot of the Google Earth tour for this exercise.

Views 1 & 2

e. Look at View 1. Are the layers of rock horizontal or tilted? _______________

i. If the layers are titled, in what direction are they dipping toward. (HINT: the red triangle on the
compass on the bottom-right always points north.) _______________

f. Look at View 2. Are the layers of rock horizontal or tilted? _______________

i. If the layers are titled, in what direction are they dipping toward. _______________

g. What geologic structure are you looking at? _______________


h. Look at Rock Layer 1 and Rock Layer 2. Determine and explain which one is older.

i. Look at Rock Layers 3, 4, and 5. What is the age relationship between these layers?

j. Explore the map. Locate similar structures.

Geologic Map and Cross-Section

Figure 9.13 is a geologic map of the region. The location of the structure is right near the town of Imilchil,
toward the center of the map. Figure 9.10 is a cross-section along the line labeled Section 2b. There are a cou-
ple of items that you need to know to be able to read this map and cross-section. First is that this region was
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 277

originally mapped by French geologists. So, they used European names for geologic time such as Lias and
Dogger. Secondly, a shallow basin formed during Triassic rifting of Pangea; these sediments include salt and
red beds (hematite-rich sediments). After the basin filled this salt rose, forming diapirs and pillars in a wide
variety of shapes. Subsequently, the area was deformed by folding and thrusting during the continental colli-
sion between the African and European plates.

k. What is the age of the predominate sedimentary rocks exposed along section 2b? _______________

Figure 9.13 – Geologic map of the west-central Atlas Mountains in northern Africa. Insert shows location of this geologic
map in the Morocco. The terms Lias and Dogger are European lithostratigraphic units for rocks deposited in the upper
Triassic through middle Jurassic. Image Credit: Teixell et al. 2003.

Figure 9.14 – Cross-section of line labeled “Section 2b” on Figure 9.9. The yellow star is the approximate location of the
Google Earth tour for this exercise. The terms Lias and Dogger are European lithostratigraphic units for rocks deposited
in the upper Triassic through middle Jurassic. Image Credit: after Teixell et al. 2003.

l. Find three different sedimentary units, following them across the cross-section. Which unit is the most con-
tinuous? Which unit is the least continuous?
278 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

m. Is there evidence of an unconformity? If there is, mark on the cross-section where it is. Identify the type of
unconformity.

n. On the cross-section, label the anticlines and synclines.


o. Label and identify which kind of faults are present in this area.

p. What type of structure formed the Toumliline diapir?

q. Give a generalized geologic history with the relative stratigraphic, igneous, and deformation events.

The Grand Canyon and adjacent U.S. National Parks cover the entire stratigraphic section from Early Pro-
terozoic to Holocene. This sequence records almost continuous sedimentation with gaps in time (uncon-
formities). This entire section is not exposed in one National Park. In addition to having an awesome
exposure of sedimentary rocks, the area has also undergone deformation with faults and folds; this
occurred between 70 and 30 Ma ending with the uplift of the Colorado Plateau. If you want to keep track,
there are ~40 major sedimentary units in the Grand Canyon.

Exercise 9.4 – Grand Canyon

One of the most complete records of geologic time is in the Grand Canyon. So, this is an excellent area to learn
how to read geologic maps. Before beginning this exercise about the Grand Canyon, you should review the
stratigraphic section in Figure 2.25 or look at the sketch by Powell in Figure 0.3.

Getting the Map

If your department has a poster printer, you can print out a large version of the geologic map of the Grand
Canyon from this file (opens in a new window). This map is from a USGS poster, but we have removed much
of the extra material. Otherwise, you can use this interactive online version of the map.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 279

For fun, you want to look at a 3D geologic map. Another way to look at the Grand Canyon is to use Google
Streetview. Start at the South Rim on this map, and drag the orange person icon onto one of the trails to view
the rocks from inside the canyon! Or, if you want to do more, you can look at an immersive virtual field trip to
see more of the canyon for yourself!

Figure 9.15 – Screenshot of the geologic map of the Grand Canyon. Image credit: Screenshot from USGS map, Public
Domain.

Part 1: Map Reading

a. What is the dominant rock type in this map? ______________________________


b. What is the age of about 80% of this map? Give its geologic period. ______________________________
c. Find an example of each type of structure including three types of folds, two types of faults, dipping and
horizontal beds, sinkhole, collapse structure, and pyroclastic cones.
d. Which stratigraphic unit is the oldest? Which is the youngest? Give their geologic period and name.
______________________________

Part 2: Timing

e. Locate on the map the Great Unconformity and estimate its time gap.
_______________________________________
f. On the map, locate the Greatest Unconformity and estimate its time gap.
_________________________________
g. Which of these has the greatest time gap?____________________________
h. Critical Thinking: The Great Unconformity is an iconic geologic feature that marks the boundary between
non-fossil-bearing Precambrian rocks and fossiliferous Phanerozoic strata. In North America, it spans
from Mexico to Canada. Can you explain why there is this unconformity? What would you look for in
either the map pattern or an outcrop (Figure 9.16) to support your hypothesis?
280 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

Figure 9.16 – The Great Unconformity in the Grand Canyon with Vishnu schist overlain by Tapeats sandstone.
Image credit: Al_HikesAZ CC BY-NC 2.0.

Part 3: Map Patterns

i. Describe or sketch the map pattern for a nonconformity.

j. Describe or sketch the map pattern for an angular unconformity?

k. Using the Rule of V’s, describe or sketch the map pattern for flat-lying sedimentary units.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 281

l. Critical Thinking: Can you tell from the map pattern which is a resistant rock unit?

Part 4: Critical Thinking about the Geologic History

m. What is the age of the Bright Angel fault? _______________________________


n. There are a series of parallel faults such as the De Motte and Uncle Jim Faults. Using Rule of Vs
and their map pattern, are these vertical, horizontal or dipping faults?
____________________________________________________________________________
o. Is the Kaibab anticline older or younger than these faults? Explain.

p. Some think the Grand Canyon started to form only 5 million years ago when the Colorado River started
to cut through the stratigraphy. Others say the Grand Canyon started to form in the Laramide 50 to 70
million years ago. Are there any features on this map that support either hypothesis? Look at the ages
of the rocks and faults that are cross-cut by the Colorado River to support your answer.

q. Critical Thinking: Using this map, construct a geologic history beginning in the Cambrian and ending
in the Holocene. Here is a youtube video of past plate configurations. Anytime the area is blue, it was
underwater. Green and brown indicate was landscape and subject to erosion and/ or tectonic uplift.

The Burgess Shale in Yoho National Park in British Columbia, Canada, is a world-famous fossil locality.
This locality is famous for its 508 Ma fossils that mark the Cambrian explosion. These are one of the ear-
liest fossil beds that contain soft-bodied imprints of a wide variety of life forms. Many of these are ances-
tors of higher life forms such as chordates, while others appear unrelated to any living forms. Charles
Walcott, Secretary of the Smithsonian Institution, first discovered these in 1909. Walcott collected over
65,000 specimens. Nearby there are other quarries that are still yielding new specimens. There are over
282 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

seventy similar localities for soft-bodied Cambrian fossils! They were initially described in Vermont 25
years before the Burgess Shale discovery.

One example of a bizarre animal from this location is Opabinia which had five eyes, a backward-facing
mouth as well as a proboscis to scoop out prey from the marine muds. Only twenty specimens of this
animal have ever been described; they all come from one layer within the Walcott quarry site. No one
agrees on what family this carnivore belongs to; some say primitive arthropods and others trilobitoids.

Figure 9.17- Opabinia regalis. (A) Fossil of Opabinia regalis from the Burgess Shale originally collected by Charles
Walcott. (B) Restoration of Opabinia regalis. This carnivore was 4-7 cm in length. It is difficult to restore this fossil
since it was flattened during fossilization and lithification of the host shale. Image credit: (A) Smithsonian Museum
Public Domain. (B) N. Tamura CC BY-SA.

Exercise 9.5 – Burgess Shale

Before you begin to look at the next map, it is useful to remember some of the principles of using a topo-
graphic map that you learned in Physical Geology. In this exercise as you’ll be using a Canadian map, which
uses different scales compared to maps from the U.S. Geological Survey. So, the contour intervals will be dif-
ferent as well as its scale. Contour intervals generally are inversely proportional to the map scale and also
depend on whether the area has flat or steep topography. So, by just determining this, you’ll be able to judge
whether or not you’ll be able to easily hike through the area or not.

Look at Lake Louise, west half map by Price et al., 1980. The geologic map (opens in a new tab) and cross-
sections (opens in a new tab) can be downloaded and printed on a poster to be able to better view geologic
features. This area is very mountainous and has small valley glaciers
glaciers.
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 283

Figure 9.18 – Screenshot of part of Lake Louise (west) map, British Columbia-Alberta Canada. This is the part of the map
you’ll use for this exercise. Image Credit: Geological Survey of Canada.

Part 1: Map Information

a. What are the contour intervals on this map? ___________________________________


b. How would you identify a glacier versus a lake on this map? _______________________
c. What is the scale of this map? ____________________________

Part 2: Cross-Sections

Find the two cross-section lines. Look at the northernmost cross-section between points labeled “6.”

d. What is the highest point along this line? ________________________________


e. What is the name of one prominent mountain along this cross-section?
284 | CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING

___________________________________________________
f. What is the elevation of the lowest point along this line? __________________________________________
g. What is the total relief along the line of this cross-section?
____________________________________________________________

Part 3: Stratigraphic Column

g. The Burgess Shale quarry locality is labeled with an F. It occurs north of the town of Field, to the west of
Mount Field.

i. What is the name of this formation? _______________________


ii. What is its color? __________________________
iii. What is its map symbol? __________________________
iv. What is its rock type? __________________________
v. What is the general strike of this unit? __________________________
vi. What is the general dip of this unit? __________________________

h. Look another unit either stratigraphically below or above the Burgess Shale location.

i. What is the name of this formation? __________________________


ii. What is its color? __________________________
iii. What is its map symbol? __________________________
iv. What is its rock type? __________________________
v. What is the general strike of this unit? __________________________
vi. What is the general dip of this unit? __________________________

i. Which of these two formations is thicker? __________________________


j. Which of these is made of resistant rocks? __________________________

Part 4: Structures and Correlation

p. What type(s) of brittle structures are in the general vicinity of the Walcott quarry in the Burgess Shale
unit? ___________________________________________________________________________
q. What type of natural disaster do you think a gravity fault is associated with?
___________________________________________________________________________
r. Using the schematic stratigraphic relationships, what formations are time correlative with the Burgess
shale? ___________________________________________________________________________
s. On the eastern edge of the map lies the oldest formation in this map area, the Proterozoic Windermere
Group. You may remember a correlative unconformity, the Great Unconformity above the Vishnu Schist
and Zoroaster Granite in the Grand Canyon (Figure 2.25) or a similar unconformity in Central Texas seen
at Slaughter Gap. What type of unconformity occurs between the Windermere Group and the overlying
lower Cambrian Gog Group? ___________________________________________________________________________
t. Critical Thinking: Looking at this map, are there any areas you think might be good to find additional
fossils equivalent to the Burgess shale fauna? When you propose an area, you need to consider several
factors such as ease of access as even Walcott quarry is about 8 miles from a road and involves a
steep climb. Also, avoid outcrops that may be close to faults. In addition, locations down low may
be forested and not have any exposed rocks.
___________________________________________________________________________
CHAPTER 9: GEOLOGIC STRUCTURES AND MAPPING | 285

Additional Information

Exercise Contributions
Virginia Sisson, Daniel Hauptvogel, Carlos Andrade, and Joshua Flores

References
Billingsley, G. H., 2000, Geologic Map of the Grand Canyon 30’ x 60’ Quadrangle, Coconino and Mohave
counties, northwestern Arizona, U. S. Geological Survey, Geological Investigations I-2688 Map and Pam-
phlet. Public Domain

Price, R.A., Cook, D.G., Aitken, J.D., and Mountjoy, E.W., 1980, Geology, Lake Louise (west half), west of
fifth meridian, British Columbia-Alberta. Geological Survey of Canada, “A” Series Map 1483A, 1980, 2
sheets, https://doi.org/10.4095/108873 (Open Access)

Teixell, A., Arboleya, M.‐L., Julivert, M., and Charroud, M. (2003), Tectonic shortening and topography in
the central High Atlas (Morocco), Tectonics, 22, 1051, doi:10.1029/2002TC001460, 5.

Google Earth Locations


CHAPTER 10: INTERPRETING EARTH'S HISTORY
USING MAPS

Learning Objectives

The goals of this chapter are to:

• Understand how fossil locations can be found using geologic maps


• Interpret geologic provinces using maps
• Construct geologic histories using maps

Some of the maps in this chapter can be printed on poster paper from large PDF files found here (opens in a
new tab).

10.1 Introduction
Throughout this lab manual, you’ve been looking at geologic maps for many purposes, such as locating
different types of rocks, finding plate boundaries, investigating stratigraphic relationships, and more.
However, there are few things you have yet to do. One important skill is using geologic maps to find
fossils. This involves using your knowledge of the environment fossil used to live such as deep marine,
swamp, or deserts, and the age of the stratigraphic unit. You will need to use your knowledge gained
from Chapters 5 to 8.

The other is relating many different features such as magmatic activity, paleoenvironment, deformational
history, metamorphism to create geologic provinces that share a common Earth history. Typically a geo-
logic province is many 100’s of kilometers in size. They can be a single geologic feature (for example, a
sedimentary basin or thrust belt) or a combination of contiguous geologic features. To define geologic
provinces, you will use almost all of the concepts you’ve learned over this course. Why is this important?
Well, you can use a geologic province to find resources such as oil and ores. They also will share a similar
story of Earth.

You may also identify provinces on other planets, such as Mars, which has three geologic provinces; 1)
Tharsis, which is volcanic, 2) northern highlands which have a smooth topography, and 3) southern high-
lands, which are rugged and heavily cratered (see Golombek and Phillips, 2010). If you were going to go
put a settlement on Mars, which of these would you pick to settle just knowing this information?

10.2 How to Locate Fossils


When scouring an area for fossils, the easiest to find are those exposed by weathering from wind or
water. In 2020, a California state park ranger was checking out a watershed near the Mokelumne River.
He spied a fossilized tree. Then he came back and found that erosion over the year exposed a whole
petrified forest and fossilized bones from Miocene mammalian megafauna such as ancestors of horses
and rhinos, gomphotheres (similar to elephants but with twice as many tusks), giraffe-sized camels, and
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 287

many other creatures. Now a team of paleontologists and geologists is excavating the site. You may ask
why so many fossils are in one place and why they eroded away so quickly. Well, they were either buried
in a mudslide or volcanic ash such as a lahar. Both of these are easy to erode rapidly and expose all the
bones of the trapped creatures. So, keep on looking, and perhaps you’ll find some of your own fossils.

Once you have found a fossil, what comes next? There are many resources available for amateur fossil
collectors. After your initial discovery, if you think you might have an interesting find, you can contact
other paleontologists for help or use web resources. If this is a new discovery or unusual specimen, pale-
ontologists will excavate the fossils by removing any overburden and bring them to a museum or other
collection. Small fossils such as mammalian teeth are sometimes collected by dry sieving sands.

Many think this is glamorous work similar to what is seen in movies such as Indiana Jones or Jurassic
Park. In reality, it often involves tedious work sitting in a dry, hot, dusty quarry removing overburden for
many days, especially to expose large fossils. Or you may have to dig a trench around a large specimen
to successfully get all of the fossil (Figure 10.1). These large fossils need to be encased in a plaster jacket
before bringing them back to the lab. In contrast, delicate fossils are often wrapped in toilet paper to pro-
tect them. Plus, Indiana Jones’s job performance would not have been viewed favorably by his university.

Figure 10.1 – A paleontologist works to encase Cretaceous dinosaur footprints in a plaster jacket. A plaster jacket is
like a cast that a doctor uses to fix a broken bone. This plaster jacket is made of sheets of burlap soaked in
plaster-of-Paris. This shows a yellowish sandstone layer below the iron-rich clay layer in which the dinosaur
footprints were preserved. Image Credit: NASA/GSFC/Rebecca Roth.

Fossil collecting must be done with proper permission and possibly permits, especially if you are looking
for fossils on public lands in the United States and Canada. It is important to know the local laws. For
example, you can’t collect anything in a National Park. In a National Forest or Bureau of Land Manage-
288 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

ment land, it is illegal to collect fossil bones but okay to take up to 25 pounds of fossilized wood per day,
up to 250 pounds per year. So, before you start your own collection, be sure you have proper permission
from either the landowner or the proper authorities to collect fossils.

Now that you are using fossils for more than just curiosity, you need to think about how to best preserve
them and their future usefulness for others. Many amateur fossil collectors only record the fossil name
and the formation. What kind of information do you think is needed?

1. Location from GPS with latitude/longitude


2. Do you need elevation?
3. Additional descriptive details such as collected near a stream
4. Do you need stratigraphic information?
5. Age of formation?
6. Orientation of the fossil relative to layering
7. Labeling – should this be on a separate piece of paper or on the specimen, or both?
8. What other fossils were found at this location?

Exercise 10.1 – Using Maps to Find Fossils

Now that you have all of the basics for fossil collecting, it’s time to go on a search. Fossils in museums are iden-
tified by location data and geologic age. Below is a map of all the fossil localities known in the United States
from PaleoDB.

Figure 10.2 – Map of North America and the northern Caribbean showing fossil locations indicated by colored dots. The
color correlates with geologic age according to the geologic time scale at the bottom. The size of the dot indicates how
many species are found at specific locations. On the left side of the image are the percentages of different fossil phyla for
all locations. Image credit: map and data from PaleoDB Navigator world map CC-BY.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 289

a. On the map, mark a few locations that you think would be interesting to go visit. What is the age of the
fossils you would find there? _______________________
b. What type of fossils do you think you might find in these locations? Use your knowledge of fossil diver-
sity through time shown in Chapter 7, such as Figure 7.7 or 7.18. ____________________

If you search online for the best places in the world to go look at fossils or collect them, the number one local-
ity is the Burgess Shale (Exercise 9.4) to see the first soft-bodied animals (but you can’t collect here as it is a
protected site). If you want to be in the United States, go to Calvert Cliffs in Maryland to collect shark teeth and
more. However, if you want to combine a trip to Europe with this search, go to either the UNESCO World Her-
itage Site in the Jurassic Coast of the United Kingdom (Exercise 5.3) to collect ammonites or take a trip off the
beaten track to Stevns Klint in Denmark. We’ll explore what fossils can be found here as well as their geologic
history.

To start your exploration of Stevns Klint, first look at the geologic map (Figure 10.3) and then the stratigraphic
column (Figure 10.4). To the east and north is Sweden, a part of the Precambrian Fennoscandian Shield made
of Archean and Proterozoic gneisses and greenstones. The latest geologic event in southern Sweden was the
Sveconorwegian orogen active between 1140 and 960 Ma; too old an area to have any fossils. North-south
rifting associated with the break-up of Pangea in the Triassic created the Danish Basin, an extension of the
Oslo Graben in Norway. After rifting, marine incursion can be seen as a section of carbonates. A marine set-
ting can host many animals and might be a good place to collect fossils. During the Upper Cretaceous, deep
marine carbonates composed of coccolithophores (algae) settled in the rift basin as a carbonate ooze and are
now lithified as chalk
chalk. As the basin filled in, carbonates were deposited in bryozoan mounds in shallow water.
Today, sea cliffs at Stevns Klint expose chalk at the base, then a clay layer rich in iridium
iridium, and finally a bryozoan
limestone with chert nodules. This boundary between the Cretaceous and Danian (Paleogene) was initially
described in 1847 by Desor.
290 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

Figure 10.3 – Geologic map of Denmark and Danish Basin. Circle with SK shows the location of Stevns Klint UNESCO
World Heritage site. Some geographic location names are given in both Danish and English. Legend indicates the
sedimentary units. The gray color represents the Fennoscandian Shield. Image credit: VB Sisson adapted from Adolfssen
and Friedman (2017) with index map from NuclearVacuum CC BY-SA 3.0.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 291

Figure 10.4 – Stratigraphic column for the Danish Basin. This figure has a
lithologic pattern you’ve not seen before that represents chalk
(carbonate made of coccolithophores). Note that the geologic time scale
uses different colors from the geologic map. Image credit: VB Sisson
adapted from Adolfssen and Friedman (2017) CC-BY.

Use the geologic map and stratigraphic section to address the geologic history of this region.

c. What types of unconformity separates the Swedish Fennoscandian Shield from the Danish Basin?
___________________________________
d. What major geologic event occurred in Stevns Klint? Hint: think through all of Earth’s history.
292 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

_______________________
e. This boundary is seen at the Stevns Klint location and elsewhere. Circle several other locations on the
map that you may go to look for this boundary.
f. What type of fossils would you find in the Mons Klint Formation? ______________________
g. Do you think the same fossils occur in the Stevns Klint Formation? Explain. _____________________________
h. Would you find Mosasaurs or Plesiosaurs in the Rodvig Formation? _______________
i. In which formation would sample to collect echinoderms? _____________________
j. Draw the boundary between the Cretaceous chalk and the Danian bryozoan mounds on this image. If
you could get close to this cliff, this would be marked by a distinctive clay layer rich in iridium.

Figure 10.5 – Photo of Stevns Klint showing the Upper Cretaceous and chalk and Danian bryozoan mounds. Image
credit: Lisa Risager CC.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 293

Figure 10.6 – Topographic profile across the Danish Basin. from A-A’ . Vertical exaggeration is 1:50. Image Credit:
VB Sisson using Google Earth Pro for topography. CC-BY

k. What type of contacts are there between the sedimentary units? _______________________________
l. The Danish Basin geologic map only has formations and does not have any strike and dip symbols (Fig-
ure 10.2). The topography in the area is very subtle; even with a very large vertical exaggeration of 1:50,
you do not see much in the way of mountains on the cross-section. To make a cross-section, you will
need to locate the boundaries between stratigraphic units along the cross-section. Then, using the out-
crop picture, the rule of V’s, and the vertical exaggeration, you can determine the dip of the units. Draw
a schematic cross-section along A-A’. This might be tricky as some of the units are not very thick.

10.3 Geologic Provinces


Have you ever looked at a geologic map of North America? You may not remember it, but part of the map
you saw during the first week of lab was North America (Figure 0.9). The age and character of the rocks
give rise to landscape characteristics (physiographic) and the types of soil and sediments that cover the
surface. The modern landscape of the United States has been shaped by tectonic, climatic, sedimentary,
and erosional processes.

On a grand scale, the contiguous United States is divided into eight physiographic regions which are then
subdivided into 25 provinces. These are further subdivided into sections. We won’t worry about the small-
est subdivisions. In comparison, Africa has eight regions with 17 provinces, and South America has only
four regions with 27 provinces.

Exercise 10.2 – Geologic Provinces of the United States

People tend to be either “lumpers” or “splitters”. The “lumpers” like to group things together and join areas
with similar characteristics. On the other hand, “splitters” like to subdivide characteristics into smaller and
smaller groups. This is why you might find more or fewer physiographic provinces on maps from different
geologists.

a. Write a general statement about the topography of North America from the west coast to the east
coast. Use Figure 10.7 as your reference.
294 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

Figure 10.7 – Shaded relief map of the United States. Red and white colors represent high elevations, greens
and yellow represent lower elevation, and purple and blue represent areas below sea level. Map scale
reference is 40° latitude. Image credit: Daniel Hauptvogel, CC BY-NC-SA

b. On Figure 10.7, roughly draw how you would divide North America into different physiographic
provinces based on topography. Use as many provinces as you think necessary. A marker might
be best for drawing the boundaries.

Another way to look at geologic provinces is to see the general distribution of rock types. Figure 10.8 shows
how igneous (volcanic and plutonic), sedimentary, and metamorphic rocks are located throughout North
America. Many sediments and sedimentary rocks are exposed at the surface. These materials typically cover
the surface and overly older rocks. Sedimentary rocks exposed throughout North America are of all geologic
ages and were often deposited in paleoenvironments different from what exists in the region today.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 295

Figure 10.8 – The distribution of rock types on the surface throughout North America. Volcanic rocks are extrusive
igneous rocks, and plutonic rocks are intrusive igneous rocks. Image credit: From Kate Barton, David Howell, Jose Vigil,
USGS, Public Domain

c. Why do we consider plutonic and volcanic rocks separately when evaluating provinces? Hint: think about
the cooling ages of the two types of igneous rocks.

d. Why do we consider the youngest exposed crustal rocks when evaluating provinces (plutonic and/or
metamorphic rocks)? Hint: Consider what processes are necessary to expose these deep rocks.

e. Table 10.1 lists 10 major provinces of the United States. Use the map and key in Figures 10.9 and 10.10
to draw where these are located. You may be familiar with the names of some of these provinces, but
feel free to search for their general locations to help you.
296 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

f. How do the province boundaries you identified in Figure 10.7 (relief map) compare to the ones you iden-
tified in Figure 10.9 (geologic map)?

g. Use all the maps provided (Figures 10.7 to 10.10), compile all the different features for each of the tec-
tonic provinces listed in Table 10.1 to look at the characteristics of the ten provinces. Make sure to note
observations in the “Topography Patterns” column, such as if the area is mountainous, or what direction
are the mountains oriented, or if the hills are flat or gentle.

Figure 10.9 – The geologic map of the United States, southern Canada, and northern Mexico. Image credit: From Kate
Barton, David Howell, Jose Vigil, USGS, Public Domain.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 297

Figure 10.10 – Geologic ages for the map in Figure 10.7. Numbers are in millions of years. Image
credit: From Kate Barton, David Howell, Jose Vigil, USGS, Public Domain.
298 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

Table 10.1 – Worksheet for Exercise 10.2

Youngest
Dominant plutonic or Topographic Geologic
Province
rock type metamorphic Patterns Structures
rock

Sierra-Nevada

Columbia
Plateau

Basin and
Range

Colorado
Plateau

Rocky
Mountains

Ozarks/
Ouachita

Great Plains

Canadian
Shield

Coastal Plains

Appalachians

One major issue for society is potential sources of energy. Traditionally this has been coal and petroleum, as
well as hydroelectric. Now, there is a push for alternative sources. Have you ever considered just using the
temperature of the Earth as an energy source? This can be done using a heat pump to warm or cool a house.
Since Vikings first landed in Iceland, they used geothermal energy for bathing, baking bread, and washing
clothes. Now, nine out of ten homes in Iceland are powered by geothermal energy. In the United States, the
Basin and Range province has great potential for geothermal energy (NREL).

h. Critical Thinking: Figure 10.11 is a theoretical cross-section of the Basin and Range province. Why does
the Basin and Range Province have a great potential for geothermal resources?
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 299

Figure 10.11 – Theoretical geologic structure of the Basin and Range Province. Image credit: from MiraCosta
College, CC.

Exercise 10.3 – Northern New Jersey Map

The geology of northern New Jersey contains a record of the Appalachian Orogenies, including the assembly
of Pangea and its subsequent breakup. You can print the geologic map (opens in a new tab) and cross-sections
(opens in a new tab) on poster paper. Both maps contain descriptions of the bedrock lithology. The bedrock
is either exposed at the surface or sits beneath loose, unconsolidated sediment. Alternatively, you can access
digital versions of both maps from the USGS. Use the red arrows in the Map Preview to cycle between each
map. The files available for download are on the right-hand side.

Part 1: Separating Geologic Provinces

a. Observe the patterns of colors on the main bedrock map. Divide northern New Jersey into three different
regions based on the colors and patterns on this map. Describe the three regions that you can identify
using the colors on the map or ages of the rocks.

Part 2: Valley and Ridge Province

The Valley and Ridge Province formed as a result of deposition and slight deformation of sediments associated
with the Taconic and Acadian Orogenies.

b. Do the sedimentary rocks become older or younger as you move northwest? ___________

i. On a regional scale, what does that tell you about the direction these rocks are dipping towards?
300 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

c. The Basin and Range province of Nevada is formed through a series of normal faults where basins (val-
leys) are the hanging walls that dropped down relative to the ranges (mountains) on the footwall. Look
at the structures that form the Valley and Ridge Province in New Jersey; how are they different from the
Basin and Range Province of Nevada?

i. For both locations, describe the types of plate boundaries or stresses that would cause these struc-
tures.

d. The rock types give geologists clues to what the tectonic setting was. Limestones are typical examples
of passive times with little to no tectonic activity. Sandstones, shales, and conglomerates are indicative
of active tectonic margins, indicating uplift and then erosion of mountains. Table 10.2 contains a gener-
alized stratigraphic column of the Valley and Ridge Province in northwestern New Jersey.

i. In Table 10.2, identify the type of rock(s) that comprises each unit. Use terms limestone, shale,
sandstone, conglomerate, and quartzite. Quartzite tells you the rock used to be sandstone. You
can find all of this information on the 2nd map page.
ii. Indicate whether the area was an active or passive margin for each unit. Some units contain evi-
dence for both.
iii. The Valley and Ridge Province contains sedimentary rocks derived from two Appalachian moun-
tain-building events: the Taconic Orogeny and the Acadian Orogeny.
iv. In Table 10.2, label the rock units you think best represent when these orogenies were active. An
orogeny can span more than one unit.
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 301

Table 10.2 – Worksheet for Exercise 10.3

Map Active or
Lithology Orogeny
Symbol Passive

Dm

Db

Ds

De

Do

Dp

Dmn

Dkc

DSrd

Sbv

Sp

Sb

Ss

Omr

Omb

Oj

Ow

Obu

Obl

O?a

?l

?h

Yu Igneous and Metamorphic Grenville

Part 3: The Highlands Province

The New Jersey Highlands lie between the Valley and Ridge Province and the Newark Basin.

e. What two types of rocks are the Highlands primarily made of? _________________________
f. Are the igneous rocks primarily intrusive or extrusive? ________________
g. Are the metamorphic rocks primarily high, medium, or low grade? ________________
h. Most of the rocks in this region of New Jersey have the same age. What is it? (Hint: Look for the Mount
Even Granite and the Hornblende granite of the Byram Intrusive Suite) _______________
i. Critical Thinking: Come up with an explanation of how you get these different rocks forming simultane-
ously. (Hint: Think about the tectonic setting that could have done this)
302 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

j. In the middle of the Highlands lie Silurian and Devonian sedimentary rocks (blue and purple colors). You
can look at cross-section A-A’ to help.

i. Do you think this isolated group of sedimentary rocks could be related to the Silurian and Devon-
ian sedimentary rocks from the Valley and Ridge Province? (Hint: Look at some of the rock descrip-
tions) _______________
ii. Explain your answer.

iii. If you thought they were similar, why aren’t these two areas connected?

iv. If you thought they were different, explain how you would get these two sets of sedimentary rocks
deposited at the same time.

Many of the rocks in this region have an age that corresponds to the Grenville Orogeny and the assembly of
the supercontinent called Rodinia. This orogeny is seen on many of the world’s continents, including North
America, South America, Antarctica, Europe, and Africa, but because this event happened so long ago, much
of the evidence has been altered or destroyed by tectonic processes.

You may have noticed there is no evidence of the Alleghenian Orogeny, which was the assembly of the Pan-
gaea about 250 Ma. Any rocks from this orogeny have all been eroded in New Jersey, but much of the fault-
ing and folding of rocks throughout the Valley and Ridge Province and the Highlands occurred during the
Alleghenian Orogeny.

Part 4: The Newark Basin

The Newark Basin (Figure 10.12), which is part of the Piedmont Province of the eastern U.S. that extends from
New York to Alabama, formed during the breakup of Pangaea during the Jurassic, which eventually led to the
opening of the Atlantic Ocean.

k. There is a stark transition from the Highlands Province to the Newark Basin.

i. What is the name of the feature that separates the Newark Basin from the Highlands?
____________________
ii. What type of structure is this? Use the cross-sections to help. ____________________
iii. Critical Thinking: Look at the cross-sections associated with this map. Why are the sediment layers
dipping to the NW?
CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS | 303

l. What types of igneous rocks would you expect to find in a rift valley? ____________________

i. On Figure 10.12, find and mark these igneous rocks. (Hint: look for the Watchung Mountains)

Figure 10.12 – Screenshot of the Newark Basin from the bedrock map of northern New Jersey. Image credit:
USGS, Public Domain.

m. Critical Thinking: Since this province was associated with rifting and tensional stress, you would not
expect to find evidence of compressional stress in a divergent tectonic setting. But if you follow the
Ramapo Fault closely, you will see anticlines and synclines connected to it perpendicularly. Explain why
these compressional features are here.

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson
304 | CHAPTER 10: INTERPRETING EARTH'S HISTORY USING MAPS

References
Adolfssen, JS and Friedman, M, 2017, Review of the Danian vertebrate fauna of southern Scandinavia,
Bulletin of the Geological Society of Denmark, 65, 1-23. doi: 10.37570/bdsg-2017-65-1

Desor, P.J.E., 1847, Sur le terrain Danien, nouvel étage de la craie. Bulletin de la Societé Géologique de
France, 2 (4), 1779-182.

Golombek, M.P., and Phillips, R.J., 2010, Mars Tectonics in Planetary Tectonics, edited by Watters, T.R. and
Schultz, R.A., Cambridge University Press, 181-232.

Some background info from https://whc.unesco.org/uploads/nominations/1416.pdf

Google Earth Locations


CHAPTER 11: PALEOCLIMATE

Learning Objectives

The goals of this chapter are to:

• Understand the different sources of evidence used to study past climates


• Recognize the processes influencing natural climate change
• Use geologic evidence to interpret past climates

11.1 Introduction
Earth’s climate has constantly changed throughout its existence, with periods that have been much
warmer and cooler than today. Evidence of these climate variations is found in the geologic and fossil
records, and geologists use them to interpret how climate has changed through time. The study of past
climates is called paleoclimatology.

Geologists can’t directly observe past climates because they already happened; no temperature or pre-
cipitation can be measured. Instead, geologists turn to the geologic record to study things dependent on
climate, in a sort of second-hand view. These secondary sources of information are called proxies. The
most common proxies used to interpret past climate are fossils, but ice cores, tree rings, and sediment
cores are also used. For example, a particular species of foraminifera coils its shell (technically called a
test) to the left or right, depending on whether the seawater temperature was cooler or warmer. While
this does not give an exact temperature, it gives you a relative idea of the conditions. If you were to find
mostly right-coiling fossils in a layer of sediment, you would know that seawater temperature was rela-
tively warm during the time those organisms lived.

Exercise 11.1 – Foram Morphology as a Climate Signal

Foram fossils are an excellent proxy for paleoclimate because their growth is influenced by climate and ocean
conditions. Some species of foraminifera are so sensitive to seawater temperature that they will alter the
way they grow their test (hard shell). Neogloboquadrina pachyderma, for example, is a species of planktonic
foraminifera that will coil its test to the right when ocean water is warmer and coil to the left when ocean water
is cooler.

In a sediment core recovered from the North Atlantic Ocean, you counted the number of left- and right-coiling
varieties of N. pachyderma to help interpret paleoclimate for the past 160,000 years. Samples of the sediment
core were taken at 10,000-year intervals.

a. Complete Table 11.1 by calculating the values for the total number of N. pachyderma for each sample,
the percentage of the right-coiling variety, and the percentage of the left-coiling variety. The first row is
306 | CHAPTER 11: PALEOCLIMATE

done for you.

Table 11.1 – Worksheet for Exercise 11.1

Age
Right-coiling N. Left-coiling N. Total number N. % Right-coiling N. % Left-coiling N.
(years
pachyderma pachyderma pachyderma pachyderma pachyderma
ago)

0 230 50 280 82% 18%

10,000 220 75

20,000 70 230

30,000 45 300

40,000 50 302

50,000 65 389

60,000 20 140

70,000 56 287

80,000 63 267

90,000 212 56

100,000 120 23

110,000 87 45

120,000 203 66

130,000 56 205

140,000 45 332

150,000 89 135

160,000 123 166

b. Plot the percentage of right-coiling N. pachyderma on the blank graph in Figure 11.1 (or make your own
graph in Microsoft Excel). The x-axis is the percent of the right coiling variety, and the y-axis is the age of
the sediment sample the forams were in.
CHAPTER 11: PALEOCLIMATE | 307

Figure 11.1 – Graph for Exercise 11.1.

c. Has climate changed in the last 160,000 years?

d. On your graph in Figure 11.1, indicate where warm and cold periods occur.
e. How many times has it been warm with a high percentage of right coiling forams? You should group sev-
eral points as one time interval. For example, you could group 0 and 10,000 years as one time interval.
308 | CHAPTER 11: PALEOCLIMATE

f. Has the length of time that it has been cool (with a low percentage of right coiling forams) been the
same each time there has been an ice age?

g. Why does the total number of forams vary between samples? Does the number of forams vary with cli-
mate?

h. Critical Thinking: How does the length of time of the ice ages compare to the length of time it was hot?
Can you think of a reason for this difference?

Exercise adapted from Hilary Clement Olson, Learning from the Fossil Record.

11.2 Causes of Natural Climate Change


Climate change is complex, and many factors affect Earth’s climate system. These include Earth’s orbital
patterns, plate tectonics and the position of continents, changes in solar output, changes in atmospheric
greenhouse gases, and others. All operate at various time scales that alter Earth’s climate over hundreds,
thousands, and millions of years.

On a time scale of millions of years, the movement and location of continents is the biggest influence on
Earth’s climate system. The position of continents affects ocean and atmospheric circulation, which are
the two conveyor belts responsible for taking excess heat from the equatorial regions and moving it to
the polar regions.

On time scales of thousands to hundreds of thousands of years, Earth’s orbital patterns have the greatest
effect on climate. Earth’s orbit varies on three different cycles, called Milankovitch cycles. The shape of
Earth’s orbital path around the sun changes from more circular to more elliptical, completing an entire
cycle about every 100,00 years. This is known as eccentricity. Eccentricity changes how close Earth is to
the sun, affecting how much solar radiation the planet receives.

A video element has been excluded from this version of the text. You can watch it online here: https://uhli-
braries.pressbooks.pub/historicalgeologylab/?p=92

Earth has a tilt to its axis, and it moves back and forth between 22.1 and 24.5 degrees. This cycle, known
as obliquity, takes about 41,000 years. Obliquity changes increase seasonality when the angle is higher
and decrease seasonality when the angle is lower. Seasonality refers to the difference between summer
and winter temperatures. Higher seasonality means hotter summers and colder winters; decreased sea-
sonality means mild summers and mild winters.
CHAPTER 11: PALEOCLIMATE | 309

A video element has been excluded from this version of the text. You can watch it online here: https://uhli-
braries.pressbooks.pub/historicalgeologylab/?p=92

The last orbital change is called precession, best described as a wobble around Earth’s axis. If you’ve ever
spun a toy top as a kid, you may remember that as the top slows down, it begins to wobble. That wobble
is the same as Earth’s precessional cycle, which takes nearly 26,000 years. Precession affects the timing
of seasons because it changes what time of year the northern hemisphere points toward the sun. Today,
the northern hemisphere experiences summer during June-September, but in 13,000 years, the northern
hemisphere will have winter during those months.

A video element has been excluded from this version of the text. You can watch it online here: https://uhli-
braries.pressbooks.pub/historicalgeologylab/?p=92

These three cycles can work together to amplify climate change or work against each other to dampen it.
They also have much longer versions that affect climate on longer time scales. For example, eccentricity
also has 400,000-year and 2.4 million-year cycles and obliquity has a 1.2 million-year cycle.

It’s important to note that these natural causes of climate change cannot explain the magnitude of warm-
ing and climate change that we see happening today. Current climate change can only be explained by
human influences, particularly from the release of greenhouse gases from burning fossil fuels.

11.3 Oxygen Isotopes as a Climate Proxy


There are many ways fossils can be used as proxies. One of the most common uses is stable oxygen iso-
topes in organisms like forams and diatoms. Forams and diatoms are shelled organisms found in aquatic
and marine environments. There are both planktonic forms floating in the water column, benthic forms
that are bottom-dwellers. Foram shells are made up of calcium carbonate (CaCO3), while diatom shells
are composed of silicon dioxide (SiO2). Do these formulas look familiar? They are essentially calcite and
quartz. These organisms extract these elements from the seawater to build their shells. Oxygen, in partic-
ular, is sensitive to global climate and has different isotopes that vary in abundance based on how much
ice is on the planet.
16 18
Oxygen-16 ( O) is the most common form of oxygen, with 8 protons and 8 neutrons. Oxygen-18 ( O) is
16 16
less abundant but has two extra neutrons, making it slightly heavier than O. Since O is lighter, water
18
molecules containing it are preferentially evaporated from the oceans than its heavier counterpart ( O is
evaporated too but in smaller quantities, Figure 11.2). When the water vapor begins to condense into rain
18 18
or snow, heavier O isotopes preferentially do so. The remaining water vapor then has even less O. If
temperatures in the polar regions are cold enough when precipitation falls, the water turns to snow and
ice and becomes stuck on the continent in ice sheets instead of flowing back to the ocean. This is impor-
18 16
tant because, during cold times, oceans become relatively enriched in O because O is being extracted
and stored on continents in the ice. When these organisms build their shells, the chemistry of the shell
reflects the chemistry of the seawater. Suppose you have a sediment core containing an abundance of
18
these organisms. In that case, you can look at their shell chemistry for changes in O to interpret how
much ice was on that planet.
310 | CHAPTER 11: PALEOCLIMATE

Figure 11.2 – Oxygen isotope changes during a) warmer and b) colder climates. Oxygen-16 preferentially
evaporates because it is lighter. During colder times, the ratio of oxygen-18 to oxygen-16 in seawater increases
because the evaporated oxygen-16 gets stored in ice sheets on continents.

Exercise 11.2 – Using Oxygen Isotopes to Interpret Climate

One of the most famous oxygen isotope records in paleoclimatology is from Drs. Lorraine Lisiecki and Mau-
reen Raymo, who published an extremely detailed record of the past 5.3 myrs from foraminifera. Figure 11.3
is a portion of this oxygen isotope record that spans the last 800 kyrs. The notation used for oxygen isotope
18 18 16
records is δ O (usually pronounced as delta-O-eighteen), which is the ratio of O to O. The higher the num-
CHAPTER 11: PALEOCLIMATE | 311

18
ber, the more O there is. The unit is per mil (‰), which looks like a percent sign with an extra 0 in the denom-
inator because it is out of a thousand instead of a hundred.

Figure 11.3 – Oxygen isotope record for the past 800,000 years from foraminifera shells. Data from Lisieckie and Raymo,
2005. Image credit: Daniel Hauptvogel, CC BY.

a. How many distinct cold periods can you identify in Figure 11.3? ____________________
b. About how far apart are these cold periods? ____________________
c. What do you think has been controlling natural climate change over the past 800,000 years? Explain.

d. Compare and contrast the transitions from warm to cold periods and from cold to warm periods.

e. Critical Thinking: If we assume that much of this oxygen isotope signal is a proxy for ice volume, do you
think it takes longer for ice sheets to grow or melt? Explain.
312 | CHAPTER 11: PALEOCLIMATE

Exercise 11.3 – The Climate Record of the Past 65 myr

For Earth’s climate record, we know the most about the past 65 myr than we do for the rest of Earth’s history.
Figure 11.4 contains oxygen isotope and sea-level data for the entire Cenozoic.

Figure 11.4 – Sea-level and oxygen isotope records for the past 65 myr. 0 in sea level represents modern sea level,
positive numbers mean sea level was higher than today, and negative numbers mean sea level was lower. The light blue
in the oxygen isotope data is the complete data set, the dark blue line is a 25-pt smoothing. Data from Miller et al. (2020).
Image credit: Daniel Hauptvogel, CC BY-NC-SA.

a. Table 11.2 contains major climate events of the past 65 myr. Identify these events in Figure 11.4.
b. Explain why there is a steep drop in sea level during the Eocene-Oligocene transition.
CHAPTER 11: PALEOCLIMATE | 313

Table 11.2 – Major Cenozoic climate events

Event Description

Paleocene-Eocene Thermal Maximum A short, 200,000 year period where massive amounts of carbon dioxide were
(PETM) released into the atmosphere, and global temperatures increased by 5-8 °C.

Early Eocene Climatic Optimum (EECO) A 2 myr period of the warmest temperatures of the past 65 myr.

Late Eocene Cooling A global cooling driven by lowering CO2.

Permanent ice sheets form in Antarctica, driven by CO2 lowering past a


Eocene-Oligocene Transition
critical threshold and changes in ocean circulation from moving continents.

Middle Miocene Climatic Optimum


A prolonged warm period whose cause is still debated.
(MMCO)

Middle Miocene Climate Transition Antarctic ice sheet expansion and global cooling driven by lower CO2 and
(MMCT) changes in ocean circulation.

Repeated glacial cycles in which northern hemisphere continental glaciers


Pleistocene Glaciations
reached as far south as 40° latitude.

11.4 Tree Rings as a Climate Proxy


Every year, trees grow taller, and their trunks grow wider. The new growth that makes the trunk wider
is called a tree ring. Tree rings are a valuable tool to climate scientists because they contain a record of
temperature and precipitation. During dry years, trees hardly grow at all and create narrow rings. How-
ever, during rainy years, trees will grow more than usual and have wide rings. Because of this difference
in ring size and connection with precipitation, scientists can use tree ring patterns to learn about the cli-
mate during the time the tree lived. The science of studying tree rings to see what they tell us about past
climates is called dendrochronology.

Each tree ring has a wider light-colored portion and a narrow dark-colored portion (Figure 11.5). The light
coloring, called the earlywood, forms during the winter and early spring precipitation as trees grow at
their greatest rate. The darker color, called the latewood, forms in late summer during slower growth in
anticipation of the dormant season to come. A full year’s growth includes both a light, earlywood ring and
a dark, latewood ring. Together, they make up one tree ring.
314 | CHAPTER 11: PALEOCLIMATE

Figure 11.5 – A photomicrograph showing tree rings for years 1504-1520 on a Douglas fir core specimen. The black
and white bar across the bottom illustrates which components would be measured as earlywood (white) and
latewood (black). Image Credit: USGS, Public Domain.

Data from multiple trees can be combined to create a longer history than a single tree can make. To do
this, dendrochronologists match patterns in the rings from each tree (Figure 11.6). For example, if tree A
started growing in 1830 and died in 1930, and tree B started growing in 1910 and died in 2010, scientists
can combine the records of both trees by matching the patterns for the 20 years that both trees lived.
Dendrochronologists rarely cut down trees to analyze rings. Instead, they take a thin coring tool to extract
a core about the thickness of a pencil without harming the tree.

Figure 11.6 – Long tree ring chronologies can be constructed by matching overlapping patterns of different trees.
Image credit: Andreas Schmittner, CC BY-NC.

Exercise 11.4 – Investigating Tree Ring Growth and Climate

You have been given four simulated tree ring cores (Figure 11.7) from the following sources:

• Core Sample 1: From a living tree from the Pinetown Forest in the fall of 2018
• Core Sample 2: From a tree from the Pinetown Christmas Tree Farm a few years ago
• Core Sample 3: From a log found near the main trail in Pinetown Forest
• Core Sample 4: From a barn beam removed from Pinetown Hollow
CHAPTER 11: PALEOCLIMATE | 315

a. Count the number of tree rings in each core


sample. That number is the age of the tree in
years. Record the ages in Table 11.3. Don’t Figure 11.7 – Simulated tree ring cores for Exercise 11.3.
count the pith, which is a special plant tissue Image credit: adapted from the USGS and UCAR.

found at the center of tree trunks.


b. Determine the year that tree from core sample 1 started growing. Fill this out in Table 11.3.
c. Cut out the 4 tree ring records. Match the patterns in core sample 1 and core sample 2 by lining up the
records and looking at what years overlap in both records. Determine when the tree from core sample
2 started growing and when it was cut/cored.
d. Follow the same steps for core sample 3 and core sample 4.

Table 11.3 – Worksheet for Exercise 11.4

Core Age of Year Growth Year Cut or


Sample Tree Began Cored

1 2018

e. Which tree ring matches the year you were born? ____________________
f. Was the weather in Pinetown the year you were born good for tree growth or not? Explain.

g. Over the entire record, how many years were good for tree growth, and how many were not?

11.5 Fossils as a Climate Proxy


The foraminifera fossils used to construct oxygen isotope records are not the only use of fossils for pale-
oclimate. Many fossils in the geologic record can provide at least some information about climate. For
example, fossils of palm trees from the Green River Formation in Wyoming (Figure 11.8) suggest climate
there was subtropical during part of the Eocene 50 Mya, similar to what Florida is like today. If you look
at the oxygen isotope record for the past 65 myr in Figure 11.4, you’ll notice that at 50 Ma, we had the
warmest climate of the Cenozoic. This is an excellent example of using multiple sets of data to support
ideas in science.
316 | CHAPTER 11: PALEOCLIMATE

Figure 11.8 – A palm fossil from the Green River


Formation at Fossil Butte National Monument in
Wyoming. Image credit: National Park Service,
Public Domain.

Pollen fossils are widely used to interpret paleoclimate. Pollen grains (Figure 11.9) have a resistant outer
shell that helps to keep them preserved in the geologic record, and they are very common. The study
of pollen is called palynology, and palynologists use these fossils from sedimentary records to interpret
what climate was like. Each plant species produces a different sized and shaped pollen grain, and paly-
nologists can determine which type of plants were present or trace the migration of plants as climate
changes.
CHAPTER 11: PALEOCLIMATE | 317

Figure 11.9 – Pollen from a variety of common plants: sunflower (Helianthus


annuus, small spiky sphericals, colorized pink), morning glory (Ipomoea
purpurea, big sphericals with hexagonal cavities, colorized mint green),
hollyhock (Sildalcea malviflora, big spiky sphericals, colorized yellow), lily (Lilium
auratum, bean-shaped, colorized dark green), primrose (Oenothera fruticosa,
tripod shaped, colorized red) and castor bean (Ricinus communis, small smooth
sphericals, colorized light green). The image is magnified some x500, so the
bean-shaped grain in the bottom left corner is about 50 μm long. The coloring
is artificial. Image credit: Dartmouth Electron Microscope Facility, Dartmouth
College, Public Domain.
318 | CHAPTER 11: PALEOCLIMATE

Exercise 11.5 – Using Fossil Pollen to Track Ice Sheet Movement

During the late Quaternary, our planet experienced large, rapid


changes in climate associated with the growth and retreat of
continental ice sheets. These changes were related to major
reorganizations of terrestrial habitats, especially in the northern
hemisphere. Ice age mammals like the classic mammoths and
mastodons had to cope with these changes, as did our own
species, Homo sapiens, as we expanded out of Africa during this
time.

Because plant and animals species often have specific climate


requirements, we can use the geographic distribution of fossil
species to reconstruct ancient climate parameters and docu-
ment how these parameters changed over time. The coniferous
tree species Picea mariana (Figure 11.10) and Picea glauca, com-
monly known as black spruce and white spruce, are the two
most abundant and widely distributed Picea species in North
America. Their migration with changing climate is well-docu-
Figure 11.10 – Black spruce (Picea mariana)
mented through pollen fossils. Their modern habitat, like all trees at an Arctic home south of Inuvik, NT,
Canada. Image credit: Daniel Case, CC BY-SA.
spruces, is in boreal forests or mountains, most commonly
found in Canada and some parts of the northern U.S.

Part 1

You will document the modern distribution of Picea mariana and its preferred climate conditions.

a. Take a close look at Figure 11.11. In terms of geography, what is the extent of Picea mariana’s habitat?

b. Figure 11.11 also shows Picea mariana’s modern temperature and precipitation ranges. Describe the
type of climate this tree grows in.

c. Predict how the distribution of this tree taxon has changed over the last 21,000 years. Do you think it
has remained in the same position, or has it migrated through time?
CHAPTER 11: PALEOCLIMATE | 319

Figure 11.11 – This map shows the regional extent of Picea mariana. The graph shows the range of annual
precipitation and average July (warmest month) temperature Picea mariana grows in. Picea glauca has similar
distribution and climate parameters. Data from Thompson et al. (2015). Image credit: adapted from Thompson
et al. (2015), Public Domain.

Part 2

In this part, you will look at the distribution of all Picea species over time using the Neotoma Database, which
will show you the locations where its pollen grains have been found. The Neotoma database is a repository for
paleoenvironmental data, including pollen fossils.

Follow these steps:


320 | CHAPTER 11: PALEOCLIMATE

• Go to the Neotoma Database and click on the Explorer icon. The search window should automatically
pop up; if it doesn’t, click on the binoculars icon.
• For dataset type, select “pollen” from the dropdown menu.
• Click on the gears symbol next to the “Taxon name” box and select “vascular plants” from the Taxa group
pull-down menu.
• In the “Search for” box, type “Picea” and click “Go”.
• Select all species of Picea and click close (We are using all species because not all scientific studies have
separated Picea fossils into different species. Instead, they simply reported the fossil as Picea sp. There
are only a few species of Picea in North America, and their habitats and climate conditions are similar in
boreal forests or mountainous areas.)
• Click the checkbox next to “Abundance/density” and set the abundance to “> 20%”.
• Click on time and set the age range to 21,000 (oldest) to 18,000 (youngest) and hit search.

After a few seconds, your map will update with the locations of pollen fossils for Picea sp. at the indicated time
range.

d. Where in North America did Picea sp. exist? How does that compare to the modern distribution of Picea
mariana?

e. Based on the modern climate conditions and habitat for Picea, what was the climate like for the U.S. dur-
ing this time?

Repeat the search for the following time ranges; you will only need to change the time range, all other options
should not have to be selected again:

• 15,000 to 12,000 yr BP
• 10,000 to 7,000 yr BP
• 5,000 to 1,000 yr BP

Observe the patterns for spruce at each time slice on the map. You can change the symbols and colors,
rearrange the plotting order, and turn layers on and off using the View Current Search Layers tool (it’s next to
the binoculars icon)

f. Describe the distribution of Picea sp. through time.

g. What climate parameter (temperature or precipitation) do you think is driving the distribution of Picea
through time? Explain your reasoning.
CHAPTER 11: PALEOCLIMATE | 321

h. Compare the distribution of Picea through time to the location and size of the Laurentide Ice Sheet in
Figure 11.12. How does the position of Picea relate to the position of the ice sheets?

Figure 11.12 – Evolution of the Laurentide Ice Sheet during deglaciation. Numbers represent different ice streams
(regions of fast-moving ice) and letters represent ice drainage divides; not relevant to this exercise. Image credit: adapted
from Margold et al. (2018), CC BY.
322 | CHAPTER 11: PALEOCLIMATE

i. Critical Thinking: This exercise focused on one species of a conifer. Most conifers are restricted to either
high elevation or high latitudes. What types of trees could you use for tracking glacial history in temper-
ate or tropical locations?

This exercise was adapted from Samantha Kaplan, Margaret Yacobucci, and John Williams.

11.6 Sediment as a Climate Proxy


Sediment on the continent or at the bottom of the ocean can give geologists a lot of information about
what the climate was like when sediment was deposited. Not only does that sediment contain fossils that
provide clues, but the grain size and provenance of the sediment can help determine temperature and
precipitation patterns. In polar regions, geologists can use the sedimentary record to determine how far
ice sheets advanced and when they retreated. Those IODP Expeditions we talked about in Chapter 5 are
among the primary ways sedimentologists study past climates.

Exercise 11.6 – Using Sediment Provenance to Determine Paleoclimate in Antarctica

All rocks have a distinct chemical makeup; even rocks of the same type can have slight variations. When sedi-
ment is weathered and eroded from rocks, the sediment retains this chemical makeup. Thus, when geologists
recover sediment, like drilling a sediment core from the ocean floor, they can analyze the sediment chemistry
and match it up with known sources on the continent.

Ice sheets are large ice masses that sit on top of continents. Antarctica, for example, is a continent a bit larger
than the U.S., and 98% of it is covered by an ice sheet that is two to three miles high. Ice on continents flows,
like water flowing through a series of rivers but at a much slower pace. As the ice moves across the rocky sur-
face, it weathers and erodes sediment. Most of this erosion occurs at the edge of the ice sheet, where the ice
is moving fastest. When ice sheets reach the coastline, the sediment is then carried by ocean currents to even-
tually be deposited. When geologists investigate sediment eroded by glaciers, they can use the geochemical
fingerprint of the sediment, match it to rocks that exist on the continent, and determine where the edge of
the ice was because that is where the most erosion was taking place.

Figure 11.13 is a map of Antarctica with a focus on a region called the Wilkes Subglacial Basin. A marine sed-
iment drill core was taken from a location just off the coast; the site is labeled U1361A. Layers of sediment
recovered from this core were analyzed for two trace element geochemical signals, neodymium (Nd) and
strontium (Sr). These trace elements have several uses in geology, but we will use them as indicators of prove-
nance.
CHAPTER 11: PALEOCLIMATE | 323

Figure 11.13 – (A) Topography and bathymetry (meters above sea level, m asl) of the Antarctic continent, highlighting
areas below sea level, including the three largest subglacial basins in East Antarctica: the Aurora Subglacial Basin, the
Recovery Subglacial Basin, and the Wilkes Subglacial Basin. The location of IODP Expedition 318 drill site U1361A offshore
of the Wilkes Subglacial Basin is marked in red. (B) Geological map of the area around the Wilkes Subglacial Basin with
simplified lithologies and their isotopic characteristics. Areas of inferred Jurassic Ferrar Large Igneous Province (FLIP)
from airborne geophysics are shown in hatched marking. CB denotes the Central Basin. Image credit: Bertram et al.,
2018, CC BY, adapted from Cook et al., 2013.

Figure 11.14 contains five different types of data collected from the layers of sediment in this core: sediment
type, natural gamma radiation, diatom concentration, Sr, and Nd. Diatoms are single-celled, photosynthetic
algae that build a body out of silica. Generally speaking, diatoms are more productive when water tempera-
ture increases. Use these data sets to answer the following questions.

a. Describe any trends or patterns that you can see in the data in Figure 11.14.
324 | CHAPTER 11: PALEOCLIMATE

Figure 11.14 – Lithostratigraphy (Escutia et al., 2011), natural gamma radiation, diatom valve
concentrations, and Nd and Sr isotopic compositions from Site U1361A. Lithostratigraphy
and natural gamma radiation are from Escutia et al. (2011), diatom, Nd, and Sr data are from
Cook et al. (2013). mbsf = meters below seafloor. Image credit: Daniel Hauptvogel, CC
BY-NC-SA.

b. Table 11.4 contains the Sr and Nd chemical analyses collected from sediment in the drill core (the same
data is plotted in Figure 11.14). Using Figure 11.14, note whether each Sr and Nd analysis was from
diatom-rich or diatom-poor layers.
CHAPTER 11: PALEOCLIMATE | 325

Table 11.4 – Sr and Nd isotopic data

87 86 Diatom-rich or
Depth (mbsf) Sr/ Sr εNd
Diatom-poor

74.09 0.73215 -12.9

76.07 0.715075 -6.9

76.27 0.716604 -6.9

76.31 0.716901 -7.4

77.77 0.72563 -11.1

79.47 0.725313 -11.2

83.05 0.72871 -14.7

84.45 0.71985 -11.4

86.75 0.734461 -12.3

88.17 0.716557 -9.5

89.55 0.731468 -12.8

92.07 0.715194 -8.2

92.95 0.712514 -7.3

94.03 0.717386 -7.5

96.7 0.738815 -14.2

100.75 0.719381 -9

103.76 0.715787 -8.1

106.94 0.729142 -12.9

109.65 0.718272 -8.5

109.74 0.71947 -8.4

115.45 0.72711 -14.5

115.67 0.732095 -14.4

120.55 0.716649 -9.3

c. Figure 11.15 is a graph comparing Sr and Nd values for the four potential source areas of the sediment,
with colors matching the areas from the regional map in Figure 11.13. Plot the Sr and Nd data on this
graph using different colors/shapes for samples from diatom-rich and diatom-poor layers.
326 | CHAPTER 11: PALEOCLIMATE

Figure 11.15 – Characteristic geochemical signatures of the different lithological terranes in the
Wilkes Subglacial Basin vicinity. Image credit: Daniel Hauptvogel, CC BY-NC-SA, adapted from
Bertram et al. (2018).

d. Why do some of the samples plot inside the known chemistry of potential source areas (the colored poly-
gons) and some plot outside of them?

e. Which source(s) are primarily contributing sediment to Site U1361A?


________________________________________
f. When the edge of the ice sheet retreats inland, which source area should become more abundant?
____________________
g. Combining all the data, explain whether diatom-rich or diatom-poor layers represent ice sheet retreat
or ice sheet advance toward the coastline.
CHAPTER 11: PALEOCLIMATE | 327

h. Critical Thinking: Site U1361A is closest to the Archean to Proterozoic source. Why don’t we see that
chemical signature in the sediment?

Additional Information

Exercise Contributions
Daniel Hauptvogel, Virginia Sisson

References
Bertram, R.A., D.J. Wilson, T. van de Flierdt, R.M. McKay, M.O. Patterson, F.J. Jimenez-Espejo, C. Escutia,
G.C. Duke, B.I. Taylor-Silva, and C.R. Riesselman, (2018). Pliocene deglacial event timelines and the biogeo-
chemical response offshore Wilkes Subglacial Basin, East Antarctica. Earth and Planetary Science Letters
494, 109–116, DOI:10.1016/j.epsl.2018.04.054.

Cook, C., van de Flierdt, T., Williams, T. et al. (2013). Dynamic behaviour of the East Antarctic ice sheet dur-
ing Pliocene warmth. Nature Geoscience 6, 765–769. DOI:10.1038/ngeo1889

Escutia, C., Brinkhuis, H., Klaus, A., and Expedition 318 Scientists, (2011). Proceedings of the Integrated
Ocean Drilling Program, 318, Integrated Ocean Drilling Program Management, Inc., Tokyo, DOI:10.2204/
iodp.proc,318.2011

Margold, Martin & Stokes, Chris & Clark, Chris. (2018). Reconciling records of ice streaming and ice margin
retreat to produce a palaeogeographic reconstruction of the deglaciation of the Laurentide Ice Sheet.
Quaternary Science Reviews. 189. DOI:10.1016/j.quascirev.2018.03.013

Miller, K.G., Browning, J.V., Schmelz, W. J., Kopp, R.E., Mountain, G.S., and Wrigth, J.D., (2020). Cenozoic
sea-level and cryospheric evolution from deep-sea geochemical and continental margin records. Sciences
Advances, 6(20).
DOI: 10.1126/sciadv.aaz1346
18
Lisiecki, L. E., and M. E. Raymo (2005), A Pliocene-Pleistocene stack of 57 globally distributed benthic δ O
records, Paleoceanography, 20, PA1003, DOI:10.1029/2004PA001071

Thompson, R.S., Anderson, K.H., Pelltier, R.T., Strickland, L.E., Shafer, S.L., Bartlein, P.J., and McFadden,
A.K., 2015, Atlas of relations between climatic parameters and distributions of important trees and shrubs
in North America—Revisions for all taxa from the United States and Canada and new taxa from the west-
ern United States: U.S. Geological Survey Professional Paper 1650–G, DOI:10.3133/pp1650G.
328 | CHAPTER 11: PALEOCLIMATE

Google Earth Locations

You might also like