... when people thought the Earth was flat, they were wrong. When people thought the Earth was spherical they were wrong. But if you think that thinking the Earth is spherical is just as wrong as thinking the Earth is flat, then your view is wronger than both of them put together.
–Isaac Asimov,The Relativity of Wrong, Kensington Books, New York, 1996, p 226. (Note that the Earth bulges a little around the equator in response to the planet’s rotation, but the Earth is still much closer to being spherical than to being flat!)
Before delving too much into the world of geology (broadly, the study of the Earth), this first unit will focus on a larger question that is a more fundamental issue for most people; namely, what is science? Why pay for it? Why do it? Why trust it? Why learn it? Why bother?
Unit 1, SCIENCE!!, will introduce you to this issue, and help you develop the perspectives to answer these questions. This unit will also introduce you, very briefly, to the field of geology, giving you an overview of what it's about, why it's important, and how it benefits people. And finally, Unit 1 is our entry point to the magnificent environmental legacy that is our National Parks—a system of parks and monuments that lets us enjoy today, and preserve for the future, so many of our society's geological, biological, cultural, and historical records and artifacts.
You will have one week to complete Unit 1. See the course calendar for specific due dates.
As you work your way through these online materials for Unit 1, you will encounter a video lecture, additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the instructional materials |
No, but you will be tested on all of the materials. |
Begin Exercise #1: Scientific Literature | Yes, this is the first of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 1 "RockOn" quiz | Yes, this is the first of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #2" survey | Yes, this is the second of 12 weekly surveys and is worth 1% of your total grade. |
Science is the most successful way humans have developed to learn how things work, and to use that knowledge to do and predict things.
Science is not a magic bullet to the ultimate truth; science is humans keeping track of what works and what doesn't, and trying not to fool themselves in the process. The "scientific method" is just common sense, dressed up with fancy words and expensive machines. Science doesn't tell us what we should do or why we're here, but it makes us a lot healthier and more comfortable while we try to figure out those really big issues.
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 1 - including the online textbook, a video lecture, a supplemental enrichment article, and two supplemental Virtual Trips (vTrips).
Students who register for this Penn State course gain access to assignments and instructor feedback and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Your Geosc10 instructional team is made of people who love Science, Geology, and National Parks. We hope you, do, too, and if not, we'll try to show you why we do. These “big-picture” questions are probably more important than anything else covered in this class.
We humans have always had a love-hate relationship with our “tools.” Cars are great, but getting run over by one isn’t. Televisions are great, until you really want to have a heart-to-heart discussion with someone who is deeply engrossed in a playoff game. Science collects the wisdom of the world’s peoples, their experiences and insights, and then tests that wisdom repeatedly, revising and improving, to help us learn to do things we want. This may be humanity's greatest tool... but that also means that occasionally someone may not like it.
Sometimes, a person becomes unhappy when their idea loses to a better one. In the early 1600s, when Galileo advocated the idea that the Earth orbits the sun, Pope Urban VIII saw conflict with certain verses in the Bible (e.g., Psalm 93, “The world will surely stand in place, never to be moved”, or Psalm 104, “You fixed the earth on its foundation, never to be moved”; both quoted here from the New American Bible, although Urban VIII would have read them in Latin.) Some religious authorities of the time did not see any necessary conflict between these verses and Galileo’s ideas, and the Pope initially had been at least somewhat open to Galileo’s ideas. But, the Pope eventually turned Galileo over to the Inquisition over this supposed heresy, and the Inquisition forced Galileo to recant, sentenced him to house arrest, and banned his book and future publications. (It is an interesting question whether the problem was really Galileo’s sun-centered view, or whether the Pope got mad because Galileo's book featured a dialogue with the Pope's favored views spoken by the "loser.")
The papacy subsequently decided that the reality of the Earth orbiting the sun did not undermine scripture, and astronomers could do their job while the religious leaders did theirs. Indeed, rather interesting scientific discussions have been hosted by subsequent popes.
It remains that sometimes conflict arises between some members of some religious or other groups and some aspects of science. In 2005, for example, the state school board in Kansas changed their definition of science, apparently to enable teaching in science classes of ideas that repeatedly have been rejected as being nonscientific by courts and scientific organizations. After another election that changed membership, in 2007, the board restored to their definition that science is a search for natural explanations for what we observe in the world around us. We will have a chance to discuss these ideas later in the course, because Kansas and other states have continued to fight over the issues.
So, let’s look a little more carefully at what science is, and isn’t.
The reason for science is clear—tightly coupled to engineering and technology, science really works. The products of scientists and engineers are tested in the real world every day. Oil companies hire new geologists, geophysicists, and petroleum engineers when the old ones retire because those scientists and engineers really do find oil and make money for the oil companies. Congress funds biomedical research because it keeps lengthening our lives and curing diseases. You read this on a computer, designed using the principles of quantum mechanics, and using the remarkable discoveries of materials scientists and engineers.
High school teachers like to expound on the scientific method. Scientists do have a method of sorts, and it helps them achieve their results. But lots of people—astrologers, palm readers, telephone “psychics”—have methods that don’t get funded by industry and Congress. The typical industrial officer could not possibly care less how a scientist achieves a result, but only that the result is achieved.
Across campus, scientists are sometimes viewed as just another group for the sociologists to study. Scientists have their own tribes, mating rituals (!), and other social interactions. Scientists seek fame and fortune, lie, steal, and violate their mating rituals in much the same way that other humans do. The extremists in sociology have gone so far as to argue that science is only a social construct, one of many possible ones. This, however, is the kind of intellectual exercise that gets academics in trouble with the real world. Anyone with a little common sense knows that it is possible to have a cruise missile deliver a small exploding device to a selected building in another continent using some clever applications of Newtonian physics, and that no other human social construct can make a similar claim. Mere social constructs do not design new antibiotics that save millions of lives, either.
Science differs from other human endeavors in that its disputes are appealed to nature. In art, you cannot judge whether Picasso or Rembrandt was a “better” painter. You can study the brush work, perspective, social context, or whatever, and learn a tremendous amount about art from the discussion, but you cannot reach an objective decision on who is better. But if asked whether Aristotle’s or Newton’s physics work better, we can answer the question.
This is where the scientific method comes in. We study Aristotle’s ideas and Newton’s ideas until we figure out some way that they differ. This allows us to propose an experiment: if we do A, Aristotle expects B to happen, and Newton expects C. Then, we do A, and see what happens. If it comes out C, Aristotle is wrong. In reality, one test is never definitive—the fans of Aristotle might claim that the experiments were rigged, or the experimenters didn't really understand Aristotle's ideas and so did the wrong test. But after many tests, the answer becomes obvious. Science has then progressed—we’ve gotten rid of something that was wrong.
Science remains an exercise in uncertainty. If Newton “beats” Aristotle, that means Aristotle is wrong, but it doesn’t mean that Newton is right—maybe he’s just lucky, or pretty close, but not quite right. As it turns out, Newton’s ideas fail for things that are really small, really large, or moving really fast, and we have to turn to quantum mechanics and relativity. (But all that fancy physics reduces almost exactly to Newton’s description for things of size and speed that we usually deal with—bigger than atoms, smaller than galaxies, and much slower than the speed of light—so, Newton was and is useful.) Science thus cannot give the ultimate answers to anything, because we’re never sure whether we’re right, close, or lucky. We can only say that, if we act as if the scientific results are true, we succeed (in curing diseases, finding oil, etc.).
Science is an expensive way of learning about the world. Suppose you’re a farmer and you’re trying to feed yourself. You try an idea (say, burying fish heads with your corn seeds, or planting during the dark of the moon), and the corn grows well. So you do that every year. If it works, great. If it doesn’t work but doesn’t hurt, no big problem. If it actually makes things worse, well, you might starve, but not many other people are bothered.
Now, suppose you’re a modern farmer trying to feed 100 people. If you try something that actually makes things worse, many people may starve, and some of them may get really mad at you before they do. So, you start asking whether the fish-head works, and whether two fish-heads would work better, or whether other parts of the fish would be better, and so on. One test doesn’t do it—crops grow well most of the time, so most things you test (such as planting in the dark of the moon) will seem to work even if they really don’t help.
The modern solution is to have a scientist helping the farmer, trying things carefully, and trying them many, many times, figuring out which ones really work better, and communicating those results to others who are interested. All that testing takes a lot of effort, but it is cheaper in the long run for important things. Rather than 100 people each trying to feed themselves, and some failing and starving, we have a scientist, a farmer, a tractor manufacturer, a trucker and a grocer feed all hundred, freeing 95 to do something else. (Enjoy! You probably don’t have to spend the summer hoeing corn to keep from starving over the winter.) So, although science is expensive, for important things it is cheaper than ignorance. For unimportant things, living with a little more uncertainty may be easier.
Science has been wildly successful on simple questions: If I drop a rock, how fast will it fall? If I put a lot of a certain isotope of uranium in a small area, what will happen? If I use steel beams this big, in this pattern, how heavy a truck can drive over the bridge without breaking it? Most of physics, much of chemistry, and some of medicine fall in this “simple question” part of the world.
Science is gaining ground on some harder questions. Predicting weather or earthquakes, understanding and curing cancer, understanding and managing ecosystems and biodiversity—these are more complex, involve more interactions, and may have limits on predictability (chaos), but real, useful progress is being made. The research frontiers lie in these complex systems. Much of geology lies in complex systems, and we are in the midst of some great advances in geology.
Science has a ways to go on really hard questions, such as predicting how various actions will impact the working of society and the health and happiness of people. And science cannot address many questions—“How should society work?” is a value judgment, not a question of reality, and is not part of science.
Science is restricted to the search for natural explanations of the world around us. This does not mean that science opposes religion or claims that there is no God. (Some scientists may do such things, but many other scientists don’t.) Quite simply, no experimenter knows how to guarantee the cooperation of an omnipotent deity. A miracle, by definition, cannot be repeated reliably by anyone in any lab anywhere in the world, and so must fall outside of science.
In short, science is a human social activity but differs from other human social activities in that the ideas of science must be tested against reality. Science enjoys a special place in society because science is so successful. Science shows which ideas are wrong, and also identifies ideas that scientists cannot disprove. If we act as if these not-yet-disproven ideas are true, we are successful in doing things. These not-yet-disproven ideas remain conditional because we might find better ideas in the future. Science keeps track of what works and what doesn’t, to save future workers trouble. Science is a meritocracy—good ideas tend to rise to the top, no matter who originated those ideas. (This may take a while because scientists are human with human failings, but the triumph of merit is more likely and faster in science than in most human activities.) Science tests the structure of knowledge continually—a good scientist does not tiptoe around the tower of knowledge put up by earlier scientists, but tries to tear that tower down. Only those ideas sturdy enough to survive such attacks are saved, so the scientific edifice is exceptionally sturdy.
Societies have tried many different ways to deal with private versus group ownership. Private ownership often raises ethical questions—did you really come by that piece of land fairly? Can you claim for your king some land that was already occupied by other peoples? Do other species have land rights? Public ownership often raises the “tragedy of the commons”—if I can sneak a few more of my sheep onto the public green, I’ll gain in the short term, even if, in the long term, we all lose because the extra sheep kill the grass.
The US tradition has focused on private ownership, but we’ve also recognized the benefits of public ownership. The idea of a National Park—taking the really choice pieces of the country and placing them under public control—is a US idea, developed by the Washburn expedition to Yellowstone in 1870 and eventually enacted by Congress in 1872.
Since then, the idea of national parks has spread across the nation and worldwide. This is surely one of the great ideas of the modern world, to save key scenic environments in the public domain.
However, the national parks of the United States, and the world, face a grave dilemma. The act establishing Yellowstone and the concept of national parks specified “conservation... unimpaired for...future generations” and “to provide for the enjoyment” of the parks. Saving a wild region for the future while having it enjoyed by millions of visitors each year is perhaps the largest of many difficulties facing the parks today.
Most of the national parks were founded to preserve geologic features—the geysers of Yellowstone, Crater Lake, the Grand Canyon, the Badlands, etc. Many national parks were founded when they were, biologically, small pieces of a vast, unbroken range of similar habitats. Today, the parks are often becoming islands of natural environment in a sea of human-controlled and human-altered land. Thus, much focus on the parks today involves biodiversity. We will revisit the questions of biodiversity and island biogeography later. (Yes, this is a geology course, but some things are too important to pass up just because they belong in a different department.)
Geology, broadly, is the study of the Earth. Geologists and friends—geophysicists, geochemists, geobiologists—study the rocks that make up the Earth, the history of the Earth as recorded in those rocks, and the processes that change those rocks. This includes oil and ores, landslides and volcanoes, dinosaurs and meteorites, and much more. Most geologists are involved in one of four areas: i) finding valuable things in the Earth (gold and silver, diamonds, oil, building stone, sand, and gravel, clean water, etc.); ii) warning of geological hazards (volcanic explosions, earthquakes, landslides, groundwater pollution, etc.); iii) building an operators' manual for the Earth (Earth System Science); and iv) informing/entertaining (What killed the dinosaurs? How has the Earth changed over time?).
Historically, most geologists have worked at finding valuable things. These geologists have been truly successful, too successful for their own good, in fact. Some of the things we extract from the ground are very cheap today (after you subtract off inflation and taxes), so there have been fewer jobs for geologists with mining companies than in years gone by. Oil prices have gone up recently, and oil companies have started a hiring boom (talk with any of us in case you're interested in a career move). The US used to spend a lot more money on cleaning up groundwater pollution than we have recently, but it turns out that an immense amount of that money was spent on lawyers arguing about paying for cleanup rather than on scientists and engineers cleaning up. A lot of geologists are not happy with this situation and hope that finding and restoring clean water will be more vigorously pursued in the future.
Warning of geological hazards is also a growing field. As more and more humans build houses on floodplains, debris-flow deposits, and other indicators of past disaster, these people become more dependent on someone to tell them if and when the trouble will return. Many geologists favor a different approach—find out where the dangers are, and then don’t build in those places—but real estate developers often don’t listen. (In the spring of 2012, a bill was introduced to the North Carolina legislature—although not passed in its original form—to make it illegal to use the best science to tell coastal people the regions that might be attacked by the sea. This echoed efforts a century before by developers in San Francisco to discredit scientists who correctly argued that the earthquake that had just devastated the city meant that additional earthquakes were possible.)
The disaster of Hurricane Katrina in New Orleans and surroundings in 2005 really showed the dangers of building in harm's way. With over 1800 dead, and over $100 billion in damages (that is $300 for every person in the US!), Hurricane Katrina definitely caught the attention of many people. Interestingly, geologists had known of the impending disaster, and warned of it, for decades, as the city slowly sank beneath river-level and sea-level. Thousands of Geosciences 10 students had studied this issue in the years before the storm struck (and you’ll get to look into the issue soon).
The operators’ manual for Earth is a new idea. It may be the most important thing geologists can do for the future of humans. We humans are everywhere today—living on every continent, tilling more and more of the land, claiming as our own more and more of the productivity of the planet. We have changed the forests, changed the soils, changed the atmosphere, changed the waters—nowhere on Earth remains free of our imprint. Credible estimates indicate that we and our close friends—cows and corn and chickens and house-cats and Chihuahuas—are using roughly half of everything made available by the planet. We are managing to support roughly 5 billion people pretty well (out of the 7-plus billion of us here), with population projected to reach 9 or 10 billion in a few decades, so we are planning on doubling the number of people we support well.
Given that we are doing this, and we will continue to do so, many thinkers believe that it would be wise to have a better idea of how all of this works and what we are doing. You would not try to repair a fine watch without knowing how it works—take a few pieces out and you may never get it running well again. We are doing precisely that to the planet, changing a lot of things we don’t understand. Earth System Science is the attempt to understand the planet, its water, air, ice, rock, and life, well enough to learn the consequences of our actions so that we can make wise decisions. Earth System Science is in its infancy, and we do not even know whether we will ultimately succeed, but many of us believe that it is an incredibly important effort.
And, there is always education and entertainment. Some people really like to know things, and geologists have some of the most interesting stories to tell. Perhaps you will find some of the stories here to be interesting.
This text will try to show you not only what geologists learn, but how we learn it. For the first few chapters, however, we ask you to take our word for some things. You will see statements such as “The Earth is 4.6 billion years old.” Before we’re done, you will also see where that number comes from, how good it is and much more about it. But we can’t do everything at once.
Anyway, we believe that the universe started in a “Big Bang” about 14 billion years ago. The Earth is “only” about 4.6 billion years old. We live in a second-generation solar system because the planets and the sun contain abundant chemicals such as iron that was first formed during the death of older stars. So, there were some stars, and they exploded and generated gases and dust, and something (another nearby star exploding?) caused that gas and dust cloud to be compressed a little. Once the dust and gas started falling together, gravity took over. Eventually, most of the mass went to form the sun, which was squeezed enough under gravity that the sun’s hydrogen began fusing to form helium, in the process releasing energy—sunlight! Some of the gas and dust collected into planets.
Assembly of the Earth involved falling together of lots of big and little chunks. The largest chunk was probably about the size of Mars. It hit the Earth after most of the assembly was finished and blasted enough material off the Earth to make the moon. (Note that there is still a little rumbling in the scientific literature about this Mars-sized moon-forming collision, so stay tuned...)
The falling-together of pieces makes heat. That heat partially or completely melted the planet. The melting allowed the planet to differentiate, or become layered. The denser material sank to make a core, mostly of iron, with some nickel and a few other elements. The lowest-density material rose to the top to form a silicate scum, or crust, floating on a vast mantle of denser silicate (see the sidebar on chemistry). The Earth is hottest in the middle, coldest on the outside. Heat favors melting, but higher pressure tends to make most liquids turn solid. These two effects compete in the Earth, so you find both solid and liquid down there. Going down in the Earth, the crust and the upper part of the mantle are solid (together forming the lithosphere) except in special places where volcanoes occur; the deeper part of the mantle is solid but soft, and has a zone about 100 km (60 miles) down in which a little melting occurs. The soft zone in the mantle is the asthenosphere--we won't learn a huge number of new words in this class, but we do get a few great ones! The core has two layers, a solid inner core and a liquid outer core.
Some of the Earth’s heat is left over from when the planet formed, and a lot comes from the decay of naturally radioactive materials in the Earth. As the early heat has escaped and the radioactive materials have decayed, the Earth has slowly been cooling off, but plenty of heat remains to drive geologic processes. The Earth has developed an atmosphere, oceans, and life, and a rich sedimentary history of how those developed. The atmosphere and oceans spend their time wearing down mountains, but the heat of the Earth keeps driving processes that build mountains up, so there is a near-balance. And all of this should become clear as we tour the national parks.
Most students reaching the university have taken a chemistry course somewhere along the way, but a few of you haven’t. Here is a BRIEF summary of chemistry, as a refresher for those who have had a chemistry course and as a teaser for those who have not. We do not use a lot of chemistry in this course, but it comes up often enough that you may find a summary to be helpful.
If we pick up anything around us (water, chewing gum, rocks, whatever) and try to divide it into smaller and smaller pieces, we will find that it changes as it is divided. A tree becomes a log as soon as we cut it down. If we dry the log before burning it, we find that it contained lots of water, plus other things that are not water. If we then use fire to break the log into smaller pieces, we find that we can do so, while releasing energy. Using tools and energy levels that are easily available to us, we will find that we can continue dividing something until we get to elements, but that we cannot divide the elements.
The “unit” of an element is called an atom. There are 93 naturally occurring elements, plus others that humans have produced. If we use even higher energies, such as those achieved in nuclear accelerators, we find that we can take atoms apart.
Each atom proves to have a dense nucleus, surrounded by one or more levels where electrons are found. It may prove helpful to think of electrons circling a nucleus the way planets orbit the sun, although this simplified model doesn't capture all the features of an atom.
The nucleus contains smaller particles called protons and neutrons. Protons have a characteristic that we call positive charge, electrons have an equal-but-opposite negative charge, and neutrons are uncharged or neutral. A neutral atom of an element contains some number of protons and the same number of electrons, with their positive and negative charges just balancing each other.
The type of atom, or element, is determined by the number of protons; add one proton to a nitrogen atom, for example, and it becomes an oxygen atom. (Breathing nitrogen without oxygen would cause you to die quickly; they are different!) The positively charged protons packed tightly in a nucleus tend to repel each other, but the neutrons act to stabilize the nucleus.
Some elements come in different “flavors,” called isotopes, which have different numbers of neutrons and so different weights. Slight differences in the behavior of isotopes allow us to use them to learn much about certain processes on Earth, as we’ll see later. All atoms of an isotope are identical, and all atoms of an element are nearly identical.
Chemistry includes all of those processes by which plants grow, we grow, wood burns in a fireplace, etc. Chemistry involves changes in how electrons are associated with atoms. An atom may give one or more electrons to another, and then the two will stick together (be bonded) by static electricity, the attraction of the positive charge of the electron-loser for the negative charge of the electron-gainer. An atom that gains or loses one or more electrons is then called an ion. Atoms may also share electrons, forming even stronger bonds and making larger things called molecules.
Most Earth materials are made of arrays of ions, although some are made of arrays of molecules. The ions or molecules usually form regular, repeating patterns. For example, in table salt, a lot of sodium atoms have given one electron each to a lot of chlorine atoms, making sodium and chloride ions. Then these stick together. One finds a line of sodium, chloride, sodium, chloride, sodium, chloride, and so on, and a line next to it of chloride, sodium, chloride, sodium, ..., and above each sodium there is a chloride, and above each chloride a sodium, in a cubic array. A grain of salt from your salt shaker will be a few million sodiums and chlorides long, and a few million high, and a few million deep.
The properties of the grain of salt—how it tastes, and dissolves, and breaks, and looks, etc.—are determined by the chemicals in it and how they are arranged. We call such an ordered, repeating, “erector-set” construction a mineral. Almost all of the materials in the Earth are minerals. Liquid water is not a mineral because the water molecules are free to move relative to each other, but liquid water becomes a mineral when freezing makes ice.
When the Earth formed, we received a few elements in abundance and only traces of the other naturally occurring elements. More than half of the crust and mantle is composed of two elements–oxygen and silicon. Most minerals and rocks thus are based on oxygen and silicon, and we call these rocks silicates. In silicates, the silicon and oxygen stick together electrically, with each little silicon surrounded by four oxygens. Silicon, oxygen, and six others–aluminum, iron, calcium, sodium, potassium, and magnesium–total more than 98% of the crust and mantle. Geology students used to be required to memorize the common elements, and their abundance, in order; for our purposes here, know that only a few are common, and we’ll come back to them later.
The silicon-oxygen groups form minerals either by sticking to each other by sharing oxygens, or by sticking to iron, magnesium, or other ions. Minerals that contain a lot of iron and magnesium are said to be low in silica, even though they may still contain more silicon and oxygen than iron or magnesium. In general, minerals high in silica are light-colored, low-density, have a low melting point, often contain a little water, and occur mostly in continents; minerals that are lower in silica usually are dark-colored, high-density, melt at high temperature, and occur on the sea floor or in the mantle more often than in continents. You may hear “basaltic” used for low-silica, because basalt is the commonest low-silica rock at the Earth's surface; similarly, “granitic” means high-silica because granite is a common high-silica rock.
It would be fun to take a tour of all the national parks, learning a little about each. But Penn State would not award you General Education credit for such a course—you are supposed to be taking a tour of a field of knowledge, in this case, geology.
So, we will take a tour of geologic ideas. But, some of the best geological features of the world are enshrined in the US (and other!) national parks. We will use national parks as illustrations, delving into park history and culture when we can, but concentrating on those things that illustrate how the Earth works.
Scientists communicate in a lot of ways, but the most important is through the refereed scientific literature. Any scientific paper is first submitted to a learned journal, and the editor sends the paper out for peer review. In this, several recognized world experts read the paper and make sure it is “good.” Are the methods described well? Are uncertainties given? Is proper credit assigned to other sources? Do the equations make sense? Are substantive conclusions reached? If there are obvious errors, then the paper is sent back to the author or authors for revision. If the paper is unclear, or can’t be read well, or information is omitted, or if unsubstantiated claims are made, or anything else is wrong, the paper is sent back for revision. Only when the paper clearly and logically presents new results will it be published.
Peer review takes a lot of time and effort. Peer review also slows down publication of important results. (The papers authored by Drs. Alley and Anandakrishnan that have been of greatest use to other people were also the ones that had the hardest time gaining approval from reviewers, who check especially carefully on the big stuff.) And, there is no guarantee that the reviewers will get everything right; errors do sneak by. But peer review really raises the quality of the scientific literature above the quality of other sources that are available to you.
You can find information in many places—books, magazines, newspapers, the Web, speeches by public officials, graffiti in restrooms, etc. Some of this information is more reliable than others. In general, the more permanent a publication is, and the more expensive it is to get you the information, the better the information. (So the Web, which is cheap and has a huge turnover in websites, includes an immense amount of nonsense as well as some good stuff.) But, there are surely exceptions to this “rule.”
If something really matters, the refereed scientific literature, with its long traditions, its focus on accuracy, and its appeal to nature to test ideas, is the most reliable source available. Textbooks, lectures by professors, and other ways we give you information aren’t bad, but the refereed scientific literature is still better. You’ll have the opportunity to explore this in Exercise 1.
Join Dr. Alley and the Geosc10 team for "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered. For the rest of the course, the VTrips will present important material that may show up on the quizzes. In general, the slide shows start with pretty pictures to introduce you to a park, and end with pictures focused on scientific ideas. The more scientific, the more likely to be on a quiz! Here, mostly for your enjoyment, are pictures of two gloriously beautiful places, the world's oldest national park (Yellowstone) and the world's largest national park (Northeast Greenland). Have fun!
Please watch the 34-minute Unit 1 lecture featuring Dr. Richard Alley.
Check out the Unit 1 PowerPoint Presentation used in the online lecture here.
We’re off on a great journey through the National Parks, around the world, and deep into history. Here’s a preview of some of the things we’ll visit before the semester ends. These "Rock Videos" are mostly parodies of famous songs. Later in the semester, they really will be useful reviews of the units.
In this course, we will deal with a number of ideas—the age of the Earth, the occurrence of evolution, the prospect of global warming, and others—about which there are heated public debates in the US. These debates have persisted in the US, and in some other places, long after scientists reached consensus and moved on, using the science to help people. These debates have persisted here long after most people in many other countries accepted the scientific results, and began helping the scientists use the information to help people. Why is this? Are scientists and people in some other countries just stupid, ignoring common sense? Are many regular people in the US just stupid? Are US politicians cynically exploiting subjects for personal gain? You will get many different answers to these questions from different people!
Here, we’d like to give you something to think about. The discussion here in no way proves that scientists are right about the age of the Earth, or evolution, or global warming—we’ll discuss the evidence about these later in the course. But, the discussion here may help you to think about thinking about these issues, and to examine your own ideas on the topics. So…
Have you ever visited an old European city and tried to drive through the downtown? Or have you watched from the sidewalk, or in a movie, while others tried to drive the winding streets of Rome or London? People who have done so quickly form opinions about the experience, and those opinions are very seldom, “Wow, what an efficiently designed road system, ideal for moving traffic.” Far more common is “Wow, what a mess!” And yet, although almost everyone knows that the roads are a mess in the downtowns of major old cities, those roads are still there. Why?
The answer is fairly simple. The roads were built when the city was tiny, centuries or even millennia ago, to serve that proto-city. Then, as the city grew, it grew around those roads. Buildings went up, and museums and castles and theaters and sewers and all the other stuff of a city. Would you move Big Ben, or tear down the Louvre and start over, to straighten out a winding road? When faced with that choice, people usually keep the old roads. Careful checking will show you that people actually have put a lot of effort into improving the roads over time, moving things, tunneling under or bridging over, adding subways to take off some of the strain, patching and fixing and repairing, spending billions of dollars (or Euros, or pounds, or whatever), but always starting from the existing system rather than starting over.
There is a useful analogy here in considering how humans learn things, and in particular how we learn science. I (this is Dr. Alley writing) have had the joy of watching closely as our two daughters grew from babies to toddlers (and on to remarkable young women), and there is a good chance that you have either closely observed growing babies, or will. Scientists watch babies, too, and are learning a lot about learning.
By the time a baby is a year old, he or she knows an amazing amount about the world. The baby knows that some things are inanimate and others animate—rattles don’t walk away, but parents do. Many things are predictable for a baby—a rattle released in midair always falls down, not up, unless grabbed by a mother or father or other living thing. A rattle placed properly on the railing of a crib will stay there.
In gaining this knowledge, the baby is putting down “roads” in the brain, wiring in information that later will be called “common sense”. When artificial-intelligence researchers have tried to get computers to do human jobs, perhaps the biggest difficulty has been that the computers lack this “common sense”—teaching the computer all the things that a baby learns proves to be quite difficult, because the baby learns so much.
Notice, however, that this “common sense” is often not really correct. For example, babies do not start off with a modern view of the shape of the Earth and the physics of gravity. Whatever a baby does think, a “round, round world”, with people and rattles pulled toward the center by the warping of space-time by mass that causes gravity, is not in the original common-sense picture.
Careful studies show that, when children are finally told about gravity on the spherical Earth, they initially resist the idea. They may deny it, or they may try to modify it to fit with their “common sense”. (If asked to draw the world, they may add a flat spot or divot just where they live in an otherwise spherical Earth, or they may draw people living inside a sphere.) Often, it takes until age nine or so before children really say that they accept the idea that they are held to a spherical Earth by gravity, and they draw pictures properly illustrating the idea.
Furthermore, studies find that children show this sort of resistance to most or all new ideas that conflict with the “naïve physics” or “naïve psychology” formed in the cradle. Like a city preserving its old roads, a child preserves the initial ideas—perhaps adding or patching, building new paths in the “suburbs” of the mind or building “subways” that take the new idea around the old one, but getting rid of the old idea only when absolutely necessary.
Now, almost all children eventually come to accept the spherical Earth. (There actually are a few people out there who still argue against the spherical Earth, just not very many.) But, typically all of the authority figures in a child’s life agree in telling the child that the Earth is spherical—teachers, parents, preachers, TV, books, and more—with almost no disagreement. (And yes, even young children have a ranking of reliability of information sources, picking which “authority figures” to believe more.) Yet with all of the authority figures agreeing, years may still be needed to convince a child to replace the old “road” in the brain with a new one (and that old road might even still be in there somewhere).
What about, say, the age of the Earth? The scientific evidence indicates that the Earth formed about 4.6 billion years ago. But 4.6 billion years is not “common sense” to most people—we don’t imagine such large numbers very well. If you ask a very young child how old the Earth is, you are almost guaranteed to get an answer that is not “4.6 billion years”.
In many parts of the world, all of the authority figures in a child’s life agree in telling a child that the Earth actually is 4.6 billion years old—teachers and parents and preachers and politicians and more—and almost all children in those places eventually come to believe that the Earth is 4.6 billion years old. But in some places, including the USA, many authority figures do not agree on this. Some preachers, some parents, some politicians, and even a few teachers express doubt about this scientific idea, and some may be actively hostile to it, claiming a very different answer. When faced with a scientific idea that runs against the “common sense” developed at a young age, and with a split opinion among authority figures, a typical child will keep the old “road” through the “city” of their mind. And that child is likely to grow up into a parent, or preacher, or politician who disagrees with the science.
As noted above, this does not in any way prove that science is right and young-Earth politicians are wrong, or that young-Earth politicians are right and scientists are wrong. We’ll address those questions later, after looking at how the Earth works. But, lots of evidence (and common sense!) shows that people do arrive at ideas much as described here, and you can go watch babies, visit other places, and convince yourself. (You might also look at the article listed at the end of this text for a little more information; some of the ideas and examples given here were taken from that article, which provides additional insights.) So, when you visit Paris or Madrid, please safely enjoy walking the narrow, winding streets. Then, on the plane home, ask yourself what the “roads” look like in the mind of the person in the seat next to yours—or in your seat.
Reference:
Childhood Origins of Adult Resistance to Science
Paul Bloom and Deena Skolnick Weisberg
Science, 2007, v. 316, p. 996-997
You have reached the end of Unit 1! Double-check the list of requirements on the Unit 1 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Exercise #1 opened this week. You have two weeks to complete it, but please start early to avoid technological difficulties. See Course Calendar for specific dates.
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
“If they’d lower the taxes and get rid of the smog and clean up the traffic mess, I really believe I’d settle here until the next earthquake.”
— Groucho Marx, about Los Angeles, California
“We learn geology the morning after the earthquake.”
—Ralph Waldo Emerson
“It’s snowing still,” said Eeyore gloomily. “So it is.” “And freezing.” “Is it?” “Yes,” said Eeyore. “However,” he said, brightening up a little, “we haven’t had an earthquake lately.”
—A.A. Milne, The House at Pooh Corner
As you may recall in the beloved Christmas special about Rudolph, at a key point in the story, Yukon Cornelius hacks off a chunk of sea ice—frozen ocean water—so that he, Rudolph, and Hermey the dentist elf can drift away over the ocean and escape from the Abominable Snowman. If you've ever neglected to check the spaghetti sauce you were heating on the stove while watching Rudolph because you were just dying to know whether 'Bumbles bounce', you may have noticed that the stiff scum that forms on top can break into chunks and drift around on the liquid beneath—much like Yukon C's ice block.
In this unit, and the next two, we'll explore the equivalent activity in the Earth, with the hard-frozen upper layer breaking into drifting chunks, melted lava leaking up in cracks to feed volcanoes (vaguely like the water in the crack that the Bumble fell into when trying to catch the heroes), collisions causing earthquakes when blocks run together ("Land, Ho"), and more. Hold onto your teeth, and let's get started for the land of misfit plate boundaries.
You will have one week to complete Unit 2. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 2, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the instructional materials |
No, but you will be tested on all the materials. |
Continue working on Exercise #1: Scientific Literature | Yes, this is the first of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 2 "RockOn" quiz | Yes, this is the second of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #3" survey | Yes, this is the third of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 2, including the online textbook, a video lecture, several vTrips and animations, and two overview presentations.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Death Valley National Park of California and adjacent Nevada sits as deep as 282 feet below sea level near Badwater, the lowest land in the western hemisphere. Yet, Telescope Peak in the Panamint Range, 11,049 feet high, is in the park less than 20 miles west of Badwater, and Mt. Whitney, at 14,494 feet the highest peak in the continental United States, is only about 80 miles away.
Death Valley is the hottest, driest place in the United States. On July 10, 1913, the temperature reached 134oF (57oC), and 125oF is common. All of the rainfall (about 2 inches or 5 cm per year) evaporates quickly.
Occasionally, water will sit on mud flats for a while (especially in the colder winter) before evaporating to leave salt deposits. Sometimes in winter, the water freezes on top, and strong winds blow the ice, dragging rocks that are frozen in it. The tracks of such rocks, at the Devil’s Race Track, long puzzled people before the answer was discovered. (And you’ll still find people who argue about the tracks and rocks out on the flat surface—no one has ever watched the rocks when they were moving!) During the ice age, more rain and cooler temperatures than today caused the valley to fill with water, forming Lake Manly, which was about 90 miles (150 km) long and almost 600 feet (200 m) deep at its deepest.
Salts left on mud flats by evaporating waters often contain boron, a rare element that has leached from volcanic rocks nearby and been carried into the valley by streams. Boron is used in soaps, water softeners, pottery glazes, and other things. Mining of the boron-bearing borax salts began in the early 1880s, and they were hauled out by the famous 20-mule teams. Mining was allowed to continue in the park after it was established (as a national monument in 1933, and a national park in 1994), but is being phased out. Copper, gold, silver, lead, and talc has been mined in or near the park.
The great depth of Death Valley was not carved by a river or glacier or the wind, and is not from a cave-like Mammoth Cave collapsing. Instead, as shown in the diagram below, labeled Pull-apart fault at Death Valley (which is more-or-less what you would see if you could take a slice across the valley from the air at the top down into the Earth at the bottom), Death Valley has been dropped along faults as the region has been pulled apart. Similar down-dropped and upraised blocks extend across much of Utah, Nevada, and neighboring areas, forming the Basin and Range region. A little bit of such faulting even extends up into Wyoming, and the great front of the Teton Range is geologically a close cousin to the west side of Death Valley.
The geologic evidence of faulting is clear—layers of rock and even recent stream deposits are offset across the faults, as you'll see in a VTrip at the end of this unit. The faulting occurs in fits and starts rather than smoothly, producing earthquakes. The Basin and Range regions remain active in producing earthquakes. (More about earthquakes is coming soon.)
Because of the way that the faults are angled, the dropping of blocks requires that the region is getting wider! (Look again at the diagram to the right labeled Pull-apart fault at Death Valley.) It has recently become possible to measure this widening using GPS or very-long-baseline-interferometry techniques. The widening is not fast—maybe an inch or two (a few centimeters) per year across the whole Basin and Range—but the widening is occurring.
Go south from Death Valley and you will splash into the Gulf of California. The geologic record shows that all of Baja California was attached to the main part of Mexico, but has been moved out to create the Gulf, and is still moving out. Eventually, the Gulf may extend up into Death Valley as the spreading continues and the west “unzips,” although the unzipping may stop before the whole west is opened up (see figure below — Pull-apart fault in the Gulf of California).
Next, boat across the Gulf with a depth-finder running, and you will see that there is a ridge down the middle of the Gulf. Get in your submarine and dive on that ridge, and you will find that it is a volcano. Lava leaks up along a crack in the middle of this ridge volcano, hardens, and then moves slowly away as new lava leaks up along the crack. Interestingly, lava also leaks into Death Valley along its faults, with eruptions within the last 200 to 300 years.
Next, follow this ridge volcano south. You will find that it runs into the open Pacific Ocean and then wraps around the globe through all of the world’s oceans, like the seam on a baseball. The ridge runs right up through the middle of the Atlantic (coming to the surface in Iceland). (Because the ridge is in the middle of the Atlantic, and the middle of the Gulf of Mexico, and the middles of some other ocean basins, it is often called a mid-oceanic ridge, although in some places the ridge is most assuredly not in the middle of the ocean.) Study the seafloor, and you will see that it is made of rocks that are young—they hardened as the lava cooled recently—near the ridge, with older rocks—they hardened as the lava cooled a long time ago—as you move away in either direction. We will discuss how we learn the ages of rocks later; for now, please accept that the rocks, and the sediments just above the rocks, are progressively older with increasing distance from the ridge.
Everywhere we meet the ridge, it is shallow and hot. The rocks that form at its volcano slowly cool as they move away, contracting and sinking. The ridge is a bizarre place in many ways—water circulates through cracks in the hot rocks, dissolves many chemicals in the rocks, and emerges as “black smokers” where the minerals reprecipitate or “rainout” in the colder ocean as a sort of black dust or as a cone around the emerging water. Many valuable ore deposits were originally formed from the chemical precipitates around these black smokers. An entire ecosystem of tube worms, clams, etc., lives on bacteria that use these chemicals for a fuel source, rather than relying on the Sun for energy, as do most living things.
We just discussed a lot of observations—a baseball-seam volcano through all the world’s oceans that is making seafloor and splitting and spreading continents and ocean basins. Explaining these observations will tell us much about the Earth and how it works.
The existence of volcanoes, bringing melted rock up from below, tells us that the Earth is hotter inside than at the surface. We learn the same in deep mines and drill holes—the Earth is warmer towards the center (once you get below the top thirty feet or so that are warmed a little by the summer and cooled by the winter). All rocks contain radioactive elements (mostly uranium, thorium, and radioactive potassium, but with some others). Radioactive decay of these makes heat in the Earth, much as radioactivity produces heat in nuclear power plants. Most of the Earth’s heat comes from this radioactive decay, although some heat is left over from when the Earth formed, or is being released as the core freezes, or the core, mantle, and crust continue to separate and things sink and give up heat from friction.
Heat is just the vibration of the atoms from which everything is made. More heat causes more vibration. Temperature is a measure of these vibrations; more vibration gives a higher temperature.
Heat is moved around in three main ways: radiation, conduction, and convection. Vibrating atoms give off electromagnetic radiation. The electromagnetic radiation that comes from very hot atoms is visible to us and is called light; cooler atoms give off infrared or other electromagnetic radiation we cannot see. When an atom gives off electromagnetic radiation, the atom cools and slows down, unless it receives other energy to speed it up again. You soak up electromagnetic radiation if you lie out in the sun (and may get skin cancer as it damages the DNA in your skin), and you can feel electromagnetic radiation if you hold your hand off to one side of a hot stove burner. Electromagnetic radiation is a very efficient way to move heat from the sun to the Earth through space and our atmosphere, but radiation is a very inefficient way to move heat through rocks—the radiation doesn’t get very far before it is absorbed. (Thus, you cannot see far through most rocks.)
If a rapidly vibrating atom (a hot one) sits next to a slowly vibrating atom (a cold one), collisions between the two will tend to slow down the fast one and speed up the slow one. This process is called conduction and moves heat energy from atom to atom. Conduction is a very rapid process over short distances. (If you foolishly touch a hot stove burner, you will almost instantly realize how quickly it makes the atoms in your skin vibrate rapidly, and how much damage can be done if they vibrate too rapidly and jump out of those places where they are supposed to be in your skin.) Conduction is a very slow process over long distances. Think of standing in a line of 1000 people. If you push your neighbor, the response is almost immediate. But if you wish to disturb someone at the other end of the line, pushing your neighbor who pushes her neighbor who pushes his neighbor... it becomes very inefficient. For the Earth, the distance from the center to the surface is about 10,000,000,000,000,000 atoms. The Earth is not old enough for heat trapped at its center when it formed to have been conducted to the surface.
Convection is the third option for moving heat. Take something hot, and move it from here to there. To get heat from the stove to your dinner table, you cook things on the stove and then carry them to the table, which is much more efficient than putting the food on the table and waiting for conduction or radiation to bring heat from the stove.
Nature has a special way to arrange this motion in many things. Heating causes almost all materials to expand because hotter molecules vibrate more rapidly and tend to bounce farther away from each other. This lowers the density of the material; its mass is the same, but it takes up more room. Low-density materials tend to rise, and high-density materials tend to sink. This leads to convection currents—a material is heated, rises, cools, and sinks. Typically, rising occurs in one place, the material flows along while it cools, then it sinks and flows back to the rising point (see the convection figure above). (Technically, moving heat by rising of materials that expanded when they were heated is called “convection,” and moving heat by transferring hot materials in other ways is called “advection,” but introductory texts usually call it all “convection.”) You should have seen the effects of convection currents in the air (we’ll talk about them later, but they produce wind, rain, etc.), and in the boiling water of a pot of spaghetti on the stove. We believe that convection currents occur in the Earth as well. This may seem odd at first because most of the Earth is solid rather than liquid. (Volcanoes come from a depth where there is a little bit of liquid, but even there most of the rock is solid.) However, sufficiently hot rocks are soft enough to flow slowly. Again, we will discuss this later, but it is the same principle that allows a blacksmith to work hot iron into horseshoes, or allows a chocolate bar to droop on a hot day.
Our modern picture of the Earth, then, is that it is heated inside, mostly by radioactivity. That heat cannot escape easily by radiation or conduction. When the Earth was young, the heat stayed in and warmed the planet. When the rocks became hot enough, they began to convect. The planet probably melted completely and convected vigorously, followed by solidification that slowed but did not stop the convection.
In convection, the hotter rocks rise and then spread. Rising occurs beneath the ridges in the oceans. There, the new seafloor is made and then rafted away on the spreading convection cells. Where a spreading zone passes under a continent, the continent is thinned and stretched and may be torn apart to make a new ocean. This is occurring under East Africa in the rift valleys, and in the Basin and Range of the western United States—including Death Valley—and occurred to open the Gulf of California, moving Baja away from the mainland. (There may be convection cells stacked on top of each other in the mantle, and other complexities—if we tried to cover all of the wonderful complexity in an introductory course, some of you would be overjoyed but many of you would be unhappy—but this is a good start.)
The Earth is layered chemically into a medium-to-high-silica crust, a low-silica mantle, and an iron core (well, there’s a good bit of nickel in the core, too). The Earth is also layered based on its ability to flow rather than break (see the Cross section of the Earth figure on the right). The lithosphere includes the crust and upper mantle. The lithosphere can flow a little in some places but usually breaks rather than flowing if you hit it or squeeze it or pull it with sufficient vigor. Below the lithosphere, the asthenosphere is that part of the mantle that flows rather than breaking, and from which many spreading-ridge volcanoes come. The topmost layer of the mantle is the bottom part of the breaks-rather-than-flows lithosphere. Below this in the mantle is the flows-rather-than-breaks asthenosphere, and then other layers that also flow rather than break, and that we don’t make you learn. And the core has a liquid outer part and a solid inner part.
The lithosphere is broken into a few basic pieces called plates. These float around on the convecting, soft asthenosphere. A plate may include just continental rocks, or just sea floor, or some of each. A plate map is given in chapter 3, in section 3.2 on Olympic National Park. The study of these plates and how they move and interact with each other is called “plate tectonics.”
Yellowstone: for many people, the name is synonymous with national parks. Yellowstone was the first national park, for a while it was the biggest (at approximately 60 miles by 60 miles—100 km by 100 km—it is still a big one), and it is widely considered to be the best. You could probably identify a dozen features or more that, separately, would merit protection by a national park. The Grand Canyon of the Yellowstone is a great chasm cut through volcanic rocks that have been “cooked” by hot water and steam circulating through them, in the process gaining the yellow color that gave the park its name. The Canyon contains two large waterfalls (rather unimaginatively called the Upper and Lower Falls), where the river plunges over more-resistant rocks. Mammoth Hot Springs is a mountain being turned inside-out, with acidic hot-spring waters dissolving caves underground, bringing the dissolved limestone to the surface, and depositing some of it as gleaming terraces. Specimen Ridge is home to at least 20 petrified forests complete with petrified roots, standing one on top of the other, from roughly 40 million years ago. Volcanic ash and debris flow buried the standing trees, and chemical reactions caused the silica in the ash to move into the wood, replacing it (a subject for much later in the course). A new forest grew and was buried, and this was repeated 20 or more times.
The biggest draws at Yellowstone are the thermal features. Various lines of evidence indicate that there is a body of melted rock (magma) under the park, now up towards the northeast side. The rocks under most of the park are anomalously hot at shallow depth. In addition, the park receives abundant rainfall and snowfall. The water from rain and melted snow circulate deeply through rocks broken by numerous earthquakes, and the water is heated from below. In some places, the water is heated all the way to steam, which emerges from holes known as fumaroles. In other places, hot water bubbles to the surface in beautiful springs. If the bubbling action mixes in enough mud, then paint pots, mud pots or mud volcanoes develop.
Sometimes, cold water on top holds hot water down, with the pressure preventing the boiling of the hot water in a pressure-cooker effect. Eventually, a little boiling manages to expel a little of the water above, reducing the pressure, allowing more boiling, and a geyser erupts. Geysers require heat, water, and a tight, tough plumbing system to hold the hot water and withstand the high pressures. The volcanic rocks of Yellowstone are rich in silica, which is dissolved and re-precipitated by the hot waters to seal cracks in the rocks, helping produce geysers. Perhaps half of the world’s geysers are in Yellowstone, including the largest and most spectacular ones. Yellowstone also is noted for many waterfalls besides those in the Canyon, for a number of other interesting features, and for abundant wildlife, which we'll visit later in the course.
Yellowstone itself is centered on the Yellowstone Caldera, a collapse feature related to three great volcanic eruptions, or periods of eruptions. The caldera, roughly 50 x 30 miles (80 x 50 km), includes Yellowstone Lake but extends well beyond it. (No lake in the nation is both higher and larger than Yellowstone Lake, yet it is only a piece of the caldera.) The eruptions occurred roughly 1.8, 1.2, and 0.6 million years ago. Each of these eruptions moved roughly 1000 times more material than did the Mt. St. Helens eruption of 1980 that we will discuss soon; thick deposits erupted from Yellowstone are known from the Badlands region of South Dakota. The erupted material that spread across South Dakota was removed from a magma chamber, and after removal, the “top fell in” to create the large depression that is the caldera.
Yellowstone has many lessons to teach us. (Some year, it would be fun to have a course on the geology of Yellowstone alone, and we certainly could fill a semester.) The size of the Yellowstone eruptions is of considerable interest, especially considering the likelihood that they will recur. Here, we wish to use Yellowstone to introduce earthquakes.
European exploration of the Yellowstone region probably began with “mountain man” John Colter during his return from the Lewis and Clark expedition in 1806, although Native Americans had used the region for thousands of years before. Colter brought back fantastic tales of the region, which were largely dismissed because they seemed impossible. Other travelers, and especially Jim Bridger in the 1850s, returned with similar tales, which also were discounted, in part because Bridger was a bit of a tall-tale teller. He is credited with stories of petrified birds sitting on petrified trees singing petrified songs (an exaggerated description of Specimen Ridge), of rivers that ran so fast they became hot on the bottom (the Firehole River, which actually has hot springs on the bottom), of trying to shoot an elk and missing because a mountain of glass was in the way (Obsidian Cliff, where rapidly-cooled volcanic rocks have made a glass called obsidian, which was mined, shaped, and traded by the Native Americans), and more.
To separate fact from fancy, the Washburn expedition from Montana (Washburn was surveyor-general of the Montana Territory) visited the region in 1870, and first developed the idea of a national park. The government-sponsored Hayden expedition of 1871 provided scientific documentation of the wonders of Yellowstone, supported by the artwork of Thomas Moran and photography by W.H. Jackson, which convinced Congress to found the park in 1872.
While in the park, the Washburn party felt earthquake activity. Breaks in recent stream and glacier deposits showed the geologists of the party that faulting had occurred recently, and motion on faults produces earthquakes. Since then, modern monitoring equipment has detected numerous quakes in the area.
On August 17, 1959, a Richter-magnitude 7.5 quake occurred, centered near the northwestern boundary of the park. Many of the geysers were changed, and a new one (Seismic Geyser) suddenly began to erupt. The ground over the quake (at the epicenter—the place above the center of the quake) was broken, with one side dropping roughly 6 feet (2 meters) relative to the other side, and with a little twisting and turning causing even larger drops in some places. A large landslide was triggered, burying a campground, damming the Madison River to form Quake Lake, and burying many highways. 28 people were killed. Some survivors had their clothes literally torn off by the immense blast of wind pushed out of the way by the huge landslide. The Old Faithful Inn was evacuated, and the west entrance to Yellowstone closed. The University of Utah’s Seismograph Station has a nice summary of the press reports. You may find it interesting to search for and read the report from the Billings Gazette that a beauty pageant was going on in the historic Inn with 800 people watching, and that “Moments after the queen had been crowned and she was walking down the aisle to the plaudits of the crowd, the first, mighty shock hit. Everyone in the place dashed for the door.”
An earthquake is just the shaking of the ground, and many things can cause earthquakes. Much effort has been devoted to detecting underground nuclear tests by the earthquake waves produced. Mining cave-ins, conventional explosions, and other events can cause earthquakes. The deepest earthquakes, which are very rare but often among the biggest ones, may have a phase-change or "implosion" origin, which we'll discuss later.
However, most earthquakes are produced by elastic rebound. We’ve already seen that rocks are moving around on the planet and that the pull-apart action has allowed Death Valley to drop down. We will see that other motions occur as well, with one group of rocks moving past another. Where rocks are warm and soft, they flow. Where cold and hard, they cannot flow.
Consider, for example, two large pieces of rock, such as southwestern California and the rest of the state. The southwestern part of the state, from Los Angeles to San Francisco, and the adjacent ocean floor are moving northwest relative to the rest of the state. The break separating the different parts is called the San Andreas Fault. (Both sides are moving westward, but the southwest side has an additional bit of northwesterly movement relative to the northeast side. Faults may go east-west, or north-south, or any other direction, may be vertical or angled, and the rocks may move vertically or horizontally or in-between across the fault.) The forces that move the rocks are huge and applied over large areas so that far from the fault the motion is smooth. But at the fault, rough patches can get stuck against each other and become locked for a while. The rocks then bend. This bending is elastic—it can spring back. Eventually, the stress on the rough spots becomes too great, the fault “lets go”, and the bent rocks “spring back”. The springing back is very rapid, in the same way as for a spring or a rubber band. Displacements of several feet (more than a meter) or more are possible in much less than a second. A building sitting on the rocks near the fault can be subjected to very large accelerations and may fall apart.
Such an earthquake will shake rocks beyond a fault. This is achieved through seismic waves—one piece of rock pushes another next to it, which pushes one next to it, and so on. Two major types of seismic waves are P or push, and S or shear. The P-wave is the ordinary sound wave. It represents a push-pull in the direction the wave is moving. A P-wave moving to the north will shake a mineral grain north-south-north-south. An S-wave moves slower than the corresponding P-wave. When an S-wave moves to the north, the mineral grains are shaken east-west-east-west or up-down-up-down (or some combination). A shear wave is similar to the wave you generate by shaking a rope. A piece of the rope moves up and down or back and forth, but the wave moves along the rope. The “wave” at football games is the same way—you stand up and sit down, but the wave moves along the bleacher bench. A P-wave may start a building shaking in one way, and then the S-wave hits the building and starts shaking it a different way, making the building more likely to break and fall down. (Earthquakes also make surface waves, which move more slowly than shear waves and go along the surface of the Earth like wind-driven waves on the ocean, rather than going through the Earth the way P- and S-waves do. The surface waves can also contribute to breaking buildings.)
S-waves don’t travel through liquids at all. (Wiggle one piece of liquid to the side, and the moving piece slides freely past the next piece rather than wiggling it.) Recall that earlier we claimed that the outer core of the Earth is liquid. You may have asked, “How does anyone know that?” The answer is that, after a really big earthquake, P-waves can be detected all over the Earth. But S-waves are missing across the Earth from the quake, in places reachable only by passing through the core, as shown by the figure. So we know that the outer core is liquid because it transmits P-waves but not S-waves. And the outer core is nearly spherical, because no matter where an earthquake occurs on the planet, there is a zone on the other side of the Earth in which S-waves are absent. (Learning that the inner core is solid is tougher; one of the pieces of evidence is based on wave conversions. The P-wave that passes through the outer core loses some energy in making an S-wave when the P-wave hits the inner core; this S-wave passes through the solid inner core, making a new P-wave when it hits the liquid outer core again, and that P-wave travels on to the surface. The delay associated with the slower motion of the S-waves allows this to be figured out. But don’t worry about wave conversions, or the other evidence for a solid inner core, in an introductory course such as this one.)
Earthquakes, as noted above, occur where rocks are moving past other rocks. We have seen that this happens where rocks are being pulled apart, as in Death Valley, because the breaks often are angled rather than vertical, and the upper side slides down over the lower. We will see that there are other situations in which rocks are being pushed together, or that rocks are sliding past each other as at the San Andreas Fault, and these also can make quakes. This mostly occurs near plate boundaries. Volcanoes can cause small-to-medium-sized quakes as well, when moving melt pushes rocks aside or leaves spaces into which rocks fall.
Quakes also occur at weak spots in continents. Often, when a continent is splitting open at a spreading ridge, the tear assumes a 3-armed form, something like the Mercedes-Benz logo. (Poke your finger through a piece of paper and you’ll often get the same thing.) Two arms then grow into a spreading ridge that forms an ocean, and the third arm fails. Major rivers often form in such failed rifts. When the Americas split from Africa and Europe as the Atlantic Ocean grew, failed rifts became the river beds of the Amazon, the Niger, and the Mississippi. You open a fast-food ketchup packet by tearing at a little notch cut in the foil, because the notch weakens the foil. If the notch isn’t there, you may have to poke a fork through or end up saying bad words, because the packet is much harder to tear without the pre-existing notch. In the same way, earthquakes can occur at the tips of failed rifts, which are the “notches” in the “foil” that is the lithosphere of the Earth. Some of the largest quakes known to have occurred in the U.S. were located at the northern tip of the rift along which the Mississippi flows, near New Madrid, Missouri. Quakes also are known from an old weakness near Charleston, South Carolina.
There are several ways to measure earthquake size. The commonest is the Richter scale, a measure of how much the ground shakes during a quake. Richter developed a logarithmic scale—a magnitude 2 quake shakes the ground 10 times more than does a magnitude 1 quake, and a magnitude 3 quake shakes the ground 10 times more than does a magnitude 2 or 100 times more than does a magnitude 1. You may think of the number of zeros after the 1: if the ground motion of a magnitude 1 quake is 10 (you can choose the units you use, or the distance from the quake at which you measure, to get a motion of 10), then at the same distance from the quake in the same units, a magnitude 2 moves the ground 100 (two zeros), a magnitude 3 moves the ground 1000 (three zeros), a magnitude 0 moves the ground 1 (zero zeros), a magnitude -1 moves the ground 0.1, and so on.
The ground motion can be measured with special instruments called seismographs. Scientists usually look at either P-waves or surface waves to get the size of the quake. The motion must be corrected for distance from the quake; the farther away from the quake your seismograph is, the smaller the ground motion you will measure. Distance can be calculated from the difference in arrival time between the first P-wave and the first S-wave from the quake to reach the instrument, using the difference in speed between P- and S-waves, or by timing the arrival of the earthquake waves at three or more stations, and determining where the quake must have been so that the waves arrived earlier at this station than at that one.
A Richter magnitude 1 quake is just big enough to feel if you are standing on the ground very near where the quake occurs. Magnitude 3 or 4 quakes are usually strong enough to convince some people to call the police (although it is not obvious what these people want the police to do), and magnitude 5 quakes usually cause some damage. The largest known quakes, around 9, each release about 10,000 times the energy of the first atomic bombs.
Small quakes are very common and large quakes rare—one or more years may pass between one magnitude-8 quake and the next one anywhere on the planet. Approximately, each increase in magnitude of 1 causes a 10-fold decrease in frequency of occurrence. But, moving the ground 10 times more takes about 30 times more energy, so most of the energy release and the damage is by the few big quakes rather than by the many little ones.
A tremendous amount of effort has gone into trying to predict earthquakes. This is because they are so destructive of life and property. Seventeen quakes are estimated to have killed more than 50,000 people each, and the worst, in Shaanxi, China in 1556, is estimated to have killed over 800,000. (In the U.S.A., the worst death toll was 503 in the San Francisco quake of 1906.) The magnitude 9.0 Tohoku earthquake in Japan in 2011 killed over 15,000 people, although the toll would have been far, far worse if the Japanese had not put so much effort into making strong buildings and otherwise planning to reduce the damages. Earthquakes usually kill by dropping pieces of broken buildings on people, but also by triggering tsunamis (big waves) or landslides, and by breaking gas lines or other things that cause fires.
The easiest way to “predict” quakes is to identify those places where quakes are likely to occur. This can be done from historic records, and from prehistoric geologic evidence. A pattern of landslides of a single age in a region, or of drowned forests related to land subsidence, may indicate the effects of an earthquake. Once people know where quakes are likely, appropriate zoning codes for buildings can be enacted. Spending a million dollars on special engineering for a building to survive quakes is wise indeed near the San Andreas Fault, and really helped Japan, but may not be necessary in State College, Pennsylvania, where big quakes are considered to be highly unlikely.
Other ideas have been advanced for predicting quakes. One is to use patterns of seismicity. If one section of a fault has had a quake every 20 years for the last century, you might expect another quake 20 years after the most recent one. Also, a fault that is slipping and moving may have lots of little quakes, whereas a locked fault that is building to a big quake may have no little quakes. So you could use such a no-quake “seismic gap” in predictions. Such a pattern—historical repeats and a seismic gap—was used recently to predict a quake near Parkfield, CA on the San Andreas Fault. The prediction failed miserably. The possible explanation is that there are many faults in that part of California, and several big quakes occurred on faults near the San Andreas shortly before the expected quake. These certainly perturbed the state of stress at Parkfield, at least delaying its quake. (The other quakes allowed rocks to move, and even caused some faults to run backwards compared to their normal behavior for a while, taking the load off and delaying the next quakes.) This illustrates the difficulty of predicting quakes. Perhaps, if the motions of all of the important blocks were monitored, one could model the whole system and do a better job of predicting where stresses are accumulating. Such work is ongoing, but results are not yet in.
Even if pattern-predictions of earthquakes can be made to work, the predictions are unlikely to be precise enough to really tell us what we want. Knowing that a quake will occur sometime in the next few years, or even the next few days, does not allow us to get people out of old buildings, off bridges, etc., during the quake. The best hope for predictions within hours or minutes is to find premonitory events. As the stresses build toward failure, rocks may begin to crackle, groundwater may move around in the cracks so that water rises in some wells and falls in others, electric signals may be given off by the cracking rocks, and animals may act strangely. The difficulty is that groundwater rises and falls in wells for many reasons, strange actions in an animal may indicate bad feed, mating season, or any number of other causes, and other premonitory events of earthquakes also have non-earthquake causes. The subtle clues to an earthquake may be recognized using 20/20 hindsight, but no one has figured out how to read them in advance. The possibility is there, though, waiting for brilliant insights and hard work by some interested researchers. (After a big quake, lots of people show up claiming that they predicted it, but none of these “predictions” has ever been verified. And, many people, including some scientists, have made predictions of particular earthquakes to come, but again, these predictions have not proved to be useful.)
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Death Valley National Park (Provided by UCGS, 3D version requires red/cyan stereo viewing glasses)
GeoMations are drawings that are captured as animations, narrated by Dr. Richard Alley and posted on YouTube for you to review at your own convenience.
The material in these drawings is almost never new—rather, it is a visual presentation of ideas and concepts you'll have already encountered in the readings and lectures. It's also important to note that these drawings are like the rest of the material presented to you in this class — that is, you should expect to be familiar with all the GeoMations in order to be successful in your weekly quizzes, just like you need to be familiar with the readings, the notes, the slide shows, and anything else included this week.
The great depth of Death Valley was not excavated by a steam shovel or an atomic bomb, nor were the rocks ground down by rivers, wind, or glaciers. Death Valley was dropped, as the sides of the valley were pulled apart as part of the great motions of the planet’s rocks. High drama indeed. Here is a brief description:
A person zipped into too-tight pants may “leak out” as the zipper is lowered. Baja California is being unzipped from the mainland of Mexico, and the leak is a volcano making new seafloor. If the unzipping continues, the sea might someday extend up toward or into Death Valley. Get a good grip on your zipper pull, and let’s go see.
You can push things together, pull them apart, slide them past each other - or some combination of pushing or pulling while sliding. Nature does the same, giving different types of faults, which are found in different geological settings. Here’s a quick look.
Dave Janesko, a geologist, was one of the students on the CAUSE trip. Here, he and Dr. Alley explore the great Sevier Fault just west of Bryce Canyon in Utah. The pull-apart action that opened Death Valley actually affected a lot of the west and is responsible for the Sevier Fault. The red limestone of Bryce was deposited in a lake. Much later, black lava flowed over the top and cooled. Then, the faulting occurred, and the black rocks were dropped to lie next to the red ones. Where the two meet along the nearly-vertical although slightly inclined fault, the red rocks show no sign of having been heated by the lava, so they must have been put together when cool. The red and black show near-vertical scratches, formed when the rocks slid past each other to get where they are now. So, join Dave on the fault.
Here are some optional videos and animations you might also want to explore! (No, these won't be on the quiz!)
The original "geyser" in Iceland
Yellowstone has roughly half of the world's geysers, but there are geysers in New Zealand, Iceland, and elsewhere. The original "geyser" is Geysir, in Iceland. Here is a short film clip of Geysir erupting, just for fun. Filmed by Vicki Miller.
Plate Tectonic Movement Visualizations
Watch the NSF-sponsored video on Plate Boundaries
San Francisco Earthquake Aftermath 1906
(Amazing film footage that is now over 100 years old!! Click on one of the "View Movie" options on the left side of the new page to watch the 1:39 minute clip.)
Please watch the 50-minute Unit 2 lecture featuring Dr. Richard Alley.
Check out the Unit 2 PowerPoint Presentation used in the online lecture here.
We've been having fun making Geosc 10 Rock Videos, proving that: 1) The material can't be too hard, because we can put it to music, and 2) There are really good reasons why Dr. Alley became a professor instead of trying out for American Idol. Most of the Rock Videos are written to review part or all of the material in a unit, although we occasionally toss in a bit of extra stuff to make it rhyme. Here, we've borrowed the tune from the great Johnny Cash hit "Walk the Line" to sing the praises of seismologists (Dr. Anandakrishnan happens to be a famous seismologist...). This video reviews a bit about earthquakes, and then shows some things we'll consider soon--volcanoes next week, and tsunamis the week after. (We never really worry about evil dictators in this course, but here's one for good measure.) So, rock on!
You have reached the end of Unit 2! Double-check the list of requirements on the Unit 2 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Continue to work on Exercise #1. See Course Calendar for specific dates.
Following are some supplementary materials for Unit 2. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
They tied pillows on top of their heads as protection against the shower of rock. It was daylight now elsewhere in the world, but there the darkness was darker and thicker than any night…(t)hen came a smell of sulfur, announcing the flames, and the flames themselves…he stood up, and immediately collapsed…his breathing was obstructed by the dust-laden air.
—Account by Pliny the Younger of the death of Pliny the Elder in the eruption of Mount Vesuvius in A.D. 79 that destroyed Pompeii and Herculaneum in Italy. [Living with a volcano in your backyard: an educator's guide with emphasis on Mount Rainier]
Supposedly, Atlantis was an island civilization "outside the Pillars of Hercules" and thus located in the Atlantic Ocean, where it was destroyed by an earthquake or tsunami (giant wave) about 11,000 years ago. The source of this information (according to Wikipedia) is an account that Plato wrote in 360 BC of information reportedly given to Solon two hundred years earlier by priests he visited in Egypt. Now, if someone told you that 200 years ago someone else had received information by yet another person regarding something that happened 9000 years earlier, would you immediately believe it? A lot of people apparently do; a search of Google for "Atlantis Plato" finds about 327,000 2.8 million matches, and not all of them are academic discussions.
A better question might be whether there really are islands that disappear below the sea. The answer is yes; many do. Some slide slowly downhill, at about the same rate as your fingernails grow, and disappear first beneath the waves and finally beneath the continents. Others suddenly explode, scattering themselves across the world.
Before we go any further, take a look at the following short video introduction by Dr. Anandakrishnan...
You will have one week to complete Unit 3. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 3, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read /view all of the Instructional Materials |
No, but you will be tested on all this material. |
Submit Exercise #1: Scientific Literature | Yes, this is the first of 6 Exercises and is worth 5% of your total grade. |
Begin Exercise #2: Geology is All Around You | Yes, this is the second of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 3 "RockOn" quiz | Yes, this is the second of 12 end of unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #4" survey | Yes, this is the third of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 3, including the online textbook, a video lecture, several vTrips and animations, and two overview presentations.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Crater Lake, at 1932 feet (about 600 m) deep, is the deepest and probably the cleanest lake in the United States, and surely among the most beautiful. Crater Lake sits in a great volcanic crater or caldera, 5 miles (8 km) across, formed when Mt. Mazama experienced a cataclysmic eruption about 6600 years ago. That massive eruption laid down ash that is 200-300 feet thick (almost 100 m) on the flanks of the volcano; the ash forms a recognizable layer hundreds of miles away in Yellowstone, and has been identified in Greenland ice cores.
The peak of the volcano had risen more than a mile above its mile-high base on the highlands of southwestern Oregon, but the great eruption removed about 4000 feet (1200 m) from the mountain’s height. About 16 cubic miles (40 cubic km) of rock were blown away. Glaciers had flowed down from the mountain peak; today, the glacial valleys can be followed upward until they disappear at the caldera rim. Although 50 feet (15 m) of snow falls in a typical year now, melting in the summer is sufficient to remove all this snow, and no glaciers exist. A tongue-in-cheek Christmas celebration on Aug. 25 substitutes for the snowbound December event.
After the great eruption, lava flows began building Wizard Island. If the water were removed from the lake, you could see that Wizard Island is roughly 0.5 mile (0.8 km) high. No permanent streams feed into the lake; the great rainfall and snowfall in the crater are balanced by evaporation, and by seepage through the rocks and eventually out the sides of the volcano as springs. With no streams supplying sediment, the lake is exceptionally clear and clean. Aquatic moss receives enough sunlight to grow 425 feet (130 m) below the water surface. When trout were stocked in the lake, freshwater shrimp were stocked first because otherwise biologists feared that the trout would have nothing to eat.
We will discuss volcanoes soon. For now, note that Crater Lake sits atop one of a string of volcanic peaks: Lassen Volcanic National Park and Mt. Rainier National Park preserve other peaks in the Cascades range. Mt. St. Helens, Glacier Peak, and several others are protected federally. These peaks line up in a row, called a volcanic arc, parallel to the coast. A similar arc sits along much of Central America, and forms the Andes of South America. And similar arcs also occur as island chains—the Aleutians, Japan, and others. In fact, the Pacific Ocean is almost entirely encircled by such volcanic arcs, forming the “Ring of Fire”.
Sitting offshore of the Ring of Fire is a ring of trenches, which include the greatest depths of the ocean. The trenches parallel the arcs. Some trenches, which sit near continents, are nearly filled with sediments dumped off the continents, but other trenches are almost free of sediment and so have very deep water, to almost 7 miles (11 km) deep. (Figuring out depths often is complicated by sediment. The surface of Death Valley sits more than two miles lower than the adjacent peaks of the Sierra Nevada. But below the salt flats of Death Valley there are sediments as much as three miles thick, materials that were eroded off the tops of the peaks, so the valley has dropped by much more than three miles relative to the peaks.)
The trenches and volcanoes that ring the Pacific are clues to help solve a problem that might have been bothering you from last time. If the sea floor is made at spreading ridges and then moves away, where does it go? The earth is not getting bigger. (Well, meteorites are adding a tiny, tiny bit, but not nearly enough to account for sea-floor spreading.) So, the sea floor must be disappearing somewhere. The oldest sea floor we know of is only about 160 million years old, but the continents contain rocks as old as almost 4 billion years, suggesting that the sea floor is being consumed before it gets very old. (Remember, before the class ends, we’ll discuss how geologists date rocks.) (And remember that when a geologist dates a rock, it involves physics or chemistry but not dinner or a movie.)
Now, a few more clues. Sea floor is made of basalt. This is just the kind of rock that would be made if you melted a little bit of the deep, convecting rocks of the Earth’s mantle, and let that melt float up to the surface and “freeze” (cool until it solidifies). If you take basalt, a little ocean sediment, and some ocean water, and heat them enough to cause a little melting, let that melt come to the surface and freeze, you obtain a rock called andesite with a little more silica and a little less iron and magnesium than basalt, lighter in color and lower in density than basalt. Interestingly, the dominant rock in the walls of Crater Lake, and in the other Cascades and Ring-of-Fire volcanoes, is andesite (named after the Andes, which are part of the Ring of Fire).
If sea floor were plunging under the continents and melting to make andesite, you might expect that occasionally the down going rocks would get stuck and then break free, making earthquakes. Indeed, a three-dimensional map of earthquakes shows that shallow ones occur near the trenches, and the quakes are progressively deeper inland beneath the volcanic arcs. The great 1964 Alaska earthquake was such an earthquake, which happened where rocks of the Pacific Ocean floor plunge to the north under coastal Alaska and the Aleutian chain. The earthquakes there are shallow to the south and deepen to the north, along the down going rocks. The disastrous 2011 Tohoku earthquake in Japan was of the same type.
Careful work on the speed of earthquake waves, which is affected by the temperature of rocks, even provides a picture of cold slabs going down into the hotter mantle. As these down going slabs of rock are heated, with their water and sediment, a little melt is produced. (Interestingly, wet rocks melt at a lower temperature than do dry rocks, just as adding a little water to flour and yeast speeds cooking in the oven.) When the melt rises to the surface and cools, andesite forms, such as is seen around Crater Lake, in the Andes, or in the Aleutian volcanoes.
So, sea floor is made at the spreading ridges. It is hot and low-density initially, but cools and contracts as it gets older and loses heat to the colder ocean water. When the sea floor becomes cold and dense enough, it can sink back into the mantle, and we call the place where it sinks a “subduction zone”. The sinking sea-floor slab drags along a little sediment and water, warms up and melts a little of the slab and some of the sediment, and feeds the volcanic arcs. Old sea floor is going down around much of the Pacific Ocean, and in a few other places, such as beneath the Caribbean and beneath portions of the Alps. Wherever this happens, andesitic volcanic arcs form (see Figure 1).
The Olympic Peninsula juts out into the Pacific Ocean, separated from Seattle by Puget Sound. Moisture-laden winds off the Pacific dump more rain and snow on the Olympic than anywhere else in the lower-48 United States. Great old-growth forest trees—Sitka spruce, Douglas fir, etc.—tower up to 300 feet (almost 100 m) above the forest duff, where butterflies flit past crystalline streams and cascading waterfalls. Along the coast, sea lions bask on offshore stacks, while urchins and starfish populate tidal pools. On the “high peaks,” numerous glaciers form and flow downhill. More snow accumulates than melts above about 6000 feet (2000 m). On most mountains, you have to go much higher to find summertime snow, but the huge winter snowfall on the Olympic allows the peaks to be snow-clad year-round despite rising less than 8000 feet (about 2500 m) above sea level.
Olympic National Park is a bit unusual in that it was established as much for biological reasons as for geological—to protect the Roosevelt elk that live on the peninsula. (The elk, named after Theodore Roosevelt, were critical in obtaining national monument status, which was signed by President Theodore Roosevelt. Later, the upgrade to national park status was signed by President F.D. Roosevelt. The Roosevelt elk is the largest of the elk subspecies in the country. Some consideration was given to naming the park Elk National Park before Olympic was chosen.)
The geologic story of the Olympic is somewhat shorter and less dramatic than for most of the national parks. The rocks of the Olympic are almost all young--less than 40 million years. (Again, please bear with us—we will justify these numbers before the course ends!) Before that, the coastline must have been farther to the east, perhaps in North Cascades National Park, and before that even farther east.
You’re standing at the grocery-store checkout. You put a bag of potato chips on the conveyor, and off they go, followed by a case of Pepsi, three loaves of bread, a watermelon, a box of Ho-Hos, and a sack of potatoes. Then, you realize that there is no bagger working, and that everything is piled up at the end, in a BIG mess. That mess is a good model for the Olympic Peninsula, and the whole coast from there up to Alaska.
The rocks of the Olympic Peninsula are a mixture of sea-floor basalts and of the sorts of sediments that accumulate today off the coast and fill the trench there. Rivers draining the peninsula, and much of the West Coast, carry great loads of sediment down to the ocean. Much of that sediment piles onto sea floor that is slowly moving beneath the continent, a conveyor belt that tries to pull the sediment down to melt and be erupted. Most of those sediments are “scraped off” on the way down, just as at the grocery store. The Olympic Peninsula is the off scrapings. Most of the rocks have been bent and twisted from the attempt to shove them under the continent (think of the potato chips after the Pepsi hits them!). Some of the Olympic rocks have been heated a good bit—the conveyor belt took them down a ways, but they then were squeezed back out.
Our emerging picture of plate tectonics is that the earth is heated inside, softening the deep rocks of the asthenosphere enough that they can move in great, slow convection currents that transfer heat from deep in the earth to near the surface. Heat is conducted through the upper rocks, or is erupted through them by volcanoes, and eventually is lost to space. But, the upper rocks in most places are cold enough that they tend to break rather than flow—they are brittle. These brittle rocks form the lithosphere, which includes the crust and the uppermost mantle. The rocks of the crust in continents are rich in silica (often like andesite in composition), making them light in color and low in density, so that they float on the deeper rocks and are rafted around on them by the moving convection currents. The sea floor rocks in the crust fall between the continents and the mantle in composition, and typically are basalt. The sea-floor rocks are usually intermediate between continents and mantle in density as well, but if the sea-floor rocks are cold enough, they will be slightly denser than hot mantle. Then, the sea-floor lithosphere consisting of the sea-floor crust plus a little attached mantle will sink into the asthenosphere of the deeper mantle.
The lithosphere is broken into a few big rafts, called plates—eight big ones plus some smaller ones, depending a little on how you define “big” and “small”—that float around on the convection cells below. Plate boundaries include spreading ridges where the plates move apart (remember Death Valley and the mid-ocean ridges), and subduction zones where the plates come together and one side sinks under the other. You might imagine that if plates can come together or pull apart, they must be able to slide past each other as well, which is what happens at the San Andreas Fault in California (we met it when we were discussing earthquakes); such slide-past boundaries are often called transform boundaries or transform faults (see Figure 1). You might worry that sometime, two continents would run together; we’ll meet that soon when we visit the Great Smoky Mountains.
The lithosphere and asthenosphere are solids, but a little melted rock may occur in the asthenosphere, and leak out where plates are pulled apart, feeding basaltic volcanoes. And the water taken down subduction zones can stimulate a little melting, feeding andesitic volcanoes that line up in arcs above the downgoing slabs of the subduction zones; examples of these volcanic arcs include the Cascades, Aleutians, and Andes. Continents are a collection of scum from melting of the mantle, too low in density to sink back into the mantle. Continents grow as the conveyor belt from the mid-ocean ridge to the subduction zone brings in sediments and islands and what-not, or when andesitic volcanoes are erupted on continents, or when andesitic volcanoes form an arc offshore that then collides with a continent (sometimes the site of subduction moves, and the volcanoes find themselves on the conveyor belt, or they hit a different continent). Because much of the sediment comes from the continents themselves, the growth of continents is not fast—material eroded from the continents falls on the conveyor and is added back at Olympic or erupted back at Crater Lake.
Mt. St. Helens, in southwestern Washington, was in some ways the queen of the Cascades Range. Beautifully symmetric, snow-capped, it had been called the Fujiyama of the Pacific Northwest. Scores of people flocked to St. Helens’ flanks to hike, camp, ski, and generally enjoy the environment. And all that changed in 1980.
Mt. St. Helens has also been the most active of the Cascades volcanoes over the most recent centuries. In early 1980, the volcano clearly was “waking up”. Earthquakes shook it almost continuously, including special “harmonic tremors”, similar to those sometimes caused by fluid flow in pipes, which showed that liquid rock was moving up from below. Small eruptions occurred, and hot springs and fumaroles (steam or gas vents) became increasingly active. The north side of the mountain was bulging, blowing up like a balloon as the magma moved into it. Scientists were scrambling to study the volcano, and predict its course. They recommended evacuation for safety, and most people (but not all, including some scientists) were moved out of the way. Penn State professor Barry Voight warned that the huge bulge on the north side of the volcano would fail, unleashing a giant landslide and a devastating eruption.
On the morning of May 18, 1980, Professor Voight’s prediction came frighteningly, awesomely true. The bulge failed. A large earthquake either caused, or was caused by, failure of the north side of the mountain in a giant landslide. Like pulling the cap off a hot, well-shaken soda bottle, the liquid beneath flashed into froth, driving an eruption 12 miles (20 km) high. A shock wave knocked over full-grown trees in an area 20 x 10 miles (32 x 16 km). The landslide eventually poured more than 100 million cubic yards of rock material down the Toutle and Cowlitz Rivers, raising the floor of the North Fork of the Toutle as much as 600 feet (200 m), and sweeping roads and houses downstream, with the debris reaching and clogging the shipping channels of the Columbia River. The Toutle floor now sat higher than the smaller streams that fed it, and lakes began to form; only quick work by the Army Corps of Engineers prevented those lakes from overtopping the mud damming them, cutting quickly down through it and releasing further floods.
All told, the Corps of Engineers spent $250 million clearing shipping channels and doing other critical work. Approximately 60 people were killed in the blast and landslide; some were buried under hundreds of feet of steaming mud and their bodies were never recovered.
President Jimmy Carter scowled at the disaster from a helicopter. Disaster planners pontificated. And in the shadows of the other Cascades volcanoes, people continued building houses in regions of known volcanic hazard.
The Mt. St. Helens Volcanic Memorial today has little in common with conditions pre-1980. The center of the volcano was more than 1/2 mile (nearly 1 km) lower after the eruption than before, with the missing rock spread over the surrounding countryside, forming a visible layer as far as 900 miles (1500 km) away. (Professor Alley and his wife Cindy were driving in Alberta, Canada during the summer of 1980, on a great, seven-week, see-the-national-parks-in-a-Chevette-with-a-tent honeymoon. A secondary eruption of Mt. St. Helens put enough ash in the air to halt traffic because of reduced visibility, hundreds of miles from the volcano.) Many of the trees knocked over by the blast still lie there—hundred-foot-long toothpicks pointing in the direction of the searing winds of the blast. Among these dead trees, however, salmonberry and fireweed and young firs are pushing skyward, elk are grazing, and coyotes search for rodents. In some places, salvage-logging of the downed trees was allowed. In some of those places, it appears that erosion has run amok, large gullies have developed, and the return of vegetation has been greatly slowed. In the crater of the volcano, a new lava dome is forming, squeezing slowly upwards like toothpaste from a very hot tube, and amazingly, a glacier sits behind it, fed by the great snowfall and the avalanches down into the crater. At night, the dome glows dull red. (It may seem weird that we focus on an event from before most of you were born, from 1980, when larger eruptions have happened more recently. But, St. Helens is the largest recent eruption in the lower-48 of the US, the easiest eruption site to get to and observe, and it really is awesome. The elder Alley daughter, Janet, was a ranger there one summer, and recommends that you take in Ape Cave if you visit, but the real goal is to see just how immense the eruption's effects really were.)
Volcanoes occur where melted rock rises to the Earth’s surface. Almost all volcanoes are associated with one of three settings—pull-apart margins (spreading ridges), push-together subduction zones, and hot spots. We’ve already met the volcanoes at spreading ridges, where low-silica basalt is erupted, and those producing higher-silica andesite at subduction zones.
Hot spots are creatures of another type in the zoo of Earth’s features, and especially interesting creatures at that. Deep in the warm, soft, convecting mantle of the planet, in some places a rising tower of hot rock forms and then lasts for quite a while. Some of these rising towers may come all the way from the base of the mantle where it meets the core; others may start shallower. (To see something that looks vaguely like the formation of such a hot spot, go back and view again the “lava lamp” film of Dr. Anandakrishnan in the introductory material to this unit.)
As the lithosphere drifts overhead, the hot spot may “punch through” to make a volcano. Then as the lithosphere carries that volcano away, the hot spot punches through a new place to make a new volcano. Rising melt behaves a little bit like people driving cars, who use one road or the other but not the lawn in-between; hot spots often make a string of separate volcanoes rather than a continuous line or ridge, by coming up through one hole in the lithosphere for a while and then switching to a different one. Hot spots bring melt from deep in the mantle, and so normally make basaltic volcanoes (it takes fairly subtle and sophisticated chemistry to tell the difference between hot-spot basalt and sea-floor basalt from spreading ridges). However, where a hot spot pokes through a continent rather than through sea floor, silica from the continental rocks may mix with the melt to increase its silica content, as at Yellowstone.
When a new hot-spot first rises from below, the top must push through the mantle and crust, and the resistance of the stuff in the way of the rising column causes its top to spread out like the head of a thunderhead, or of a mushroom cloud from an atomic bomb, or of a blob in a lava lamp, and for the same reasons. When that wide head reaches the surface, immense lava flows can be produced that spread across state-sized areas and bury them hundreds of feet deep. Much of central and eastern Washington and Oregon is buried by the “flood basalts” from the head of the Yellowstone hotspot. As the continent has moved across the hot spot after it reached the surface, a string of volcanoes erupted, including Craters of the Moon National Monument in Idaho. The hot spot now fuels Yellowstone (which is why it is called the Yellowstone hot spot…a lot of this stuff isn’t that difficult!).
The flood basalts from the Hawaiian hot spot have been subducted beneath the Aleutians, but the string of volcanoes that has formed since from the hot spot can be traced from the Aleutians south all the way to the volcanically active big island of Hawaii. A new volcano, Loihi Seamount, is growing undersea just off Hawaii as the big island slowly drifts away, and Loihi will someday (not in our lifetimes!) be the new, volcanically active “big island” while the volcanoes of the current big island die and the island slowly erodes away.
So, melt can leak up from below to feed volcanoes at spreading ridges, at hot spots, and above subduction zones. But very different volcanoes develop: sea floor from spreading ridges; flood basalts and then wide, not-very-steep Hawaii-shaped volcanoes from hot spots; and, steep Mt. St. Helens-type volcanoes above subduction zones. The type of volcano that develops at a place depends on a host of factors: temperature, composition, supply of melt, duration of supply, and several others. We will focus on two: composition (how much silica) and volatile content (mostly how much water, although carbon dioxide, hydrogen sulfide, and other compounds that are gaseous under earth-surface conditions may be present and classified with volatiles). Silicon and oxygen get together in melt to form the material we call silica. Left to itself, each silicon atom will be surrounded by four oxygen atoms, which form a tetrahedron (a little pyramid). But, give them a little time, and the tetrahedra will start sticking together, or polymerizing, into chains and sheets and bigger clumps, with some oxygens serving in more than one tetrahedron. If these lumps get big enough, they are minerals and the melt has solidified.
When the lumps are present but not too big, the melt is like lumpy oatmeal—it doesn’t flow very well. There are three ways to get rid of the lumps: make the melt really hot; fill the melt with iron, magnesium and other elements that interfere with the tetrahedra polymerizing; or, fill the melt with volatiles that interfere with the tetrahedra polymerizing. When polymerization is low, the melt flows easily. Lava comes out of the volcano quietly, without making big explosions, and flows easily and far from the mouth of the volcano. In extreme cases, flows may be nearly horizontal and cover much of a state, as in the flood basalts. If the melt spreads almost as easily as flood basalts, the lava will have very slight slopes of only a few degrees, forming shield volcanoes (they look like a warrior’s shield lying on its side) such as Hawaii. Hawaiian lavas and flood basalts flow easily because they are hot and are high in iron and magnesium.
When volatiles remove the lumps, a different situation develops. This is because the volatiles will only stay in the melt under high pressure. Just as a bottling company can force CO2 into the water of a soft drink under high pressure, but the CO2 escapes as the pressure falls when you open the can, the water and CO2 and other volatiles stay in the melt under high pressure down in the Earth but escape when the melt gets close to the surface.
Silica-rich melts usually form with many volatiles. Remember that in subduction zones, wet sediment dragged down the trench releases water (and carbon dioxide and others) that promote melting. When the melt (called magma when it is in the Earth and lava when it reaches the surface) nears the surface, the lower pressure allows the volatiles to bubble off and escape into hot springs, geysers, etc. (Note that most of the fluids that come out of such hot springs are rainwater that has circulated down into the earth, but some of the fluids may be “juvenile” waters from the magma below.) Silica-rich, relatively cool lava that has lost its volatiles flows only with great difficulty. It may emerge from the volcano and flow a little ways as a very thick, slow-moving, steep flow. It may not even flow, but simply form a dome directly over the volcanic vent. And, it may “plug the system” when it solidifies. Then the stage may be set for a big explosion.
The next melt that rises in the volcano cannot follow the same path, because hardened lava above prevents escape. The gases are trapped, and pressure builds up. The volcano is like a hot, shaken pop bottle. If the top is removed, either by an opener (say, a landslide as in the case of Mt. St. Helens, or a crack opened by an earthquake) or just because the pressure becomes great enough to blow the top off, the sudden release allows the soda, or the magma, to come foaming out. A good champagne may fountain to many times the bottle’s height, and blast the cork across the room. A powerful volcano may blast ash higher than jet flight paths. The melt really does get foamy, and that foam hardens into little glass shards. The ash layer deposited by Mt. St. Helens, which stopped drivers hundreds of miles away, was mostly of such little glass shards, although torn-up bits of the former volcano were also included.
The andesitic volcanoes of the Ring of Fire are typically stratovolcanoes, formed of alternating layers of thick lava flows and of pyroclastics—things thrown through the air by the volcano. The steepness comes from the flows, which cannot get far from the vent. Some of the andesitic volcanoes, including the rebuilding Mt. St. Helens, include plug-dome elements, the oozing lava staying right above the vent.
So, the major volcanoes for our purposes are the quiet, basaltic shield volcanoes of hot spots, the quiet basaltic rift volcanoes of spreading ridges, and the steep, scenic, explosive, andesitic volcanoes of the Ring of Fire. Other types exist, notably cinder cones thrown up by typically minor eruptions tossing pyroclastics short distances. Also, hot spots or rifts trying to poke through continental rather than oceanic crust may produce explosive silica-rich volcanoes. But if you understand shields and stratovolcanoes, you will be a long way toward understanding volcanism.
People who live near volcanoes should be worried about them. Volcanoes can do much damage. The volcanic-triggered landslide that buried Armero, Colombia in 1985, and the eruption of Mt. Pelee on the island of Martinique in the Caribbean in 1902, each killed about 30,000 people. Other volcanic disasters bring the human death toll to perhaps 200,000 over the last few centuries. Compared to war, disease, or even automobile accidents, this is not a terribly high toll; however, the 200,000 people directly involved almost certainly would have appreciated enough warning to get out of harm’s way. One of the goals of modern geology is to predict volcanic hazards and to save lives and property by doing so.
There are many hazards to worry about. These include:
So, we wish to predict when and where volcanoes will endanger people. Various things can be done. For problems such as climate change, the best we can do is to know that every few years or decades some region is likely to experience difficulties with crop production because of eruptions. The solutions are either to maintain a little excess food to feed those endangered people, or to ignore them and figure that some will starve to death. (Many other climate changes, including droughts, give us the same choice. Despite the apparent silliness—either we stockpile food and figure out how to distribute it to the needy, or we let people starve to death—it is surprising how often starving to death is the outcome.)
For tsunamis, an operational warning system now exists around the Pacific and in some other places. If a tsunami is detected, or if seismographs detect the shaking caused by a large earthquake, landslide, or volcanic eruption, warnings are relayed to coastal regions likely to be affected, in time to allow evacuation. Such a system is being assembled for the Indian Ocean following the tragic 2004 earthquake-generated tsunami that killed approximately 250,000 people.
One way to avoid volcanic hazards is to stay out of harm’s way. Geologists can map regions where large pyroclastic chunks have fallen, or where landslides have occurred, with great confidence. Using carbon dating of logs caught in debris flows, or tree-ring dating of those growing on landslides (just hang on; explanation of such things will come later), we can determine the recurrence interval—how often do such disasters happen? Today, whole subdivisions are being built around Mt. Rainier National Park in the growing Seattle-Tacoma region that have a danger of destruction by landslide many, many times higher than their danger of destruction by fire. Roughly 200,000 people work, and more than 100,000 people live, on debris-flow deposits less than 10,000 years old, with more people coming. The homeowners will all carry fire insurance, but few if any are insured against the volcano; presumably, if they survive the next volcanic disaster, these people are counting on disaster aid from the rest of the country to bail them out.
(Much argument is attached to sending disaster aid for predictable events, even if they are not very common. Should those who wish to live in beautiful but risky areas carry insurance to pay for their gambles? Increasingly, planners are saying “yes,” and much effort is being devoted to quantifying the hazards so that insurance rates can be set wisely. This applies to such things as hurricanes along coasts, earthquakes along faults, and floods along rivers. Geologists have an important role to play in learning hazards and thus setting rates.)
With sufficient care, volcanic eruptions can be predicted with some confidence. Volcanoes usually give off many signals before an eruption: the ground swells as magma moves up; the moving magma and the swelling ground create earthquakes and especially the distinctive harmonic tremors of fluid flowing in a pipe; small eruptions occur; gaseous emissions increase as the magma nears the surface and then cease if the system becomes plugged and builds up pressure for an explosion. A monitoring program of seismographs to detect earthquakes, repeat surveying of laser reflectors set on the mountain to watch for deformation patterns, gas sampling, and perhaps photographic or other sensors to watch for landslides, can track a volcano’s behavior and allow timely warning. Monitoring of ground shape from space can even see the changes in volcanoes as magma moves under them. The eruption of Mt. St. Helens was predicted well enough to save hundreds of people including the residents of a YMCA camp. The eruption of Mt. Pinatubo in the Philippines in 1991, which heavily damaged the U.S. military bases there, was predicted accurately, allowing timely evacuation and saving tens or hundreds of thousands of lives of residents and military personnel.
The burden of predicting eruptions is very high. Imagine telling an Air Force general to abandon his or her assigned duty post, spend a few hundred thousand dollars to move tens of thousands of people, and then having nothing happen—the general, and all of those people, would be very unhappy. Imagine instead deciding to wait another day to be sure, and having all of those people (possibly including you) killed. Important as this is, predicting disasters is not for the faint of heart.
The Mt. St. Helen’s eruption was a small one compared to many others. Each of the major eruptions of Yellowstone moved about 1000 times more material than Mt. St. Helens did, and Yellowstone’s eruptions were not the largest known. Small eruptions are more common than large ones. But, eruptions ten times as big as Mt. St. Helens are perhaps five times as rare, but not ten times as rare. This means that, as for earthquakes, most of the “work” done by volcanoes is achieved by the few big ones, not the many little ones.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 3.
TECH NOTE
Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Hawaii Volcanoes National Park and a second Crater Lake Slideshow
(Provided by Dr. Alley)
Hawaii
(Provided by USGS)
Crater Lake National Park
(Provided by UCGS)
Crater Lake National Park - 3D version
(Provided by UCGS - red/cyan stereo viewing glasses required)
Mount St. Helens National Volcanic Monument
(Provided by UCGS)
Olympic National Park
(Provided by UCGS)
There are three Unit 3 GeoClips (movies) linked below. We hope they help you understand Unit 3 just a little bit better, and that you enjoy them.
The hot spot of Hawaii erupts runny lava to the surface, giving some very interesting features, such as the lava tubes you will see forming in the first video, and formed in the second video. The hike out to the flowing lava was, in spring of 2007, over three miles across rough, often broken and glassy lava that solidified from glowing hot flows over the last couple of decades. Whales were spouting offshore when Dr. Alley and family made the trip. Tag along, and see what they saw.
Lava was erupting in the Southwest Rift of Kilauea not that long ago. Sometimes, the lava erupts with a little force, throwing pieces that freeze to glass in the air and rain down. Other times, the lava flows even more quietly along the surface. Here, you can see evidence of both.
Optional Videos, for your enjoyment (and education, but you won't be quizzed on them.) Volcanoes are just too interesting to leave so quickly, so here are some more looks at these important, and dangerous, pieces of our planet. First, visit Hawaii again, and see some strange things. Then, head over to Sunset Crater Volcano National Monument, Arizona, with the CAUSE class. Have fun, and keep your feet cool!
Hawaiian lava flows engulf whatever is in their way, including trees. What happens when hot lava hits a cold, wet tree? Find out here, your chance to look down on Dr. Alley.
Kilauea Volcano is a wonderful place to visit. Stay in the lodge on the rim, and you'll wake up to the view shown here behind Dr. Alley.
Cinder cones are rather odd volcanoes, formed of pyroclastic bits tossed through the air to pile up near the vent. If you let the spaghetti sauce boil on the stove, without a lid, you would soon have a lot of tomato-sauce blobs around the pot. Let those build up, and you are heading for a cinder cone. Here, see three different versions of the cinder cones at Sunset Crater.
An explanation of cinder cone volcano formation by CAUSE student Sam A.
Another, slightly "dramatized" explanation of cinder cone volcano formation by CAUSE students Stephanie S. and Raya G.
A third explanation of cinder cone volcano formation, by Dr. Alley himself.
There are TWO Unit 3 lectures both featuring Dr. Sridhar Anandakrishnan.
Start with the first lecture. It is 40:50 minutes long.Now watch the second lecture (36:46 minutes long).
Check out the Unit 3 PowerPoint Presentation used in the online lecture here.
Did you catch all of that? Review the chapter with another Johnny Cash tune not sung by Johnny Cash, "Ring of Fire.", Mt. St. Helens by the subduction zone—it really is a burning thing!
The eruption from Mt. St. Helens in 1980 was not especially big—each of the main Yellowstone eruptions moved about 1000 times more material, for example. But, the Mt. St. Helens eruption killed more people (57; before that only 2 deaths in the US were blamed on volcanic eruptions, with none since), and did more property damage (almost $1 billion), than any other eruption in the United States since the country was formed. Novarupta, in Alaska in 1912, blasted more material than the Mt. St. Helens eruption, but was so far from most people that no one died and damages were small; also, Alaska, at that time, was a territory of the US but not yet a state.
The 1980 Mt. St. Helens eruption probably involved the largest landslide ever observed by humans, in the US or anywhere else, at least since we started writing down what we saw. So, by many measures, in 1980, Mt. St. Helens gave us "The Biggest Eruption in the Whole USA." Here's a Rock-Video parody to show you what happened. If an eruption this big happened from Mt. Rainier, which is close to many more people, the damages would be far, far greater. And, Rainier might just do it some day...
We've seen that most of the Earth's crust is made of minerals with silica (silicon and oxygen) in them. Melting these minerals feeds volcanoes, and freezing the melt clogs volcanoes and makes new minerals in new rocks. When we get to the Redwoods and the Badlands in a couple of weeks, you'll see how the weather attacks the new minerals, and you'll meet a few of the minerals being attacked. Here, just for fun, you can see some truly beautiful minerals, and learn a bit about the wonderful ways they are put together—Lego blocks and erector sets have nothing on nature! Meet the main minerals of the crust, in a parody of Bill Haley and the Comets' Rock around the Clock. This Rock Video has a bit more detail than we'd ever ask in this class, so don't worry about actually learning the difference between nesosilicates and inosilicates unless you're really interested.
You have reached the end of Unit 3! Double-check the list of requirements on the Unit 3 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Following are some supplementary materials for Unit 3. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TAs through Canvas conversations with any questions. Failure to email ALL teachers and TAs may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions.
"No matter how sophisticated you may be, a large granite mountain cannot be denied—it speaks in silence to the very core of your being."
—Ansel Adams, The Spirit of the Mountains
Old, cold sea floor goes down subduction zones beneath warmer sea floor or continent, but what happens when a high-floating continent or island arc tries to go down a trench under another continent or island arc?
The answer is obduction, a BIG collision. The Great Smokies, Mt. Nittany near Penn State's University Park campus, and all of the Appalachians were formed by just such a collision when Africa and Europe hit the Americas, causing a long, thin slab of crust to become a short, thick one by folding and thrust-faulting. Higher mountains have deeper roots (for the same reason that toy boats can float in less water than aircraft carriers, and the iceberg that sank the Titanic stuck farther down in the water than did the ice cubes in the drinks on deck). When erosion lowers a mountain range, the root floats up, bringing metamorphic rocks to the surface that have been "cooked" by heat and pressure deep within the Earth.
Before we go any further, take a look at the following short video introduction by Dr. Anandakrishnan...
Pull-apart, slide-past, push-together obduction and push-together subduction plate boundaries plus hot-spots make earthquakes, volcanoes, and steep slopes that can landslide. If any of these happen underwater, great waves called tsunamis can be generated, with catastrophic consequences. Fortunately, warning systems can be devised to reduce the loss of life, and building with a little foresight can reduce property damage. We'll be looking into these as we wrap up our multi-week exploration of Plate Tectonics and Mountain Building.
You will have one week to complete Unit 4. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 4, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials for Unit 4: | No, but you will be tested on the material found in the textbook. |
Take the Unit 4 "RockOn" quiz | Yes, this is the fourth of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Continue working on Exercise #2: Geology is All Around You | Yes, this is the second of 6 Exercises and is worth 5% of your total grade. |
Complete the "StudentsSpeak #5" survey | Yes, this is the fifth of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 4, including the online textbook, a video lecture, several vTrips and animations, and two overview presentations.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
The Great Smoky Mountain National Park of North Carolina and Tennessee includes 16 mountains over 6,000 feet (about 2,000 m) high, making this generally the highest region in North America east of the Mississippi River. Gatlinburg is a mile (1.6 km) lower than Mt. Le Conte, a relief not much smaller than in many of the great mountain parks of the west, where the peaks are higher but so are the valleys. The Smokies were preserved in a park in 1926, with much of the funding for land purchases provided by J.D. Rockefeller. The Great Smokies, today, are the most-visited national park, because they combine spectacular scenery, rich biological and historical diversity, proximity to major population centers, the lure of a quick stop-off on the drive from the northeast to Florida, and a shortage of other nearby national parks to draw off the crowds. (Although we should not forget Shenandoah, connected to the Smokies by the Blue Ridge Parkway, a beautiful park in its own right.)
In case you’re interested, the top-ten most visited national parks in 2012 included great Smoky Mountains with 9.7 million, Grand Canyon with 4.4 million, Yosemite with 3.9 million, Yellowstone with 3.4 million, Rocky Mountain with 3.2 million, Zion with 3.0 million, Olympic with 2.8 million, Grand Teton with 2.7 million, Acadia with 2.4 million, Cuyahoga Valley with 2.3 million. Cuyahoga Valley may seem the odd-one out; it is a recent addition, seems to have been justified as a national park in part so that Ohio would have a national park, and seems to have a lot of day-picnickers from nearby Cleveland who increase the attendance a whole lot, based on Dr. Alley’s observations during a recent visit. But, Dr. Alley also believes that the Park Service is doing great things with it, and it is well worth the visit! Over 282 million people visited US national parks in 2012 (bear in mind that is with ~2 million fewer visitors than expected as a result of park closure caused by Hurricane Sandy). Note that this is nearly 90% of the whole US population. Although some people will have visited a few parks whilst others visited none, overall people are enjoying their parks!
Much interest in the Smokies centers on the historical aspects. For example, how did the early European settlers survive and flourish in this region? At Cades Cove, wonderful relicts of a bygone lifestyle are maintained in a living museum. Many visitors are also seeking to learn about the earlier Native Americans. Biologically, the Smokies host an amazing array of tree species, flowering bushes (azaleas, rhododendron, and mountain laurel in particular), wildflowers including many orchids, and more. Approximately one-third of the park is covered with "virgin" timber that was not cut by European settlers, and the regions that were logged are growing back rapidly.
Abundant rainfall and snowfall “scraped” from the sky by the high peaks feed numerous cascades and waterfalls, with trout in the pools and kingfishers by their banks. Rainfall is roughly 50 inches (1.3 m) per year in the valleys, and more than 85 inches (2 m) per year on the peaks, so the Smokies share some characteristics with temperate rainforests of the west such as in Olympic. Especially during “off-peak” times, you can get lost in the Smokies, and imagine what the Appalachians must have looked like without humans; approximately 3/4 of the park is wilderness.
The Smokies are a small part of the great Appalachian mountain chain, which extends along the coast of North America from Newfoundland through the Smokies, and then bends westward into Oklahoma. In the Great Smokies, the mountains display a truly remarkable feature—older rocks sit on top of younger rocks! (See the diagram below.) The very high peaks are composed of hard, resistant, old metamorphic rocks (which we will explain soon), of the sort that one finds deep in a mountain range. Beneath them are younger, sedimentary rocks that were deposited in shallow seaways. Between these is a surface called a thrust fault or push-together fault. Thrust faults often show evidence of sliding—scratches and polish indicating motion in one direction, crushing or breaking of rocks, etc. In some cases elsewhere in the world where deformation is still active, thrust faulting has been observed during earthquakes. In the Smokies, the older rocks have been shoved as much as 70 miles (110 km) to reach their present position on top of the younger rocks. The picture below the diagram shows two very much smaller thrust faults, with the upper rocks shoved up to the right only a few inches, but the idea is the same.
You may recall that we started with pull-apart faults at Death Valley. As shown in the diagram above, thrust faults are of the push-together type. Squeeze from either side, and one set of rocks will be pushed over another set. Each set is right-side up, but where they meet, the older rocks are on top of younger. This is seen clearly in the Great Smokies.
Farther north, in the State College, Pennsylvania area, where Drs. Alley and Anandakrishnan teach and where Dr. Alley wrote most of this text, we see a different way that rocks can respond to push-together stresses. There, in addition to some push-together thrust faults, many folds occur. Take a piece of paper, lay it on your desk, and squeeze the opposite sides towards the center. The paper will buckle into a fold. You may achieve the same effect by trying to push a carpet along the floor. Clearly, there are push-together forces involved here.
Just as pull-apart forces occur at spreading ridges, we should expect push-together forces at subduction zones, or at other collision zones. Today, the Appalachians and the east coast of South America look across the quiet sea floor of the Atlantic, across the spreading center of the mid-Atlantic ridge, to the coastlines of Africa and Europe. The coastlines on either side of the Atlantic are parallel to each other and to the mid-Atlantic ridge—slide the new and old worlds back together again, and they fit like a jigsaw puzzle. You can put all of the modern continents back together jigsaw-puzzle style. This fact, and especially the wonderful fit across the Atlantic, has figured prominently in suggesting the idea of drifting continents to scientists and other observers almost since the first decent maps were available of the Atlantic coasts. More importantly, putting the continent shapes back together jigsaw-puzzle style puts the “picture”—the geology—back together as well for events that happened while the continents were joined. For example, the tracks of a glacier run out to sea from Africa, and glacier tracks run from the sea into South America; put the continents back together, and the tracks fit together in showing the path of a single ice flow.
The oldest rocks on the Atlantic sea floor are about 150 million years old, approximately the same age as sediments that were deposited in a Death-Valley-type setting in the Newark Basin of New Jersey and elsewhere along the U.S. east coast. Evidently, the modern situation of a spreading Atlantic began about then, splitting apart a supercontinent to form the Atlantic Ocean in the same way that Baja California is being split off to open the Gulf of California.
But, the Appalachian Mountains are much older than that. A story begins to emerge of a cycle—older push-together forces led to closing of a proto-Atlantic ocean that produced the Appalachians. When the proto-Atlantic was closing, subduction-zone volcanoes formed and spread ash layers across the land, much as Crater Lake/Mt. Mazama and Mt. St. Helens did more recently. (You can find some of those ash layers in many places, including the road cut along the Route 322 expressway just south of East College Avenue in the State College, PA area; see the picture at left.) Sometimes, the proto-Atlantic subduction zones formed offshore and then their volcanoes collided with the North American continent.
When the pushing stopped, the giant pile of the Appalachians, with deep, hot rocks beneath, began to fall apart in Death-Valley style, and red sandstones and mudstones accumulated in Death-Valley-style valleys along parts of the U.S. east coast. The drop in pressure as the Appalachians fell apart probably caused a convection cell in the deep mantle to rise right there, eventually forming the Atlantic Ocean.
Such a cycle, with subduction or continental collisions building mountain ranges that then spread Death-Valley style and eventually split to make ocean basins, has been played out many times over the history of the Earth. And, the fun isn’t over yet; North and South America are cruising westward toward Australia and Asia, as the Atlantic widens and the Pacific narrows. Africa is still bumping into Europe and pushing up the Alps, and India has not yet grown weary of ramming Asia to raise the Himalaya. At fingernail-growth speed, the next 100 million years or so should lead to a lot of geological high drama, but the next 100 years won’t see a whole lot of change. A good animation of this cycle can be found on YouTube.
The key to most of this is that you can sink old, cold sea floor, but you can’t sink a continent. Island arcs and continents float on the mantle too well. So, rather than going down the subduction zone with the oceanic lithosphere, the island arc or continent rides across the subduction zone for a major collision. In such a collision, called obduction, layers of rock are bent into folds such as those of the State College area, or broken into thrust faults such as those under the Blue Ridge and the Great Smokies. In the case of the Appalachians, the thrust faulting was very efficient, with older rocks sliding tens or hundreds of miles (or kilometers) over younger ones.
You have now seen, at least briefly, the three structural styles that are possible: pull-apart (Death Valley, spreading ridges); push-together (Crater Lake and Mt. St. Helens subduction, State College and the Great Smokies obduction); and slide-past (the San Andreas Fault in California). Pull-apart behavior involves stretching of rocks until they break, forming pull-apart or gravity faults (after being pulled apart, gravity pulls one block down past the other). Pull-apart action occurs at the spreading centers, probably where the convection cells deeper in the mantle spread apart. Push-together behavior occurs at subduction and obduction zones, and produces squeeze-together folds and faults, with the faults also known as thrust faults. Slide-past boundaries, also called transform faults, occur where two large blocks of rock move past each other but not toward or away from each other. Slide-past motion produces earthquakes without mountain ranges
Now, you might imagine that we have oversimplified just a little. There is no law that rocks must move directly toward each other (push-together) or exactly parallel to each other (slide-past); sometimes you see an oblique motion with rocks approaching on a diagonal. Or, the rocks may pull apart on a diagonal. And a bend in a slide-past boundary may produce pull-apart or push-together features, depending on which way the bend goes relative to the motion, as shown in the diagram above. A large bend in the San Andreas Fault just north of Los Angeles gives push-together motion and some impressive mountain ranges.
Mountain ranges correspond directly to the main boundary types. Fault-block mountains—the Sierra Nevada, the Wasatch Range, the flanks of the great rift valleys of Africa, and the mid-oceanic spreading ridges—form at pull-apart margins. The mountains are high because the rocks beneath them, in the mantle, have expanded because they are the hot upwelling limbs of convection cells. Volcanic-arc mountain ranges form over subducting slabs, where some of the downgoing material melts and is erupted to form stratovolcanoes; smaller ranges (such as the Coast Ranges of the Pacific Northwest, including Olympic National Park) may form from the sediments scraped off the downgoing slab just above the trench. Continent-continent or continent/island-arc obduction collisions occur at convergent or push-together boundaries as well, producing folded and thrust-faulted mountain ranges.
Remember that crustal rocks are the low-density “scum” that floats on the denser mantle. When obduction occurs, this crustal scum is crunched—it goes from long and thin to short and thick, in the same way that the front end of a car is changed when it runs into a brick wall. Then, just like an iceberg floating in water, a mountain range is a thick block of crust floating in the mantle, with most of the thickness below and only a little bit sticking above.
With an iceberg, about 9/10 of the thickness is below the water and 1/10 above. If you could instantly cut off the 1/10 that is above water, the iceberg would pop up to almost as high as before. A 100-foot-high berg would have 10 feet above the water and 90 feet below. Cut off the top 10 feet, and it is a 90-foot berg with 9 feet, or 1/10, above the water and 81 feet below. So removing 10 feet from the top shortens the ice above water by 1 foot, and the ice below the water by 9 feet. With mountain ranges, the density contrast between crust and mantle is larger than that between ice and water—only about 6/7 of a mountain range projects down to form the root, and 1/7 projects up to form the range.
It remains, however, that if you erode a mountain range, some of the root is freed to float upward. Only by eroding the equivalent of 7 mountain ranges can you eliminate the mountain range entirely. So the Appalachians, despite having been deeply eroded, are still high because they still have a root.
The idea that things on the surface of the Earth float in soft, denser material below is called isostasy, which means “equal standing”—that each column of rocks on Earth has the same weight or standing. Lower-density columns then must be thicker to weigh as much as thinner, higher-density columns. The continents stand above the oceans because the silica-rich continental crust is lower in density than the silica-poor sea-floor crust. The mountain ranges stand above the plains because the thick, low-density roots of the mountains have displaced some of the high-density mantle that is found beneath the plains, or because the rocks beneath the mountains are especially hot and so low in density.
Put a big weight on a piece of crust (say, an ice sheet, or the Mississippi Delta, or a mountain range) and that piece of crust sinks, pushing up material around it in the same way that the surface of a water bed sinks beneath your posterior when you sit down, while the surface is pushed up around you by the water that is shoved sideways. The rising and sinking of the land are slower than for a water bed—thousands of years rather than seconds—because the hot, soft, deep mantle flows a lot slower than water does. But for a mountain range over 100 million years old, a few thousand years doesn’t mean much.
Notice something else fascinating; when a mountain range is being eroded, the top is taken off, and rocks below bob up. Those are taken off, with their place taken by more rocks from below. Pretty soon, the rocks at the surface have come from far down in the Earth, where temperatures and pressures are high. And as you might imagine, high temperatures and pressures change rocks. The rocks around State College, PA have not been “pressure-cooked” much, but the rocks around Philadelphia have been; they tell the story of a great mountain range that fell apart, leaving the remnant that we know as the Appalachians. The rocks in Rocky Mountain National Park are similar to those in Philadelphia, having been deep and now occurring at the surface. Let’s go take a quick look.
Rising high above Estes Park, Colorado, and almost within shouting distance of the population centers of Boulder and Denver, Rocky Mountain National Park is a natural destination for the crowds that throng to this mountain playground. Long’s Peak, at 14,256 feet (about 4300 m), dominates the south-central part of the park; the peak was first climbed in 1868, by a party that included John Wesley Powell, the man who later commanded the first boat passage of the Grand Canyon and then led the United States Geological Survey. Numerous peaks over 13,000 feet (4000 meters) in Rocky Mountain lure climbers.
Small and rapidly shrinking active glaciers still carve the mountains, and much greater glaciers of the past left the numerous tarn lakes, moraines, and other features that decorate the park. Trail Ridge Road surmounts the high tundra of the park, giving the visitor a first-hand look at periglacial processes and ecosystems (those of cold regions; more on this later). The Colorado River rises on the west slopes of the park, and lovely little trout streams such as the St. Vrain flow down the east slope. Bighorn sheep and elk attract traffic jams in Horseshoe Park.
It is a tad embarrassing to say that we don’t fully understand the Rockies yet, including those of Rocky Mountain National Park. Oh, the long history of mountain building, erosion, glaciation, etc., is well-known—we can tell the story. But most mountain ranges hug coasts, or are trapped inland only by the destruction of the ocean to which they once were coastal, whereas the Front Range of the Rockies is as far as almost 1,000 miles (1600 km) from the coast, yet the Rockies are not the direct result of obduction.
The U.S. West is a complicated region (see the Enrichment section for a little more on this). The continent has been approaching and overriding the East Pacific Rise spreading ridge. The San Andreas Fault is the product of the rise running into the trench. Before these met, subduction had been occurring from push-together motion, but with a little slide-past motion thrown in. After the meeting, the subduction stopped, so the push-together stopped, but the slide-past remained to make the San Andreas. To the north of the San Andreas Fault, subduction is still active, forming the Cascades including Mt. Rainier and Mt. St. Helens.
Long ago, the west-coast subduction zone started in the usual way, with old, cold ocean floor going down into the deep mantle. But as the continent approached the spreading center, the down-going ocean floor became progressively warmer and more buoyant. The ocean floor didn’t “want” to go down, but it was still attached to the older, colder floor ahead of it that was going down. So, the ocean floor went under the continent, but stayed high rather than sinking, and rubbed along the bottom of the crustal rocks rather than plunging steeply into the mantle. The friction between this buoyant subducted ocean floor and the crust above, in turn, caused thrust-faulting and crustal thickening far inland (see the figure just below). Because the western part of the country has been built up of many old rock bodies and sediment piles bulldozed from the Pacific, there are scars of many old faults and other geological features that have been reactivated by the recent events, so mountains and valleys have formed along the old weaknesses in response to the new pushes.
As to exactly how this came to produce 14,000-foot (4300-m) peaks in the Rockies, geologists can tell the story, but it isn’t clear that any geologist could have predicted this story without seeing the rocks first. Science moves from explanatory (easier) to predictive (harder), so we still have some work to do. (And, we've oversimplified a bit here; see the Enrichment for more.)
If you drive west from Boulder up the slope to Rocky Mountain National Park, you will go through sedimentary rocks, made from sediments brought down from the current Rockies and from earlier versions of the Rockies. As folding and faulting pushed up the mountains farther west, and as erosion of those peaks allowed “bobbing up” of their thickened deep root, the sedimentary rock layers were tilted (see the figure below), so you actually will be driving into older and older rocks as you go. Eventually, you will reach the heart of an old (Precambrian) mountain range that also forms the heart of the modern one. Even to a casual observer, the rocks here have been “beaten up.” Follow a single layer in a rock, and you will see that the layer twists and bends, doubles back on itself, or even pinches out in places. Analyze the rocks chemically, and you will find a composition similar to the common sedimentary “mud” rock called shale, but the rock clearly is not shale.
Think about cooking. If you mix up a bunch of ingredients to make cake batter, throw the mixture into a pan and put it into a warm oven, the cake you obtain will not be very similar to the mixture you started with. Grill a steak, and the original cow part will come out quite different. Marinade the steak before grilling, and more differences appear. It is common knowledge that a material that is stable in one environment will change if it is placed in a different environment. This is true of everything (and everyone!) on Earth.
The Earth clearly has a great range of conditions. The inside of a mountain range is hotter, higher-pressure, and less affected by acidic groundwaters than is the surface. Materials that are stable at the Earth’s surface (such as the clays in a piece of shale) are not stable deep in a mountain range. The minerals change, grow, and produce new types even without melting. This process is called metamorphism. Metamorphism makes rocks that many people consider to be especially pretty, produces some wonderful gems, and contributes rock names that make good puns. (The Geoclub at Wisconsin used a metamorphic rock, a volcanic rock, and a sedimentary rock in claiming that geologists are “gneiss, tuff, and a little wacke.”) You can read a little more about rocks and minerals in the Enrichment section.
We’ve been looking at the ways the planet moves rocks around and makes mountains, and some of the ways that mountain-building can be dangerous to humans. Volcanoes and earthquakes are sometimes truly dangerous and damaging. But it is worth remembering that, in the developed world, only about 1% of us die in “accidents,” and car crashes greatly dominate those deaths (so 99% of us die of other things, such as heart disease, cancer, etc.). With good scientific warnings, good zoning codes, trained medical personnel, hospitals and ambulances to take care of us, nature just doesn’t kill that many of us. (In the less-developed world, this is, sadly, less true.) For the developed world, things we do to ourselves (smoking, eating and drinking too much, not exercising enough) are far, far more destructive of health and life than anything the planet does to us.
But, it is still wise to know about the dangers from the Earth. And now that we’ve completed the tour of mountain-building, we will look at another hazard. Tsunamis are not directly related either to the Great Smokies or the Rockies, but anything that makes earthquakes, volcanoes, or really steep slopes in or near the sea might be involved in a tsunami. And tsunamis can be truly horrific. We’ll start with a surprising hot-spot tsunami, and then look at some others.
On the flanks of many of the Hawaiian Islands, including Lanai, Molokai, and Maui, to at least 1,600 feet (500 m) in elevation, there are deposits of broken-up, mixed-up, battered corals, other shells, and beach rocks. Corals are undersea creatures, and surely don’t grow 1,600 feet above sea level. Some corals grow just below sea level and later are raised above the water by mountain-building processes; however, these Hawaiian deposits occur on islands that are sinking as they slide off the “hill” made by the rising motion of the hot rock of the Hawaiian hot spot, and the deposits are geologically too young to have been raised so far by mountain-building processes. Clearly, something strange happened.
One of the deposits in particular is the same age as a nearby, giant underwater landslide, as nearly as the age can be measured, which may suggest something. The Hawaiian volcanoes have rather gradual slopes above the water, where the hot, low-silica lavas spread out to make shield volcanoes. But when lava hits the water, the hot flows cool and freeze very quickly, and can make steep piles. Too steep, and eventually the side of the island fails in a great landslide, perhaps when melted rock is moving up in the center of the island and shoving the sides out to make them steeper. Surveys with side-scanning sonar have shown where several such slides have slipped. (Try saying the previous sentence five times fast!) Such slides can be miles thick, tens of miles wide, and over 100 miles long.
Now, if a chunk of rock miles thick and tens of miles wide suddenly starts moving, maybe at hundreds of miles per hour, it will shove a LOT of water out of the way. Where will the water go? The answer is that it will make a huge wave, or tsunami, that will race up any land it encounters after crossing the ocean. Imagine a wave that would run far inland and reach heights of 1,600 feet above sea level. Fortunately, the highest deposits in Hawaii are from a tsunami about 110,000 years ago, long before people were in the way there. Although many such tsunami-generating landslides have occurred, they typically are spaced thousands of years apart or more. But, we can’t absolutely guarantee that there won’t be another one.
The word tsunami comes from two Japanese words, for harbor and wave, a sort of shorthand for a wave that devastates a harbor. Most tsunamis are generated by undersea earthquakes, but undersea landslides or volcanic eruptions, and even meteorite impacts in the water, can generate tsunamis.
Tsunamis move rapidly across the deep ocean, with speeds of 300 to 500 miles per hour (600 to 1000 kilometers per hour). In the deep ocean, the “bump” of water that is the wave of a big tsunami may be only a very few feet high, but may extend well over 100 miles in the direction it is moving. Waves slow down as they enter shallower water, and the leading edge of a wave hits shallow water before the trailing edge. So, the leading edge slows as it nears the coast, the trailing edge catches up, and the wave goes from being long and low to being squashed and high. Even so, the tsunami wave is usually not a towering wall of water, but a strong surge, something like the tide coming in but higher (hence the mistaken name “tidal wave”).
An especially nasty feature of a tsunami is that the water often goes out before it comes in (waves consist of troughs and crests, and some places get the crest first while other places get the trough first). The sudden retreat of water and exposure of the sea floor tempts people to walk out and look around. Then, the ocean returns faster than a person can run. The outcome is predictable and unpleasant.
Terrible tsunamis have occurred. Still horribly fresh in our memories is the Indian Ocean tsunami of 2004, which was triggered by the second-largest earthquake ever recorded, and which killed over 300,000 people. Japan was much better prepared for the 2011 Tohoku earthquake, but almost 16,000 people still died, mostly from the tsunami. The massive 1883 explosion of the volcano Krakatau in Indonesia essentially destroyed the island, with tsunami waves observed as far away as England. Floods raced miles inland on Java and Sumatra, with probably tens of thousands of deaths. Another volcanic eruption, likely in the 1600s BC, of the Greek island volcano Santorini, pushed a tsunami perhaps 300 feet (100 m) or more high across the coast of Crete, and may have contributed to eventual demise of the Minion civilization there. Many commentators have suggested that this is the source of the myth of Atlantis. The great 1964 Alaska earthquake generated a deadly tsunami that killed 118 people, with deaths as far away as California. In 1958, an earthquake-caused landslide in Lituya Bay, Alaska, caused a tsunami that included a wave 50-100 feet high out in the bay, which a father and son safely rode out in a boat. They watched in awe as the wave then ran 1,800 feet up an adjacent coast; five people were killed in the event.
There isn’t a whole lot that can be done to stop tsunamis, but loss of life and property damage can be limited. Tsunami warning systems are functioning in many places, and are being extended rapidly. When instruments (called seismometers) sense the shaking of the Earth from a large undersea earthquake, volcano, or other disturbance, if characteristics suggest that a tsunami is likely, communications are sent out to various agencies concerned with safety, and sirens or other warnings on beaches are activated to get people away from the coast before the tsunami arrives.
The Indian Ocean tsunami seems to have been especially deadly in places where human activities had caused damage to the coral reefs and coastal vegetation that would have blunted the strength of the wave, so maintenance of these natural buffers can help the people living there. And, scientists can figure out where tsunamis are likely, how big and how frequent they are likely to be, and then zoning codes can be enforced so that people build in appropriately safe ways on appropriately safe land if they want to live in an area.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 4.
TECH NOTE - Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Death Valley National Park
(Provided by UCGS)
Death Valley National Park - 3D version
(Provided by UCGS - red/cyan stereo viewing glasses required)
There are two Unit 4 GeoMations (animations) and three GeoClips (movies) linked below. We hope they help you understand and enjoy Unit 4.
In the text, you read about changes in subduction as North America neared and overran the spreading ridge in the Pacific, with the increasingly warm downgoing slab rubbing along the bottom of the continental lithosphere and squeezing and wrinkling the rocks far inland. The Front Range of the Rockies is the most dramatic evidence of the "wrinkling" from that squeezing, but many other ranges and smaller features have the same origin. All sorts of different shapes of folds and wrinkles are observed. One spectacular one, which few relatively few people visit, is the Waterpocket Fold in Capitol Reef National Park. The diagram here, from the Capitol Reef website of the National Park Service, shows the fold (of a type called a monocline, although you don't have to worry about memorizing fold types) both before and after erosion along the lovely Fremont River. You will see more of the fold with Kym Kline and Dave Janesko in the short video. (Note that the Park Service diagram and Dave-and-Kym’s demonstration are of the same thing, but viewed from opposite sides, so that Dave-and-Kym’s fold slopes down to your right and the Park Service’s slopes down to your left.)
Metamorphic rocks—those cooked and squeezed deep inside a mountain range—are often especially pretty. At the bottom of the Grand Canyon, you can see such rocks. They were formed long ago and many miles down, and then reached the surface as erosion removed the mountains above and the deep roots of those mountains floated upward. Later, these rocks were buried again under sediments from oceans, rivers and wind, and finally revealed to us as the Grand Canyon was carved by the Colorado River. Some people—including Dr. Alley—think that these rocks are so beautiful that they're worth the overnight hike into the canyon all by themselves!
The Great Smokies are geologically attached to the whole Appalachian mountain range, including the ridges near Penn State’s University Park campus. There, if you’re so inclined, you can visit the beach in the mountains—all thanks to Africa.
Here are some optional animations you might also want to explore! (No, these won't be on the quiz!)
Tsunami Visualizations
(An extensive collection of animations on this subject)
Mountain Uplift and Erosion
(An extensive collection of animations on this subject)
There are TWO Unit 4 lectures both featuring Dr. Sridhar Anandakrishnan.
Please watch the unit 4 lecture #1, Plate Tectonics III: Making Mountains & Obduction (36:10 minutes).
Now watch part 2, Plate Tectonics III, Mountain Building & Obduction (34:47 minutes)
As noted in the text, the history of the West is quite complex, with some big questions not yet answered, and more work to be done. The most widely accepted history, with the most scientific support, is sketched in the text and given in a bit more detail just below and in the Optional Rock Video Review, The Hanging Wall.
Much evidence indicates that beginning about 100 million years ago, the subduction zone in the west grew shallower. The Pacific sea floor that was sinking under what is now Seattle is called the Farallon Plate, and rather than going down somewhat smoothly into the deeper mantle, the Farallon began to move along just beneath the lithosphere of the North American plate. The text notes that the plate was slowly getting warmer over time and thus more buoyant because the ridge and trench were getting closer together, so the plate was having less and less time to cool off before it went down the subduction zone. Another reason, and perhaps a more important reason, may be that a huge volcanic outpouring on the sea floor more than 100 million years ago made a thick layer of not-very-dense rock, which began going down the subduction zone about 100 million years ago. Where smaller bumps on the sea floor are going down subduction zones today, as off Costa Rica, features similar to those seen in the US West are forming, so the bigger features of the West are easy to explain this way.
Anyway, friction between the top of this buoyant Farallon Plate and the rocks above it—the rocks that we see in the Rocky Mountains and elsewhere in the US West—began to squeeze, bend and break those rocks above. The breaks are thrust faults, with older rocks thrust upward and over younger rocks to make the mountains. You will recall that the rocks above the fault are the “hanging wall”, which gives rise to the pun in the Rock Video.
Where a fault didn’t break all the way to the surface, it often raised the rocks above (think of lying on your back in bed under covers, and then raising your knees—the cover draping over your knees are the unbroken rocks). If the rocks were raised in a more-or-less circular pattern when viewed from above, we call the feature a dome; if shaped more like a US football when viewed from above, it may be called an arch, and a few other names are sometimes used. Such uplifts gave us the Black Hills of South Dakota, the Waterpocket Fold of Capitol Reef National Park, the Kaibab Uplift that the Grand Canyon cuts through, the beautiful San Rafael Swell of central Utah, and many other features of the West. These events mostly happened soon after 100 million years ago, during what we now call the Sevier Orogeny, and somewhat more recently in the Laramide Orogeny—we see more faults from the Sevier, and more uplifts from the Laramide.
The squeezing and thrusting of mountains caused some downwarping as well as upwarping. One of the downwarps held the lake in which the pink limestones were deposited that we now see at Bryce and Cedar Breaks, and similar lakes gave us the Green River limestones that produce such fantastic fish fossils from farther north in Utah. Oil is found in the west in rocks that accumulated in such downwarps, too.
The Farallon Plate probably eventually broke, and a new, “normal” subduction zone started up, feeding Mt. St. Helens and Mt. Rainier in the west. The Farallon Slab is now sinking beneath the eastern part of the US. As it sinks, it creates space into which hot rock flows slowly, and this may be helping “rejuvenate” the Appalachians, so the Great Smokies are a bit higher than they would be otherwise because of processes traceable back to the Rockies and the Farallon. Why the Farallon “decided” to sink eventually may be because, as it ran into thick rocks beneath the west, either some of the low-density parts were peeled off so the higher-density ones could sink, or the low-density ones were shoved deep enough that the pressure changed the mineral structures, making denser things that can sink.
Anyway, there is still much to discover about the western US. But, to catch a light-hearted version of what we know, check out the tale of a geology student trying not to be hung on the Hanging Wall.
What happens when you "cook" a rock metamorphically depends on how hot, and what chemically active fluids are present, and the pressure and deviatoric stress. Both pressure and stress have units of force per area, and represent a "push" on a material. Pressure is the part of the stress that is the same in all directions--it squeezes rocks or fluids to make them smaller, but doesn't tend to change their shape. Deviatoric stress, sometimes just referred to as stress, is an extra push or pull in one or more directions, and does change the shape. Deviatoric stress has been involved in aligning the mineral grains of most metamorphic rocks into layers, or folia. Chemically active fluids—water, carbon dioxide, methane, etc.—can add or subtract chemicals, lower melting points, and dissolve and reprecipitate chemicals.
Materials that are stressed deviatorically can have one of three responses. The materials may bend and, when you release the stress, snap back (elastic deformation), they may bend permanently (plastic deformation or creep), or they may break. We have already seen examples of all of these. Earthquakes are caused by the snap-back of rocks near faults following bending. Faults such as the San Andreas are the result of breakage, and folds such as those around State College, or those in the rocks exposed in the heart of Rocky Mountain, are the result of plastic deformation.
Whether a rock bends elastically or plastically, or breaks, depends on the rock itself and on several other factors: heat (or more properly, how close a material is to melting), pressure, deviatoric stress, and even the chemically active fluids, acting over time. Elastic deformation is favored by low stresses and high pressures (to prevent breakage) and low temperatures (to prevent creep). Plastic deformation is favored by low stresses and high pressures (to prevent breakage) and high temperatures (to allow creep). Fracture is favored by high stresses and low pressures (to allow breakage), together with low temperatures (to prevent creep). Pressure matters in breakage because, to break a material, the two sides must be pulled apart, which increases the size of the material. When pressure is high, this is difficult to achieve. Fluids generally soften rocks and promote creep, although the details depend on the fluid and the rock involved. You will often see that the rocks in a region will have undergone both permanent folding and breakage, and may be bent and ready to snap back, with different modes of deformation more important in different places. Figuring this all out is fascinating, as well as useful (miners, well-drillers, and many others want to know what they'll hit in the rocks, whether the rocks are about to break, what kind of cracks the oil or gas or ores may be hiding in, and much more).
If you take water and freeze it, you obtain ice that is made of the same stuff as the water. Melt the ice, and you have the water back. Melting and freezing without separating chemicals is easy to understand, but is more the exception than the rule. If you take beer and start to freeze it slowly, the crystals that form will be almost pure water ice, although each crystal typically will start to grow on a small impurity particle in the beer. Filter the crystals out, and you will have cleaned the remaining beer marginally, and you will have increased its alcohol content. Hire some advertising agents, and you have a “new” product to pitch: ice beer. (If you're under age, please substitute iced root beer. You'll end up with a lot of ice and a little sugar water.)
In the world of rocks, things are even more complex than with (root) beer. Suppose you take a piece of granite (containing quartz, mica, potassium feldspar, and a sodium-rich sodium-calcium feldspar) and melt it by heating it hot enough to melt basalt. Suppose you then start to cool it. The first minerals to crystallize may be a little bit of olivine, and a calcium-rich sodium-calcium feldspar. Cool the melt a little more, and the olivine and remaining melt react to make pyroxene, while the feldspar and melt react to make more more-sodium-rich feldspar. Keep cooling, and eventually you will get the sodium-rich feldspar and the mica, followed by the potassium feldspar and the quartz.
This ideal sequence may not be observed in many situations, but portions of it are well-known in laboratory experiments and in nature. In general, the first things to crystallize are poorer in silica, sodium, potassium and aluminum, and richer in iron, magnesium and calcium than the melt from which they grew. (Note that there are MgSiO3 pyroxenes and Mg2SiO4 olivines, or mixtures in which the Mg and Fe substitute for each other because they are almost the same size and have the same electric charge.) As the temperature drops, the early minerals react with the melt to make new minerals that are more like the original melt in composition. However, if the early minerals are removed from the melt, perhaps by settling to the bottom, they may be preserved, and exceptionally silica-rich rocks will be formed from the minerals that grow from the remaining melt. For more on the minerals, take a look at our "Sidebar" next.
If you throw a bunch of typical Earth chemicals into a pot, melt them, and cool them slowly, you will find that only certain things grow. You might, for example, find the mineral quartz (SiO2), or the mineral pyroxene (FeSiO3) or the mineral olivine (Fe2SiO4). You will not find something midway between olivine and pyroxene; it doesn’t exist. Nature puts the chemicals together in certain ways, and only certain ways. It is a little bit like building with Tinkertoys—there are only certain holes you can put the sticks into, which fix the angles at which you can build things.
Minerals are orderly—the same basic structure is repeated over and over and over (say, a silicon surrounded by four oxygens, each oxygen in contact with an iron that then contacts another oxygen that is one of four around another silicon, which is the structure of olivine). When minerals are allowed to grow freely, they assume certain shapes that look as if a gemstone-cutter had shaped them. The faces on such crystals are controlled by the underlying order of the chemicals. The classification of minerals is based on the chemical composition, and on the structure in those cases when a single composition can assume one of two or a few different structures.
Rocks are collections of minerals. One can have an all-olivine rock, or an all-pyroxene rock, or a mostly-olivine/some-pyroxene rock, or any other possible combination. We humans have chosen to classify rocks based first on their origin, and then on other characteristics such as their grain size, or their composition, or more details of their origin. The main subdivisions are igneous (rocks that formed from cooled magma or lava), sedimentary (those formed from pieces of pre-existing rocks, or from such pieces that dissolved in water and then crystallized from it), and metamorphic (those formed from igneous, sedimentary, or older metamorphic rocks by the action of heat, pressure, stress and chemically active fluids).
The classification of igneous rocks is next (we'll do igneous and then metamorphic, and save the classification of sedimentary rocks for later in the semester). We distinguish coarse-grained rocks that cooled slowly from magma deep in the Earth, and fine-grained rocks that cooled rapidly from lava at the surface; the extreme case is obsidian, a glass that cooled too rapidly to allow the high-silica types. Low-silica rocks contain the minerals olivine, pyroxene, and calcium-aluminum-rich feldspars. High-silica types include quartz, potassium- and sodium- rich/aluminum-poor feldspars, and mica. Putting these together (grain size and composition) allows us to draw the following grid:
-- | Low Silica | Medium Silica | High Silica |
---|---|---|---|
Small Grains | Basalt | Andesite | Rhyolite |
Large Grains | Gabbro | Diorite | Granite |
Of these, basalt dominates the sea floors, andesite dominates island arcs, and granite to diorite are common in the hearts of mountain ranges. Many other types occur, but these are the most important ones.
Good question. We’re not sure. But we do know that mid-ocean ridges are high because they are hot. Plates will slide off high spots, heading toward low places. So, North and South America are being pushed westward because they are sliding off the mid-Atlantic spreading ridge. If the spreading center in the Pacific simply sat still for a long while (which it probably did, more-or-less), then eventually the Americas would get to it, which they seem to be doing now.
The Eagles got a peaceful, easy feeling while going into the desert, but they didn’t tell us whether they planned to do a lot of geology on the way. Whatever; in this parody, you can get a peaceful, easy feeling about a whole mountain range, the beautiful Appalachians. The Proto-Atlantic Ocean really did close before re-opening as the modern Atlantic, acting like a very slow accordion. While the proto-Atlantic was closing, three "collisions" happened, raising mountains near the modern east coast of the US, and erosion of the mountains produced sand that now is the sandstone of the ridges of central Pennsylvania and other places along the Appalachians. The first two collisions—small obduction events—involved North America hitting island arcs containing explosive volcanoes formed at former subduction zones. The third collision was the major obduction event as the Proto-Atlantic disappeared, with Africa and Europe meeting the Americas as the great supercontinent of Pangaea was assembled.
Geologists long ago decided that the place where subduction occurs is a "subduction zone", but have been slow to adopt "obduction zone" for the place where obduction occurs. We just couldn't make the rhyme scheme work with "the mountain range formed where an obduction event happened", so we call it an obduction zone. No one is quite sure whether the mountains in our "obduction zone" were really as high as the Himalaya, or how many glaciers eroded the top of our obduction zone, but the mountains probably were quite high and at least some of them may have been glaciated. The diagrams in this rock video have been simplified a bit, to make it easier for you. You can see slightly more-complicated versions of the second and third collisions, and see why we simplified them.
So, relax, get out your Irish flute, and let’s go obducting.
You have reached the end of Unit 4! Double-check the list of requirements on the Unit 4 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Continue to work on Exercise #2. See Course Calendar for specific dates.
Following are some supplementary materials for Unit 4. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
Fans of old-fogey rock music may recall that Paul Simon was "slip-sliding away," while Harry Chapin (when not slip-sliding on 30,000 pounds of truck wreck bananas) was convinced that "All my life's a circle". Mother Nature plays the best rock, and the Grand Tetons slip-sliding away are part of a circle that remakes the mountains from shells and soil. So crank up the tunes and the worm poop, and let's get rolling.
You will have one week to complete Unit 5. See the Class Calendar for specific due dates.
As you work your way through the online materials for Unit 5, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Review the Unit 5 Overview | No, the overview outlines the main topics and ideas that you will encounter in Unit 5. You will, however, be tested on the material found in the overview. |
Read/view all of the Instructional Materials for Unit 5 including: | No, but you will be tested on the material found in the Unit 5 Instructional Materials. |
Take the Unit 5 "RockOn" quiz | Yes, this is the fifth of 12 end of unit RockOn quizzes and is worth 4.5% of your total grade. |
Submit Exercise #2: Geology is All Around You | Yes, this is the second of 6 Exercises and is worth 5% of your total grade. |
Begin Exercise #3: The Age of Nittany Valley | Yes, this is the third of 6 Exercises and is worth 5% of your total grade. |
Complete the "StudentsSpeak #5" survey | Yes, this is the fifth of 12 weekly surveys and is worth 1% of your total grade. |
Read the Optional Enrichment Article | No |
Review the Unit 5 Wrap Up including the Supplemental Materials | No, but you may find them helpful in preparing for the quiz. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants: through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 5 - including the online textbook, a video lecture, a supplemental enrichment article, and two supplemental Virtual Trips (vTrips).
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
There are certain other forms of waste which could be entirely stopped—the waste of soil by washing, for instance, which is among the most dangerous of all wastes now in progress in the United States, is easily preventable, so that this present enormous loss of fertility is entirely unnecessary. The preservation or replacement of the forests is one of the most important means of preventing this loss."
—President Theodore Roosevelt, Seventh Annual Message to Congress, Dec. 3, 1907
The foggy drizzle of Redwood National Park may cause you to want a jacket for warmth as well as dryness, yet, above the clouds, Redwood gets about the same amount of sunshine as toasty Death Valley. Much of Death Valley's warmth can be traced to Redwood rain releasing heat that was stored during evaporation of water from the hot tropical ocean. Rain and other weather phenomena break down rocks, making soil, while washing some materials to the ocean to be used for shells or other things. Landslides, rivers, and other processes eventually remove soil from the land about as rapidly as the soil is produced, and deliver the soil to the ocean, where subduction and volcanoes recombine the soil, shells and other washed-away chemicals to make new rocks for weather to attack.
Redwood National Park has the feel of a soaring, gothic cathedral—only more so. One of the great Sequoia sempervirens trees may live for two millennia, but when it falls, new trees will grow from the fallen trunk. (Hence the scientific name, which means “ever-living sequoia” in Latin.) The redwoods are the tallest trees on Earth, commonly more than 200 feet (60 meters) and with the very tallest soaring above 379 feet (more than 115 meters). If such a tree growing on the goal line fell over, it would stretch the length of a football field and the top branches would extend into the stands at the end. Ferns growing in the duff beneath the redwoods are lost, inconsequential, despite occasionally standing shoulder-high or above. Reports from early loggers included trees even taller than any known today.
The redwoods lie in the southern part of the great, coastal, temperate rainforest that extends from San Francisco north along the Pacific coast to southeastern Alaska, and includes the Olympic National Park discussed earlier. The redwoods actually grow in soils that came from rocks much like those of the Olympic.
Redwoods are mostly restricted to a narrow band along the coast with 50-100 inches (1.25-2.5 meters) of rain per year, and with frequent to continuous fogs that slow the drying-out of the trees. The trees, in turn, help maintain the fog; cutting the trees may decrease the fog, making it difficult or impossible for the redwoods to re-grow. The wood of redwood trees is highly resistant to fire and rot, and so is greatly sought after. Logging of old-growth redwoods, thus, is a contentious issue; an estimated 96% of the old-growth forest has been cut already, and with typical ages of 500-700 years and maximum ages of about 2000 years, a new "old-growth" forest is not returning soon, but some people want to cut the remaining 4% or so of the original old-growth redwood forest. Fossil evidence shows that redwoods once were much more widespread. U.S. parks that preserve fossilized redwood include the Petrified Forest, Yellowstone, and the Florissant Fossil Beds.
A bit farther south and higher on the slopes of the Sierra, but still in a wet zone, are the great sequoias of Yosemite, Kings Canyon and Sequoia National Parks. These Sequoiadendron gigantea are close relatives of the redwoods. Although not as tall (“only” up to 311 feet, or about 95 meters), the great sequoias are more massive. The General Sherman tree, at 275 feet tall and 102.6 feet around, is generally considered to be the largest single-trunk tree on Earth (and much larger than whales and such). The great sequoias are extremely fire-resistant, and require fire to clear-out competing trees and trigger sprouting of sequoia seeds. Fire suppression instituted after the parks were established led to a period with few or no new sequoias sprouting; now, prescribed burns and procedures that allow some natural fires to burn are returning the forest to a more-natural state. Sequoias can live 3500 years.
A few hundred miles to the east of the redwoods and sequoias lies the U.S. Great Basin. All across this region of Nevada and adjacent states, rainfall typically is less than 10 inches (25 cm) per year, with Death Valley receiving just 1.6 inches (4 cm) per year on average. Sage and cactus grow here instead of towering trees - in response to the very hot, dry conditions. Yet, above the clouds, this region receives the same amount of sunshine as do the redwoods. Clearly, the climate is much more than just the amount of sunshine above the clouds. We now take a brief tour through the basics of climatology. This should help you to understand climates, weathering, erosion, and glaciers, and a number of important near-surface processes.
The wind blows, and the deep mantle convects, for more-or-less the same reason. Both the air and the mantle are capable of flowing, and both are heated from below and cooled from above. The amount of heating and the rate of flow are VERY different, which helps make the world interesting. But you might see a thunderstorm and imagine hot-spot formation, and there is at least a little similarity between a cold front and a subduction zone.
Anyway, on Earth, the equator receives more sunshine than the poles. Some of the energy reaching the Earth is reflected back to space from clouds or snow and ice, especially near the poles. Much more sunshine is absorbed to heat the Earth at the equator than at the poles.
The equator receives more sunshine because of simple geometry. Imagine for a moment that Dr. Alley’s head is the Earth, with his nose on the equator and the bald spot on top the North Pole. (See the picture; you can congratulate yourself on good taste if you didn’t imagine a bad joke about the South Pole here.) If he stands in front of a sun-lamp “sun,” he’ll never get a sunburn on his North-Pole bald spot, but he will on his equatorial nose.
The Earth works the same way—at the top of the atmosphere, the amount of sunlight passing through a square meter is the same at the equator as at the pole (see the diagram below). But because of the Earth’s curvature, the light passing through a top-of-the-atmosphere square meter at the equator illuminates a square meter at the surface, whereas the light passing through a top-of-the-atmosphere square meter near the poles is spread over many square meters on the surface. (Additionally, in both cases, the rotation of the Earth spreads the light over a larger area than at the top of the atmosphere.)
This means that land at the equator becomes hotter than the poles. If we had no atmosphere or oceans, the equator would become too hot for life as we know it, and the poles too cold. However, the atmosphere and oceans take some of the excess heat from the equator to the poles, making both habitable to humans.
As the sun heats the land at the equator, the land heats the air above, and the air expands, rises, and then moves poleward in convection currents. Some of the energy in this hot air goes to warm the regions along the way, but eventually all the energy is radiated back to space. The sunlight that comes in is called shortwave radiation (because its waves are short—this stuff really isn’t that complex a lot of the time!), and the radiation going out has longer waves and is called longwave radiation. We will see later that the difference between shortwave and longwave is important in understanding the greenhouse effect. For now, note that the global energy budget is balanced—the total amount of energy brought in by short waves and absorbed in the Earth system is very nearly equal to the total amount of energy taken out by long waves to space. (At the moment, we are sending out a little bit less than we receive, and the Earth is warming, because we humans are changing the composition of the atmosphere. But, once we quit doing that, the Earth will get back to balance.) A factory balances the total amount of stuff coming in and going out, but little auto parts come in and big cars go out; the earth balances the total energy coming in and going out, but short waves come in and long waves go out. And, the uneven heating drives the wind, and the wind plus uneven heating of the oceans drive ocean currents.
Because the Earth rotates, the winds end up turning rather than going straight from equator to pole, and this makes the weather much more interesting than it would be on a non-rotating planet. If you want to explore this a little more, see the optional Enrichment.
Some of our friends in meteorology just hate it when we say that warm air can hold more water than cool air. They like to say that the saturation water-vapor pressure in the atmosphere increases with temperature. But this means that warm air can hold more moisture than cool air.
Air is cooled in two major ways—by losing energy to its surroundings (longwave radiation to space, or warming the polar regions), or by being lifted. As air is lifted, it expands and cools. You can experience this by letting air out of a high-pressure bicycle tire, and feeling how cold the air becomes as it expands. Lifting may be caused when one air mass moves over another along a front, or when air moves over mountains. In either case, higher elevations have lower temperatures.
Cooling causes condensation of the water vapor in the air because cooler air can hold less water, and the condensation makes clouds and then rain or snow. Our friends in meteorology would point out that condensation requires cloud condensation nuclei—little dust particles and such that water drops can grow on—plus a little supercooling to get things started, or else extreme cooling if the nuclei are absent, but if you make a parcel of air sufficiently cold, clouds form and then rain or snow. (And if it bothers you that cooler temperatures exist higher, but that cold air sinks, see the optional Enrichment—there is really no problem, and it all does make sense.)
Evaporation of water requires energy. Remember that everything we see—including water—is made of fast-moving particles (atoms, or groups of atoms called molecules). The faster-moving particles are the hotter, higher-energy ones. In a pool of water (the ocean, or a drop of sweat on your brow) the faster, hotter water molecules break the attraction to their neighbors and escape, or evaporate, leaving the slower, cooler ones behind. Evaporation thus cools the remaining water. More heat then is conducted or radiated into the water from its surroundings (because heat flows from warmer to cooler places, and the evaporation has left the water drop cooler), cooling the surroundings while fueling more evaporation. This is why sweating cools a body. Condensation is the opposite of evaporation, and when condensation occurs, the heat that was added to the water to cause evaporation is released back to the air.
Now, consider a parcel of air rising up the Coast Ranges above the redwoods, and up the Sierra Nevada above the giant sequoias. As the air rises, the pressure on it (the weight of the air above) becomes smaller, and the air expands and cools. Once the air has cooled enough, it becomes saturated with water, and further cooling causes condensation. But the condensation releases some heat that partially counteracts the cooling from expansion. Air in which condensation is not occurring cools by about 1ºC for each 100 meters it is lifted, but air from which clouds and rain are forming only cools about 0.6º C for each 100 meters it is lifted (5ºF per 1000 feet dry, and 3ºF per 1000 feet wet). The difference represents the heat released by the condensation. The heat released by condensation originally came from the sun, and was stored in the air in the water vapor when water evaporated, cooling the ocean. As the vapor condenses, the sun’s energy is turned into heat that one can feel.
Strong moisture-laden breezes from the Pacific rise as they hit the redwood-clad Coast Ranges of California, cooling and raining to nourish the rainforest. The winds continue upward over the Sierra, cooling by 3ºF per 1000 feet upward, as they precipitate. By the time the wind has risen 15,000 feet to get over the high peaks, it has cooled about 45ºF (25ºC), and has rained or snowed almost all of its moisture. When this wind continues on down the other side, it is dry, and warms at the dry rate of 5ºF per 1000 feet. When this wind returns to sea level, the air has warmed about 75ºF (42ºC) on the way down. Hence, the wind comes down about 30ºF (17ºC) warmer than when it went up, and the difference is the heat that was stored when the rainwater evaporated from the ocean. A comfortable onshore breeze at 70ºF on the Pacific Coast will be 100ºF when it reaches the valleys on the other side. Add a little solar heating through the cloudless desert air, and it is no wonder that Death Valley is hot! (And yes, the wind usually goes around the Sierra rather than over, so you haven’t learned everything about meteorology in one short chapter of a geology book; but it’s a start.)
The Badlands of South Dakota are much more than just the land to the south of Wall Drug Store (“Have you dug Wall Drug?”). Today, the Badlands may be most valuable as a wonderful piece of the shortgrass prairie that once covered much of the western Great Plains. The sea of grass and flowers that nourished the bison and the native Americans of the plains has been almost entirely broken to the plow. But in the upper prairie of the Badlands, the grass still waves in the breeze like an ocean, the pronghorn still course the grass, and one can, perhaps, imagine what the prairie once was.
The Badlands are a creature of the Rockies. During one of the phases of uplift of the Rockies (especially about 30 million years ago), the exposed rocks were first weathered (changed into other forms, such as clay), and then transported from the mountains onto the plains to the east, or on down the Mississippi River system to its delta. Those sediments deposited on the Great Plains piled up until the mountains were almost completely buried, and an unbroken ramp of sediments extended to the Mississippi. The sediments were mostly river deposits, on flood plains or in channels of rivers, although some ash-falls from volcanoes and other wind-blown sediment were present.
Eventually, renewed uplift of the Rockies rejuvenated the streams draining through the Badlands, causing these streams to cut down through the older deposits. In these deposits are uncounted fossils of the animals and plants that lived on those floodplains. The fossils tell a story of gradual drying from wetter times of the past, and of changes in the types of things living there.
The types of clays and volcanic ashes in the Badlands, combined with the climate in which they now occur, do not make fertile, stable soils. The clays expand and shrink on wetting and drying, breaking roots, triggering small landslides, and helping keep the slopes unvegetated. Water doesn’t flow easily through clay, so rainfall doesn’t soak in but flows rapidly across the surface during the rare but intense rainstorms, eroding channels in the clays and washing away vegetation; this also leaves the soils dry between rainfalls so that plants have difficulty sustaining themselves.
That a lush prairie can grow on the flatter regions is testament indeed to the hardiness of the prairie ecosystem. But, on steep slopes, the pastel-colored muds eroded from the old Rockies are exposed, along with the fossils they contain. We will explore stories of this sort over the next few weeks. We start with weathering, the changes that can turn hard rocks to soft clays of the type observed in the Badlands.
Squeezing mud, perhaps with a little heating, produces a sedimentary rock called mudstone or shale, and with further heating and squeezing this rock changes to various other things on the way to melting (you may have met shale→slate→schist→gneiss somewhere; we saw gneiss in Rocky Mountain, and all of them in the Enrichment, last time). The materials in mud are stable (or at least nearly so) under Earth-surface conditions but not deep in the Earth. And minerals produced deep in the Earth usually are not stable under surface conditions. Compared to deep in the Earth, the surface is wetter, has more oxygen, has a wider range of acid/alkaline conditions (with acid especially common at the surface), and has many more organisms trying to break down the minerals to extract fertilizer.
As a general rule, the further surface conditions are from conditions under which a mineral formed, the more rapidly that mineral changes at the surface. (This “rule” has many exceptions, but it is often useful, especially with silicate minerals.) The changes that occur at a place are called weathering. Moving the products of weathering is called transport. Weathering and transport together make erosion.
Weathering in turn is divided into mechanical and chemical weathering. Mechanical weathering is the making of little pieces from big pieces; chemical weathering is the making of new types of materials that were not there previously.
To turn big pieces into little ones, the big ones need to be cracked. Cracks in rocks are caused or enlarged by processes including:
Probably the most important mineral that grows in cracks is ice, but others do too. For example, the mineral thenardite, Na2SO4 (no, you don’t have to memorize the formula!) can add much water to its structure (10 molecules for each Na2SO4), expanding in the process. If thenardite exists in an environment that alternately is wetted and dried, the resulting expansion and contraction from adding and losing water may break the rocks around it. Increased humidity will cause the mineral to rearrange and take up water, expanding and wedging open cracks. Continued wetting may dissolve the mineral, which will move deeper in the crack in the rock and then be redeposited and dehydrated during a subsequent dry time. Then further wetting will cause re-expansion and more wedging. This process is breaking many of the ancient monuments of Egypt as increased irrigation and other activities increase humidity in some places.
Chemical changes are often more interesting and more complex than physical ones. There is a great range of possible changes, and you must know a lot of chemistry to really appreciate all of them. In general, the main agents of weathering are weak acids. Rainwater picks up carbon dioxide from the air and becomes a weak acid called carbonic acid. In soils, water may pick up more carbon dioxide plus organic acids from decaying organic material, becoming a stronger acid.
When acid attacks a rock, the results depend on what minerals are present, and how warm and wet and acidic conditions are, among other things. We can sketch some general patterns. Suppose we start with granite, a silica-rich rock that forms in a few continental and island-arc settings, including when bodies of melted rock solidify beneath the andesitic volcanoes of subduction zones.
Granite usually is composed of four minerals: quartz (pure silica, which is silicon and oxygen), potassium feldspar and sodium-calcium feldspar (mostly silica, with a little aluminum replacing some of the silicon, and the potassium, sodium or calcium added for balance), and a dark silica-bearing mineral containing iron and magnesium (often a dark mica called biotite). Interestingly, almost 99% of the atoms in the rocks of the crust of the Earth are oxygen, silicon, aluminun, iron, calcium, sodium, potassium or magnesium, the common elements in granite. We discuss granite here because it does so well representing the whole crust.
When granite interacts with carbonic acid, several things happen.
One can write a sort of equation:
Granite → quartz sand + clay + rust + (dissolved-and-washed-away Ca + Na + Mg).
The calcium and silica dissolved and washed to the ocean are used by sea creatures to make shells, the dissolved magnesium washed to the ocean usually ends up reacting with hot rocks at spreading centers to make new minerals in the sea floor, and the dissolved sodium accumulates in the ocean to make it salty. (Eventually, evaporation of ocean water in restricted basins causes deposition of salts including table salt, which is sodium chloride, written NaCl. Also, some salt is taken down subduction zones with the water in the spaces in sediment. So, the ocean does lose sodium as well as gaining sodium.) The rust, sand, and clay left behind, plus a little organic material often including worm poop, become the indispensable layer we know as soil.
You should recognize that this is a very general description of what happens; were it this easy, there would not be hundreds of soil scientists working to understand this important layer in which most of our food grows. In general, the hotter and wetter the climate, the more stuff is removed—rust and quartz sand can be dissolved in some tropical soils, leaving aluminum compounds that we mine for use in making aluminum. In dryland soils, the Ca and Mg may be left behind (forming special desert soils), or even the Na may be left behind, forming salty soils in which little or nothing will grow.
You also should recognize that the “chunks” in soil – rust, clay, sand and organic materials – can be carried away by streams or wind, but as chunks rather than invisible dissolved materials. We discuss this loss of chunks in the next sections. If chunks are carried away more rapidly than new ones are formed, the soil will thin and we will find it difficult to grow food to feed ourselves.
When we were looking at subduction zones, we noted that old, cold sea floor goes down. We now see that it takes some extra magnesium, added from sea water by reactions with the hot rocks at the mid-ocean ridges. Sediment goes down subduction zones, including the calcium- and silica-bearing shells of sea creatures, the rust and clay washed from the continents, and some sea water. When these melt and feed volcanoes, andesite and similar rocks are made during eruptions, or their underground equivalents including granites are produced. The calcium-bearing shells include carbon dioxide, and this plus water escape from the volcanoes to make carbonic acid to attack the new granite. The Earth really does cycle, and recycle, everything! But, going around this loop once takes at least millions of years, and may take a lot longer than that, issues we'll discuss later.
The Grand Tetons tower above Jackson Hole, Wyoming, the epitome of western scenery for many people. A steep pull-apart fault lies along the front of the range and slopes downward beneath Jackson Hole. From the highest peaks to the Hole is well over a mile (roughly 2 km), but the total offset on the fault (including material eroded from the top of the range and deposited in the valley) is almost 6 miles (10 km). The uplifted block is primarily old metamorphic rocks that erode only slowly. The faulting is probably related to the Basin and Range extension that also gave us Death Valley, although the complexity of the region makes any interpretation difficult. Dr. Alley recalls huddling next to an overhanging rock, far up on the steep front of the Tetons, watching hailstones rattle off the trail from a black deck of clouds barely over his head. It is a truly awesome place.
A few miles (few km) east of the park is another interesting feature: the Gros Ventre slide. There, the rock layers slope steeply under a long ridge or mountain, with the layers almost parallel to the mountainside, down to the Gros Ventre River. Strong, resistant sandstone rests on weak, slippery shale. The river has eroded down through the sandstone and into the shale, leaving the toe of the sandstone unsupported. In June 1925, after a particularly wet spring, the entire mountainside let loose, sliding down and across the river. The slide mass made a dam, trapping a lake miles long and 200 feet (60 m) deep. The entire slide probably required only seconds to occur, and moved cubic miles (many cubic kilometers) of rock.
Such a debris dam is not very strong; water flow through its pore spaces or over it can remove rocks and weaken it greatly, until it collapses catastrophically. When we were discussing earthquakes, we saw that a similar slide happened during an earthquake in 1959 just northwest of Yellowstone, damming a river to form a new lake, and that the Army Corps of Engineers had rushed in to prevent collapse of the dam. The Corps had learned from previous experience, including events at Gros Ventre. In 1925, efforts were not made to stabilize the dam there. It collapsed two years later, washing out a small town downriver and killing six people.
The Gros Ventre slide is an especially dramatic example of an important but usually-more-boring process: mass movement. This is the name give to the downhill motion of rock, soil, debris, or other material, provided the flow is not primarily water (if material is washed along by a river, we call it a river), or a glacier or wind.
Water is usually involved in mass movement, however, because most mass movements occur when soil or rock is especially wet. This happens for four reasons: water makes the soil heavy; water lubricates motion of rocks past each other; water partially floats rocks (a rock pushes down less in water than in air) so they are not as tightly interlocked and can move more easily past each other, and saturation of a rock mass removes the effect of water tension.
This last deserves a bit more explanation. Think about going to the beach and building sand castles. Dry sand makes a little pile with sides rising at maybe 30 degrees. Totally saturated (wet) sand flows easily, forming a pile with a much more gradual side. But damp sand can hold up a vertical face. You can even make and throw damp sand balls (be careful where you throw them).
The details of the surface physics involved are a bit complicated, but basically, a drop of water will sit at the junction of two sand grains. Pull the sand grains apart, and some water will end up on each grain. This requires “breaking” the water; what had been one continuous film of water is now broken into two. Next, think of a dripping faucet. A water drop doesn’t fall off immediately, but first becomes large and heavy. Clearly, water has strength, and does not “like” to be broken. (This strength arises from the attraction of water molecules for each other, and is usually called surface tension.) Damp sand thus is strong—a landslide would require some sand to move rapidly past other sand, breaking the water bonds between the grains. In fully wet sand, however, the grains move about in the water without ever breaking it, so motion is easy.
There are elaborate classifications of mass movements, depending on how fast, how wet, how coarse, how steep, and how "other" they are. Most of the names make sense: falls are rocks that fell off cliffs, topples are rocks that toppled over from cliffs, landslides and debris flows and debris avalanches are fast-moving events, and slumps are failures of blocks of soil along concave-up curved surfaces, something like a person slumping down in a chair.
One fascinating, and scary, type of mass movement occurs on “quick” clays. These are clays usually deposited beneath the sea, in which the playing-card-shaped clay particles exist in a house-of-cards structure stuck together by ions of the sea's salt, which sit in the spaces between clay particles. When raised above sea level, ordinary water begins to replace the ions of the salt, and the structure becomes weaker. Eventually, it may fail catastrophically, going from a solid to a very thin liquid almost instantaneously, and generating a flow. Flows from such clays are known especially from parts of Canada and Scandinavia. Such a flow crossing a road can pose a problem, because bulldozing the clay out of the way does little good; more just flows across. But throwing a bag of salt into the flow near the road and driving a tracked vehicle through to mix the salt and clay may cause the flow to solidify so that it can be bulldozed away. Tragically, large portions of some towns have been destroyed, with much loss of life and property, when such quick clays have failed under buildings. Fortunately, these are quite rare.
The most important mass-movement type in terms of transferring material downhill is soil creep, the slow (typically inches, or centimeters, per year or less) downslope motion of soil. Creep may be a very slow landslide. It may occur from freeze-thaw processes—a column of ice that grows under a small pebble on a cold night pushes that pebble out from the hillslope, and the pebble falls straight down when the ice melts, effectively moving a tiny distance down the hill (see the figure above). When trees fall over and uproot soil, or when groundhogs and even worms dig out rock grains and allow them to move downhill, creep is occurring. If you look at a typical hillslope, streams on the lower slopes are present to move water and rock downhill, but the upper slopes lack streams. There, soil creep serves to move material downhill.
Hillslopes typically reach a balance, in which weathering breaks down rocks about as rapidly as mass movement and streams take the broken rocks away. The balance may occur with bare rock sticking out (as in some cliffs, for example), or with a lot of soil covering the rock. If soil creep dominates the mass movement, the hillslope may always be close to balance. If landslides dominate, then the soil will build up for a long while before suddenly slumping off, and you have to watch for a long time to see that formation and transport of soil are in balance. And, over very long times, the hill will eventually flatten, causing the mass movement to slow. However, soil will very gradually thicken to slow the weathering as the hillslope is reduced, and near-balance will be maintained.
Humans are greatly upsetting this balance worldwide. Our activities—bulldozing, cutting trees whose roots held the soil, plowing, and more—are moving more material than nature moved before we were involved.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 5.
TECH NOTE - Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Death Valley National Park
(Provided by UCGS)
Grand Tetons National Park
(Provided by UCGS)
There is one GeoMation and there are four GeoClips available for your viewing in this unit. Hopefully, you'll find this helpful in understanding Unit 5 just a little bit better.
We can measure the uplift of mountains, which may occur slowly, or suddenly in earthquakes, and we can watch volcanoes erupt. But overall, nature tears down mountains about as rapidly as they form, and we can watch and measure the tearing-down, too. The slow disappearance of names from old tombstones, the hubcap-rattling holes in late-winter city streets, and the maintenance budget for university buildings all attest to the effects of nature on human-made things. Here, Dave Witmer takes you to Bryce Canyon, one of the many, many places where you can see nature removing natural things.
Geologists observe the wear-and-tear of nature on human-made and natural things, gaining clues to help understand how mountains are torn down. When climbing the sheer cliffs of Zion National Park into the mysterious crevice of Hidden Canyon, the intrepid hiker clings to a rather precarious-looking chain to avoid falling into the stream-carved potholes just beside the trail, and on down to the Virgin River, in the Canyon a hair-raising drop below. In these two clips, Dave Witmer and Dr. Anandakrishnan show how rocks are worn away, a little at a time, and what this has to do with south-Indian cuisine. You might begin thinking about what this wearing-away of rocks has to do with the Virgin River in the Canyon far below.
Rocks and soil on hillsides really do slide down, whether rapidly or slowly, in big chunks or small ones. When a river erodes downward to make cliffs in the riverbank, or when a volcanic eruption or an earthquake makes a cliff, rocks fall or creep or slump or are washed off the steep slope, smoothing it over time. Here, Dr. Alley shows this for a tiny "canyon" in the bottom of the truly Grand Canyon.
Enrichment: More movies for you to enjoy—and these two won't be on a quiz. Erosion has carved fantastic forms from rocks, and weathering is important in loosening pieces to be transported away. Here, National Park Service Ranger Jan Stock and the CAUSE students explore weathering in Bryce National Park, and then Dr. Richard Alley explains changes at Delicate Arch in Arches National Park.
Bryce Windows and the Freeze/Thaw Cycle, (Transcript)
Here are some optional animations you might also want to explore! (No, these won't be on the quiz!)
Soil Erosion
(An extensive collection of animations on this subject)
Mass Wasting/Landslide Animations
(An extensive collection of animations on this subject)
The Unit 5 lecture features Dr. Sridhar Anandakrishnan and is 57:43 minutes long.
Check out the Unit 5 Overview Presentation used in the online lecture here.
The rotation of the Earth is surprisingly fast. If you buckled a belt around the Earth at the equator, the belt would need to be 25,000 miles (about 40,000 km) long. The Earth spins once every 24 hours. This means that the Earth’s rotation is moving a point on the equator just over 1000 miles per hour (1600 km/hr). In comparison, a point near the pole is essentially stationary (the pole itself stays in place as it rotates). Dr. Alley has walked around the South Pole in three steps, but he couldn’t walk around the equator in a day. If this doesn’t make sense, get a tennis ball or the head of a friendly classmate, draw an equator, and try it out!
If you stand on the equator, the wind is not blowing 1000 miles per hour. This is fortunate indeed, because the “mere” 150-mile-per-hour winds occasionally reached in hurricanes can cause immense disasters. Rather, the air near the surface of the Earth on average is moving with the surface, because of the drag of mountains and trees and such on the air. This is somewhat akin to driving down the highway at 65 miles per hour with a bug stuck to your windshield. The wind at 65 miles per hour is plenty strong enough to blow a bug, but the wind very close to the windshield is very much slower than 65 miles per hour—really strong winds don’t blow close to surfaces, because of the friction of the surface, and instead the air close to the surface does almost the same thing that the surface is doing.
Now, suppose that you watch a parcel of air that rises at the equator and begins moving towards the North Pole in a convection cell. Once the air has moved a ways north, the Earth under it is no longer rotating at 1000 miles per hour, but is somewhat slower, perhaps 900 miles per hour and dropping northward toward zero at the pole. But aloft, there aren’t trees and mountains to slow the air down from the 1000 miles per hour it had at the equator. Thus, this air “gets ahead” of the Earth. The Earth rotates to the east—you see the sun rise in the east as the rotation of the Earth brings the sun into view—so, the equatorial surface air is moving east at 1000 miles per hour. As this air moves northward over slower-moving land, the wind will appear to turn to the right or east as it blows, getting ahead of the ground. Wind heading south from the equator will also move east ahead of the surface, making a left turn.
Similarly, wind moving from the pole to the equator in the returning limb of a convection cell will lag behind the rotating surface of the Earth, again seeming to turn to the right in the Northern Hemisphere (and to the left in the Southern Hemisphere). Thus, the wind cannot go directly to where it “wants” to go; instead, it turns and tends to go in circles. The circular air flow around low-pressure systems and hurricanes is unavoidable on a rotating Earth.
More-precise definitions are possible of this “Coriolis effect,” the turning of flows on a rotating Earth. The intuitive explanation given above will fail you if you think of a wind moving due-east or due-west, because those also turn. Starting from conservation of angular momentum might be better, and you can find that in any good meteorology textbook. Notice, however, that the explanation given above will get the right answer for you, in how much the wind will turn, and which way.
Notice also that Coriolis turning affects large, fast flows. The convection in the Earth is too slow to feel Coriolis much. The difference in rotation speed between opposite sides of a sink or toilet is so tiny that Coriolis turning has no significant effect on the direction that water swirls as it goes down. That is controlled by the design of the sink or toilet, and by any motion in the water at the time the drain is opened; get the water swirling in a sink and then pull the plug, and the water usually will keep swirling in the way you started it as it goes down. Dr. Alley has seen both clockwise and counterclockwise flows in Pennsylvania and Greenland, and in New Zealand and Antarctica.
When air moves up, it expands, which requires that work is done in pushing away other air to make room for the expansion. The work requires energy, which comes from the heat energy in the air, so the rising air cools. Similarly, when air moves down, it contracts as the surrounding, higher-pressure air squeezes the sinking air parcel, and this squeezing is work that is done on the sinking parcel, and warms the parcel. If this is happening near the surface of the Earth, and the air is dry, the change in temperature is about 1oC per 100 m of vertical motion. This applies everywhere, at all times. So, in the 2004 movie, The Day After Tomorrow, when huge storms brought air down from above so rapidly that the air didn’t have time to warm up, that was complete hooey.
Now, imagine that a “chunk” or parcel of air, sitting somewhere, cools a little (say, by radiating energy to space as the sun goes down in the evening), and you want to know whether that air parcel will sink now. If your air parcel sinks, it must warm by about 1oC per 100 m of sinking, and it must push other air out of the way. You measure the temperature of your parcel, add 1oC for the warming from sinking 100 m, and if your air is still colder than the air it must displace 100 m below, then your parcel will sink. (Friction comes in as well—if your parcel plus 1oC would be only a tiny bit colder than the air it must displace, motion is unlikely; you need a notable difference.)
Overall, a balanced, stationary atmosphere will cool upward by about 1oC per 100 m under dry conditions, and slightly less under wet conditions (when cooling causes condensation that releases heat which offsets some of the cooling, or warming causes evaporation that uses heat which offsets some of the warming). Vertical motions will be triggered when cooling or warming creates air that is anomalously cold or warm relative to this stationary profile. So, cold air on a mountaintop won’t necessarily sink, unless that mountaintop air is anomalously cold. Perhaps on an October evening in the Appalachians, when a fog develops and holds heat in the valleys while the mountaintop radiates heat to space, then the mountaintop air will become anomalously cold and sink to the valleys.
If you want another look at the weather system, and the difference between the Redwoods and Death Valley, the Wizard of Odd takes you Somewhere Over the Puddle in this review revue. (The Sierra tops out over 14,000 feet but in most places is lower, so don't let it bother you that the air in the GeoClip went a little higher than the air in this song--both are right, depending on just where the air goes over.)
You have reached the end of Unit 5! Double-check the list of requirements on the Unit 5 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Exercise #2 is due and Exercise #3 opens this week. See Course Calendar for specific dates and times.
Following are some supplementary materials for Unit 5. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
(with apologies to John Fogerty & Creedence Clearwater Revival)
Left a big cliff in a landslide, Loosened by a rain and an earthquake, too
And I never lost one minute of sleepin', Worryin' 'bout that trench 'neath the ocean blue.
River keep on movin', Bed load keep on groovin'
Rollin', rollin', rollin' with the river.
Shot through a braid in the mountains, Wrapped around a big old meander bend
But I cannot see the good side of the river, A reservoir trapped me away from my friend.
River can't keep groovin', When dams stop the bedload movin'
Holdin', holdin', holdin' from the river.
If you go down to the river, Bet you're gonna see some houses too near
They might want to worry, storms are in a hurry, When the levees fail, there is something to fear.
River will get movin', Wildness will be provin'
Rollin', rollin', ever-rollin' river.
You will have one week to complete Unit 6. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 6, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Review the Unit 6 Overview | No, the overview outlines the main topics and ideas that you will encounter in Unit 6. You will, however, be tested on the material found in the overview. |
Read/view all of the Instructional Materials for Unit 6 including: | No, but you will be tested on all of the material found in the Unit 6 Instructional Materials. |
Take the Unit 6 "RockOn" quiz | Yes, this is the sixth of 12 end of unit RockOn quizzes and is worth 4.5% of your total grade. |
Continue working on Exercise #3: | Yes, this is the third of 6 Exercises and is worth 5% of your total grade. |
Complete the "StudentsSpeak #7" survey | Yes, this is the seventh of 12 weekly surveys and is worth 1% of your total grade. |
Read the Optional Enrichment Article | No |
Review the Unit 6 Wrap Up including the Supplemental Materials | No, but you may find them helpful in preparing for the quiz. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to complete Unit 6, including the online textbook, a video lecture, several VTrips and animations, and two overview presentations.
Students who register for this Penn State course gain access to assignments and instructor feedback and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
It was kind of solemn, drifting down the big, still river, laying on our backs, looking up at stars, and we didn't even feel like talking aloud..."
— Mark Twain, Adventures of Huckleberry Finn, Chapter 12.
Southwest of Moab, Utah lies Canyonlands National Park. A “rough” park with few services available, little water (except for the rather muddy rivers), pot-holed roads, and awesome mountain biking, Canyonlands preserves the confluence of the Colorado and the Green Rivers. Many “visitors” to the park never actually enter it, choosing to gaze down on the Colorado from the vantage point at Dead Horse Point State Park. The two rivers of the park are incised a third of a mile (half a kilometer) or more into the red sedimentary rocks of the Colorado Plateau. Those rocks, mostly sandstones (made from sand) and shales (mud rocks, made from pieces smaller than sand), give the distinctive cliff-slope pattern of the canyons—resistant sandstones form cliffs and cap the flat-topped mesas, while softer shales form slopes.
The great Colorado Plateau, flanked by the spreading regions of the Basin and Range to the West, and the Rio Grande Rift to the east, occupies large parts of Utah and Arizona plus some of Colorado and New Mexico, and includes Zion, Bryce, Capitol Reef, Arches, Grand Canyon, Petrified Forest and Mesa Verde National Parks as well as Canyonlands, and many national monuments and other public treasures. The Colorado Plateau is noted for reddish, flat-lying, rocks from the Paleozoic of a few hundred million years ago.
The details are not all known of how the Colorado Plateau avoided extreme deformation for hundreds of millions of years, while most of the rest of the west was being bent, broken, or erupted. The silica-rich continental rocks of the Plateau may be a bit thicker than in its surroundings, so the spreading of Death Valley was unable to tear the Plateau apart and instead jumped across to the east side of the Plateau to continue as the Rio Grande Rift (the valley in which the Rio Grande flows). The spreading even seems to have nibbled away at the Plateau, with a few big pull-apart, Death-Valley-type faults visible in places such as Red Canyon just west of Bryce (see the picture below). However, our concern here is slightly different— the role of streams.
The rivers of the Colorado Plateau are nearly as well known as their parks, for great rafting, incredible scenic views, and deep canyons. The Colorado, the Green, the San Juan, the Fremont, the Virgin and others have taken their place in history. But what are those rivers doing down there in the canyons?
Simply put, a stream or river is a conduit to take excess water, and sediment, from high places to low ones and usually to the ocean. Looking first at the water, rain or snow falls on the ground. Averaged around the world, rainfall (plus snowfall after it melts) is about 3 feet per year. Pennsylvania's annual rainfall is also right near the global average, as is much of the tree-covered eastern US. Some of this water evaporates directly, but most is used by plants and then is transpired (evaporates) from them. The evaporation from plants and from other surfaces usually is lumped together and called “evapotranspiration.” In a humid temperate climate such as central Pennsylvania, roughly two-thirds of the rainfall is involved in evapotranspiration and returned directly to the sky; in dry climates, a larger fraction of the rainfall—maybe almost all of whatever rain falls—may be returned to the air by evapotranspiration.
Of the water that avoids evapotranspiration, a little actually falls on lakes or streams, and some may fall on the land surface and then flow directly and immediately over the surface into lakes or streams, especially from the surfaces that humans have covered with buildings, roads, and parking lots, thus keeping the rain from soaking in. But most of the rain that avoids evapotranspiration soaks into the ground to form groundwater.
Soils and most rocks include interconnected spaces, either gaps between grains of sand, the cracks in the rock called joints, caves, or other openings. The ground acts a bit like a sponge, with water soaking in and then slowly draining out to the rivers. We will discuss this groundwater flow a bit more when we visit Mammoth Cave, next; for now, simply note that because gravity pulls water down, rocks near the surface usually have some air in the spaces even where conditions are damp, and deeper rocks usually have all their spaces filled with water. The surface separating the rocks with water-filled spaces from those containing some air is called the water table, and where the water table intersects the surface of the Earth, a stream or lake occurs.
Rivers flow even when it isn't raining because water is slowly draining through the ground from beneath hills to the rivers. The water table rises in elevation during wet times as the “sponge” of the Earth fills up with rain, and the water table falls during dry times as the sponge drains to keep the streams flowing. And, the water table is just below the surface in valleys, actually hits the surface at streams, but you must drill deeper under ridges to penetrate the water table and complete a water well.
As we saw back at the Badlands, weather attacks rocks to produce loose pieces through the processes of weathering. And as we saw at the Gros Ventre slide in the Tetons, processes on hillslopes including soil creep and landslides deliver the loose pieces (which we can call sediment) to rivers. A river is then faced with a balancing act; it must transport both the water and the sediment delivered to it.
You may be able to think of many ways for the river to adjust if more or less water, more or less sediment, or bigger or smaller or “stickier” pieces (those more likely to clump together) of sediment are delivered. For example, if more sediment is delivered to a river than it can remove, then the sediment will pile up, raising the elevation of the bed of the river. This steepens the river—the elevation of the ocean where the river ends has not changed, but the elevation of the river bed away from the ocean is now higher—so the river flows faster and is better able to move the sediment. Deliver more water and less sediment, and the river will wash away all the sediment and have energy left over to carve into the river bed. This cutting downward will make the river less steep—because the river ends at sea level and can’t lower that, lowering the upstream reaches of the river must make the slope to the ocean less steep. A less-steep river will carry less sediment and so quit cutting its bed—the river tends to reach a balance in which it just removes the water and sediment supplied to it. In the process of reaching this balance, rivers also may adjust the width and depth of their channels as well as the steepness.
Rivers are diverse; a white-water rafter on the Youghiogheny River sees a very different setting than greets a river passenger on the Mississippi River going to New Orleans! Of the many river patterns, we often focus on two: meandering and braided (the Mississippi is a meandering river; the "Yough" has some meanders, but not nearly so well developed as the Mississippi). Whether a river meanders or is braided depends on the sediment supplied to the river as well as the water.
A river moves small sediment particles up in the water (suspended load), and larger pieces by rolling or bouncing them along the bed (bed load). If the river receives mostly small particles—what geologists call "fine-grained sediment"—the river typically will form a deep channel. Such a channel is especially efficient at transporting water, and the sediment in the water, because most of the flow is kept away from the river bed and river banks where friction with tree roots and other things slows the water. Very fine sediment—clay—can make fairly deep and steep river banks without collapsing, because the clay particles stick to each other—if you’ve ever worked with clay in art class, you know how well the clay particles stick together, and how much easier it is to make a pot out of clay than out of gravel or sand!
Such deep streams typically curve back and forth, or meander, along their paths. Put a tree’s roots in the way of the stream, and the water flow will curve around the obstacle. In doing so, the stream will race along the outside of the curve and erode it, and a meander bend will grow. (Meandering has been observed to occur without any tree roots in the way, and so is a bit more complex than just curving around obstacles.) Meandering streams usually occur in relatively flat, lowland regions towards the coast.
If a stream receives lots of sand and gravel or even bigger chunks, the large blocks will tend to plug, or dam, a single, deep and narrow channel. The stream then assumes a wide, shallow pattern that is efficient at rolling debris. When sand and gravel get really wet along a river bank, they cannot form steep slopes, as we saw with the mass movements in the Tetons, and the collapse of any steep slopes that start to form along a river bank contributes to having a wide, shallow stream. Within this broad stream bed, gravel (or boulder) bars often form when floods are ending and losing their ability to roll lots of sediment; the water then must flow around these bars. The splitting and rejoining of channels around bars, when viewed from above, looks something like ropes of water that have been braided together, so these are called braided rivers. They are common in upland regions, where steep mountain slopes shed landslides of coarse rocks into the channels.
Again, please recognize that this is a very simple view of stream processes. You can find meandering streams in the flat bottoms of valleys in which beaver dams have trapped mud, for example. Streams with bedrock rather than sediment in their beds often have relatively straight single channels, as do many small streams with sediment beds. The key thing to remember about a stream is that it exists to move sediment as well as water.
A tremendous example occurs a little ways downstream of Canyonlands and upstream of the Grand Canyon. The Glen Canyon Dam was built on the Colorado River in the 1960s. The dam stopped floods coming through Canyonlands from reaching the Grand Canyon—water from floods that had raged through the canyon now is stored in the reservoir and then released gradually. Several things began happening to the river once the dam was completed. The dam trapped the sediment carried by the river, and released clean water, so the reservoir is filling with sediment, and in a few centuries or less will be full. Unique fish species that thrived in the muddy waters of the Grand Canyon suddenly were easy prey for clear-water species that were introduced, so many of the native species are endangered and disappearing now.
With no floods raging down the canyon, high water no longer piled up sand in corners of the canyon to make sand bars. The clean water was able to carry some sand and smaller pieces of sediment, so the existing sand bars below the dam were slowly washed away. The many types of wildlife that depend on sand bars thus were harmed—cottonwood trees that rooted in the sand, birds that lived in the cottonwoods, deer that came down to drink from the river by standing on the sand bars rather than the rocky cliffs. Floods on un-dammed side streams continued to dump large rocks into the Colorado, but the Colorado lacked the high flows to move this material onward, so the rapids at the mouths of side canyons began to steepen.
In the spring of 1996, an artificial flood was released from the dam to attempt to re-balance the system. Indications are that this flood was a partly successful experiment, rolling some of the big rocks out of the rapids at the mouths of the side streams, freeing sand trapped beneath, and forming that sand into bars. But, those bars weren’t very big and didn’t last very long. Additional human-made floods were released in 2004 and 2008, timed to occur when natural floods coming down side streams were delivering additional sediment, to help make bigger sand bars. Additional attempts may be made. Such human-caused floods cost money (lost hydroelectric power when extra water is routed around the hydroelectric plant in the dam to get water into the river in a hurry) and require lots of planning (you need to warn people camping or hiking in the Canyon before you suddenly flood them out!), and they can never get the Canyon back the way it was originally, but the artificial floods do seem to have helped restore some of the sandbars that are so critical for wildlife.
Meanwhile, upstream of the dam, as sediment builds up to fill the reservoir, sediment will also accumulate along the river upstream of the reservoir. There aren’t many people living there, but if there were, their fields and houses would begin to be buried by mud. A river slows down as it enters a reservoir, or any other lake or the ocean, and sediment is dropped from the slowing water. Unless strong waves and currents in the reservoir or ocean take that sediment away, a pile called a delta forms. But, the delta cannot be perfectly flat on top. If it were, then the stream would drop its load when it hit the flat spot and slowed down, and that would raise the flat spot. So, as the delta grows into the lake, the upstream end of the delta must build up so that the river still flows downhill, and that, in turn, will cause sediment to build up for some distance upriver (see the figure below).
The troubles with the dam on the Colorado at Glen Canyon are not unique. Reservoirs immediately start filling with sediment, and so cannot last forever. This filling contributes to deposition of sediment upstream, which may bury houses or fields. The clean water released tends to pick up fine-grained sediment below the dam, but the lack of floods means that the river is no longer able to move coarse-grained sediment. Hence, the river bed is changed, affecting species that lived there. The lack of floods often has huge effects on the plants and animals that live along the river; what had been a region reserved for wildlife adapted to the changing water levels and channels becomes a place for humans to settle in the shadow of the dam.
When two dams were built on the Elwha River, which flows north from Olympic National Park, the dams contributed to blocking salmon runs, to washing away the sand and gravel in which the salmon had spawned downstream of the dam sites, and to turning an annual “flood” of 300,000 salmon into a trickle of barely 3,000 salmon. (A fish hatchery was built instead of fish ladders around the first dam, but the hatchery was quickly abandoned.) Once the spawning bars were washed away below the dam, the river quit delivering sediment to the beaches of the Strait of Juan de Fuca (an arm of the Pacific Ocean); as the beaches washed away, the native peoples were no longer able to carry on their traditional shell-fishing, and engineering was required to protect the nearby harbor of Port Angeles, which was no longer guarded by sediment-fed bars. In an ambitious plan to help the beaches, the river, the salmon and the park, the federal government purchased the dams. One was removed in spring of 2012, with the second slated to follow quickly. The river, beaches, harbor and more will take a while to get back to "normal," because so much was changed by the dams, but there is much optimism about the recovery. You can read more, and track the progress, at the official website for Olympic National Park.
At the tip of the Mississippi Delta lies the Delta National Wildlife Refuge. This is one of several wildlife refuges along the Gulf Coast that, in addition to spawning a great range of resident wildlife, draw migrants from the north. Ducks and geese, herons and cranes, gallinules and rails, the wetland birds of most of a continent stream in through the fall, and then spread north again in the spring. (Yes, technically, a National Wildlife Refuge is not a National Park, but it is a national park, so we’ll cheat a little and use it—it makes a good story. And, it is just down the river from the wonderful bayous of the Barataria Reserve in the Jean Lafitte National Historical Park and Preserve—be sure to stop if you're in the area!)
Unfortunately, these wetlands are disappearing at an astonishing rate, because of the indirect effects of human activities. Estimates are that every year Louisiana is losing over 100 square kilometers of wetlands (equal to loss of a square with sides more than six miles long). Whether the wetland birds will continue to stream north for generations to come may depend on how humans respond to the challenge.
The Mississippi Delta is a massive pile of mud and sand from the Rockies and Appalachians, transported by the river and dumped into the Gulf of Mexico. The Gulf used to extend far into the heartland of what is now the U.S.; over the last 70 million years, the delta has grown southward from near Cairo, IL (up by St. Louis), until now the former embayment has been turned into a projection from the end of Louisiana into the Gulf. There, the delta is as much as seven miles thick. If you have ever watched mud settle in a bottle of water, or if you have observed how your boot packs mud down if you step in it, then you know that, over time, mud will compact under its own weight or under the weight of anything placed on it.
As the delta grew into the Gulf over the millennia, a natural balance was reached. The compaction that occurred during a year would leave a little space at the top, but the springtime floods would bring new mud to fill the space. Trees and other vegetation would grow up through the new sediment, or re-seed on top, and the system would continue, wildly productive and vibrantly green.
There is a problem with this system and humans, however. Many people have settled near the river. Plants can grow up through the mud of floods, but people don’t enjoy having their houses slowly fill up with mud. So, humans have built control structures. We built dams upriver, which trap sediment behind them and which hold some floodwaters in check. Because large floods threaten dams and must be let through, we also built levees along the river in its downstream reaches, great walls that hold the river in. We also dredge the river, deepening it to carry the water—and shipping. The great floods that shoot down the river then do not spread over the floodplain and the delta, depositing fertile sediment to fill the space left by compaction of mud, but instead are piped to the Gulf, where the sediment jets off the edge of the delta to settle in mile-deep water.
Way back in 1996, when the very first edition of this textbook was written, it read:
"Today, much of New Orleans, which does lie on the delta, is well below sea level. A tanker in the river between its levees is higher than the playing field of the Superdome. Rainfall, and water seeping from the river, must be pumped out so that the city doesn’t fill with water. If the pumps were to fail, the city would become a lake. The city steadily sinks deeper, and the levees are steadily raised by the Army Corps of Engineers, as instructed by Congress, to keep the river caged. Meanwhile, the wetlands of the delta, unnourished by new sediment, are sinking beneath the Gulf..."
After 2005, we know how terribly accurate the geological understanding of New Orleans really was. Where natural wetlands should have slowed the waves from Hurricane Katrina (which was not a really big storm by the time it got to New Orleans!), the high waters of the storm surge roared unimpeded from the Gulf. Parts of the levees failed. The pumps failed. The city filled with water, as much as 20 feet deep.
The hurricane showed what geologists (and emergency planners!) had long known, and had long told Congresspersons and Senators and students and others—New Orleans was a disaster waiting to happen, not “if” but “when.” With almost 2,000 people killed, hundreds of thousands of people displaced, and damages approaching $100 billion (that is, more than $300 from every single person in the United States), the danger of ignoring solid scientific evidence and hoping for the best is clearer than ever.
But, the city is rebuilt where it was, the sinking will continue, the loss of wetlands will continue unless many things are changed, and the levees will need to be raised. With the likelihood that the strongest storms will get stronger and sea level will rise in the future (we’ll revisit this later in the semester), the scene will be set for an even more horrific disaster at some future date. Many options are available, including restoring wetlands, filling parts of the city with debris or other materials, moving construction to higher parts of the city, moving out entirely, and more; it will be interesting to see how much of this will be done. But primarily, the donations and tax dollars from the rest of the country after the 2005 disaster were used to rebuild the city directly in harm’s way, with the knowledge that the rest of the country will once again foot the bill when disaster strikes.
Another story is being played out in this region as well. The river wants to leave New Orleans. The city has a love-hate relation with the river, fearing the floods but needing the drinking water and the shipping channel. The river can harm the city rapidly by flood, or slowly by leaving.
To understand this tendency of the river to leave New Orleans, note that especially large, muddy, flood-prone rivers normally have natural levees (which are much lower than the human-made ones). When a flood happens, the water spreads out of the main channel onto the flood plain, the flattish region of river-deposited muds next to the main channel. As the water spreads out into the trees or houses of the flood plain, the flow slows, and the water drops some of its muddy load. Just as the water leaves the main channel, the water is carrying the most load, and the slowing is most prominent, so most of the load is deposited right there next to the channel. Hence, the mud layer from a flood is thicker next to the river than farther away, forming a natural levee. Humans have raised these natural levees in many places.
When we discussed reservoirs, we saw that the delta of sediment formed when a river enters a lake must build up as well as out, so that the river still flows downhill into the lake. The same is true for a river entering an ocean. The Mississippi River, with its levees, naturally dumps mud into the Gulf of Mexico, slowly lengthening and raising the river bed. After a while, the river is a bit like a log flume in an amusement park, following a long path to the Gulf; a break in the levee wall would allow a much steeper, shorter, and more exciting downhill trip. The recent history of the Mississippi Delta is that, roughly every 1,500 years, the main outlet of the river has broken through the natural levee, like a log full of park-goers breaking through a curve in the ride, and the river has then followed that new shortcut. But, as mud is deposited along that new shortcut, it lengthens until it is like a long log flume, and then the river breaks through a side again.
During the 1940s and 1950s, the Mississippi started to break out, into a side stream called the Atchafalaya River. To save the shipping channel and the water supply for New Orleans, the Army Corps of Engineers has used the Old River Control Structure and other dams and levees to allow some water to go down the Atchafalaya while keeping a vigorous flow in the main log-flume channel past New Orleans. During a flood in 1973, the Corps very nearly lost the Control Structure, and the river, when a giant whirlpool undercutting the dam came close to causing it to collapse. The task of the Corps is very difficult, taming immense natural forces as the system becomes more and more out of balance.
(An excellent account of this is given in John McPhee’s book The Control of Nature, 1989, Farrar, Straus and Giroux, New York, which may be a little out of date but is still fascinating, and shows that policy-makers and others were warned about the dangers in the area long before the disaster of the 2005 hurricane.)
Deep beneath the rolling Kentucky hills lies Mammoth Cave. With about 390 miles (almost 630 km) of surveyed passageways, this is the longest known cave in the world, more than twice as long as second-place Jewel Cave (which is in a national monument in South Dakota, and has just under 160 miles or 260 km of surveyed passageways). The great size was achieved when a team of explorers showed that the historical part of Mammoth Cave and the even larger neighboring Flint Ridge Cave were actually connected, followed by a lot more surveying. There is every reason to expect that further connected passages remain undiscovered, and estimates extend as high as 1000 miles of total passageways.
The cave was mined for saltpeter (containing nitrates) for use in gunpowder, especially during the war of 1812. The source was bat guano (the polite name for it) deposited over the ages by great flocks of bats. Native Americans used the cave long before European settlement.
The cave was dissolved in limestone. The limestone was deposited in shallow seas during the Paleozoic (a few hundred million years ago), mostly as shells of sea creatures. The limestone lies beneath sandstone from old beaches. The sandstone is a rather sturdy layer, and provides a “roof” that does not collapse easily, which is important in making Mammoth Cave so big.
As we saw in discussing rock weathering to make muds for the Badlands, rainwater and soil water are weak acids. Chemically, limestone is calcium carbonate, which is especially prone to attack by acid. The usual test for limestone is to drip a little weak hydrochloric acid on a sample; limestone fizzes vigorously as the rock decomposes, freeing carbon dioxide gas, but most other rocks react much more slowly and do not fizz. Where soil waters move through limestone, the rock dissolves and washes away rather rapidly. If the rock has lots and lots of cracks, the water may follow so many different paths and spread out the dissolution so evenly that caves are not formed. But if the limestone has just a few cracks for water to flow through, all of the dissolution will be concentrated in those few places, and cave passageways may form. Then, if the water table is lowered (say, because the nearby river has cut downward through rocks and is draining water from the rocks to its new lower elevation, as Kentucky’s Green River has done near Mammoth Cave), water will drain away and the cave will fill with air.
The beautiful stalaCtites (from the Ceiling), stalaGmites (on the Ground), and other cave formations can then develop. Because processes in soil—worms exhaling, and similar activities—typically release carbon dioxide, groundwater usually contains more carbon dioxide than does rainwater, so groundwater is more effective than rainwater at dissolving limestone. Occasionally, a cave may be so isolated from the surface that dangerous levels of carbon dioxide exist in the cave, but caves usually exchange enough air with the outside world to have near-normal levels of carbon dioxide. When groundwater drips into a cave that has a near-normal carbon-dioxide level, the water loses carbon dioxide to the air. The water then cannot hold all of the dissolved limestone, and deposits some of it to form the beautiful stone features we see.
Sometime, look at a cliff or road cut (but be careful—we don't want to lose anyone to distracted driving, or being run over along a roadside!). If you look carefully, you should see that almost all cliffs and road cuts have cracks (called joints), and that some of those joints are usually vertical; very often there are two or more sets of vertical joints, perhaps with many joints oriented north-south but many others oriented east-west. Rainwater that is not used by plants will trickle down the “pipe” where the joints intersect. If the rock around the pipe is limestone, the rock will dissolve, leaving space, although that space may fill with mud. Such a hole, whether mud-filled or air-filled, is called a sinkhole.
Sinkholes formed by downgoing waters are very common in the State College, PA area, where Dr. Alley lives. Penn State’s Geosciences Department is housed in the Deike Building, which required extra funding for special strengthening because the building has sinkholes beneath—a building can rest firmly on bedrock, but tends to fall into air-filled or mud-filled holes. Extra funds were similarly expended to strengthen the nearby Mt. Nittany Middle School, the runway extension at the airport, and other construction projects in the area. A newly constructed storm-water catch basin at the airport filled with water during the first rain, and the weight of the water blasted mud out of a buried cave passageway somewhere beneath, suddenly turning Spring Creek red with trout-choking sediment.
Where sinkholes and caves are common, streams often disappear underground into swallow holes, only to re-emerge at springs. Spring Creek is aptly named, and many other similar features occur around central Pennsylvania, around Mammoth Cave, and in other such regions. Corn cobs once were dumped in a sinkhole behind a cannery at Old Fort east of State College, and after a rain would pop out of a spring in Spring Mills, a few miles away. Regions with sinkholes, caves, springs, swallow holes, etc., are referred to as karst, after a region in Slovenia with many such features. Karst features are present across 20% of the Earth’s surface, and roughly 40% of the US population obtains drinking water from karst, according to the National Park Service.
In the past, people often threw trash into sinkholes. Big pieces would sink into the mud or fall into cave passages beneath, “disappearing.” When Dr. Alley was in high school and went to Sloan’s Valley, Kentucky to go wild caving (spelunking), one of the cave entrances was known as the Garbage Pit, which led into the Tetanus Tunnel. A commercial cave near Mammoth Cave was forced to close in the 1940s because of the stench from sewage draining in.
Slowly, we are learning just how stupid it is to dump things in sinkholes. A test run by Penn State’s hydrogeologist, the great Dick Parizek, during the building of the Nittany Mall showed that a little harmless dye dumped in a sinkhole near the mall came out in a nearby trout stream in a day or two. It should be evident that anything else dumped in a sinkhole near the Nittany Mall (or many, many other sinkholes in the region and in other karst regions) would show up very quickly in the water used by people and wildlife.
Dr. Alley lives in a house served by the Lemont Water Company, which is now well-known for fine water. But many years ago, before a reorganization of the water company, the intestinal parasite Giardia showed up in Lemont well water. Giardia causes intense and possibly dangerous gastric discomfort, and is to be avoided. Giardia usually is restricted to surface water; the spaces in most rocks are small enough to filter out the Giardia cysts before they reach a well, or the water takes so long to go from the surface to a well through the small spaces that the cysts die of old age on the way. At the long-ago community meeting to discuss the water contamination, company officials noted that they had installed well filters to remove sticks, leaves, etc., that came out of the wells with the water. In karst country, surface water can become groundwater and return to the surface in hours or days. Whole streams go down and up, and if sticks can go through, microscopic cysts can, too. Clearly, contaminants dumped somewhere today can be poisoning someone tomorrow.
In some other regions, the groundwater-contamination problems are quite different. In sandstones, for example, the water moves slowly, pore-by-pore, through the rocks. In some places, the water can be shown to have first entered the ground during the ice age, more than 20,000 years ago, or even earlier. Contaminants dumped in such rocks may not bother people for a while. But, when the contaminants do start to bother people, clean-up can be very difficult.
Try this experiment. Squirt some soap on a sponge, and squeeze the sponge a few times to distribute the soap well. The sponge is our rock, and the soap is the contaminant. Now, wet the sponge, hold it up, and squeeze it. Foamy, soapy water will come out. Wet the sponge again, squeeze it again, and more soap comes out. Repeat, and repeat, and repeat. You may need ten or more times to remove enough soap that you no longer see it, and sensitive instruments would detect the soap through dozens or even hundreds of additional washings. Now, suppose that we had not soap, but a chemical that causes cancer in humans. If the water in the rocks naturally is hundreds or thousands or more years old, then nature takes a long time to wash out the rocks once, and washing them out ten or one hundred times will take much longer than all of human history.
There are things that can be done about groundwater pollution. You can pump clean water in and dirty water out, or pump steam or hot water in and out (and then try to figure out how to clean the dirty water or steam once you have them on the surface). People are experimenting with installing filters so that polluted water will flow through them, sometimes using large masses of iron filings to react with and break down some organic chemicals in groundwater. Geomicrobiologists are prospecting really dirty sites for “bugs” that “like” to eat pollutants, and then trying to introduce those microorganisms into other polluted sites; other biologists are looking into ways to design pollutant-eating microbes. But, such techniques usually are very expensive and not very effective. Most people who have thought about it agree very strongly that the best way to handle groundwater pollution is to keep the chemicals out of the ground in the first place. A whole lot of money has been spent on clean-up because we did not learn that lesson soon enough—and there are days when it appears that we have not yet learned that lesson.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 6.
TECH NOTE - Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
This week, we feature three GeoMations, and two GeoClips from the CAUSE trip out west, featuring Dr. Alley and Dr. Anandakrishnan in the Spring of 2004.
The three GeoMations could appropriately be called 1) How Deltas work, 2) How Dams work, and 3) How New Orleans Doesn't work, and in the video, you'll get to see Dr. Alley and Dr. Anandakrishnan "argue" over the Glen Canyon dam and its effect on Lake Powell above the dam and the Colorado River below.
As before, we hope you enjoy these, and find them to be useful complements to the readings, class notes, and slide shows of Unit 6.
Dams cause huge changes on rivers, both upstream and downstream. In this film clip, Drs. Anandakrishnan and Alley discuss the Glen Canyon Dam and Lake Powell on the Colorado River. Huge changes were caused by this project, including in the Grand Canyon far downstream. The CAUSE 2004 class used some clever editing to manufacture a disagreement between the professors, who are much closer to being on the same wavelength than you might imagine by watching this.
Human actions are more and more affecting the amount and quality of water and the ground, with effects that bounce back on us and on other living things. Here, Dr. Alley chats about groundwater issues of great concern to the National Park Service at the Grand Canyon.
Here are some optional resources you might also want to explore! (No, these won't be on the quiz!)
Join Dr. Alley to learn about formation of Fossil Fuels, in Barataria Reserve, Jean Lafitte National Historical Park and Preserve, Louisiana, from the PBS television special Earth: The Operators’ Manual.
Deltas and Plumes
(An extensive collection of animations on this subject)
River Systems: Process and Form
(An extensive collection of animations on this subject)
Processes of River Erosion, Transport, and Deposition
(An extensive collection of animations on this subject)
The Unit 6 lecture features Dr. Sridhar Anandakrishnan and is 1 hour and 4 minutes long.
Check out the Unit 6 Presentation used in the online lecture.
Canyonlands poses a special puzzle. The rivers in Canyonlands meander, making big, sweeping curves through the red rocks. In the Canyonlands chapter, you read that meandering streams occur in flat, lowland regions lacking supply of big rocks to the stream. This is decidedly not the situation in Canyonlands. The streams are steep, and in places (although not just where the picture below was taken) are laden with rapids, challenging for whitewater rafters. Landslides and rockfalls from the canyon walls deliver large blocks to the streams.
The likely story is that the streams once flowed across nearly flat lowlands. Then, uplift of the rocks began, giving the streams a steeper slope to the sea and so speeding their flow and causing them to erode. But, the uplift was gradual enough that the streams held their old courses. The streams cut downward without a change in pattern, which is called incision. Canyonlands contains clear examples of incised meanders. Similar features are preserved throughout the Colorado Plateau, documenting widespread uplift of the Plateau.
Much of geology involves study such as this; the history of a region produced the modern features. One can often learn much about that history by understanding the modern features and how they were formed. Clearly, if we did not know what conditions produce meandering streams today, we would not know what conditions likely occurred in the past when the meandering streams developed. The geological saying that captures this idea is “The present is the key to the past.”
Down at the Mississippi Delta, we saw that the mud under New Orleans is compacting under its own weight, contributing to sinking of the city. A couple of other things also contribute to the sinking.
As the delta grows, the weight of the mud pushes down the rock beneath, with soft, hot rock much farther below flowing away laterally, like water under your behind if you sit on a water bed. The sinking is not instantaneous. And, just as the strength of the water-bed cover spreads the dimple around your behind when you sit down, a fairly large region around the delta is pushed down by the weight of the delta. So add some mud anywhere near New Orleans, and you get a little sinking in New Orleans for a while.
Perhaps more importantly, as we saw with mass movements at the Grand Tetons (and as you can see in the picture of Canyonlands on the previous page, in the cliff on the inside of the meander), rocks tend to fall off steep slopes. Sometimes a single rock falls, or a thin layer of rocks slides. Other times, large thicknesses of material move. The Mississippi Delta is a giant pile of mud towering above the deep waters of the Gulf of Mexico, and that “cliff” is subject to downward motion of its materials. Some of this occurs along faults that are something like pull-apart Death-Valley-type faults—the fault intersects the Earth’s surface well inland, sloping down toward the Gulf, and the mud on top slides down and toward the Gulf.
After Hurricane Katrina devastated New Orleans, geologists renewed their efforts to understand what is going on geologically—such understanding should help in planning how to slow down the next disaster. An argument has erupted about the relative importance of the various reasons for sinking in and near New Orleans, with faulting probably more important (and mud compaction less important) than previously believed, and with sinking of the rock beneath minor but not zero. All are almost certainly contributing, but with a little work to figure out just how much to blame on each one. Naturally, all of them contributed to lowering of the surface, and sediment deposited from the river's floods filled that space—the river doesn't really care why the surface dropped, and deposits sediment in low places formed by any process.
You have reached the end of Unit 6! Double-check the list of requirements on the Unit 6 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Continue to work on Exercise #3. See Course Calendar for specific due dates.
Following are some supplementary materials for Unit 6. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
To read about the restoration of the Elwha River, go to the Olympic National Park website, and in particular to the Elwha River Restoration page.
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
Although not quite as large as minivans, musk oxen have better acceleration and cornering. This picture shows musk oxen thundering across the tundra of east Greenland. If you had been in central Pennsylvania's Bear Meadows with a camera 20,000 years ago, you might have taken this picture--tundra and musk oxen very similar to these existed in Pennsylvania and adjacent states back then, while an ice sheet even bigger than modern Greenland's loomed just to the north. How do we know that the ice came and went, and what caused the changes? Look both ways for musk oxen or minivans, depending on where and when you are, and let's go see.
You will have one week to complete Unit 7. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 7, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS |
SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 7 Instructional Materials. |
Submit Exercise #3: The Age of Nittany Valley. | Yes, this is the third of 6 Exercises and is worth 5% of your total grade. |
Begin Exercise #4: Which Way is Up? | Yes, this is the fourth of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 7 "RockOn" quiz. | Yes, this is the seventh of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #8" survey. | Yes, this is the eighth of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 7, including the online textbook, a video lecture, several vTrips and animations, and an overview presentation
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
When we try to pick out anything by itself, we find it hitched to everything else in the Universe.
— John Muir, My First Summer in the Sierra, 1911, p. 110
When your tour guide, Dr. Alley, was a much younger man (the year he graduated from high school, 1976), he traveled with his sister Sharon and cousin Chuck on a camping tour of the great national parks of the American west (in Chuck’s 1962 Ford Galaxy 500 land boat). At Yosemite, they hiked from the valley up to Glacier Point. The trail switch-backs up the granite cliffs, opening increasingly spectacular panoramas across the great valley of the Merced River. The view from Glacier Point, across the side of Half Dome, and the thundering Vernal and Nevada Falls, is well worth the climb. It was here that John Muir helped convince President Theodore Roosevelt of the need for a National Park Service to care for the National Parks, which were protected by law but not by rangers for some decades after the parks were established.
The hikers were a bit disheartened by the crowd at Glacier Point—the view is also accessible by the Glacier Point Road. While they sat and lunched, a tour bus pulled in. Most of the passengers headed for the gift shop, but three settled at a picnic table while a fourth strolled over to the railing to see the scenery for a few moments before joining the others at the picnic table. One of the quick-sitters asked “Anything out there?” To which the ‘energetic’ one replied “Nah, just a bunch of rocks. Let’s go check out the gift shop.” At last report, the gift shop had been removed. Regardless, it must be a sad person indeed who would not walk 50 feet to see the glory of Yosemite.
To anyone with open eyes, Yosemite Valley—the “Incomparable Valley”—is well worth inspection. It is carved from the granites and similar rocks of the high Sierra Nevada of California. Once, this granite was magma (melted rock below the surface), far beneath an earlier mountain range. The magma may have fed subduction-zone volcanoes much like those of the Cascades, which continue to the north of the Sierra. However, stratovolcanoes along this part of California have died as the East Pacific Rise spreading center ran into the trench along the west coast, forming the San Andreas Fault but ending subduction. Such a fate eventually awaits the Cascades volcanoes, some millions of years in the future.
The Sierra Nevada was raised and tilted along the great fault to its east, and looks down on Death Valley and the rest of the Great Basin. Earthquake activity, and faults cutting recent sediments, show that the mountain range is still being lifted above the still-dropping Great Basin.
The tough granite of the Sierra Nevada is more resistant to weathering and erosion than are most rocks, but granite does eventually break down, and some streams have managed to exploit weaknesses and cut deep channels through the range. These streams include the Tuolomne River, which carved the mighty Hetch Hetchy valley, now dammed so that a valley the equal of Yosemite is lost under water. The Merced River, which runs through Yosemite Valley, also cut into the range.
The stage was then set for the ice ages. Glaciers gathered on the high peaks, flowed into the valleys, and began to change the landscape.
A glacier is a mass of snow or ice that deforms and moves. Glaciers form wherever snowfall exceeds melting over enough years to make a pile big enough to flow. In places with extremely high snowfall, this can occur where average temperatures are near or even slightly above freezing, such as on the mountains of the Olympic Peninsula. In dry places, glaciers may be absent even if average temperatures are well below freezing. Such places have frozen ground instead, called permafrost (because the frost is permanent).
A pile of pancake batter spreads across a griddle, moving away from where the upper surface of the pile is highest. In the same way, a glacier moves from where its upper surface is highest to where its upper surface is lowest. In the figure below, the pressure at point A (the weight of the material above point A) is larger than at point B, because there is more ice above A than above B. The higher pressure at A gives a net push from A to B. Whether the diagram shows pancake batter, or the ice sheet on Antarctica, this push causes the material to deform and flow.
If you make a pile of pancake batter on a waffle iron, some of the batter may flow along the low grooves and then move up to cover the bumps, but the flow will always move away from the place where the upper surface of the pile is highest. In the same way, ice can flow up a hill in the bedrock if the flow is going in the “down” direction of the upper surface. For example, pieces of Canada are strewn across northeast Pennsylvania, and were brought across Lake Ontario and New York by the ice-age ice sheet. Ice at the bottom of that glacier climbed out of the low spot that now is the lake basin, driven by the upper surface of the ice sloping down from Canada to the U.S.
The total imbalance in pushes is larger for thicker ice than for thin. A very thin ice mass will not deform fast enough for the motion to be measured, and so is not considered to be a glacier. Typically, ice thicker than about 50 m (150 feet) will deform and flow, making a glacier.
Look at the second part of the figure. Glaciers move in one or more ways. All glaciers deform internally, like your slow pancake batter spreading on the griddle. A vertical hole drilled in a glacier will deform as shown in the figure. The stresses are largest, causing most-intense deformation (the permanent bending of the hole shown in the figure), in the deepest ice. The upper ice rides along on the deeper ice, so the velocity is fastest at the surface. Some glaciers are at the melting point at the bottom, warmed by heat flowing out of the Earth beneath. These melted-bed glaciers may move over the material beneath them, either by sliding over those materials (shown in the figure) or, if the materials are soft sediment, by deforming those sediments in a sort of slow landslide (not shown).
Recall that rivers adjust to move sediment and water from one place to another. So do glaciers. The water is supplied, frozen, in the accumulation zone, where snowfall exceeds melting, causing snow and ice to accumulate. The frozen water flows to the ablation zone, where melting exceeds snowfall (ablation means wearing away), or else flows to where icebergs break off (called calving) and drift away to melt elsewhere. For ice sheets covering continents or for smaller ice caps covering plateaus or mountain tops, the ice forms a dome and spreads out in all directions. For glaciers on the sides of mountains, the ice flows down the mountain—the upper and lower surfaces of the ice slope in the same direction.
When we talk about the advance and retreat of a glacier, we are referring to the position of its terminus, where the glacier ends by melting or calving. A glacier is advancing when it is getting longer, and retreating when it is getting shorter. Notice that ice almost always continues flowing from the accumulation zone to the terminus whether the glacier is advancing or retreating—retreat occurs when ice loss by melting or calving is faster than new ice is supplied, and advance occurs when ice is supplied more rapidly than it is removed by melting or calving.
Permanent deformation—flow—within ice may seem strange—after all, ice is a solid. But, as for the soft rock of the asthenosphere down in the mantle, or the soft chocolate bar in a hot pocket, or the red-hot horseshoe in the blacksmith’s shop, ice in all glaciers on Earth is nearly warm enough to melt, and so can flow slowly. As a general rule, materials heated more than halfway from absolute zero to their melting point can flow slowly, and flow becomes easier the closer the temperature is to the melting point. For ice, the coldest yearly average temperature on Earth is about eight-tenths of the way from absolute zero to the melting point, so ice at the Earth’s surface is “hot” and is able to flow. For more on this, and on the occurrence of crevasses as well as flow, see the Enrichment.
A glacier frozen to the rock beneath does not erode much. However, thawed-bed glaciers, especially those with surface meltwater streams draining to their beds through holes (something like cave passages, although formed in different ways), can erode much more rapidly than streams or wind erode. Consider for a moment the Great Lakes of the U.S. and Canada. The lakes were carved by glaciers. The bedrock beneath Lakes Superior and Michigan is well below sea level, and was carved that deep by glaciers, not rivers! Today, rivers carry sediment into the Great Lakes, slowly filling them up. We will see later that over the last million years, times when glaciers were eroding have alternated with times when streams were filling the lakes back up with sediment, and the streams have had more filling-up time than the glaciers had eroding time. And yet, there are the lakes, not the least bit full of sediment. Evidently, the glaciers have been much better at their “job” than the streams have been. The same can be said for many other places. It is not too extreme to say that the regions glaciated 20,000 years ago and free of ice today still preserve a glacial landscape.
Ice moving over bedrock “plucks” rocks free, and then uses those rocks to abrade or “sandpaper” the bedrock, scratching and polishing it. As ice flows over a bedrock bump, the side the ice reaches first is abraded smooth while the other side is plucked rough. Subglacial streams sweep away the loose pieces, and may cut into the rock as well.
Plucked and abraded rocks show clearly that glaciers were present, but so do big features, as seen in Yosemite and elsewhere. If you could make a cut across a typical stream valley in the mountains, you would see that it usually is shaped like the letter “V”. The narrow stream cuts downward, and then mass-movement processes remove material from the walls, giving a “V”. (Where a V-shaped stream enters an ocean to make a delta, the outward and upward growth of the delta over time may eventually fill the bottom of the V with mud to make a flood plain, as we saw with the Mississippi, but initially, when a stream is cutting down, it makes a V.) However, glaciers are quite wide, and can erode across a broad region. Glaciated valleys exhibit a characteristic “U” shape. Yosemite Valley, with its near-vertical walls and near-horizontal floor, is a classic “U”, not a “V”.
The steeper a stream is, the faster it erodes. If a main river cuts down rapidly, then the side streams that flow into it will become very steep, and will cut downward very rapidly. In this way, even a small side stream can “keep up” with the main stream as it erodes downward, and stream processes usually produce “rapids,” rather than waterfalls where the side streams must plunge over cliffs to reach the main stream. Glaciers are different. A main glacier often fills its valley, the ice burying most or all of the rock. The ice from a side glacier then does not drop steeply down into the main glacier because there is no drop. So the side glacier is not steeper than the main glacier. The main glacier has more ice and rock and water than the side glacier, and so the main glacier erodes down more rapidly. When the ice melts, a “hanging valley” remains—a small stream that replaces the small side glacier must plunge over a glacially carved cliff and then flow across the bottom of the “U”-shaped valley to reach the main stream. Eventually, the side stream will wear away the waterfall. But today in Yosemite, numerous streams emerge from small “U”-shaped hanging valleys to cascade down the glacially carved cliffs—the landscape is pretty much what the glaciers left. (Piles of rocks at the bottoms of waterfalls show that the streams are indeed changing things, but slowly.)
Glaciers make many other erosional features. At the head of a glacier, where freezing and thawing can break rock that is hauled away by the glacier, a bowl can be carved into the side of a mountain. If bowls chew into a mountain from opposite sides until they meet, a knife-edged ridge is left—the Garden Wall of the continental divide in Glacier National Park, which we’ll meet in the next chapter. Where three or more bowls intersect from different sides, a pinnacle of rock is left, such as the Matterhorn of Switzerland. Mountaineers have dubbed the bowls cirques, the ridges aretes, and the pillars horns, and geologists continue to use these terms.
Glaciers also leave distinctive deposits. Streams, waves and wind all sort rocks by size, leaving too-big ones behind and carrying away smaller ones. Glaciers don’t care how big the rocks are that the ice carries, so a deposit put down directly from ice may have the tiniest clay particles mixed in with house-sized boulders. Such a deposit is called glacial till. Till plus glacial outwash (sediment washed out of a glacier by meltwater) may be piled up together in a ridge that outlines the glacier, called a moraine.
Pennsylvania has a Moraine State Park, which features glacial moraines. Cape Cod is a moraine, and a moraine is draped across Long Island, showing some of the places where glaciers from the ice age ended.
Glacier National Park is the southern half of the Glacier-Waterton Lakes International Peace Park, extending north-south across the Canadian-U.S. border and east-west across the great Front Range of the Rockies. Glacier is wolves and grizzly bears, mountain goats balanced on cliffs, moose munching on water plants, beargrass and avalanche lilies. Going-to-the-Sun Road winds past Going-to-the-Sun Mountain, among the best-named features of the park system. The continental divide snakes along the Garden Wall, in many places a knife-edge ridge left as glaciers gnawed into the backbone of the continent from the east and the west. Long, narrow lakes lie along the valleys, which sometimes host lines of lakes strung like beads on the string of the connecting river. (Such glacier-carved strings of lakes are called paternoster lakes, after a resemblance to the beads of a Catholic rosary.
Glacier National Park had roughly 150 active glaciers a century ago, but that is down to 25 or so very small ones, and many of them may be essentially dead now, as modern warmth melts many away (see the changes shown by the older and more-recent photos below). Glacier National Park thus is now more noted for the tracks of past glaciers than for the activity of present ones. But, we suspect that “Ex-Glacier National Park” would not have made the Great Northern Railway happy when they were promoting tourism in Glacier National Park (via the Great Northern Railway, of course). When the last glacier has melted away, perhaps within a few decades (we'll return to this with the Arctic National Wildlife Refuge, near the end of the course), there are no plans to change the name of the park.
Today, permanent ice covers about one-tenth of the land on Earth, mostly in Antarctica and Greenland, with a little ice in mountainous regions. We saw at Yosemite that glacier erosion and deposition produce features that differ from those produced by mass movement, rivers, wind or coasts. Geologically recent examples of those features, from roughly 20,000 years ago, are spread across almost one-third of the modern land surface—in places such as Wisconsin, and northern Pennsylvania, and Yosemite, and Glacier, and many others, the mark of the ice is unmistakable. The 10,000 lakes of Minnesota, the Great Lakes, the gentle moraines of Illinois, and many more features reveal a landscape that is glacially dominated. Such features in Europe first motivated the hypothesis that ice ages have occurred.
This ice-age hypothesis makes many predictions, which allow testing. In times before modern geology, the glacial deposits were called “drift” because they were thought to have drifted into place in icebergs during Noah’s flood. Other people have suggested that the deposits were splashed into position by a giant meteorite that hit Hudson Bay, and still other hypotheses have been advanced. But, the ice-age hypothesis makes predictions that differ from the Noah’s-flood hypothesis or the meteorite hypothesis in many ways. (The biggest difference is that icebergs and meteorites simply do not make features that even vaguely resemble those actually observed, but let’s look at some other differences.)
If huge ice really existed, its great weight must have pushed down the land beneath—recall that the deep rocks are hot and soft, with a “water-bed” cover of stiffer rocks on top. If the ice age peaked only about 20,000 years ago, the slow flow of the soft, deep rocks should mean that the land would still be rising after the melting of the ice, while land around the former ice would be sinking as the soft, deep rocks return to their pre-ice-age positions. The global-flood hypothesis and the meteorite hypothesis do not predict such a bulls-eye pattern of rising and sinking centered on the regions with features known to be made by glaciers today—the flood would have spread evenly across the land, and so would not have concentrated its weight in one place, and the sudden blast of the meteorite would not have left its weight long enough to push the slow-flowing deep rocks far. Measurements by GPS and other techniques show just the pattern expected from the ice-age hypothesis, a pattern that was not predicted and cannot be explained by the other hypotheses.
The water for huge ice sheets would have been supplied by evaporation from the oceans, with the water getting stuck in the ice rather than returning rapidly to the sea in streams. Hence, if ice ages occurred recently, there should be evidence of lower sea levels at the time the ice was big. No such prediction comes from the meteorite or big-flood hypotheses (the meteorite might have made a wave but otherwise would not have affected sea level; the big flood would have raised sea level). Again, the ice-age prediction is borne out by the evidence, and the predictions of the other hypotheses are wrong. For example, some corals grow only in shallow waters where there is much sunlight. Dead samples of such corals from about 20,000 years ago can be found where they grew, down the sides of islands and now under more than 300 feet of ocean water. Other evidence also points to lower sea level in the recent past—the Chesapeake Bay, for example, is a river valley that was drowned by rising waters.
So, much evidence shows that ice ages occurred. Piled tills separated by soils demonstrate that the ice has come and gone many times. But how many times? On land, the glacial record is somewhat confused—often, an advancing glacier will erode the evidence of a previous one. A pile of four tills separated by soils may record four advances, or forty, with some of the record having been eroded away. In many places in the deep oceans, sediment has been piling up without erosion for millions of years. If there were a marker of glaciation in the marine sediments, we could tell how many glaciations have occurred. If there were a way to date these sediments, we could tell when the glaciation happened. Fortunately, we can identify glaciations using shells in marine sediments, and we can date them. Identifying glaciers from shells is covered here, and learning the age of the shells is coming in the next few units.
Water in the oceans is not all the same—roughly one molecule in 500 has an extra neutron or two in one or more of the oxygen or hydrogen atoms. Such “heavy” water is still water, but weighs a little extra. (If you don’t remember isotopes, go back and have a quick look at the introduction to chemistry near the start of the course.) Not surprisingly, light molecules evaporate more easily than heavy molecules. Water vapor, rain, and snow thus are slightly “lighter” than the ocean; that is, the ratio of light water molecules to heavy ones is larger in vapor, rain, and snow than in the ocean from which the vapor, rain, and snow came.
When sea level drops during an ice age - as water vapor is changed to snow and then to ice sheets, the oceans have lost a lot of water. More light water than heavy water has been lost from the oceans, so the oceans are left a little bit isotopically heavier than normal. When ice melts, that light water from the ice sheets is returned to the ocean and makes it lighter.
These changes are very small. If we round off the numbers a little, we can say that in the modern ocean, 1 of each 500 water molecules is heavy, which is the same as saying that 1000 of each 500,000 water molecules are heavy. When the ice sheets were big, had you weighed a whole lot of molecules in the ocean, you would have found that about 1001 of each 500,000 water molecules were heavy. This is a tiny change, the water was still water, but sophisticated modern instruments are so good that such a change is very easy to measure. (And yes, the instruments actually measure the weight of waaaaaaay more than 500,000 molecules, to obtain good statistics.)
Many plants and animals that grow in the ocean build shells of calcium carbonate (the stuff of limestone) or of silica, both of which contain oxygen. These shells record the isotopic composition of the water in which they grow because the oxygen in the carbonate or silica is obtained from the water. Critters growing during big-ice times grow shells in which the oxygen is isotopically slightly heavy, and critters growing during small-ice times grow isotopically light shells. When the critters die, their shells pile up in layers on the sea floor with the youngest ones on top. A core collected from these sediments is a history of the ice volume on Earth. Just date the core, pull out the shells, analyze them isotopically, and there is the answer. With enough care, knowledge, and instrumentation, dedicated workers can obtain consistent, reproducible data that tell a wonderful, clear story. (There are a few additional details, but the main story is this simple.)
Over the most recent 800,000 years, ice has generally grown for about 90,000 years, shrunk for 10,000 years, grown for 90,000 years, shrunk for 10,000 years, etc. Superimposed on this are smaller wiggles, with a spacing of about 19,000 years and 41,000 years.
More remarkable, these cycles were predicted, and not until decades after the prediction did technology become good enough to test the prediction and show that it worked. During the 1920s and 1930s, a Serbian mathematician named Milutin Milankovitch calculated how the sunshine received at different places and seasons on the Earth has changed over long times. As the sun, moon, Jupiter and other planets tug on the Earth, the orbit changes a bit. Earth wobbles with a 19,000-year periodicity, the north pole tilts a little more and then a little less with a 41,000-year periodicity, and the orbit changes from more-nearly round to more squashed or elliptical and back with a 100,000-year periodicity. With modern computers, these changes are relatively easy to calculate for many millions of years; for Milankovitch, the calculation was the labor of a lifetime. (He did it very well, though, even correctly noting that the 19,000-year periodicity really goes from 19,000 to 23,000 years and back, a pattern that is indeed observed in the data testing his prediction!)
These orbital wiggles have little effect on the total sunshine received by the planet, but they do move the sunshine from north to south, poles to equator, or summer to winter in various ways. For example, today the northern hemisphere is farther from the sun in northern summer than in northern winter. (Remember that summer is controlled by the tilt of the planet’s spin axis relative to the plane in which the planet orbits, not by the distance from the sun!) In the few millennia centered on 9000 years ago, the northern hemisphere had slightly warmer summers and cooler winters than recently, because the Earth was closer to the sun during northern summers and farther from the sun during northern winters than today. (Meanwhile, the south had slightly cooler summers and warmer winters than recently, because the Earth was closer to the sun during southern winters and farther from the sun during southern summers than today.) The intense summer sunshine in the north 9000 years ago made mountain glaciers smaller then. As the summer sunshine decreased in the north, those glaciers expanded slowly for several thousand years, culminating in the Little Ice Age of the 1600s to 1800s; strong melting of glaciers since then is probably mostly the result of human-caused warming. (We will discuss this later in the course.)
Summer in the northern hemisphere appears to be key to controlling ice ages, probably because the northern hemisphere is mostly land and can grow big ice sheets, but the southern hemisphere is mostly water, already has ice on Antarctica, and so can’t change its land ice much more. In the north, even during warm winters the highlands around Hudson Bay are cold enough to have snow rather than rain. Survival of this snow requires cool, short summers. As summers have cooled around Hudson Bay, ice has grown; as summers have warmed, ice has melted. The way the various cycles interacted led to larger or smaller changes, and thus to the ice ages we know.
You may guess that this is slightly oversimplified so far. For example, during times when Canada has received reduced summer sunshine, allowing ice to grow, the southern hemisphere or the tropics often were receiving extra sunshine, yet they cooled during many of those times. How Canada told the glaciers of Patagonia and Antarctica to grow was for a long time a great puzzle. The answer involves the global warming from atmospheric carbon dioxide. The growth and shrinkage of the vast ice sheets, the changes in sea level, and other changes had the effect of shifting some carbon dioxide (CO2) from the air into the deep ocean during ice ages, and bringing the CO2 back out to the air during warm times. The orbits affected the ice, which affected currents and sea level and plants and other things, which affected CO2. But, as we will discuss later in the semester, CO2 in the air tends to warm the Earth's surface no matter how that CO2 got into the air. And, the changing CO2 explains why, when the ice was growing, places getting more sunshine still got colder, and why, when the ice was shrinking, places getting less sun still got warmer.
Climate records show many other types of changes. Very large, rapid changes have been caused by sudden surges of ice sheets, and by jumps in the way the ocean circulates. We do not understand these faster changes well enough to know whether they could happen again, although we're cautiously optimistic that we won't have any for a while. Naturally, the Earth’s orbit right now is in an intermediate state, and we should be looking forward to another 20,000 years or more with little change before we begin the slide into a new ice age. (See the Enrichment for a little more on this.) However, humans almost certainly are now more important to the climate than are such slow changes, as we will see later.
Meanwhile, what of things back in central Pennsylvania and in the many other places that were not quite reached by the ice-age ice? As you might imagine, with the world cold enough to grow ice to cover New York, Chicago, Minneapolis, and Seattle, as well as much of Europe, the climate was colder everywhere then than it has been more recently.
If you climb the ridges of central Pennsylvania, perhaps up in the Seven Mountains just southeast of State College (go up Bear Meadows Road past the ski area, for a start), you may notice several interesting things geologically. Beneath the hemlocks and rhododendron, the soils and streams and hillslopes have more in common with the high meadows of Trail Ridge Road in Rocky Mountain National Park, or with the coast of Greenland, than they do with the modern climate of State College. Trail Ridge Road crosses tundra, where small, hardy plants grow atop permafrost. Although the uppermost soil along Trail Ridge Road thaws during the brief summers, and the deep Earth is thawed by the heat of the Earth, the materials between are frozen year-round in permanent frost. (These areas are also called “periglacial,” because they may occur around the glacier, or on its perimeter.)
Consider the following features of the Seven Mountains.
The conclusion is nearly inescapable—Trail Ridge Road in Rocky Mountain today is an excellent picture of what the Seven Mountains looked like during the ice age. Permafrost is common across much of northern Canada and Siberia and around the coast of Greenland, and in high-altitude regions. Permafrost poses grave problems for construction—the heat of a building can melt permafrost beneath, causing uneven settling that breaks the building. Permafrost also records the climate changes that have come to central Pennsylvania and other regions.
Perhaps more meaningful than the conclusion of past Pennsylvania permafrost is the underlying reasoning. Some people today, including important government officials, claim that “historical” geology is not really science, does not use the scientific method, does not produce scientific results, and so should be ignored. (Many commentators have noted that these government officials typically dislike policies that might be motivated by the science, and are probably really more interested in complaining about the policies than the research. But, let's look at the research.)
Consider how the process works. Go up to Bear Meadows, start up toward the ridge above, and look around carefully. You see that big rocks are present, of a type that is quite different from the bedrock directly beneath.
Many hypotheses are possible to explain this observation—space aliens dropped the big rocks; or bulldozers pushed the rocks into place; or, the rocks came screaming down from uphill in a giant landslide; or, they came creeping down slowly; or, … you could think of others. Each hypothesis leads to predictions. If a bulldozer pushed the big rocks in, we should find the bulldozer tracks, and we should be able to trace back in historical records to who was driving the bulldozer, and why. The first settlers, who arrived before bulldozers were invented, should have found hillslopes free of big rocks. If the big rocks came from uphill, we should be able to find a source of such rocks uphill. Landslides start with big falls or slumps from particular places, so a landslide should have a big scar at its head, whereas creep slowly collects rocks as they are worked loose and carries them along, lining them up as they go.
So, you look for evidence that supports or refutes each of your hypotheses. The early settlers complained about the big rocks, old cabins are built on the big rocks, so the bulldozer hypothesis won’t work. There is no evidence for a landslide scar anywhere, despite evidence for lots of different “stripes” of big rocks extending downhill from a ridgetop source where bedrock of the same type as the big rocks sticks out. You quickly come to the realization that the rocks look like a soil-creep deposit extending down the ridge crests; the predictions from each of your other hypotheses fail, but each of the predictions from the soil-creep hypothesis is supported by additional data that you collect for testing purposes.
Then, you note that the material is not now creeping—roads and trails are not being slowly buried by big rocks today, the trees are not knocked over, etc. Tree roots hold many of the rocks in place and prevent motion. So you look for a time in the past when tree roots were not holding the rocks in place. You collect more information—the big rocks are on top of smaller rocks and soil, not on the bottom, the big rocks are often standing on edge, the rocks show patterning of coarse and fine, etc. Other geologists are scanning the whole planet, laboring over centuries, and among the many things these geologists report are the conditions of creeping hillslopes in the tropics, the deserts, the temperate zones, and the poles. You talk to other geologists, devote a decade of your life to careful study, and eventually learn that the things you see on the slopes of the Seven Mountains resemble features of permafrost, and not features of any other modern setting.
But, if you are correct and these are permafrost features, there should be other evidence of cold conditions in the past, at the time that these features were active. So you take a core in the bog, and find that the bog started in a very cold time (the deepest pollen you find is from plants that today are found only on the tundra), and the bog seems to be dammed by one of the soil-flow lobes, linking the soil-flow lobes to the time of the tundra cold. (It is true that no one has used a backhoe to take the dam apart to look for a space-alien-constructed dilithium-crystal foundation, so maybe the space-alien hypothesis has not been completely falsified and the science could be improved; but, there comes a point of diminishing returns….)
Next, you ask whether this makes sense. You have tentatively concluded that the hillslopes of Pennsylvania record cold conditions at a particular time in the past. Is there a reason why cold should have been here at that time? Well, just to the north, glaciers were pushing up moraines at the same time. And astronomers making orbital calculations find that the high northern latitudes were receiving about 10% less sunshine than today during that glacial age. Climate modelers who test whether such a drop in sunshine would be sufficient to grow glaciers and make conditions very cold find that cold indeed makes sense, especially when the modelers include the effects of the drop in atmospheric CO2 levels that was triggered by the change in sunshine and that is recorded in ice-core bubbles from the time.
Now, a modern geologist who tells the “story” of this chapter—Pennsylvania hikers twist their ankles on permafrost deposits—actually has a lot more evidence than the little sketch provided here. The libraries of information collected by centuries of Earth scientists are woven together in a sophisticated, carefully tested, highly reliable whole. This great tapestry of knowledge still has gaps, dropped stitches and moth-bitten places, and the ragged edge where knowledge runs out into the unknown that so excites us as scientists. But the science of the tapestry is well-woven and exceptionally strong. We can only hope that the misguided attacks on this science come from ignorance and not malice, because ignorance is more easily changed.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 7.
TECH NOTE
Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
This week, we feature two GeoClips, both featuring Dr. Alley. As before, we hope you enjoy these, and find them to be useful complements to the readings, class notes, and slide shows of Unit 7.
The Bear Meadows National Natural Landmark, just over the ridge from Penn State’s University Park campus, was recognized by the National Park Service in 1966 as a site that “possesses exceptional value as an illustration of the nation’s natural heritage.” Although many guide books somehow have decided that Bear Meadows is 10,000 years old, the Meadows are clearly much older, having formed during the last ice age. Here, take a walk just above the Meadows, and learn why Pennsylvania hikers, like those in the high country of the Rocky Mountains, are wise to wear sturdy shoes. Then, see what this has to do with the Formation of the Meadows—they really are related.
Here are some optional animations you might also want to explore! (No, these won't be on the quiz!)
Glacier Physics
(An extensive collection of animations on this subject)
Glacial Landforms Resulting from Erosion and Deposition
(An extensive collection of animations on this subject)
Examples of Deglaciation
(An extensive collection of animations on this subject)
The Unit 7 lecture features Dr. Sridhar Anandakrishnan and is 64 minutes long.
Check out the Unit 7 Presentation used in the online lecture.
The most recent ice age may have ended, but there is still a lot of ice remaining in Antarctica and Greenland. Here’s a little more about the Antarctic ice, who lives around it, how it behaves, and why we might care. We’ll talk about the warming effect of rising CO2 in Unit 12; for now, just note that we are raising CO2 in the air, and it does have a warming influence, based on fundamental physics discovered in part by the Air Force for military applications.
Many types of glaciers exist, with fairly loose or imprecise definitions. An ice sheet is a continent-scale mass of ice that spreads in all directions. An ice cap or ice dome is a smaller version of an ice sheet, sitting on a mountain top or high plateau, and also spreading in all directions (or at least in several directions). For pretty glaciers flowing down from the mountains, different people may use different terms: mountain glacier (it flows down a mountain), valley glacier (it flows down a valley on the side of a mountain), and plain old glacier. An outlet glacier drains an ice sheet or ice cap between rock walls, and an ice stream is a fast-moving “jet” of ice within an ice sheet or ice cap flowing between slower-moving regions of ice. But if an ice sheet is drained by a fast flow with ice on one side and rock on the other, is that fast flow in an ice stream or an outlet glacier? Classifications such as this help us talk about things, but are not precise.
Dr. Alley has spent months of his life living on the great ice sheets of Greenland and Antarctica. (And Dr. Anandakrishnan, who has worked so hard on this course, has spent a lot more time on the ice sheets than Dr. Alley has!) Eating and sleeping and working at -30º, it is hard to think of ice as being a hot material, but that is exactly what it is.
Recall that heat is the vibration of atoms or molecules in a material, and that in most solids including ice, the atoms or molecules are arranged in regular, repeating patterns. Melting of ice occurs when the typical molecule vibrates fast and hard enough to break free from the bonds that tie it to its neighbors and escape from that regular arrangement. When a material is almost hot enough to melt, the atoms are vibrating almost hard enough to break free from their neighbors and move around, so it is relatively easy with a little extra push to move a few molecules at a time past their neighbors. The gravitational stresses caused by the surface slope of a glacier supply that little extra push, and the ice deforms (primarily by dislocation glide, for those of you with materials-science backgrounds). When a material is not nearly hot enough to melt, the molecules are not even close to vibrating hard enough to break free from their neighbors, a whole lot of extra push is required to move molecules, and moving even a few molecules at a time is very difficult. The material then deforms elastically, or it breaks, but it does not creep and deform permanently in the way that a glacier does.
Most people measure temperature on a scale that gives “nice” numbers (something between 0 and 100) for typical daytime temperatures, so that talking about the temperature is easy for us. But, there are other temperature scales that make more sense in physics. If you slow the vibrations of molecules by cooling them, you can imagine that there must be some temperature at which vibration stops because all the heat has been removed. We call that temperature “absolute zero” or just zero in an absolute temperature scale. (Yes, in a quantum world, the Heisenberg uncertainty principle means that the last tiny bit of vibration can’t really be removed, but absolute zero comes darn close, so live with it.) If we set the zero on our temperature scale to this “absolute zero,” and then use degrees that have the same size as in the commonly used Celsius or Centigrade scale, we get the Kelvin scale. Ice melts at 273ºK and water boils at 373ºK; there are 100 degrees between melting and boiling in Kelvin, just as in Celsius. (The Rankine scale uses Fahrenheit-sized degrees and absolute zero as zero, with ice melting at 460ºR and water boiling at 640ºR, but almost nobody uses Rankine any more, so you are welcome to forget you ever heard about it. Or, you can practice sniffing disdainfully.)
As a general rule, little or no permanent deformation (creep) occurs when the temperature (in Kelvin or Rankine!) is less than about half the melting temperature, and creep occurs rather easily when temperatures exceed about three-quarters of the melting temperature. The coldest mean-annual temperature on Earth today is about eight-tenths of the melting temperature of ice (that is 217ºK, which is also -56ºC or -69ºF, in case you still like old-fashioned thermometers). Most ice is as close to melting as is red-hot or even white-hot iron being worked by a blacksmith. This is why glaciers usually flow rather than breaking—although breaking is still possible where deformation is very fast and where the pressure is very low, producing crevasses. So, the next time you are tempted to “pull down your pants and slide on the ice,” remember that ice is a “hot” material, even if you may not look very hot when you’re through. (We recommend that you don't "pull down your pants and slide on the ice," for many reasons related to public decency, avoidance of frostbite, and not sliding over a cliff or falling into a lake.)
As noted in the chapter, glaciers that are frozen to their beds don’t erode much, but if the basal ice is at the melting point, glaciers can erode very rapidly. Such thawed-bed glaciers have three ways to erode: plucking, abrasion, and subglacial streams.
Ice is an unusual material—higher pressure lowers its melting point rather than raising it (because ice becomes smaller when it melts; the tinker-toy-structure of ice has much open space, and squeezing ice tends to force it to become denser water, whereas most materials contract as they freeze so that higher pressure favors the solid). If a glacier is sliding across a bump in a bed, ice will tend to melt on the upglacier side of the bump where the pressure is higher. The meltwater will flow around the bump to the downglacier side, where the lower pressure will allow the water to refreeze. The heat given up by the refreezing will be conducted back through the bump, to allow more melting. But, you may remember that melting and freezing can open cracks in a rock. So, a glacier sliding over its bed can work rocks loose, and then freeze those rocks onto its base, in a process known as plucking. (When water spreads over the bed of a glacier in the spring as melting on the surface starts to feed water downward, the friction with rock that holds the ice back becomes concentrated on smaller regions of the bed not lubricated by the water, and this stress concentration breaks rocks, causing most plucking.)
Once glacier ice contains rocks at the bottom, it is like sandpaper—it drags those rocks over other rocks, scratching and polishing and knocking loose smaller rocks. This process is called abrasion. If you examine rocks on the walls of Yosemite, many still retain a polished appearance with parallel scratches or striations, showing where abrasion was active. Bumps are polished on one side—the upglacier side—but may be rough and jagged on the downglacier side where rocks were plucked off of them.
Melting of glaciers can produce a lot of water. The toe of a fast-melting glacier may supply more water to streams than does a similar-sized region in the rainiest place on Earth. The glacier acts to collect snowfall from a big area and take the snow to melt in a much smaller area, and trees and grass do not grow on glaciers to use the melt but they do grow on ground to use rainfall. Glacier melt usually flows down holes in the glacier, called moulins, that often form at the bottoms of crevasses. (Some brave or foolhardy people like to go caving in moulins after they drain during the winter.) The moulins eventually reach the glacier bed, where they feed large, steep, fast-moving streams. These erode in the same ways as streams outside of glaciers. Glaciers with much melt water usually cause erosion to be faster than in nonglaciated regions. Fluctuations in water pressure, as moulins fill with water during daytime melting and drain as melting slows at night, contribute to cracking rocks for plucking.
In the text, we noted that the history of ice ages generally has involved 90,000 years of cooling, followed by 10,000 years of warming, then repeat. The rate of cooling initially is slow, and some people prefer to refer to 10,000 years of warmth followed by cooling. The northern hemisphere has been in the not-much-change/slight-cooling phase for almost 10,000 years already, and you might expect that we are ready to drop into the next ice age. Some people have suggested that humans have already headed off that ice age, or that global warming is a good thing because it will head off the ice age.
The 100,000-year pacing of a 90,000-year/10,000-year world is linked to interaction of the different orbital cycles, but the 100,000-year cycle in the out-of-roundness of the orbit is important. The orbit goes from nearly round to more squashed and back in about 100,000 years. But, there exists a slower modulation that takes about 400,000 years. The orbit goes nearly round, a little squashed, nearly round, more squashed, nearly round, even more squashed, nearly round, not as squashed, nearly round, barely squashed, repeat, with the nearly-rounds spaced 100,000 years apart. We are in the barely-squashed part now, and the last time that the orbit was in the barely-squashed mode, the warm time of the ice-age cycle lasted 30,000 years rather than 10,000 years. Climate models have confirmed that this should be our natural future; another 20,000 years of warmth (or maybe 40,000 years) before the next ice age starts. However, human burning of fossil fuels may extend the warmth beyond the next 20,000-40,000 years.
Also, note that the 19,000-year cycle noted in the text is an oversimplification. There is instead a “quasi” periodicity ranging from 19,000 to 23,000 years, as we mentioned briefly, and this was calculated by Milankovitch and is observed in the data collected to test Milankovitch's calculations, beautifully confirming his predictions.
During at least one old glaciation (probably over 1 million years ago), ice dammed the West Branch of the Susquehanna River and formed a lake in the Lock Haven area of Pennsylvania. If that lake filled to the next lowest bedrock outlet (into the Juniata River along the Bald Eagle Valley at Dix), then the water would have lapped at the steps of Old Main on Penn State’s University Park campus. There is no evidence of such a large lake, and before the lake filled all the way, it probably drained through failure of the ice dam, but we’re not sure. With ice so close, however, the State College area was cold during the ice ages.
In the main text, you learned how the changes in ice volume control the isotopic composition of water in the ocean, and how we can reconstruct the ice-age cycle from the history of shell isotopic compositions in a sediment core because the shells record the water isotopic composition. As usual, things are a bit more complicated than that. Shell isotopic composition also is affected by temperature. Bigger ice gives heavier isotopic ratios in shells, and colder temperatures also give heavier isotopic ratios in shells. (At high temperature, both heavy and light atoms have plenty of energy to get up and go; at low temperatures, the heavy ones tend to get stuck in shells while the light ones can jump out.) Because both colder and bigger ice favor isotopically heavier shells, it is hard to tell how much of the signal in a shell is from temperature or from ice volume.
One way around this is to go to a place that is really cold today; the water was above freezing during the ice age (there were shells living in it…), so there the signal must be primarily one of ice volume. Other approaches include finding additional paleo-thermometers, such as estimating the temperature from the species living in a place and leaving their shells, or using changes in other “contaminant” ratios in shells that depend on temperature. Yet another way is that there is water in spaces in mud, and the water in some sediments is from the ice age, so just measure the isotopic composition of that water.
The result of this is that isotopic ratios did change because there was much more ice during the ice age than today, and because most places were much colder during the ice age than today.
You have reached the end of Unit 7! Double-check the list of requirements on the Unit 7 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Exercise #3 is due and Exercise #4 opens this week. See Course Calendar for specific dates and times.
Following are some supplementary materials for Unit 7. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
The cod has been an extremely valuable resource for several centuries in Massachusetts. Its extensive use as a food dates back to the earliest period of European settlement in coastal New England. In colonial times, it was deemed so important that in 1693 the General Court of the Massachusetts Bay Colony ordered that farmers could no longer use cod as fertilizer. This action was one of the first recorded attempts at natural resource conservation and management on this continent.
Although one of the earliest fisheries resources to be broadly utilized after European settlement in New England, cod populations along the US coast proved to be very resilient. Cod apparently withstood more than 3 centuries of harvest without displaying major, long-term regulations in abundance. However, mid-twentieth century advances in fishing technology and the introduction into the northwest Atlantic of distant-water foreign fishing fleets during the late 1950's led to a period of reduced abundance and major annual fluctuations in population size. During the mid-1980s commercial vessels captured mostly 3 to 5 year old fish, indicating that few larger, older individuals remain along the North American coast.
—From the Massachusetts Division of Marine Fisheries
Will Cape Cod be there in the future? Looking out a few millennia, the answer is probably "no." Beaches, like rivers, are interplay of water and sediment. Sand is supplied, and sand is lost. Interrupt this process, and the coast must change. And in the distant future, Cape Cod is likely to be the new Georges Bank, an underwater place that could be home to great masses of fish, if we leave enough fish to get along and get it on.
You will have one week to complete Unit 8. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 8, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 8 Instructional Materials. |
Submit Exercise #4: Which Way is Up? | Yes, this is the fourth of 6 Exercises and is worth 5% of your total grade. Please see the course calendar for specific open and close dates and times. |
Begin Exercise #5: Puzzling Out Relative Time | Yes, this is the fifth of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 8 "RockOn" quiz | Yes, this is the eighth of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #9" survey | Yes, this is the ninth of 12 weekly surveys and is worth 1% of your total grade. |
Please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations with any questions.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Cape Cod girls don't have any combs
Look away, look away
They comb their hair with codfish bones
We're bound for Australia.
Cape Cod boys don't have any sleds
Look away, look away,
They slide down dunes on codfish heads
We're bound for Australia.
Cape Cod brides don't have any veils
Look away, look away
They shield their eyes with codfish tails
We're bound for Australia.
—Traditional sea shanty
The author, Dr. Alley, is a lucky fellow. Through a fortuitous sequence of events, which involved marrying the right woman who had the right grandfather who was related to some people who have roots in the right place and worked to preserve those roots, our family has been able to occupy a room in the smallest of three houses in a “compound” in Eastham on Cape Cod for a couple of weeks during many summers. The land now belongs to the National Park Service as part of the Cape Cod National Seashore, but the houses were “grandfathered” and remain in the extended family. The bicycle trail to the Coast Guard Beach crosses the driveway, the Salt Pond Visitor Center is just around the corner, and the boat houses face Salt Pond Creek that opens on the Nauset Marsh.
Cape Cod is a glacial moraine. The part that attaches to the mainland marks the end of a lobe of the Canadian ice sheet. The “forearm” where the Cape points north is an interlobate moraine—while one glacier lobe filled Cape Cod Bay between the Cape and Boston, a second lobe lay farther east in what is now the Atlantic Ocean, and built a moraine that is now the fertile fishing grounds of the Grand Banks. (You will see this when you reach the geologic map later in this section.) The forearm of the Cape was deposited mostly by meltwater rivers flowing off these two ice lobes into the narrow space between them (more till is found along the “upper” arm of the Cape where it first projects out from the “mainland”).
This forearm part of the Cape thus is an outwash plain, and most of the numerous freshwater ponds of the Cape began their lives as ice blocks buried in outwash gravels. Melting of the ice later allowed collapse of the gravels on top to form these kettle ponds. Soon after it formed, the Cape had more ponds than we see today, but many have been filled with logs, sticks, leaves, peat, and other organic material. The logs and other materials in these former kettle ponds can often be seen weathering out of the bluffs along the Atlantic beach, including along the Coast Guard Beach that is so easily reached from the Salt Pond Visitor Center of the National Seashore. Radiocarbon dates on such deposits show that the ice was retreating from the Cape Cod region by roughly 15,000 years ago.
The sandy soils of the Cape, and its numerous kettle ponds, are home to cranberries, blueberries, and wild orchids. Beach plums produce their small but sweet fruit late each summer, and blackberries and dewberries vine across old dunes. Deer and grouse still inhabit the uplands, and turkey are multiplying rapidly as oaks spread across regions that were logged by humans but are being allowed to regrow.
Much of the interest at the Cape focuses on the sea. Nauset Marsh, for example, is miles long and a good chunk of a mile wide. The marsh is protected from the open ocean by the “outer beach,” a long sand bar that is split by one inlet or occasionally more inlets (depending on what year you visit). Behind the outer beach, the marsh is a tidewater environment, filled and then largely emptied by the flow through the inlet. Deep channels are home to crabs, scallops, starfish and striped bass. Between the channels, great flats of marsh grass flood during high tides and dry as the tide falls. Legions of herons and egrets stride the grass, and myriad shorebirds roam the beaches and mud flats. Osprey survey the marsh from high nests on perches constructed by the Park Service.
One year, a hurricane to the south had spun unusually warm waters and great swells to the Cape, and the marsh filled with ctenophores. Also called comb jellies, and looking something like jellyfish, these creatures appear opalescent in the sun because the light is broken on the cilia or hairs that the creatures beat to move themselves about. At night, this particular type (Leidy’s comb jelly) is bioluminescent. Imagine, if you can, kayaking into the marsh after dark, with golf-ball-sized fireflies of the ocean glowing and swirling with every wave. Another year, phosphorescent dinoflagellates were blown into the marsh, and the glow was everywhere rather than just in the ctenophores, with every drip from the kayak paddle lighting up, and fiery baitfish skittering across the surface, pursued by the glowing bulk of predatory stripers. Indeed, a wonderful place.
The Cape is changing rapidly, however. The large Nauset Marsh, between Orleans and Eastham, has been narrowing as the outer beach moves steadily into the marsh. Two of the Cape’s lighthouses were moved during the summer of 1996, barely in time to avoid collapsing over the bluffs into the sea as the bluffs have been eroded away.
At the ends of the Cape, to the north and south, new land is being formed as sand is deposited (yellow on the map below). But, more than twice as much land is lost each year as is formed, and the Cape is slowly but surely disappearing. With the next ice age to rebuild the Cape still far in the future, the Cape appears destined to become islands and then an undersea bank over the next few thousand years. The long-term retreat rate of the Outer Beach in Eastham has been about 3 feet per year over more than a century, with fluctuations but perhaps with a little recent acceleration.
There are many types of coasts. North of the Cape at Acadia National Park, Maine, strong igneous and metamorphic rocks make sea cliffs. South, in the Virgin Islands National Park, are coral reefs. Built from the skeletons of millions of tiny animals, the reefs rise from the sea bottom and flourish in shallow, clear, sunlit, oxygenated waters far from sediment that would bury and choke them. Around into Louisiana, you remember the Delta National Wildlife Refuge, and the miles of waterfowl-filled wetlands on mud delivered by the Mississippi River. At Cape Cod, we find barrier beaches or barrier islands offshore of marshes, or we find sandy beaches with the bulk of the Cape just behind.
The type of coast depends on the supply of sediment to it, the wave energy, the tidal range, the type of rocks, and many other factors. Here, we will concentrate on the Cape Cod sandy beaches.
Watch a wave on the Cape, or any other sandy beach, and you will see that the wave moves a lot of sand. Dig a hole just above the water level during a rising tide, and within a few minutes the hole will be filled with wave-carried sand. Go to the beach during a big storm and you will see immense amounts of sand moved. Hundred-foot-high bluffs may be eroded back several feet, or layers of sand many feet thick may be added to the beach or eroded from it in hours. A movie at the Salt Pond Visitors Center includes a series of photos taken of one section of beach before and after a string of storms one winter. The summer beach is an unbroken expanse of sand, but boulders many feet across are buried in it, and were completely covered and uncovered several times during that one winter as the sand was moved on and off the beach.
The energy for all of this comes from the wind driving waves, and to a lesser extent from the tides. Most of the sand transport is simply onshore and offshore with each wave—the sand is carried in as a wave comes in, and out as the wave goes out. Most transport is to and from the beach, rather than along the beach, because most waves turn so that their crests are almost parallel to the beach, and their water motion is almost directly towards and away from the beach. The turning happens because waves go slower in shallower water. If a wave approaches a beach at an angle, the first part to near the beach will slow down, allowing the rest of the wave still in deep water to nearly catch up, as shown in the diagram to the right.
Every wave moves sand up and down the beach. On even rather quiet days, if you sit down on a Cape Cod beach in shallow water, you soon will find that sand is piling in places around you, and being eroded in other places, and that you have sand in your swim suit, and possibly even in your hair. This in-and-out movement of the sand with every wave dominates the sand transport, allows for very efficient sorting of the sand by size, and serves to knock the sharp edges off sand grains, sea shells, old soda bottles, and other material on the beach.
Look out to sea during a winter storm, and you’ll see high-energy breakers coming at you. The white caps of the waves rise high, curl over and crash down, so that some of the water arrives on the beach after coming in through the air rather than washing along the sand. But the water then rushes back toward the ocean along the sand, carrying some sand seaward. There thus often is a little net movement of sand from the beach into slightly deeper water during storms, which are more common during winter than during summer. This transport may remove enough sand to lower the beach many feet or tens of feet during the winter, exposing buried boulders. The waves of summer are on average lower in energy, and don’t break and travel through the air as much (occasional hurricanes change this story, but most of the time the story is fairly accurate). The surge of summer waves up the beach is slightly faster than the return flow down the beach, and may carry a tiny bit more sand up than back; the net effect is to bring sand from just offshore back to the beach, burying any beach boulders that were exposed during the winter.
But, the incoming waves do not turn quite enough to come straight in, as shown in the diagram—they still have some angle. If you ride the waves in, swim out, ride in, swim out, ride in... after a while you will find you have drifted down the beach away from the lifeguard—even though you were mostly going toward the beach and back out to sea, the waves also were pushing you sideways. In such a situation, we say that you are experiencing longshore drift—you, and the water, and sand, are moving along the shore. Eventually, when the water and sand reach the end of the Cape (at the Provincelands to the north, or Monomoy to the south), some of the sand builds a spit or extension of land, but some of the sand is dumped off into deeper water beyond the reach of waves. This sand is then lost from the Cape, and the Cape has gotten a little smaller. Most references say that the great beach facing the Atlantic is retreating at about 3 feet (1 m) per year, although it may have been a little faster recently, and the panicked light-house rescues were required because retreat for a few years was much faster. We’ll look at these issues, and what might be done, after visiting Acadia in the next chapter.
The subduction and collision with Europe during the closing of the proto-Atlantic made great granite bodies draped in metamorphosed sediments, and we find these exposed along the Maine coast. Much later, the glaciers scoured those rocks clean, leaving the bald mountain tops of what is now Acadia National Park, staring out at the storm-tossed north Atlantic across Somes Sound, the only fjord on the east coast of the U.S.
The Wabanaki tribe of Native Americans probably reversed the modern tourist pattern, summering inland and then moving to the relatively more moderate coast of Acadia in the winter. Nasty winter storms do run up the coast, but the temperatures plunge far lower inland than they do on the coast. Thick piles of discarded shells of nutritious sea creatures dating back 6000 years attest to the importance of the sea to these early people.
On September 5, 1604, the French explorer Samuel de Champlain landed on Acadia’s island. Impressed by the bare, rocky, deserted appearance of the glacially scoured granite mountain peaks of the island, he named it Isles de Monts Desert.
French influence was important until the end of the French and Indian War, after which English and then U.S. activity came to dominate. In the mid-1800s, Mount Desert Island attracted the art world, and was featured in paintings by Frederic Church, Thomas Cole and others of the Hudson River School (for example, have a look at Twilight: Mount Desert Island, Maine by Frederic Edwin Church. The artists, in turn, attracted the “rusticators,” tourists who gradually were replaced by much wealthier tourists who built summer “cottages”—the Rockefellers, Fords, Vanderbilts, Carnegies, etc. These wealthy patrons in turn invested resources and political capital in preserving most of the island as a national park.
Today, over 4 million visitors per year flock to Mount Desert Island, with most of them getting out of the gift shops and into the national park. The visitors enjoy the history, the beauty, the rather chilly ocean waters (based on personal observation by the author, even on hot summer days, there is more sitting by the sea than swimming in it!), the outstanding network of paths for bicycling, superb kayaking on lakes and sea, and much more.
Change is the only constant on coasts. The Sea Grant Program at the Woods Hole Oceanographic Institution, on Cape Cod, reported that about 75% of the U.S. coastline is eroding, with only about 25% stable or advancing. For Massachusetts, 68% of the coastline is listed as eroding, 30% advancing, and a mere 2% stable.
Up in Maine, the rocky parts of Mt. Desert Island are among the few places that would be classified as “stable,” although very slow erosion is occurring as the sea pounds the granite headlands. But if you look further back in time, the size of the changes becomes evident. Glacial ice overran the highest peaks in the park during the ice age. Sea level was lowered 300-400 feet (100 m or a bit more) at that time to supply the water that grew the ice sheets, but the land of Mt. Desert Island was pushed down 600-700 feet (roughly 200 m) or even more by the weight of the ice.
Ice farther south began melting before the ice on Maine, so the sea began rising, but then loss of ice on Maine caused its rocks, including Mt. Desert Island, to begin rising faster than the sea. The rising of the pushed-down rocks is slow, and in fact is still going on today. Thus, as the ice retreated, the already much-raised sea first flooded in across broad regions of Maine and adjacent parts of the east coast. Beaches and sea caves formed, and deltas were deposited. Then, these coastal features were raised out of the ocean as the land rebounded. Such coastal features can be found today in Acadia to almost 300 feet (almost 100 m) above the modern sea level, and similar features occur all along coastal Maine. We include a picture of a similar delta from Greenland; the features in Maine are covered with trees and houses and roads and such, and although the features are quite easily identified by experienced geologists, the features are not as clear as those in Greenland to the beginning geologist.
Regions that were slightly beyond the reach of the ice-age glaciers were pushed up in a forebulge, when the hot, soft, deep rocks pushed out from beneath the sinking ice sheets bulged up the land just before the ice. In those forebulge regions, the land now is sinking, as the deep, hot rocks flow back to their starting point; where forebulge sinking has combined with rising sea level as the ice melted, the total sea level rise has been especially large. Far from the ice sheets, sea-level rise has been about what you would expect based on the amount of water returned to the ocean by the melting ice sheets (although the entire surface of the Earth was warped by shifting water from the oceans to the ice sheets and back, so the changes are all a bit more complicated than you might expect, just as a wine glass balanced anywhere on a cheap water bed will tip over if you sit anywhere on the bed).
By now, you may be getting the idea that what happens to a particular coast is fairly complex. If mountain-building is pushing the coast up, it rises; if mountain-building is pushing the coast down, it sinks. Where plates meet, when the edges are locked and building toward an earthquake, the motion may drag one side down and push the other side up; the earthquake that follows will suddenly reverse the offset—in the great Tohoku earthquake of Japan in 2011, parts of the Japanese coast moved as much as 8 feet toward North America, and offshore the largest motions of the sea floor were more than 150 feet horizontally and more than 20 feet vertically. Where cities are built on deltas, as at New Orleans, the compaction of the mud causes sinking. Much additional sinking is caused by pumping water or oil or gas out of the ground; as the fluids are removed, the ground compacts. This is happening a little on Cape Cod, and is quite dramatic in some places. Groundwater pumping may have contributed to problems at New Orleans, in Venice, and elsewhere. (Reinjection of fluids can partially offset this problem, and is being used in some places.)
Additional complexity comes from the issues of sediment supply. Beaches inevitably lose a little sediment to deep water, somewhat like losing socks behind a clothes dryer, because it is easy to drop something that falls way down there, and hard to get it back. Waves affect sediment below the surface, but only down to a depth that equals roughly half the distance between a wave’s crest and the next one. If sediment happens to slide or bounce deeper than the depth reached by the deepest-reaching waves, then that sediment cannot be brought back easily. (The sediment can go into a subduction zone, be taken down and scraped off, or else melted and erupted and loosened by weathering and landslided and transported in a river and then in longshore drift to the beach, but that takes a long time.) And, longshore drift moves a lot of sediment, but eventually the sediment often encounters an undersea canyon and drains down to deeper water below the reach of waves.
Thus, a “happy” beach requires a supply of sediment to balance the loss to deep water. Normally, that supply comes from the material delivered to deltas by rivers, and carried by longshore drift. But if there is not enough sediment coming in this way, the beach will narrow as it loses sediment to the deep ocean, and the waves will reach across the sand to erode the material behind, gaining sediment in this way.
In those cases when longshore drift does not supply enough sand to make a beach, and the waves are battering on the coast, if the rocks are really hard, like the granite of Acadia, then there won’t be a beach at all, with the tiny bit of material eroded from the granite quickly dumped into deep water. (There are a few small “pocket” beaches at Acadia in protected places, but most of the coast doesn’t have beaches, with granite sticking right out into the waves.) If the coast is sand and gravel left by the glaciers, as seen at Cape Cod, the waves cut the coast back and supply that material to the beaches.
In some places, dams on rivers have greatly reduced delivery of sediment, so the coasts are eroding. You may recall that the dams on the Elwha River, draining Olympic National Park, caused loss of beaches along the coast. At Cape Cod, there really aren’t any rivers to dam. The glaciers made a big pile of sediment in a place where rivers are not supplying much sediment to deltas, and so the Cape eventually will be lost to deep water.
So coasts change for many different reasons, and in many different ways. But, recently, most of the U.S. (and world) coasts have been retreating because sea level is rising, and that rise is accelerating.
The rate of rise has been about 2 mm/year (just under an inch per decade) over the last century, but seems to have sped up to about 3 mm/year (a bit over an inch per decade) recently. That isn’t much if you’re at the top of a cliff in Acadia, but if you are on a sandy beach that slopes very gradually, the inch of sea-level rise may cause many feet, or even a few tens of feet, of coastal retreat. That in turn means that a whole lot of houses and property can be lost in a single lifetime.
The ongoing sea-level rise is being caused primarily by the global average warming of the last century, which is being driven primarily by human activities (we’ll return to this later, but we have high scientific confidence that it is correct). Most of the world’s small glaciers have been melting this century, returning water to the oceans. Also, as the ocean itself warms, the water expands and takes up more room.
Two other possible causes of sea-level change are humans "mining" groundwater, and melting or faster flow of the polar ice sheets. We build dams on rivers, and the water that fills the reservoirs is taken out of the ocean and stored on land, causing sea-level fall. But, we “mine” groundwater by pumping it out of the ground faster than nature puts it back, in places such as Phoenix, Arizona where green yards and golf courses and fountains seem to “need” LOTS of water, and that water eventually reaches the ocean to raise sea level. Today, groundwater pumping is probably more important than dam building, contributing a little to sea-level rise.
The polar ice sheets contain a huge amount of water—if they melted, they would raise sea level nearly 250 feet (roughly 70-80 m). Philadelphia and the other great port cities of the world would become undersea hazards to shipping but really great places for fish to hide out, and the southern coast of Florida would be somewhere up in Georgia. We do not expect such a fate within at least the next few human lifetimes, but we cannot rule out the possibility that a dynamic collapse of the West Antarctic ice sheet could raise sea level 10-20 feet (3-6 m) in a human lifetime or two, and if we don’t change our behavior over centuries, Greenland and its 24 feet of sea level is looking shaky. Drs. Anandakrishnan and Alley spend a lot of their research trying to reduce our uncertainty about the future of the great ice sheets, a fascinating and important topic.
The near-certainty of continuing sea-level rise has some policy implications. For example, disaster aid following hurricanes that allows people to rebuild in coastal regions will simply create the need for more disaster aid in the future. Many people believe that those who wish to build on the coasts should be required to carry insurance to cover their coming losses. Similar arguments apply to those who wish to build on earthquake faults, landslide deposits, and floodplains. Why should those living in relatively safe but less-scenic places pay for others to live in dangerous but beautiful places?
Because people love the coasts so much, and wish to live near them, all sorts of engineering solutions have been tried. One is to build “dams” (usually called “groins” if they are small, or “jetties” if larger) that stick out into the water and block the longshore transport of sand (see figure to the left). By making the coast rougher, and slowing waves, the plan is to trap sediment along the coast the way a dam traps sediment along a river. This plan sometimes works. However, recall that in a sand-bedded river, sediment is trapped upstream but eroded downstream. The same often happens along a coast; sediment is trapped “upstream” of the groin (on the side from which the longshore drift comes), but sediment is eroded on the “downstream” side (the side to which the longshore drift goes), where the sand-free waves attack the beach to pick up more sand. Saving someone’s beach while destroying the beach of a neighbor is a good way to generate lawsuits. People also spend millions of dollars to go out to sea, find sand that has fallen off into deep water, and bring the sand back to the beach. This sand usually lasts a single year or a very few years before being washed back to deep water, but in especially popular tourist destinations, the investment may pay off.
Geologists often take a “natural” view of the coasts—we should figure out where the coast wants to go, and build there rather than trying to stop the coast. But many people just don’t like that, and a lot of construction is likely to occur—and be destroyed—over the years.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 8.
TECH NOTE
Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
This week, we feature four GeoClips, all created while Dr. Alley was on vacation at the Cape in the summer of 2005. These four "home movies" provide a bit of insight into the origins of the Cape, and the forces that are continuously at work changing our coasts, shorelines, and seas. Shooting and editing credits go to Dr. Alley's wife, Cindy.
As before, we hope you enjoy these, and find them to be useful complements to the readings, class notes, and slide shows of Unit 8.
Human impacts on the land are easy to see. We have changed the oceans greatly, but the water covers our tracks. In "The Can," Dr. Alley briefly reflects on some issues of the oceans, as he watches one of the less-beautiful pieces of the Cape Cod National Seashore.
Many of the ocean’s big fish, and other denizens of the deep, rely on salt marshes as nurseries and in other ways. But, we are losing salt marshes in many places, as sea-level rise forces the “outer beach” toward the shore, but humans don’t allow the inner side of the marsh to expand into our yards or parking lots. Obvious answers are not easily available, but Dr. Alley frames the question in this next short film clip as he paddles one of the family kayaks on the Nauset Marsh of the Cape Cod National Seashore.
Waves move immense amounts of sand, primarily up and down the beach, but also with a little motion along the beach and eventually off into deep water. In "The Feet," Dr. Alley gets cold feet on Coast Guard Beach, Cape Cod National Seashore, to show you moving sand.
Cape Cod is a gift of the glaciers. The numerous kettle ponds left by the ice contribute to the biodiversity of the Cape, but are slowly filling in with sand, peat, and other things. Many of the ponds have already filled, and a walk along the rapidly eroding outer beach often reveals where the sea has cut into one of these filled ponds. In this next clip, Dr. Alley shows one such exposed, filled kettle pond, just below the old Coast Guard station in the Nauset region of the Cape Cod National Seashore.
The Unit 8 lecture features Dr. Richard Alley and is 51 minutes long.
Check out the Unit 8 Presentation used in the online lecture.
Groundwater pumping can also cause saltwater intrusion. Fresh water has lower density than salt water, and so floats on salt water in the same way that an iceberg floats on water. (Salt and fresh water will mix, but if the fresh water is renewed by rainfall, the mixed waters will be forced out through the beach to the ocean, and there will continue to be nearly pure fresh water sitting on salty ocean water.) If the water table is lowered by pumping fresh water for human use, the interface between salt and fresh water will rise in the same way that the bottom of an iceberg or a mountain range rises if the top is eroded. The difference in density between salt and fresh water is small; an iceberg floating in the ocean has 9/10 of its thickness below the surface, but the fresh groundwater lens of Cape Cod floating on ocean water has 39/40 below sea level. So if enough water is pumped out of the well to lower the water table 1 m, the salt water will have risen 39 m! If the freshwater table is lowered to sea level, the salt water will rise to sea level, and there will be no fresh water left at that point. Many wells on the very low land of Cape Cod were drilled below sea level into fresh water, but are starting to pump up salt water, causing large problems.
The Beatles' Yellow Submarine surely must be very high on any list of songs likely to get stuck in your brain and play over and over and over and... But, watching our beaches, with perhaps three-fourths of the US coastline retreating inland in response to rising sea level and other changes, maybe a really "sticky" song is what we need down by the shore! Anyway, we hope you enjoy this parody, which really does review a lot of the key points that are likely to show up on the RockOn quiz. "We need more sand or the beach will wash away..."
You have reached the end of Unit 8! Double-check the list of requirements on the Unit 8 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Continue to work on Exercise #4. Consult the Course Calendar for specific due dates.
Following are some supplementary materials for Unit 8. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
(With apologies to the great detective.) "Holmes, did you realize that this chess board is exactly 12 inches long?" I asked, as we sat across the table in our Baker Street lodgings.
"Come, Watson. The game is a foot," he replied. "We must explain the Giant Splat of Sue Matra's birthday."
"The Giant Splat? Holmes, I know that Sue Matra is a friend of your brother Mycroft, but how came she to have a giant splat on her birthday?"
"Sue returned home from her job as a scientist measuring the air speed of swallows carrying coconuts, and discovered a splat on the floor. Here is a brief diagram of the situation."
Holmes rapidly sketched the following:
"This takes the cake," I commented sweetly, after looking at the diagram carefully.
"No," he replied, "The cake was left on Sue's study floor."
"Holmes," I said, "You know that my medical training exposed me to the scientific method and that in my military experience I conducted forensic investigations, learning what people died of. I was even exposed to geology and paleontology during my studies at the college. Your science of detection shares much with forensic medicine, and with the fields of geology and paleontology. Something has happened, and you try to determine what."
"Indeed, Watson, I am aware that you are among those few individuals who have been trained to reason from effects to causes, so unusual in a world full of people who reason only from causes to effects."
"But it is more than that, Holmes. When faced with an effect, we quickly try to think of all possible causes that are consistent with the available data and with our other knowledge about the world—call these 'multiple working hypotheses'. Then, we see what each of them predicts that we might discover or measure—each possible hypothesis is a cause, and we reason to its effects, in the same way that other people reason from causes to effects. But we go further, and we make those measurements or observations that will identify the effects of each of our likely causes, using the results to eliminate some of the hypotheses. For example, if we postulate that a horse stomped on the cake to make the giant splat, then we expect the mark of a horseshoe in the cake."
"Even if the horse missed, the shaking of the ground might have caused the cake to fall. Close counts in horseshoes and hand grenades, as you know."
"I ignored his attempted witticism, and said, "Holmes, allow me to send a few telegrams to test some hypotheses, and we will soon have the answers."
An hour later, I was ready to explain the giant splat to Holmes. "Sue Matra is married to a diligent and helpful husband, who is also an amateur cook. He confirms that he was baking a birthday cake for Sue, and made it in three layers, putting down a layer of cake, then icing, cake, icing, cake, icing. Then, he put candles in, and carefully carried the cake across the study toward a table. Unfortunately, the couple owns a large and boisterous retrieving dog, which jumped up and took a large bite out of the cake."
"The curious incident of the dog in the daytime," Holmes muttered.
"Yes," I said, "the dog did much in the daytime. You will notice that your highly accurate sketch shows that the dog's bite has severed the candle, and that the candle was shoved through the icing, so the order of events is clear. The actions of the dog so surprised the husband that he dropped the cake, which was unbalanced by the dog anyway. Given the height from which it dropped, and the rotation imparted by the dog, the cake flipped halfway over and splatted to the floor. The husband, distraught, and worried lest the suddenly sweetened canine should have an accident on the floor, took the beast for a walk, and during this interval, Sue returned home and discovered the splat."
"And how," asked Holmes, "did you determine that the dog bite occurred before the splat."
"If the splat occurred first and the dog then bit the splatted cake, the bite would not run smoothly through the whole thickness of cake, or else the dog's teeth would have scarred the floor or become stuck in the wood, yet the bite through the cake runs smoothly the whole way."
"Well," said Holmes, "I hope they have some good old H2O to clean up the mess."
"Elementary, Holmes," I replied, "elementary."
Could you have explained Sue's splat? If so, you understand some of the secrets of being a geologist. If not, then stick with us, and you should have it figured out by the end of the unit.
You will have one week to complete Unit 9. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 9, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 9 Instructional Materials. |
Continue working on Exercise #5: Puzzling Out Relative Time | Yes, this is the fifth of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 9 "RockOn" quiz | Yes, this is the ninth of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #10" survey | Yes, this is the tenth of 12 weekly surveys and is worth 1% of your total grade. |
Please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations with any questions.
On the following pages, you will find all of the information you need to successfully complete Unit 9, including the online textbook, a video lecture, several vTrips and animations, and an overview presentation.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Bryce Canyon is a fairyland. Pillars, spires, and minarets, seemingly fragile and toy-like when viewed from the rim, tower far above the hiker who descends into the “canyon,” which is really a bowl carved out of the edge of a plateau. The morning sun, bouncing off the pink and pastel walls, casts a rosy glow on everything and everyone in the canyon. Gnarled trees cling to high slopes, their root structure exposed by erosion but somehow holding and nourishing the twisted branches. The old settler Ebenezer Bryce called the scenery of this tributary of the Paria River “a hell of a place to lose a cow,” but almost everyone who followed him has had a kinder assessment.
The main pink rocks of Bryce are a mixture of limestones, shales, and some coarser sandstones and conglomerates, deposited in a lake or a series of lakes, and together known as the Claron Formation or the Bryce Limestone. (Geologists studying regions have found it useful to give specific names to recognizable bodies of rock, with the names taken from geographic locations. Because the criterion for a new name is the ability to easily recognize a unit in the field, some really careful observers like to subdivide, and others like to lump, so there is often slight disagreement as to which name is most appropriate.) These Bryce rocks are from the early part of the Cenozoic, roughly 40 million years ago. The rocks record a generally wet time shortly following uplift of the land from a shallow seaway, and before continued uplift led to the high plateaus of today, and climate change dried the surroundings to the dry climate now at Bryce.
Some of this story is told in the rocks of Bryce. The rest is told nearby. Northeast of Bryce in central Utah, the rocks record the gradual transition from undersea to terrestrial. The growth and erosion of mountains to the west are marked by the appearance and then disappearance of coarse braided-stream deposits. These were replaced by slower streams, then by lakes, then another pulse of river sediments, and then the great Green River lakes that spread from Bryce to Wyoming and that record a wet, warm time in the west.
Geologists love to tell stories like this one. The seas came in, the seas went out, the mountains rose, the mountains were eroded. Such stories are useful, because they often include chapters such as “The oil formed here, and migrated there, and was trapped in those rocks, and we can drill into it this way and get rich.” Such stories tell us much about the stability of the Earth’s climate. And the stories are interesting to many people. But how do you read such stories, and gain confidence in them? One way to find out is to major in Geosciences and spend the next three years learning to do so professionally—in case you’re interested, talk to us! But if you’re really excited about the great career prospects in your current major of theatre or philosophy, at least read over the next few sections and you’ll get a brief sketch of geological techniques and what they tell us.
We saw at the Badlands that weather attacks and breaks down rocks, and we saw at the Tetons, in Canyonlands, and elsewhere that the loose pieces are transported. The processes of weathering and erosion produce new minerals from old, produce small pieces from big ones, produce dissolved salts, and move these around on the surface of the Earth. Eventually, these materials end up somewhere. The pieces derived from older rocks we call sediment. Sediment may be deposited on the bottom of a sea, or on the flood plain of a river, or at the end of a glacier. Sediment may stay for a while and then be moved on. If it stays for too long, it may be hardened.
There is no magic in making rock from sediment, nor is there an abrupt line separating the two. Many processes slowly transform loose pieces into pieces stuck together. If you don’t believe that loose pieces get stuck together, go to an old house and try to remove some piece of plumbing without using a really big wrench, or try to get pieces off an old bicycle without a big wrench. If you don’t have the time, trust us—people make big wrenches for a good reason! Nature does rust and otherwise bond things together in many cases.
The weight of additional material deposited on top of a sediment layer will squeeze it, forcing grains together and pushing water out. The resulting material is harder than what was there before. Continued squeezing and heating as sediment is buried (remember that the Mississippi delta is miles thick) produce rock from this sediment, and eventually may change these sedimentary rocks to metamorphic rocks. The clays in sediment change into different clays and then into other things as the cooking proceeds (just as flour and water and yeast make cookies in the oven), and as the crystals of the new minerals grow, they tend to wrap around each other and get “woven” together so they are hard to separate.
Groundwater circulating through sediments will dissolve minerals in some places and precipitate them in others, in response to subtle changes in temperature, pressure, or chemistry. We saw that there is dissolved rock in Spring Creek water, and in all other waters. If you have “hard” water, of the sort that causes people to buy water softeners, that means the water has a lot of dissolved rock and especially dissolved limestone. The white deposit that builds up on drinking glasses in houses with hard water is not soap scum but rock, very similar to stalactites in a cave. A lot of the sticking-together of plumbing pieces is linked to minerals precipitated from hard water. What happens in your house can happen in nature, where the precipitate serves as a cement that binds particles together. This cementation, together with the compaction from squeezing and the recrystallization from heating and changing clays, slowly turn loose sediment to sedimentary rock. Sometimes, rocks are made entirely of hard-water deposits—stalactites are an example, but so are the salt deposits left in the bottom of Death Valley when the water that runs down from the peaks quickly evaporates. Rocks made entirely of precipitate are also called sedimentary rocks.
Our classification of sedimentary rocks is a bit of a hodge-podge. We first consider whether a rock is clastic (made from pieces or clasts of older rocks) or chemically precipitated (deposited from chemicals dissolved in water). This subdivision is not always satisfactory—a sea shell is a chemical precipitate because the animal pulled the material in its shell from the water, but a limestone made up of sea shells might be called clastic because the sea shells are chunks. Usually, people consider limestones and evaporites (rocks left by evaporation of water containing salts) to be chemical precipitates, and all others to be clastics.
Clastics are classified based primarily on grain size. The very smallest particles of clay make claystone or shale. Slightly coarser pieces are silt and make siltstone. Coarser still is sand, which makes sandstone. Going to still-bigger clasts, cobbles and boulders produce cobblestone and boulder-stone, but we also call both of these conglomerate.
We learn about the past through sediment. Archaeologists dig up the garbage dumps and houses of past peoples to learn how they lived. Historians root through the intellectual sediment of people to learn what they have done. Sedimentary rocks, in the same way, tell the history of the surface of the Earth.
A typical sedimentary rock contains all sorts of clues as to how it came to be. The clast size or grain size is a measure of the energy of transport—tiny clay particles will not settle out of a swift mountain stream, so a fine-grained deposit indicates slow or no currents. Grain shape tells more—an angular clast has not been transported far in water or wind, because during transport sharp corners are knocked off and the clast becomes rounded. Sorting also tells something—wind and water sort sediments by size, but landslides and glacier ice do not, so the sand grains on a beach are all about the same size, but a landslide leaves a deposit that ranges from miniscule clay-sized particles up to house-sized boulders and houses.
Composition provides clues—olivine is a mineral that is changed rapidly by chemical weathering and occurs primarily in low-silica igneous rocks such as sea-floor basalts, so if you find olivine grains in a sandstone, you know both that the sand was derived from such rocks, and that the sand was not transported very far in a wet environment because the olivine was not broken down. Even surface textures have information—the striations and polish of a glacier-transported rock look very different from the sand-blasted appearance of a quartz grain in a sand dune.
The sizes and shapes of deposits give further information. There are several ways to tell which way a current of wind or water was moving when it deposited a particular rock. For example, with sand dunes, the wind carries sand up a gradual slope and then dumps the sand over the top where it avalanches down and forms a steeper slope. When we find the gentle and the steeper slope in the rock, we can tell which way the wind was blowing. Sand ripples and bars deposited by flowing water have similar indicators. A sand body deposited by a stream often is long and skinny in the direction that the stream was flowing. A “fossil” sandy beach may be long and skinny like a stream deposit, but the current directions caused by waves will be pointing more-or-less across a beach but along a stream, so there is not much danger of confusing one type of deposit for another.
The list of indicators is very long. When mud dries out, it cracks to form mud cracks. Such features are easily recognized in rocks, and indicate occasional drying. Raindrops that spatter down on rocks leave little pock marks that are readily recognized if they are turned to stone and preserved.
Added to this, we have the advantage of fossils. Tree trunks in growth position indicate a non-desert deposit on land—trees are replaced by cactus in desert environments, and by seaweed underwater. Coral skeletons indicate a reef in the ocean.
With such a wealth of information, we certainly should be able to read much of the story of the past. All that is required is that: 1) we study the rocks carefully; 2) we know what sorts of sediments are produced in what sorts of environments today; and 3) we make the common-sense assumption of “uniformitarianism”, that the present is the key to the past.
Suppose that you go out today and observe that streams flow downhill and carry rock and mud from the hills to the sea. You dig around in a lot of different streams, and find that they have sorted sedimentary beds, occasional fish fossils, current-direction indicators pointing more-or-less the same way in many different sand bars, sand bars of certain shapes and sizes, etc. You talk to a lot of people, and read history books, and learn that streams have been flowing downhill and carrying rock and mud from the hills to the sea for a long time. You find a place where earlier people had diverted a stream into a canal, dig up the old stream bed, and find that it looks just like modern stream beds. You find still older deposits that look exactly like modern stream beds, and you find deposits in which the sand has been cemented by hard-water deposits to make sandstone, but everything else looks just like modern stream beds.
The reasonable person would say that these “fossil” cemented-to-stone stream beds are indeed old stream beds. Of course, we cannot prove that they are. Perhaps space aliens made the world the year before you were born in such a way that it looks like there were old streams, and the aliens invented all the history books and fake memories of old people. Calvin (of the Calvin and Hobbes comic strip) once told Susie that their class was having a substitute teacher because “our REAL teacher rocketed back to Saturn to report to her superiors. They’re trying to subvert us little kids with subliminal messages in our textbooks, telling us to turn in our parents when the Saturnians attack. Earth will be rendered helpless! I’m too smart for ‘em, though! I don’t read my assignments!” (To which Susie replied, “I think one of us has been eating too much paste in art class.”) Maybe Calvin is right. And maybe the Geology-of-National-Parks professors are Saturnian aliens, too! The textbook author, Dr. Alley, has really strange-looking hair where it hasn’t fallen out, and acts a bit peculiarly, you know.
However, despite his occasional interest in the topic, Calvin had not quite mastered geologic thought. Geologists take the common-sense approach: if it looks like a duck, and walks like a duck, and quacks like a duck, is genetically identical to type specimens of ducks, has a bone structure as observed in x-ray examination that matches known duck bone structure, mates with other ducks and produces duck offspring, etc., we call it a duck. If it looks like a stream deposit in size, shape, arrangement, location, grain size, sedimentary structures, fossils, etc., we call it a stream deposit. We learn from the present—what do streams do, what do they make, where do they occur—and then we use that knowledge as the key to interpreting the past.
Learning from the present is not a casual undertaking. By working hard in one or two advanced courses, a really adept student can learn to distinguish most of the main deposits—streams versus lakes versus beaches, for example. Many geologists spend their whole lives really learning what stream deposits look like, how they form, etc., while other people are spending their lives working on beaches, or lakes, or glaciers.
When lots of these experts agree that the limestones of Bryce and surroundings were deposited in a shallow lake and adjacent streams and mud flats, you should know that the rock type (limestone, but with differences in chemistry and texture from limestones forming in oceans), the fossils (algae, fish fossils, alligator bones, bird tracks), the structures (mud cracks, raindrop imprints), etc., have been considered carefully in telling the story. We can’t rule out Calvin’s Saturnians, but a more down-to-earth explanation is more convincing.
Arches National Park is just outside of Moab, Utah, almost within shouting distance of Canyonlands. Arches contains the largest concentration of natural stone arches in the world, with dozens of major arches, and numerous other holes and interesting rock features. The longest natural arch in the world is there, Landscape Arch, spanning about 300 feet (almost 100 meters). Several rockfalls have occurred from Landscape Arch since European settlers arrived, reducing the arch to a thickness of only 11 feet in one region. The trail under the arch has been closed; the Park Service knows that eventually more rocks, and the whole arch, will fall, and the rangers do not wish to be involved in extracting compacted humans from beneath the remnants of what used to be the world’s longest arch.
The arches are all eroded in sandstone, and especially the mid-Mesozoic (deposited during the time of dinosaurs) Entrada Sandstone, which includes both marine and wind-blown sands. Arch formation started with deposition of thick beds of salt in the Arches area. Sediments including the Entrada sands were deposited on top of the salt, and cemented by hard-water deposits to make sandstone. When the weight of those sediments became large enough, the salt began to flow, squeezed from places where thicker sediments caused higher pressure on the salt, to places where thinner sediments gave lower pressure, like ketchup in a fast-food packet squeezed from one place to another. However, the foil wrapping of a fast-food packet is fairly flexible, but the Entrada sandstone was not, so as the salt moved, the tough sandstone broke. The most common joints formed by this motion are parallel, nearly vertically oriented, and a few yards (or meters) apart. Weathering processes attacked the rocks along these joints, which were slowly widened to leave vertical slabs of rock called fins.
Some weathering and erosion processes are more effective near the ground than high up. For example, you will see huge steel bolts sunk into the rock of some road cuts just above the highway to keep the weight of the rock above from popping slabs of rock loose that might blast into the road and kill people; such rock-bursts also occur in quarries. When such slabs of rock fall, they often leave a curved or arch-like top, because it is easier to break off a curve than a square-cornered piece. This process is among those contributing to arch formation, breaking through the fins to make the beautiful stone arches.
The text above says that the Entrada Sandstone is from near the middle of a time we call the Mesozoic, an Era of geologic history, and for Bryce, our text told you that Bryce Limestone is from early in the Cenozoic Era of geologic time. You have not been asked to learn these terms yet (but you will be, so might as well lodge them in your brain now). The Mesozoic is older than the Cenozoic, so the Entrada is older than the Bryce. How do we tell which layers are older or younger?
Putting rocks in order is an enjoyable puzzle, requiring the sort of logic used in sudoku or in crossword puzzles. You rely on a few general principles that really boil down to common sense, plus a lot of puzzling and looking and thinking. Start with “what’s on top”? The layer of mud that settles out of floodwaters is deposited on top of the mud and grass and linoleum and grand pianos of the flood plain. Naturally, the mud from the flood of the next year is deposited on top of the mud from this year (nature doesn’t dust the grand piano). In ordinary sedimentary rocks, the youngest ones are on top and the oldest on the bottom. We call this the Principle of Superposition.
Note that this applies to ordinary rocks. There are cases in which rocks are turned upside down. Take a sheet of paper on your desk, put fingers on opposite sides and squeeze your fingers and the ends of the piece of paper towards each other (see the figure on the left). The paper will buckle into a fold. If you get your fingers together, the fold may “fall over”. A vertical hole through your paper would now go through the top of the page and out the bottom, but then through the bottom of the page and out the top, before finally going through the top and then the bottom—going from top to bottom, the paper is right-side up, then upside-down, then right-side up (look at the figure again).
Fortunately, there are many “up” indicators in rocks that tell you which way was up when the rock was formed. Animal tracks and raindrop imprints are made down into mud, not up. If you find a track pressed up into the roof of a cave, it is reasonable to suppose that the track originally went down but then has been turned over, so it is now upside-down. Mud cracks narrow and disappear at the bottom, not the top. Curved clam shells are usually flipped by currents into the stable, hollow-side-down configuration rather than the rocking, hollow-side-up configuration. Bubbles in lava flows rise to the top where they are trapped just beneath the quick-cooled upper surface; the bubbles do not sink down to the bottom. (You will meet some of this in the second V-Trip, and in the Rockin' Review of this chapter, and in Exercise 4.) Hence, you can learn whether the rocks have been turned upside-down or not.
In clastic sedimentary rocks, the clasts must be older than the rock is. So a conglomerate containing pieces of granite and of other things tells of formation of a granite from melted older rock, followed by physical erosion to break the granite into pieces, followed by transport to mix the granite pieces with chunks of other rocks, followed by deposition of the chunks and hard-water cementation to make the conglomerate rock. A conglomerate containing pieces of sandstone tells of two cycles—make the sand from older rocks (perhaps from breakdown of a granite by chemical weathering), followed by transport of the sand to a beach, followed by cementation of the sand with hard-water deposits to make sandstone, followed by physical weathering of the sandstone to make sandstone chunks, followed by transport of the chunks and mixing with others and deposition and cementation to make the conglomerate. If a fault has split a granite clast in a conglomerate, and moved one side relative to the other so that the pieces no longer match, the fault must be younger than the conglomerate—something must exist before you can break it. If a layer of sandstone is standing up on edge, then it must have been deposited, hardened, and then tilted—because of mass-movement processes, layers are nearly horizontal when deposited (not exactly, but pretty close); if you tip a mud layer up on end before it has hardened, it spreads into a new, nearly horizontal layer.
Below is a photograph of an outcrop near I-70 in the near-desert conditions of central Utah (left), and a sketch of the outcrop (right). The rocks at the bottom were deposited on a sea coast—they contain marine fossils and other markers of the sea, and are mostly sandstone and shale. But they are standing on edge and truncated at the top. Developed on the top of them is a soil, and on top of that are lake deposits with flamingo bones, fish fossils and alligator bones.
The story of this outcrop includes:
You can postulate many other things—Dr. Alley is just lying about the rocks; or, the space aliens put the rocks there; or, a giant tornado blew the rocks there one afternoon; or…. We humans are very inventive, and can always come up with a new idea. But when you follow out the implications of these other stories, some fundamental problem always comes up, whereas the story of the rocks being deposited and hardened and tipped on end and eroded and so on works just fine, and makes predictions that are borne out over and over. Hence, the “geological” or “scientific” interpretation is the sensible one.
Notice that the erosion surfaces [events iii and vi in the list above] represent time gaps in the sedimentary sequence; the rocks at this site do not have a record of the environment during the time erosion was occurring—you have to go to where rocks were being deposited to learn that. A time gap in a sedimentary sequence, whether from actual erosion or just lack of deposition, is called an “unconformity”. Most of the land experiences erosion most of the time—the mud in the Mississippi is eroded from the lands upstream, and most of us most of the time do not have to worry about mud being deposited on our grand pianos or carpets. But, sediments are deposited on land, and sediments are deposited almost everywhere almost all the time under the sea, allowing a fairly complete story to be told.
The material above explains how geologists can put rocks in order. The job may be quite difficult, because some outcrops are covered by soil, water, or other rocks, because offsets on faults may be so large that rocks on opposite sides can’t be matched back up easily, etc. But the job can be done.
When this job is done, a remarkable result emerges. Many rocks contain fossils, and different rocks contain different fossils. Arranging the rocks in order reveals that the fossils also are arranged in order. Rocks of similar age from similar environments have similar fossils; the more different the ages of the rocks (the farther apart they are in our ordered pile), the more different the fossils appear. The younger fossils have more in common with things living today than do the older ones. This observation, that fossils succeed one another in a definite and recognizable order, is called the Law of Faunal Succession. “Law” here is not something passed in Congress; a scientific “law” such as this one successfully summarizes an immense number of different observations, and works very well in explaining and predicting many different things in many places and times.
The Law of Faunal Succession was developed in England by a 17th century canal engineer/geologist named William Smith. He faced practical problems; digging in some rocks is easier than in others, some rocks hold canal water better than others, etc. If he crossed a river or a region with no rocks showing, and then found a few rocks, he wanted to know whether they were the ones he wanted, or whether the ones he wanted were up or down in the pile. Sometimes, a few fossils would be evident in a streambed, or even would be in use as weights for balances to weigh cheese of the local farmers, and he found that the fossils made his job easier by providing shortcuts for putting things in order.
Much later, faunal succession figured in the debates over evolution. But you should remember that the Law of Faunal Succession is like so much good science—a practical idea that gives useful results and helps humans do things successfully.
Geologists learned to put rocks and fossils in order long before calendar-year ages could be assigned accurately (how many years old is that?). Lacking accurate year counts, the early geologists agreed on a “shorthand” for describing ages. Time is divided into big chunks, called “eras,” which are themselves divided into smaller chunks, or “periods,” and on into still smaller chunks. The big chunks (which you would be wise to know) are, from youngest to oldest, the Cenozoic (= new life), the Mesozoic (= middle life) and the Paleozoic (= old life) (see the table on the following page).
The oldest period in the Paleozoic is called the Cambrian, and the rocks older than it are called the Precambrian. As the Precambrian includes most of the history of the Earth, more and more workers are using “-zoic” terms for pieces of it (see the Enrichment for Unit 9). The periods are mostly named for places where the rocks are well-developed, such as the Cambrian for Cambria (also known as Wales), Devonian for Devonshire in southwestern England, and the Pennsylvanian for Pennsylvania (Europeans usually lump the Mississippian and Pennsylvanian and call that lump the Carboniferous, but we have such wonderful deposits of their Carboniferous that we divide them to allow better time resolution, and to use our names). The youngest two periods, the Tertiary (third) and Quaternary (fourth), were named in an old scheme that also included primary and secondary; history dies hard in geology.
Life appeared early in the Precambrian, and multicelled organisms were present in the Precambrian. The Cambrian began with a great diversification of organisms with shells. Older fossils are rare because soft body parts usually are not fossilized; when shells appeared, they accumulated to make limestone layers, so fossils “suddenly” (over a few million years) became common. The Paleozoic and the Mesozoic ended with mass extinctions that probably killed the great majority of organisms on Earth (we’ll return to this soon). The mass extinction at the end of the Mesozoic was caused by a meteorite impact at the end of the Paleozoic that is not fully explained, although radical changes in ocean circulation in response to global warming and massive volcanism that produced a widespread lack of oxygen in the ocean probably were involved.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 9.
Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Arches National Park
(Provided by USGS)
Bryce Canyon National Park
(Provided by USGS)
This week, we feature one GeoMation and three GeoClips. The GeoMation focuses on "detective work"—the kind employed by geologists to trace and identify the geologic histories of the features and structures they study. The GeoClips take you out to Capitol Reef and Bryce with the CAUSE class, to see a little bit about how the detective work is done in the field. The real detective work involves more rainbows and rattlesnakes than you'll see here; for that, you might have to come see us about majoring in Geosciences.
We hope you enjoy this unit's multimedia presentations and that they help you make a little more sense out of Unit 9.
Rocks reveal how and where they were formed. What is in a rock, how it is put together, whether the pieces are big or little, sorted or mixed, angular or rounded, and so much more provide clues. Fossils also provide clues. Here, Dr. Alley and the CAUSE class are out in the desert at Capitol Reef National Park, but they are also in a shallow seaway from long ago. See why.
Rocks occasionally are turned upside-down, but nature tells us when that happens. Mud cracks can show us that; they are wide at the top, narrow, and then end at the bottom. Fill mud cracks with another layer of sand or mud, and the cracks are "fossilized," to tell us which way was up when the rocks were deposited. Here, visit Capitol Reef with Dr. Anandakrishnan to see mud cracks, with a brief look at some right-side-up ones from the Grand Canyon.
Geologists read rocks, and the stories are fascinating--historical novels full of intrigue. In this next GeoClip, Dave Janesko and Dr. Alley perch high up in Red Canyon just west of Bryce, and read one of those stories of deep time.
A conglomerate is a sedimentary rock in which many of the clasts are bigger than sand. Dave and Dr. Alley are looking at a conglomerate that includes many different clast types, including one that is itself a finer-grained conglomerate. The clasts in that conglomerate-within-a-conglomerate include several types of sedimentary rocks, including sandstones that are themselves made from older pieces.
Optional Enrichment Article (no, this won't be on the quiz!): What do beauty, saving money at Las Vegas, religion, oil exploration, emerging new diseases, and the planet’s recovery from global warming have in common? All in some way involve deep time, the immense age of the Earth. Eric Spielvogel filmed a discussion of these and other issues with Dr. Alley, for a special “time” issue of Research! Penn State. The Deep Time film clips on the next page will give you something to think about, and may even help with the course. Enjoy! Also, here are some optional animations you might also want to explore! (No, these won't be on the quiz either!)
Sedimentation Models
(An extensive collection of animations on this subject)
Sequence Stratigraphy
(An extensive collection of animations on this subject)
During three weeks in May 2004, two hardy Penn State geoscientists traveled through 12 stunning National Parks of the southwestern United States with 13 lucky students. The trip was sponsored by CAUSE (Collaborative Active Undergraduate Student Experience), an annual course offered by the College of Earth and Mineral Sciences. Richard Alley, Evan Pugh professor of geosciences, led the expedition. CAUSE 2004 was an extension of his course, "Geology of National Parks," and allowed students to interact with and learn from the rocks and landscapes of Arizona, Utah, and Colorado. In the following videos, Alley explains the concept of deep time, how it tells the history of our planet, and how it affects our lives.
Richard Alley, Ph.D., is Evan Pugh professor of geosciences in the College of Earth and Mineral Sciences, rba6@psu.edu.
—Emily Rowlands
The Unit 9 lecture features Dr. Richard Alley and is 52 minutes long.
Check out the Unit 9 Presentation used in the online lecture.
The Cenozoic, Mesozoic, and Paleozoic Eras are lumped together into the Phanerozoic Eon; “Phanerozoic” is from Greek roots meaning that life showed itself. The Phanerozoic started with the fairly sudden (over a few million years) appearance of common shells, which greatly “improved” the fossil record. You have to look mighty hard to find the remains of a jellyfish in a rock, but with shells, the remains are the rock. Going back in time before the Phanerozoic are the Proterozoic (earlier life), the Archean or Archeozoic (beginning life; different people prefer Archean or Archeozoic), and the Hadean (like Hades; hellish).
The Hadean, from about 4.55 or 4.6 billion years to about 3.8 billion years ago, probably experienced such huge meteorite impacts that they vaporized the whole ocean and some rocks as well. Life would have been hard-pressed to survive such catastrophes.
Life was present by the beginning of the Archean, at about 3.8 billion years ago. The central regions of the continents had formed by the end of the Archean, about 2.5 billion years ago. Oxygen remained scarce or absent in atmosphere and ocean, and carbon dioxide helped keep the world warm under the gaze of a young sun that was not as bright as today.
In the Proterozoic, oxygen from cyanobacteria and other microbes “rusted” the oceans. Iron rusts with oxygen, but in the oxygen-poor world of the Archean, iron dissolved and washed to the ocean rather than rusting. As oxygen was released by photosynthesis, the iron was rusted and fell to the bottom to form the great banded iron formations that we mine today in Minnesota, Australia and elsewhere. Oxygen then increased in the atmosphere. First a poison to many living things because it is so highly chemically reactive, oxygen proved invaluable as life figured out how to use the “new” chemical, allowing development of larger organisms. Oxygen also shielded all organisms from damaging ultraviolet rays by giving rise to protective ozone in the stratosphere. As oxygen rose in the atmosphere and combined with gases there that were stable without oxygen but unstable with oxygen, those gases were removed. Some of those were strong greenhouse gases, and removal of them under the not-yet-modern-brightness sun allowed “snowball Earth” events that very nearly froze the planet.
Here's another look at learning which way was "up" when a rock layer was deposited. Dance on down with the dinosaur in this parody of Neil Sedaka's "Breaking Up Is Hard To Do". (3 minutes and 19 seconds long)
You have reached the end of Unit 9! Double-check the list of requirements on the Unit 9 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Exercise #4 is due and Exercise #5 opens this week. See Course Calendar for specific dates and times.
Following are some supplementary materials for Unit 9. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
Imagine that the 100 yards of Penn State's Beaver Stadium, or any other football field, are like a timeline of all of Earth's history, and you're the star of the team, driving for glory. The planet formed on your goal line, half of the Earth's history had passed as your team marched across the 50-yard line, and now the coach personally sent you, the acme of creation, to carry the ball across the opposition's goal line of today for the winning score. If you have been carrying the ball for the whole 20 years of your life, how far did you run? (If you're not 20 years old, pretend.)
Congratulations--tomorrow's newspaper will report that you gained just a shade under 0.0002 inch, or a bit less than 1/200 of the thickness of a sheet of paper. The defense was vanquished by your onslaught, and instant-replay officials weren't needed to see that you broke the plane of the goal.
Some people still insist that the Earth is no older, and looks no older, than the roughly 6000 years of written history. But, that is barely the thickness of a sheet of paper on the 100 yards of Earth's "dark backward and abysm of time," as Shakespeare called it. Geologists often feel sorry for those young-Earth supporters—they may have seen the instant-replay of the touchdown, but they missed the thrill of the game. So come along and see how the team got you the ball for those last two ten-thousandths of an inch!
You will have one week to complete Unit 10. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 10, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 10 Instructional Materials. |
Submit Exercise #5: Puzzling Out Relative Time | Yes, this is the fifth of 6 Exercises and is worth 5% of your total grade. |
Begin Exercise #6: Fossil Fuels (& Driving Hazards) | Yes, this is the last of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 10 "RockOn" quiz | Yes, this is the tenth of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #11" survey | Yes, this is the eleventh of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 10, including the online textbook, a video lecture, several vTrips and animations, and an overview presentation.
Students who register for this Penn State course gain access to assignments and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
Leave it as it is. You can not improve on it. The ages have been at work on it, and man can only mar it. What you can do is to keep it for your children, your children’s children, and for all who come after you, as one of the great sights which every American...should see.
–President Theodore Roosevelt, speech at the Grand Canyon, Arizona, on May 6, 1903
We have gotten past the stage, my fellow-citizens, when we are to be pardoned if we treat any part of our country as something to be skinned for two or three years for the use of the present generation, whether it is the forest, the water, the scenery. Whatever it is, handle it so that your children’s children will get the benefit of it.
–President Theodore Roosevelt, speech at the Grand Canyon, Arizona, on May 6, 1903
Out in eastern Nevada, a long way from almost anywhere, is Great Basin National Park. The jewel of Great Basin is Lehman Cave, one of the most "decorated" caves known, with a wide range of odd cave formations (stalactites and stalagmites, but lots more, too). Lehman Cave is dissolved into marble (metamorphosed limestone) on the side of Wheeler Peak, which rises to more than 13,000 feet (almost 4000 m), and which was glaciated during the ice age—the remnant cirque glacier has wasted away to a debris-covered rock glacier over the last decades. Fewer than 100,000 visitors make the trek to Great Basin each year, and you can find a lot of solitude and wonder in this beautiful place.
Far up on Wheeler Peak, Great Basin bristlecone pines are living. These gnarled, straggly trees grow slowly in high, cold places, whereas those bristlecone pines growing in warmer, moister, lower-elevation sites live faster and die younger. In part because of this slow growth, the high-altitude trees can be very old. The oldest known living bristlecone pine is more than 4,600 years old, in the White Mountains of California. The oldest tree known so far was cut on Wheeler Peak in 1964, when the land was still administered by the U.S. Forest Service, as part of a study to learn more about the growth and behavior of the trees. Now known as Prometheus, that tree was 4,950 years old when cut. That one old-looking tree was not notably different from many others in the large grove. Because it is so unlikely that the first such tree cut on Wheeler Peak out of the many there would happen to be the oldest tree on Earth, it is likely that there are older trees out there that have not been sampled yet.
In a seasonal environment, a tree reliably produces a visible growth ring each year. The reasons for this behavior are well-understood, and the annual nature of the rings has been checked many, many times. Rarely, there is a problem (a piece of a ring may be missing if the tree was damaged, by a fire or a burrowing beetle, or a late frost or other odd event may make a ring look strange), but tree-ring daters (dendrochronologists) have learned to recognize these events. In general, tree-ring dating can be practiced with no errors. Many, many tests have been conducted to confirm that this works, that the results match historical records, etc. Most such sampling is done using narrow coring devices, and does not harm the trees.
In studying tree rings, one sees that the width is not the same from year to year. Thick rings grow during “good” years, and thin rings during bad years. This allows tree rings to be used to reconstruct past climates. In a dry area, a good year is a wet one, so tree rings can be used to find out how much rain fell in the past. In a cold area, a good year is a warm one, so the tree rings function as thermometers.
For our purposes here, the pattern of good and bad years (fat and thin rings) is important for dating. On Wheeler Peak, and in the White Mountains and elsewhere, dead trees occur near the living bristlecones. Some of these dead trees sprouted before the living ones and overlapped in age with the still-living trees. Other dead trees can be found in archaeological sites or buried in sediments. A tree-ring specialist can date the good and bad years using living trees. The specialist can then find the pattern of thick and thin rings in the dead tree, and so use the dead tree to extend the record back to when the dead tree first sprouted (see the figure below). By overlapping a few long-lived trees, or many short-lived trees, very long chronologies can be generated.
Such techniques are used to date archaeological sites, including those of the Ancestral Puebloan peoples (also called the Anasazi; at Mesa Verde and several other national parks). Professor Peter Ian Kuniholm and his collaborators at Cornell have for decades been doing amazing work using tree rings to understand classical history in the Aegean region, the Middle East, and elsewhere, confirming, refining, and extending historical accounts. The beautiful agreement between tree-ring and historical accounts as far back as the oldest reliable written records confirms the accuracy of the techniques.
But, the tree-ring records extend well beyond reliable written histories. The longest tree-ring record in the U.S. Southwest is now more than 8000 years. The longest known of such records is from tree trunks buried along rivers in Germany, and extends to 12,429 years—before that, closer to the heart of the ice age, conditions were too cold for trees in that region of Germany. Because most trees live for “only” centuries rather than millennia, such records (and a few other really long ones, such as a 7,272-year record, as of 1984, from Irish oaks buried in bogs) represent immense investment of time and effort, but people have devoted whole careers to assembling these outstanding records. Notice that there is much older wood, including the fossil trees at Yellowstone, Petrified Forest, and elsewhere. The more than 12,000 years in Germany are the longest continuous record reaching the present, but surely do not come anywhere close to including the whole history of trees.
Several other types of annually layered deposits exist. For example, some lakes in cold regions freeze every winter. When the lake is thawed in the summer, sand and gravel are washed in by streams. When the lake and its surroundings freeze, the streams slow or stop, and the only sediment settling to the lake bottom is the very fine silt and clay particles that were washed in during the summer but require months to fall. A coarse layer capped by a fine layer marks one year. Many such layered, or varved, lakes have been studied, and found to contain thousands of years to more than 14,000 years. Many of these lakes occur in glacier-carved basins, and so their records extend only back to the melting of the ice. One lake in Japan, Lake Suigetsu, has a spring bloom of diatoms—algae with silica "shells"—that make a light-colored layer, alternating with darker mud during the rest of the year. 45,000 annual layers have been counted in that lake.
Most lakes lack annual layers. If there is a lot of oxygen in the deep waters, worms will thrive in the mud beneath, and their burrows will disturb the layers. If the lake is shallow, waves will disturb the deep muds. But enough lakes exist with annual layers to be useful. And, simply seeing layers doesn’t prove they are annual; lots of tests have to be done, as described below for the ice cores.
A difficulty in lakes—and other archives such as annually layered stalagmites—is that an annual layer must be thick enough to be recognized, but a lake or a cave will fill up quickly if layers are thick, so the records cannot be extremely long.
Longer records are possible from the two-mile-thick ice sheets. Dr. Alley has been very active in this work, and Dr. Anandakrishnan has contributed in important ways. In central Greenland and some other places on ice sheets, the temperature almost never rises high enough to melt any snow and ice. However, summer snow and winter snow look different because the sun shines on the snow in the summer, “cooking” the snow and changing its structure, but the sun does not shine on the winter snow, which is buried by new storms before the summer comes. You can count many, many layers by looking at an ice core, and Dr. Alley did so, working especially on one core called GISP2.
To verify that the layers are annual, several things were done. First, one person (Dr. Alley) looked at the core, waited a while, and then looked at it again to see that the counting is reproducible. Then, other people counted the layers visible in the core (without cheating by finding out what Dr. Alley had gotten), just to make sure they agreed.
There are many annual indicators in ice cores, probably more than a dozen. For example, the isotopic composition of the ice is a thermometer that records summer and winter. And sunshine makes hydrogen peroxide in the air in the summer when the sun shines, and the peroxide falls on the ice quickly, but there is almost no peroxide made and deposited in the dark winter. So annual layers have been counted using several different indicators, and they agree closely.
This is still not good enough. When a large volcano erupts, it throws ash and sulfuric acid into the stratosphere. These spread around the Earth. The bigger pieces of ash fall out quickly, often in days or less, while the sulfuric acid may take one to a few years to fall (and, until it falls, affects the climate by blocking a little of the sunlight). You can use electrical or chemical techniques to find the layers of volcanic fallout in ice cores. The key sections can then be cut out, melted, and filtered, and any volcanic ash that is found can be analyzed chemically and compared to that from known volcanic eruptions. So, if you count back to the year 1783 in a Greenland ice core, you are in the year of the great Icelandic fissure eruption of Laki, which spread dry fogs across Europe and is well recorded in histories—Ben Franklin commented on the fogs in Paris while he was ambassador there for the fledgling United States. In fact, ash of the composition of Laki occurs in Greenland ice cores at the level dated 1783 by layer counting—the layer counting is right (or very close—some counts missed by a year or two initially). Similarly, ash from many other historical volcanoes has been found, back as far as historically dated volcanoes are known.
Comparison of counts of strata by one person at different times, by different people, and by different methods, and comparison to volcanic fallout, yield almost the same answers, within about one year in one hundred (so one person may count 100 years, and another will count 99, or 100, or 101, but not 107 or 93 or some similarly large error).
There are a few more tests yet. There were very large and very rapid climatic changes at certain times in the past. Ice cores record the climatic conditions locally (how much snow accumulated and how cold it was), regionally (how much dust and sea salt and other things were blowing through the air to the ice from sources beyond the ice sheet), and globally (by trapping bubbles of air, which contain trace gases such as methane that are produced across much of the Earth’s surface). Changes in all of these indicators occur at the same level in the ice cores, showing that the climate changes affected much of the Earth.
These changes left their “footprint” in the ice of Greenland, and the lakes of Switzerland and Poland, and the trees of Germany, etc. So one can date such changes in the annually layered deposits of all of these different places. And, the dates agree closely. These events also have been dated radiometrically (we’ll cover this soon), and the dates also agree closely. One event, for example, was a return to cold conditions during the end of the ice age, and is called the Younger Dryas. Close agreement as to its age is obtained from all of these different layered deposits and from radiometric ages—the Younger Dryas ended about 11,500 years ago.
Thus far, the layers in the ice cores provide the longest reliable records. Over 100,000 layers have been counted. High accuracy was achieved younger than about 50,000 years, with somewhat lower reproducibility (maybe 20% or so, and with well-understood reasons for the lower accuracy) older than about 50,000 years. Still older ice exists, but the layers in Greenland have been mixed up by ice flow and no longer give a reliable chronology. Thus, we have high confidence of more than about 100,000 years from the ice cores. (Really old ice in Antarctica, to 800,000 years or so, got less snowfall in a year than the height of a snowdrift, so annual layers are not preserved and other dating techniques must be used.)
One of the great results of geology has been the concept of “deep time.” The world was once believed in some cultures to be only as old as the oldest historical records. The Archbishop Ussher of Ireland, in the year 1664, declared that based on Biblical chronologies, the creation of the Earth dates from October 26, 4004 BC, Adam and Eve were driven out of the Garden of Eden on Monday, November 10 of that year, and Noah’s Ark landed on Mt. Ararat on Wednesday, May 5, 1491 BC. Other experts obtained slightly different dates, but with broad agreement that the world was not very old.
Ussher’s date rested on a literal reading of the particular translation of the Bible he used, and on quite a number of questionable interpretations of the text—the Bible itself never gives an age for the Earth. Early geologists nonetheless struggled with the constraints provided by such chronological readings—how could all of geologic history fit into 6000 years? The early geologists ultimately reached the conclusion that the world looks MUCH older than 6000 years; either the world is older than this, or we have been deliberately fooled by some powerful being who crafted a young world to look old. As scientists, we work with the observable part of the world, and we have no way to detect a perfect fake, so we treat this as an old world. The geologic record speaks of “deep time,” billions of years, Shakespeare’s “Dark backward and abysm of time.”
Most biblical scholars have reached the same conclusion: the chronologies of Genesis do not allow one to fix the age of the Earth precisely, and are perfectly compatible with an old Earth. Most of the large Christian denominations, for example, have accepted an old Earth based on Biblical and on scientific interpretations. In 1996, the pope added the Catholic Church to the wide range of protestant denominations that accept an old Earth.
It remains that some denominations and people insist on what they call a “literal” reading of the Bible. In addition, a few very vocal people continue to argue that the Earth looks young. Many more people hear all of this commotion and figure that maybe there is something wrong with the science, because “where there’s smoke, there’s fire.” Other people take it as an element of faith to disbelieve the scientific evidence, and even to accuse scientists of being bad people for opposing the fundamentalist interpretations.
In this course, we go to some length to show you a small bit of the evidence that the Earth does not look young—it bears the marks of a deep and fascinating history. The annual-layer counts by themselves require an old Earth, because the tree rings, the lake sediments, and the ice cores all extend to older than the historical chronologies. The Irish oaks preserve rings from more than twice as many years as Archbishop Ussher of Ireland would have said were possible since Noah's flood, so his prediction was tested, and failed. Geologic and other scientific evidence from tree rings, lake sediments, ice cores, archaeological sites, and more match historical records well as far back as those historical records go. But as we shall see in the next sections, those annual layers and other “young” things are only the tip of a very old, very deep iceberg.
Please note that it is not the author’s intent to insult or belittle anyone’s beliefs here. Science, you may recall, has no way of verifying whether it has learned the truth; it is a practical undertaking designed to discard ideas that fail, save the ones that don’t fail as provisional approximations of the truth, and push ahead. The hypothesis of an Earth that is no older, and looks no older, than historical records, leads to all sorts of predictions. Geologists began seriously testing those predictions in the 1700s, and found that those predictions were not supported, whereas predictions of an old-Earth hypothesis worked well—with very high confidence, the rocks look very old.
Consider two people, Sam and Pam. Sam has decided that belief in his literal interpretation of his favorite translation of the Bible is the most important thing in his life, as it controls the fate of his eternal soul and his relation with the most powerful being in the universe. Is it possible for Sam to look at the rocks, trees, ice and lakes, and find some way to explain what he sees in the context of that literal belief? The answer, obviously, is yes; many people do so, and some of them will be mad at us for what we write here. Next consider Pam, who is working in an oil-company laboratory trying to refine dating of petroleum generation and migration. Which works best for her in making sense of the sedimentary record, Sam’s interpretation or that of the geological profession? The answer is equally clear; Sam’s view is unhelpful, and geology works. Finally, ask whether it is possible for Sam to be a geologist and use the old-Earth tools even as he believes in his book, or whether it is possible for Pam to be a religious believer while doing geology, and the answers are yes; some people can hold a variety of ideas in mind at the same time. But recognize that the scientific evidence for an old Earth (and later, for evolution) is about as clear as science gets, and that the level of scientific disagreement on these issues is about as low as disagreement ever gets in science. Within the scientific community, the argument about whether the Earth really is older than historical records is akin to the scientific argument about whether the Earth is roughly spherical. (Lively discussions clearly continue in the blogosphere and in other many non-scientific circles, but those discussions are at best rather weakly linked to the science.)
The Grand Canyon is a mile-deep, 18-mile wide, 277-mile long (1.6 km x 29 km x 446 km) gash in the Earth. The colorful spires, the rocky cliffs, the hidden pocket canyons, the pristine springs laying down lovely deposits, the roaring thunderstorms and arching rainbows are to many the quintessence of the U.S. West. The Grand Canyon is neither the deepest nor the steepest canyon of the planet, but the Grand Canyon indeed is grand, and defines “canyon” for many people.
When the author, his cousin Chuck, and his sister, Sharon, were hiking the Bright Angel Trail from the North Rim into the canyon, a snake crossed the trail and slithered into some dry grass just at the trail edge. Chuck and I, in the lead, could see quite clearly that this snake ended in a “harmless” tail. Sharon, just behind, was not aware of the snake until it stuck its head out and rattled the grass just at her feet. Deciding that discretion was the better part of valor, and that if it rattles like a rattler it might actually be one, she made one mighty leap backward, landing in a cloud of dust on a switchback below.
Sharon almost certainly was not concerned with the rocks about her at that instant, but she had leaped backward through history. And what a history it is.
At the bottom, the inner canyon is cut through the Precambrian Vishnu and Brahma Schists. The older Vishnu has the appearance and chemical composition of metamorphosed sediments. The lava flows of the Brahma preserve the pillow structure of submarine eruptions, but the interbedded volcanic airfall material shows that at times the region was exposed as dry land. The total thickness of three miles of lava flows and interbedded layers, now standing almost on end although they initially were deposited almost horizontally, speaks of an important, long-lasting interval of deposition.
These oldest lava flows and sediments of the Grand Canyon have been "cooked," and are now of metamorphic types that form only in the hearts of mountain ranges at very high pressures and temperatures. During and after the metamorphism, melted rock (magma) squirted into cracks in these rocks, and then froze to form the pretty pink Zoroaster Granite. Yet this whole package of rocks was then brought back to the surface as the rocks of the mountains above them were eroded, with the erosion producing a very smooth, nearly horizontal plain on top of them, and weathering/soil formation causing changes that extend deep into these rocks.
The sea next advanced across this plain, first picking up and carrying and rolling pieces of the rocks and soils on the erosion surface to form a conglomerate, then giving way to sandstones, shales, and limestones that piled up to a thickness of two miles or so. (Such a great thickness does not mean that the sea was two miles deep; rather, in this case, the water stayed relatively shallow, but the warping of the crust by the drifting plates and other processes caused the sea floor to sink as the muds and other deposits piled up.) These rocks include mud cracks, ripple marks, casts of salt crystals that formed when the sea water evaporated in nearshore environments, and stromatolites, which are algal-mat deposits in which the algae trap mud, grow up through it, and trap more mud. All of these are similar to modern features, and indicate gradual accumulation (a layer, then drying for mud cracks, then more mud, then ripples from water flow, then drying for salt casts, and on and on and on).
Death-Valley-type pull-apart faulting then dropped and rotated these layers, so that they now slant (see the figure below). Long-term weathering and erosion then occurred, leading to a plain broken by a few higher points where especially resistant rocks did not erode as rapidly. Again, deep weathering speaks of long exposure. In some places, the sediments were entirely removed down to the metamorphic rocks beneath, but the sediments are preserved where they were dropped by faulting.
The sea returned, again reworking materials on the erosion surface to make a basal conglomerate, followed by beach sandstone, then offshore shale, and limestone from farther offshore. As the sea deepened and the beach moved towards the land, shale was deposited on sandstone and then limestone on shale. These three layers form the Tonto Plateau that is so evident on the south side of the canyon. The rocks of the Tonto Plateau include fossils of marine animals such as trilobites, and even trilobite tracks. Again, all evidence is of deposition by processes just like those operating today, over long periods of time. A layer with a trilobite track must have been exposed long enough for a trilobite to crawl across it. The thousands and thousands of different layers in the rocks, with ripples and tracks and fossils, indicate long times.
Time then passed of which we have no record in the Grand Canyon, except that stream channels were carved on top of the limestone, indicating that the region was raised out of the sea and erosion was occurring. Fossils from two of the periods of the Paleozoic are missing, indicating that much time passed. When deposition resumed, the first rocks put down were limestones in the stream valleys, but another time gap sits on top of those in-the-channel rocks. The limestones in the channels include coral and shellfish (brachiopod) fossils, and plates from armored fish.
The marine Redwall Limestone was deposited next, so-named because it makes a red wall. The limestone is gray, with the red (rust and clay) dripping down from red rocks above. The Redwall Limestone contains fossils of corals, sea lilies (crinoids) and shellfish (brachiopods), but with notable differences from the fossils of those general types found in limestones below, and both sets of fossils differ from those in limestones above. The Redwall Limestone contains caves and sinkholes, which in turn contain sediments associated with the rocks above. Caves generally form on land or possibly very close to land under shallow water, not beneath the open ocean, so the rocks were lifted near or above sea level and eroded after Redwall deposition.
Then, the sea flooded in, at least in the region that would become the western part of the Canyon, and deposited the Surprise Canyon limestone in erosional stream channels in the top of the Redwall. These rocks were not even described until the 1980s, and are only reachable by helicopter or arduous climbing. These rocks are not indicated in the diagram, above, which is what you would see on the Bright Angel Trail in the central Grand Canyon. Erosion cut the top of the Surprise Canyon before deposition of more layers.
Next are sandstones, siltstones and shales with plant fossils, footprints, etc., at various levels through the rocks, indicating deposition on land in floodplain conditions. Insect fossils appear on the upward trip through the rocks, and then great dunes with sand-blasted, wind-frosted grains and occasional lizard footprints. You might imagine the sand dunes of the Sahara spreading across the flood plain of the Nile for these rocks. Marine conditions then returned, providing mostly limestones with sponge fossils and shark’s teeth as well as corals, crinoids and brachiopods, finally reaching the top of the canyon.
If you're on the North Rim of the Canyon, gaze farther north. The rocks you're on slant downward to the north, and you are looking at rows of cliffs with younger rocks, up through the cliffs of Zion from the age of the dinosaurs, up through the lakes of Bryce, up and up and up until finally you reach the trees and Native American sites older than the historical chronologies of Archbishop Ussher.
(By the way, if you’re interested in the carving of the Grand Canyon, have a look in the Unit 10 Enrichment.)
A pile of rocks like those in the Grand Canyon does not reveal its age easily. But we have evidence of seas, mountain building, mountain erosion, more seas, more distortion, more erosion, and more, and more, and more. The rocks involved are old friends—similar things are forming today. Using the principle of uniformitarianism—the present is the key to the past—we can make some estimate as to how long events take. The schists at the bottom were buried miles deep in mountain ranges and later brought to the surface by erosion, and fast erosion rates require a million years to strip off a mile, for example.
The geologists of the 1700s, working primarily in Europe, pieced together stories such as this. They tried to estimate the times involved. One difficulty was that they could not tell how much time was in the erosional time gaps, or unconformities—was erosion fast, or slow? And they could not really unravel all of the stories in the oldest rocks because metamorphism had erased some of the stories.
These early geologists eventually estimated that the rocks told of events that required AT LEAST tens of millions of years to hundreds of millions of years. Just depositing the sedimentary rocks would take about that long, with much more time represented by the unconformities and the oldest really-messed-up rocks. This is deep time—the Earth is not the historical thousands of years, or even the ice-layer/tree-ring tens of thousands of years. History was written and trees grew on the relics of vastly greater histories. Looking into that history is one of the great joys of geology, but it brings us to the edge of a mental cliff from which some people do not wish to look. In the next section, we will see just how high that cliff really is.
The techniques of layer counting and uniformitarianism are useful in dating, but the real workhorse these days is radiometric or radioactive dating. The Earth is host to a wide range of naturally occurring radioactive elements. An atom of a radioactive element eventually will spontaneously change to some other type of atom, by emitting radioactive energy.
Radioactive decay occurs in various ways. The easiest to understand is when a nucleus splits into two parts, kicking out a part of itself. Remember that heat causes molecules in water to bounce around and occasionally evaporate; atoms or molecules in rocks are also bouncing around, but are so tightly bound that very, very few break free at the Earth’s surface. In somewhat the same, way, the protons and neutrons in the nucleus of an atom are always wiggling and bouncing around; most nuclei are so tightly bound that this wiggling doesn’t change anything, but some types of nuclei are weakly enough bound that occasionally some protons and neutrons “evaporate.” We call those types of atoms that “evaporate” radioactive, and those that do not stable. (A real nuclear physicist would probably yell at us because we oversimplified a bit too much, but this should do for introductory geology.)
Commonly, a nucleus that “evaporates” emits a group of two protons and two neutrons, which is the nucleus of a helium atom and also is called an alpha particle, for historical reasons. Emitting an alpha particle changes the remainder of the nucleus to the type of atom that is two protons lighter. Other types of radioactive changes also occur, including splitting of a nucleus into nearly equal-sized chunks, change of a neutron to a proton plus an electron that is emitted, or capture of an electron by a proton to change into a neutron. All of these change the type of atom from one element to another. All are explainable by well-known physical principles, and all are as natural and regular as the downward fall of your pencil if you drop it off your desk.
The behavior of any one atom is not predictable, but in large groups the average behavior is easily predictable. The basic rule is that, if you watch for some specified time such as one hour, the more radioactive atoms you start with, the more radioactive atoms you will see change. If you start your stopwatch when you have some number of a given type of radioactive atom, and stop the watch when half have changed, you will have estimated the half-life of the radioactive type. Each radioactive isotope has a distinctive half-life, which can be measured in the laboratory. (Note that you do not need to wait for an entire half-life to measure it; the Enrichment section shows mathematically that you need to wait only long enough for enough atoms to change to be measured accurately—the rule that more change in some time if you start with more is all you need.)
Suppose you start with 2000 atoms of the parent type. These decay into offspring (most textbooks refer to these offspring as daughters). After one half-life, 1000 parent atoms remain and 1000 offspring have been produced. After another half-life, half of those 1000 parent atoms have changed to offspring, leaving 500 parents and giving 1000+500=1500 offspring. After a third half-life, half of the remaining parents have changed, so that now only 250 parents remain and 1500+250=1750 offspring have been produced. During the fourth half-life, half of the remaining parents decay, leaving only 125 parents and giving 1750+125=1875 offspring.
Now, we really need to deal with large numbers, so add ten zeros to the end of each of the numbers in the previous paragraph. Such numbers of radioactive atoms are common in even relatively small samples of rock; the total number of atoms in a fist-sized chunk of rock is about 1 followed by 24 zeros.
As noted, there are many different parent types with different half-lives. Some half-lives are very short—seconds or less. Others are very long—billions of years or more. Some of the radioactive parents are left over from the explosions of stars that produced the stuff of which the Earth is made. Other radioactive parents are created by cosmic rays that strike atoms on Earth. Some radioactive decays produce offspring that are themselves radioactive parents for a further generation, and several such decays may be required to produce a stable offspring. And radioactive decays may damage neighboring atoms, producing radioactive types.
Consider the example of potassium-40 and argon-40. Argon-40 has 18 protons and 22 neutrons in its nucleus, for a total of 40 particles. Potassium-40 has 19 protons and 21 neutrons, also totaling 40. Potassium-40 is a parent with a half-life of 1.3 billion years. Potassium is abundant on Earth, and occurs in many common minerals, and some of the potassium is the radioactive parent potassium-40. The offspring, argon-40, is a gas. If lava flows out on the surface of the Earth, the argon escapes. Thus, a lava flow will start with some parent potassium-40 but no offspring argon-40. As time passes, the potassium-40 breaks down to argon-40, which builds up in the rock. If today the rock has as many potassium-40 as argon-40 atoms, then one half-life has passed and the rock is 1.3 billion years old. Whatever the ratio is, the math is not that difficult and gives the age.
It is possible, of course, for argon-40 to leak out of the mineral. If it does, we will think that the rock is younger than it really is. But if leakage is occurring from a mineral grain, then the outside of the grain will contain less argon-40 than the inside does, and this can be measured, revealing the problem. A mineral grain that grew in slowly cooling melted rock far down in the Earth and that then was erupted may have begun trapping argon-40 before the eruption occurred, in which case the age obtained will be the time when the grain started growing rather than the time when the eruption occurred. Scientists do not blindly apply dating techniques; they think about what is being measured, and apply a little common sense.
But isn’t it possible that the radioactive “clocks” ran at a different rate in the past? Some of the most free-thinking physicists have suggested slight changes in physical “constants” over time; couldn’t that affect the clocks? A critical difficulty with this seemingly “easy” idea is that the forces controlling the stability of atomic nuclei (hence the rate of radioactive decay) are the forces involved in all sorts of other processes including energy generation in the sun and other stars, so if you tweak the “clock” very much, you have turned off or blown up the sun, and we know that has not happened.
There is an easier argument against changing decay rates. The techniques of radiometric dating have been tested against layer-counting ages and historically documented ages of events, and found to agree beautifully. Radiometric dating techniques also have been tested against uniformitarian results, and found to be fully consistent. The techniques have been tested against each other—one sample can be dated using several different parent-offspring pairs—and good agreement is found. (Short-lived parent-offspring pairs are used to date young things—in old rocks, all of the parents are gone—and long-half-life pairs are used to date old things—in young things, the number of offspring is too small to allow accurate measurement—but enough different types exist that intercomparisons are still possible.)
You won’t have to look very far on the web to find sites—usually attached to fundamentalist Christian ideas—complaining about errors in radiometric dating. (And Dr. Alley was once shown a published tract pointing out how stupid Dr. Alley himself must be to think that he could count more annual layers in an ice core than the total age of the Earth!) Some of the objections to radiometric dating are fairly silly, and even some of the young-Earth sites have put up notes asking followers to avoid using certain common arguments against scientists because those arguments are just wrong. The 5000-year-old living clam falls in this category. (See the Enrichment if you want to get the low-down on “old” living clams.) The bottom line is that radiometric dating is useful, practical, successful, matches written records as far back as they go, matches other indications beyond that, and reveals a deep and fascinating history. Radiometric dating is not perfect, it does include errors, practitioners have to know what they’re doing and think about it, but it works.
The oldest rocks found on Earth are about 4 billion years old, and some of those contain mineral grains recycled from even older rocks. The active Earth has almost certainly erased the record of its very earliest rocks. Meteorites probably formed from the solar nebula at about the same time as the Earth did, and since have fallen on the Earth. The oldest meteorites are about 4.6 billion years old, and that is our best estimate for the age of the Earth. Careful analyses of the changing lead isotopic ratios over time (from decay of uranium) also yield that number for the Earth. And 4.6 billion years is, indeed, is deep time.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 10.
Click on the first thumbnail below to begin the slideshow. To proceed to the next image, move the mouse over the picture until the "next" and "previous" buttons appear ON the image or simply use the arrow keys.
Word Document of Unit 10 V-trips
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Great Basin National Park
(Provided by USGS)
Grand Canyon National Park (South Rim)
(Provided by USGS)
Grand Canyon National Park (Historic - Powell Survey 2nd Expedition, 1871-2)
(Provided by USGS)
This week's GeoMations feature geologic time in all of its glory; first, you'll accompany Dr. Alley on his "climb" up out of the Grand Canyon and past 1.2 billion year's worth of geologic evidence. After reaching the top, you'll stop at the rim to hear a bit about the widening and narrowing of the Canyon. Then, it's back to Happy Valley, where you and Dr. Alley will hit the gridiron and witness geologic time as it marches down the field for a dramatic goal-line ending!
Next, this week's GeoClip will take you "live" to the Grand Canyon Rim, where you will join Dr. Alley in a firsthand look at "deep time." (If that clip leaves you wanting more, "part 2" is also available as an optional enrichment this week!) So, enjoy your visit to the Grand Canyon and your walk up through time. We hope you find Dr. Alley's play-by-play commentary and his incisive post-game analysis helpful in explaining what the Earth has been doing these past 4.6 billion years.
The vast and varied history of the planet is best experienced by hiking the Canyon, but you can see a lot of the story from the rim. Here, you'll get a very brief overview of a very long story. ( Click the link to view movies - All are in QuickTime format )
Optional Enrichment (no, these won't be on the quiz!):
Supergroup Part 2: Grand Canyon Rim: (Transcript)
Deep Time film clips - What do beauty, saving money at Las Vegas, religion, oil exploration, emerging new diseases, and the planet’s recovery from global warming have in common? All in some way involve deep time, the immense age of the Earth. Eric Spielvogel filmed a discussion of these and other issues with Dr. Alley, for a special “time” issue of Research! Penn State. These "Deep Time film clips" will give you something to think about, and may even help with the course. Enjoy!
The Unit 10 lecture features Dr. Richard Alley and is 48:30 minutes long.
Check out the Unit 10 Presentation used in the online lecture.
A vigorous river is capable of cutting downward at 1 mm/year (or more, and glaciers may cut faster than that). At 1 mm/year, it takes 25 years to cut an inch, or only about 1.6 million years to cut a mile down and make the Grand Canyon. Usually, rivers don’t cut as fast as 1 mm/year because the rivers quickly get down close to sea level, which makes the river’s slope smaller and slows the erosion.
The Grand Canyon likely owes its existence to several events. The most important may have been the opening of the Gulf of California, south of Death Valley, as we saw way back at the beginning of making mountains. Opening of the Gulf of California brought the ocean closer to the mountains, which steepened the streams heading toward the Gulf—the height of the mountains wasn't changed by opening the Gulf, but the horizontal distance a river had to flow from the mountains to sea level got shorter as the land ripped open and the ocean extended over what had been river bed on land, so the average slope of the river increased.
A “continental divide” is the line on a map separating the rivers flowing to one ocean from the rivers flowing to another ocean. As you might imagine (and as we discussed briefly back in the history of the closing of the proto-Atlantic and opening of the Atlantic), the slope to one ocean from a continental divide is often steeper than the slope to the other ocean. The steeper side generally erodes faster, which causes the continental divide to move away from the steeper side toward the more-gradual side. (Eventually, this will lead to the slopes being similar on the two sides.)
But, the continental divide is irregular, not a straight line. Where a big river forms and cuts down, the slope from the divide to the river will be steeper than nearby, so erosion will be faster there and the divide will be forced away. Sometimes, this will cause the divide to intersect and “capture” the drainage of a stream that had been on the other side of the divide. (See the figure below.)
The upper panel in the figure shows two rivers (the black lines with arrows), separated by a continental divide (the blue line), viewed from above. If the right-hand river is steeper, then it will erode back until the headwaters of the left-hand river are “captured” in an act of “stream piracy,” as shown by the magenta line.
There are lots of small rivers in the West with fairly big canyons—look at the most-of-a-mile deep canyon of the small Virgin River in Zion, or the remarkable amphitheater that tiny Bryce Creek has gnawed into the Paunsagunt Plateau to make Bryce Canyon. So, when the Gulf of California opened, the ancestral lower Colorado River steepened and cut headward (probably involving a piracy event with a stream exploiting the easily eroded San Andreas Fault), and pirated the ancestral upper Colorado River, which previously probably had drained internally (the river ran out into the desert and evaporated, the way some rivers do in Death Valley). This happened just over 5 million years ago; at that time, chunks of rocks of types that occur only up in the Rockies at the head of the Colorado River suddenly appear in sediments of the Gulf of California, where before chunks of such rock types were absent in the Gulf of California.
Why did it take the Colorado so long to cut the Canyon, then? Well, the Death-Valley-type Basin-and-Range faults associated with the opening of the Gulf of California also have cut across the Canyon, especially in the western end. Basaltic lava has come up some of the faults, much as happened in Death Valley. At several times over the last 5 million years, lava flows have dammed the Canyon, making lakes. Lakes accumulate sediment rather than eroding, while the erosive ability of the river is spent cutting through the dam. So, the river really had to cut down much more than a mile to make the canyon—cut down, get filled with lava, cut the lava, get filled again.
In the text, you saw how radioactive decay occurs and provides “clocks” for the ages of rocks. Here, we go into a little more detail on the math, strictly for your entertainment and enjoyment.
The “law” of radioactive decay says that the more atoms of some radioactive parent you have, the more atoms of that parent will decay in some time. (There are many laws of this type: hotter bodies cool faster, rooms with more cats have cats run out the door faster, etc.). In addition, each radioactive parent type decays at its own particular rate. This is easier to say mathematically than in words. Given N parent atoms of some type, the change dN in the number of that type over some interval of elapsed time dt is:
The minus sign occurs because the number of parent atoms is decreasing over time. The K is a constant, called the decay constant (and often indicated with the Greek lambda, but we’ll stick with K). The numerical value of K is different for each different radioactive parent type, and includes the “physics” of how unstable the parent type is. A large K means a very unstable parent and a very rapid change to offspring; the units of K are inverse-time (so 1/seconds or 1/years).
If you never studied calculus, or you forgot what you studied, you won't make much sense of the next little bit. Don't worry. Those of you who took a calculus course and remember it will know that you can rearrange the equation to obtain:
Integrating yields:
in which ln indicates the natural logarithm, C is a constant that we will determine, and t is the total time that has elapsed. Taking the exponential of both sides, and noting that the decay started at some time t=0 when there were N=N0 parent atoms, yields the standard decay equation:
in which exp indicates the exponential (it usually appears as ex or exp or inv ln on calculators). The negative in front of Kt is equivalent to writing N=N0/exp(Kt). As t becomes large, exp(Kt) becomes very large, so N=N0/exp(Kt) becomes very small—the equation says that after a long time, you run out of parent atoms, which is correct.
Notice that if you can measure N0, wait for some time t1 and then measure N, the only unknown in this standard decay equation is K, so K can be calculated readily. The natural logarithm, ln, reverses the exponential so that ln(exp(-Kt))=-Kt. The natural logarithm appears on most calculators as ln or ln x or possibly as inv exp. Using this,
You usually will see this written as:
using one of the properties of logarithms.
We next estimate the half-life, t1/2. Note that after one half-life, N=N0/2. (So half of the parents have changed after one half-life.) If we let N=N0/2 in the standard decay equation, take the natural logarithm of both sides, remember that -ln(1/2)=ln2, and rearrange, we obtain t1/2=(ln2)/K. This is the basis for the statement in the text that you do not need to wait for a full half-life to pass if you wish to learn the half-life; you just need to start with N0, wait for any time t1, measure N, calculate K from this, and then calculate t1/2 from K. The half-life is useful, but most professionals in the field use the decay constant K most of the time, because K is more “fundamental” (it appears in the statement of decay given first above, and does not need to be derived as for the half-life).
Skeptics about the use of scientific age dating in geology have especially focused on complaining about radiocarbon dating. This focus is odd, because radiocarbon—also called carbon-14—is not used in establishing the age of the Earth, or the age of the main geological events. The half-life of radiocarbon is only 5730 years; samples older than about 50,000 years have nearly run out of radiocarbon and so cannot be dated by radiocarbon. But, radiocarbon is used a lot in dating archaeological sites, and this may have caught the attention of people who study early written histories. In addition, as you will see, radiocarbon is more complex than many others (such as the potassium-argon system discussed in the regular text), and it is easier to argue about complex things.
Much of the complexity of radiocarbon arises because the offspring of radiocarbon (the gas nitrogen-14) is very common, and is not retained well by the samples that are dated using radiocarbon (wood, charcoal, bone, or other formerly living things—not most rocks). Thus, radiocarbon dating does not look at the parent-to-offspring ratio; instead, the starting concentration of radiocarbon is estimated, the concentration today is measured, and the ratio gives the age. Radiocarbon is mostly made in the atmosphere, when cosmic rays collide with atoms and knock off neutrons that then hit nitrogen-14 nuclei and make carbon-14. This doesn’t happen very rapidly; natural production is just about 15 pounds for the whole Earth per year, or just over two carbon-14 atoms per square centimeter (just under 1/2 inch on a side) of the Earth’s surface per second.
In the atmosphere, radiocarbon quickly combines with oxygen to make carbon dioxide. The atmosphere is well-mixed—release some gas molecules here, and within a few years they will be spread fairly uniformly around the planet—so the radiocarbon-bearing carbon dioxide is quite uniformly distributed around the globe. Green plants grow by using carbon dioxide, and roughly one of each trillion carbon atoms in the atmosphere and in green plants is carbon-14 rather than stable carbon-12 or carbon-13. Plants are eaten by animals. Most animals live less than 100 years, whereas most carbon-14 lasts thousands of years, so when plants and animals die, they have just about the same ratio of carbon-14 to carbon-12 as was in the atmosphere when they were still alive. After plants or animals die, they do not breathe or eat any more, so they don’t take in carbon-14 while the carbon-14 in them decays. Hence, the ratio of carbon-14 to carbon-12 in a formerly living material is a clock.
Whole textbooks can be written refining the previous two paragraphs, and a scientific journal, Radiocarbon, focuses almost exclusively on the topic. If you aren’t a real stickler for accuracy—if “this died sometime between 9,000 and 11,000 years ago” is good enough for you—then you really don’t need a whole journal devoted to radiocarbon. (You still need to worry about one or two things that we’ll come to, but not about too many.) But if you want to get the answer right to within a few decades, then you have to be really careful.
One problem is that production rates of radiocarbon have varied over time. When the sun is more active or the Earth’s magnetic field is stronger, they protect us more from cosmic rays and reduce production of radiocarbon. The changes are not huge, and there are ways to correct for them (strength of magnetization can be estimated from the degree of alignment of the “magnets” in lava flows or sediments of different ages, and ice-core concentrations of beryllium-10, which is also made by cosmic rays, can be used to track the sun and the magnetization).
Changes in the Earth’s carbon cycle also matter. For example, now we are pulling up immense quantities of really old fossil fuels and burning them to make carbon-14-free carbon dioxide that goes into the atmosphere, diluting the carbon-14 there. When we humans were busily blowing up atomic bombs in the atmosphere, they made a lot of carbon-14. Before we were so influential, changes in ocean circulation were probably most important—some carbon dioxide goes from atmosphere to ocean, and the ocean waters sink in certain places and spend a millennium or so down deep before coming back up to exchange carbon dioxide with the atmosphere. Because some of the carbon-14 from the atmosphere ends up decaying in the deep ocean, the ocean circulation actually reduces atmospheric radiocarbon—if water didn’t sink into the deep ocean, there would be less carbon-14 there and less carbon-14 decay there, and that would leave more carbon-14 in the atmosphere. At certain times in the past, less sinking of ocean waters seems to have occurred, allowing more carbon-14 to exist in the air.
The usual way to handle all of this is to use radiocarbon to date tree rings (which quit exchanging carbon with the atmosphere as soon as they grow) or shells in annually layered sediments, and use the layer-counted ages and the known half-life of radiocarbon to calculate the starting concentration of radiocarbon. Because radiocarbon is well-mixed in the atmosphere, and must have been well-mixed in the past, once a calibration curve is developed, it can be used for other samples. One can also date some samples, such as corals or cave formations, using accurate techniques such as uranium-series disequilibrium as well as using radiocarbon, and so obtain a calibration curve for the radiocarbon. Many different calibration studies have been conducted, and while they do not agree perfectly and research is ongoing, they agree reassuringly well. The biggest corrections are a bit more than 10%—a sample that looks to be 10,000 years old, assuming that there were no changes in radiocarbon concentration of the atmosphere, is actually about 11,500 years old, because the radiocarbon concentration of the atmosphere did change.
If you are primarily interested in the question “Does the world really look older than written records”, even radiocarbon provides a very good answer (“Yes, with very high scientific confidence”). Science has long since moved past that question, and the research frontier involves numerous fascinating questions, such as whether we can reconstruct changes in ocean circulation from the changing calibration of the radiocarbon clock after correcting for the changes in the sun and the magnetic field.
Plants actually have a slight preference for carbon-12 over carbon-13 or carbon-14 (the lighter atoms diffuse into the plant and react more easily), so the concentration of carbon-14 in a plant is slightly less than the concentration in the air. The preference for carbon-12 over carbon-13 is half as big as the preference for carbon-12 over carbon-14, so measuring the concentrations of all three types allows an accurate correction; this one is now easily done, and need not bother anyone much.
The 5000-year-old living clam raises a different but interesting issue. All of the discussion so far has assumed that the items being dated obtained their carbon from the atmosphere. But suppose that you “ate” only things that had been dead for a long time—you would not have much radiocarbon, and so you would look old to someone who assumed that you ate things containing normal concentrations of radiocarbon. Certain special ecosystems on the sea floor do just that; they live on natural oil seeps, eat the oil or eat things that ate the oil, and the oil is old and so lacks radiocarbon. If you were stupid enough to sample these and assume that they were eating “normal” foods, then you would mistakenly assume that the living creatures had been dead for a long time.
Such oil-seep ecosystems are quite rare and special. A more-common situation is a clam in a creek in a carbonate terrain. When caves are being made, the chemical equation for the water and carbon dioxide dissolving the rock is:
The rain and atmospheric carbon dioxide on the left of the equation combine with the calcium carbonate of the limestone, yielding the calcium and bicarbonate ions on the right-hand side of the equation that are freed to wash down the creek. If a clam is making its CaCO3 shell from the water, the clam just runs this reaction backward. Notice, however, that half of the carbon, C, in the water came from the atmospheric CO2 and half from the rock. The rock is almost always very old, and has no radiocarbon. So, a clam in this situation would form a shell with only half as much radiocarbon as for a clam growing in a stream that does not drain carbonate rocks and that gets all of its carbon from the atmosphere. Hence, if scientists were clever with their instruments but stupid otherwise, those scientists might end up thinking that a living clam had been dead for over 5000 years.
Scientists are fully aware of this. For decades, however, there was a convention of reporting all radiocarbon measurements as the equivalent age assuming that the sample had been in equilibrium with the atmosphere. Dr. Alley is reasonably confident that the myth of the clam that was living yet the scientists thought it was thousands of years old came from work by a distinguished senior colleague, who almost 50 years ago published papers listing dates in the conventional fashion. That colleague actually was using the results to learn about the geochemistry of the waters. As noted above, some of the young-Earth-creationist websites have asked their supporters to “clam up” about this, because using it in an attempt to discredit scientists instead makes the young-Earth-creationists look confused.
You have reached the end of Unit 10! Double-check the list of requirements on the Unit 10 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Continue to work on Exercise #5. See the Course Calendar for specific dates.
Following are some supplementary materials for Unit 10. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
Ah, memory lane. Dr. Alley's best friend in early elementary school had a father who sold "pop" (a.k.a. soft drinks) to dealers in Ohio and often had free samples sitting around, and a mother who drove a 1964-and-a-half white Ford Mustang with a black convertible top. Hard to beat. And what a beauty that car was! Dr. Alley's Hot Wheels Mustang was a 1967, recognizably different from the '65-and-a-half. In fact, he could easily put the Mustangs in chronological order based on the differences between models.
The early geologists of the late 1600s and 1700s had never heard of Ford Mustangs, but those geologists faced a Ford Mustang problem. Recognition of unconformities and other features in the rock record opened a world far older than written records. William Smith had shown the law of faunal succession—that putting the rocks in order put the fossils in order—so biological change had accompanied geological change. But was the biological change gradual, parent-to-child in a great, unbroken evolutionary chain of being? Or did unknown cataclysms, or a tinkering god, or an angry god, repeatedly replace one world with another as Ford would one day replace each model with a new type of Mustang the next year?
Erasmus Darwin was among the evolutionists, and put his ideas in verse in the posthumously published The Temple of Nature (1802):
Organic life beneath the shoreless waves
Was born and nurs'd in ocean's pearly caves;
First forms minute, unseen by spheric glass,
Move on the mud, or pierce the watery mass;These, as successive generations bloom,
New powers acquire and larger limbs assume;
Whence countless groups of vegetation spring,
And breathing realms of fin and feet and wing.
OK, maybe this excerpt wouldn't make a hit song if put to music. But when grandson Charles added observations and mechanism to Erasmus' speculations, the evolutionists won out over the Ford-Mustang "catastrophists." What convinced the scientific community (and a great slice of "polite society") of evolution? We'll try to answer that fascinating question in this lesson.
You will have one week to complete Unit 11. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 11, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 11 Instructional Materials. |
Continue working on Exercise #6: Fossil Fuels (& Driving Hazards) | Yes, this is the last of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 11 "RockOn" quiz | Yes, this is the eleventh of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
Complete the "StudentsSpeak #12" survey | Yes, this is the last of 12 weekly surveys and is worth 1% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 11, including the online textbook, a video lecture, several vTrips and animations, and an overview presentation
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
It is a century now since Darwin gave us the first glimpse of the origin of the species. We know now what was unknown to all the preceding caravan of generations: that men are only fellow-voyagers with other creatures in the odyssey of evolution. This new knowledge should have given us, by this time, a sense of kinship with fellow-creatures; a wish to live and let live; a sense of wonder over the magnitude and duration of the biotic enterprise.
— Leopold, Aldo: A Sand County Almanac, and Sketches Here and There, 1948, Oxford University Press, New York, 1987
For one species to mourn the death of another is a new thing under the sun. The Cro-Magnon who slew the last mammoth thought only of steaks. The sportsman who shot the last pigeon [this is the extinct passenger pigeon] thought only of his prowess. The sailor who clubbed the last auck thought of nothing at all. But we, who have lost our pigeons, mourn the loss. Had the funeral been ours, the pigeons would hardly have mourned us. In this fact, rather than in Mr. DuPont's nylons or Mr. Vannevar Bush's bombs, lies objective evidence of our superiority over the beasts.
— Leopold, Aldo: A Sand County Almanac, and Sketches Here and There, 1948, Oxford University Press, New York, 1987
In south-central Colorado, at Florissant Fossil Beds National Monument, you can visit a unique deposit of fossil trees, leaves, insects, birds, fish and more from about 35 million years ago. At that time, a lava flow dammed a stream to form a lake. Then, repeated volcanic eruptions dropped ash into the lake, making it silica-rich and favoring growth of diatoms, which have silica shells. Huge numbers of diatoms quickly grew in very thin layers on things that fell into the lake, protecting them from decay until they were buried by paper-thin layers of ash and mud. Like flowers pressed in a phone book, the flowers of that ancient time can still be seen clearly. So can the dragonflies, and bees, and mosquitoes. Before the National Monument was established, private collectors could pay a small fee and hunt for fossils there. The author found a few fossil bees and ants and seeds, and a friend found a beautiful cicada. (Taking fossils from the National Monument is strictly forbidden; they belong to all of us!)
There are, however, certain differences between some of those creatures and similar ones that live today. Remember, back at Arches, we learned about the law of faunal succession—when rocks are placed in order from oldest to youngest, the types of fossils in the rocks also fall into order, becoming more like things alive today in younger and younger rocks. The Florissant rocks are relatively young, and their fossils are immediately familiar to modern people, but the fossils are not identical to modern species.
The law of faunal succession suggests the possibility of evolution, but does not prove it by any means. One early theory held that many creations and extinctions occurred over geologic history, something like the history of automobiles. One type of automobile does not give birth to another; new ones are created. But you can place the automobiles in order from oldest to youngest, and the younger they are, the more they look like modern automobiles. This idea was called catastrophism in geology. The fundamentalist interpretation of biblical creation is a one-event form of catastrophism.
Catastrophism eventually lost out scientifically to evolution. Evolution does a much better job of explaining the patterns of fossils and of living things, and of predicting events such as the emergence of antibiotic-resistant microorganisms. The triumph of evolution over catastrophism owes much to biology, and more recently to genetics and molecular biology; the identification of the mechanisms driving evolution was especially important (see below).
As geologists collected more data, they also realized that evolution explained the data and predicted the next discoveries better than did any catastrophist model. The early geologists could see the clear march of types over time, but saw catastrophic elements in the record as well. In many places, geologists would find fossils of one type, and then of a somewhat different type, with no transitions between them. Evolution implies rather gradual change, not big jumps. The early geologists knew, however, that there were big time jumps in the records (remember all of the unconformities—time gaps—in the Grand Canyon sequence). So some of the jumps in the fossil record were related to the incompleteness of the rock record. Other jumps really are related to catastrophic events in the record (again, see below).
Further study has shown that many of the evolutionary changes have been geologically fast (but recall that geological time is so long that this can be biologically slow!) and often localized. Suppose that a few animals of some type colonize a small island. Then, they have babies who have babies who have babies, a generation per year, thousands of generations in a geological eyeblink. If the babies differ by just a tiny bit from the parents, eventually a new type or species may emerge. If that species then succeeds in escaping the island (say, because sea-level fell and the island became connected to the mainland), a new type would appear suddenly on the continent. Sediments from the small island may end up being subducted or otherwise destroyed, but fossils on the larger continent are more likely to be preserved. A small island may support only a few individuals, so there never would be many critters to produce fossils that humans could find. On the continent, the species might flourish and produce millions of individuals that would leave easily collected fossils. Thus, the fossil record would show a sudden jump when the actual process was gradual.
In one famous case (of many), trilobites of the genus Phacops are classified in part by the number of columns of elements in their compound eyes. In marine sediments from the Devonian (the middle of the Paleozoic) of Pennsylvania and Ohio, a species with eighteen columns in its eyes occurs for a while. Then, a time gap or unconformity occurs, from a temporary drying of the sea. When the sea returned and began depositing sediments again, the trilobites that returned with it had seventeen columns in their eyes. Not a huge jump, but an apparently sudden one. But wait—over in a small part of New York, the sea did not dry up. There, you can find the old eighteen-column trilobites, then some with seventeen columns plus a partial column containing a varying number of elements, and finally the seventeen-column trilobites. The generations of trilobites changed gradually, and you can see this where the rock record is complete in a small region of New York. In the bigger areas, the record looks more catastrophic because the seaway was dry and no fossils were produced when the changes were occurring through the generations living in New York.
Some so-called “creation scientists” still argue that no transitional forms are known, and so that catastrophism is accurate. This is nonsense; extremely fine gradations are known in many, many lineages. There is one technical sense in which, in some lineages for which fossils are scarce, there are missing transitional types. Suppose you find young fossil type 1 and old fossil type 0. They differ a good bit—you are missing the transitional form 1/2. Now suppose you find type 1/2. It is your “missing link”. You publish your results in important scientific journals, and wait for the fame and fortune to roll in. (You are likely to wait a looooooong time….) But, while you’re enjoying your discovery, someone argues that you are haven’t really found the missing link, because now there are TWO transitional types missing: 1/4 between 0 and your newly discovered 1/2, and 3/4 between your newly discovered 1/2 and 1. So, you go back to work, and after years of effort, succeed in finding both 1/4 and 3/4. Wow! Now your critics point out that you are missing FOUR transitional forms (1/8, 3/8, 5/8 and 7/8). This can be argued to absurdity; we cannot find remains of every creature that ever lived, because almost all remains of almost all creatures are recycled by the efficient ecosystems of Earth (dead things are food to scavengers, worms, bacteria, fungi, etc.). But in many, many lines now, the gaps are vanishingly small, and the transitional forms very well known.
The gaps in evolutionary lineages are especially well-filled for commonly fossilized types, such as shelly marine creatures from shallow water. Shells are hard and resistant—they’re really rocks already—and so shells are preserved well. Although sediments from deep-water sites tend to be dragged down subduction zones and melted, which messes up old records, the underwater edges of continents are not subducted and often escape obduction for a long while, preserving their records of shelly creatures. And while most of land is eroding, most of the ocean is accumulating sediments.
The fact that most of the land surface is eroding complicates study of the fossils that especially interest people. Think about central Pennsylvania for a moment, where Dr. Alley was sitting when he wrote this. The Commonwealth of Pennsylvania has about 1 million deer, and each buck has two antlers, so Pennsylvania deer drop about 1 million antlers per year. If antlers were “preserved,” then after even a few millennia of this, walking in the state should be really dangerous, and we should hear about all sorts of antler puncture wounds from the billions of antlers scattered over the landscape. Of course, we don’t—mice and porcupines eat the antlers for the minerals in them.
In central Pennsylvania now, the only places you can find sediments being deposited are in the reservoirs (which were mostly built in the 1930s, and so give very, very short records), a very few marshes such as Bear Meadows up the road from Penn State's University Park campus (this marsh formed during the Ice Age and thus is geologically very young), and in a few caves and along a few streams. But, the caves and stream deposits don’t last long—the caves are lost as the surface is lowered, and the streams sweep across their flood plains and move the sediments on. So central Pennsylvania today is not making much of a fossil record.
Despite difficulties such as this, careful study around the world has filled in many of the details of the fossil record, including many “missing links.” (In 2001, for example, road-builders accidentally discovered Riverbluff Cave in Missouri, with loads of ice-age fossils that are offering a new window on that interesting time.) For some types of creatures, such as hominids, fossils are still scarce enough that a new find often makes headlines, and may cause a small change in the prevailing view of evolutionary history. And there was lots of excitement in 2008 when fossils of transitional flatfish were found—Darwin had worried about the way flounders evolved so the adults have both eyes on the same side of their heads, so the discovery of the transitional forms was another in the long string of successes for the theory he advanced. However, you almost never read about the great changes in thinking caused by the latest snail fossil—the record is so wonderfully complete that new insights are much harder to come by than they used to be.
The basis of evolution is diversity. (Modern social scientists and politicians are about 4 billion years behind nature on this one.) We know that kids do not look exactly like their parents—offspring are diverse or different. We also know that kids share more characteristics with their parents than with less-closely-related people from the parent’s generation. That is, we look a lot like our parents, but we are not exact copies. This arises because of genetics; the biological instructions or programs that guide development of an individual are passed down from the parents, but there are many mechanisms active that serve to experiment a little with the instructions between generations.
Suppose that one of these small experiments is successful (say, it gives a young giraffe a longer neck than her neighbors, which allows her to reach leaves that are out of reach of other giraffes). The long-necked giraffe will be better-fed than others, and eventually is likely to succeed in surviving to have babies of her own. Some of those babies will grow a little taller than their mother, some a little shorter, and some the same height as their mother, but the offspring will average taller than the kids of other giraffes lacking the initial, successful change.
Those other giraffes lacking this new development will be less successful, and so will leave fewer babies who go on to have babies. Most populations are small enough that, if one individual is even slightly more successful than others, after a few thousand generations, all the survivors will be related to the one with the successful experiment; if one individual is even slightly less successful than others, after a few thousand generations it will have no survivors. (You can demonstrate this easily using mathematical models, or with greater difficulty by breeding living types such as fruit flies, but both reach the same answer.)
Once all of the members of a species contain the successful experiment, the species has been changed a little. But over those thousands of generations, other “experiments” are conducted, some successful and some not. The slow accumulation of the successful experiments is evolution. The mechanism by which the changes accumulate is called natural selection—beneficial experiments allow more survival and reproduction and so are preserved and multiplied. When enough changes have accumulated, we say that a new type or species has emerged. (If a population is split into two or more parts, those parts are called new species when they no longer can interbreed.)
Notice that things that happen to adults, such as having their ears pierced or their behinds tattooed or stretching their necks to reach leaves, are not passed on to children. The changes that are passed on occur during reproduction. Sex helps generate new combinations of genetic instructions. Even species that reproduce asexually by splitting in half have ways (proto-sex?) to exchange genetic material. Sometimes, accidents occur owing to radioactive decay or toxic chemicals damaging the genetic instructions in an egg or sperm or asexually reproducing creature; however, these often are changes that hurt rather than help.
More importantly, the mechanisms of reproduction do experiment a little by moving a few things around in the genetic instructions during reproduction. Some species, and some individuals of species, conduct more experiments than others. Overuse and misuse of antibiotics by humans are producing antibiotic-resistant disease-causing organisms. These antibiotic-resistant types are often those that experiment a lot during reproduction, and so were lucky enough to quickly find an experiment that allows survival despite antibiotics.
The virus that causes AIDS is especially hard to “beat” with a vaccine or antiviral drug because the virus experiments a huge amount. This is costly to the virus—many of the experiments are failures, which means those offspring don’t succeed. But, this high rate of experimentation allows the virus to respond quickly to challenges such as new drugs or vaccines by producing offspring with new ways to defeat those drugs or vaccines. In an AIDS patient, the virus infecting the spleen often differs from the virus infecting the liver—the virus has evolved in the person to succeed in the chemically different environments of the different organs. And, given that the AIDS viruses in just one person are so diverse, it is not surprising that the viruses in different people are different. The remarkable advances in molecular biology allow these changes to be measured now.
Evolution is a well-tested, well-established scientific theory. It makes predictions that are borne out. Partial speciation has been achieved in the laboratory in fast-breeding types such as fruit flies. The geological evidence of gradual changes is strong, and becoming steadily stronger as more and more samples are collected.
Evolution is also being used routinely in science. A search on the ISI “Web of Science” in July of 2012 revealed over 3000 scientific papers with the subject "evolution and antibiotic resistance," with an ongoing rise in the number of papers on the topic. A quick perusal of the titles and abstracts of many of those papers revealed that, as microbes evolve to defeat our antibiotics, the scientists who are trying to keep us alive are using the tools and language of evolutionary biology. Antibiotics are quickly losing effectiveness against evolving microbes, more and more people are dying of infections picked up in hospitals, and scientists are increasingly focusing on the problem, informed by a full understanding of evolution, its rates and processes.
Computer scientists even use evolution—some “artificial-intelligence” approaches have been patterned after the natural processes of evolution. Techniques such as genetic algorithms or evolutionary computation successfully solve complex problems, in essentially the same way that nature does. (For more on this, see the Unit 11 Enrichment.)
In the U.S., some groups continue to oppose evolution based primarily on religious grounds. This opposition has the good effect of keeping the experts “on their toes”—the experts work harder and do better science. This opposition has the unfortunate effect of convincing many people that something is fundamentally wrong with evolutionary theory: again, perhaps figuring that “where there’s smoke, there’s fire.” Many people believe, for example, that evolution is somehow anti-religious, when the majority of church members in the U.S. belong to denominations that endorse evolution as the best description of how the biological world works. Evolution is consistent with the major religions on Earth, and even with rather strict readings of the Christian Bible. The idea that the Earth appears young, and was created recently with all of the modern types of organisms present, was tested in the 1700s and 1800s and proved wrong.
A longer discussion of some issues—evolution and religion, second law of thermodynamics, intelligent design, etc.—is given in the Enrichment. We strongly suggest that if you are interested in this topic, you read the Unit 11 Enrichment.
Dinosaur National Monument lies in western Colorado and eastern Utah. The key rocks were deposited in swamps and and along rivers during the Jurassic in the middle of the Mesozoic, and are called the Morrison Formation. Before the modern Rockies were raised, sluggish streams flowed across the basins of this region, with numerous low, wet floodplains. Dinosaurs flourished. After some died, their bodies were washed up on sandbars, where their bones were buried before they were consumed by scavengers and gnawers. Over time, minerals carried in groundwater reacted chemically with the bone, depositing silica. (For a little more on petrification, see the Unit 11 Enrichment—no magic is involved!)
After the bone was turned to stone at what would become Dinosaur National Monument, the rocks of the region were raised and tilted during the mountain-building that formed the Rockies. Streams, including the Green River, cut through the rocks. The Canyon of Lodore on the Green is a favorite destination for serious white-water rafting. The first scientist in the region was John Wesley Powell, who went on to run the Grand Canyon. In 1909, workers from the Carnegie Museum of Pittsburgh found Dinosaur Ledge, a sandbar-turned-to-stone on which many dinosaur bones had been deposited and fossilized. Today, some of those petrified bones are on display in the Carnegie and in other great museums, but many of the bones have been left in the ledge to be viewed in the park (see the picture above).
The dinosaurs were the dominant large animals on Earth for over 100 million years. Many were quite small, but some were gigantic. They included large plant-eaters and large meat-eaters. Some spent at least part of their time flying or gliding, and others swam.
Mammals co-existed with the dinosaurs for most of the dinosaurs’ existence. However, almost all of the mammals remained small creatures—they generally could not outcompete the dinosaurs for the big-creature jobs.
Extinction is a normal process. A new species may arise and be more successful than an existing type, pushing the old type to extinction. Diseases, accidents, or other events may kill an entire population. And, extinction is forever. As populations vary owing to random factors, sometimes the population drops to zero. But when the population hits zero, the species can never come back up—you can’t just borrow a few creatures from a future generation and bring them back to fill in the gap. So all sorts of random events cause extinctions, and most of the species that have lived on Earth have become extinct. (This characteristic of extinction, that when you hit zero, you're gone, also applies to gamblers at casinos. To see why you’re likely to lose if you gamble, roll on back to the Unit 11 Enrichment.)
About 65 million years ago, at the end of the Cretaceous Period of the Mesozoic Era and the start of the Tertiary Period of the Cenozoic Era (the K/T boundary, because K is used for Cretaceous and T for Tertiary), all of the living dinosaurs died out suddenly. At the same time, many other types became extinct—more than half of the species known from fossils near the end of the Cretaceous became extinct at the end of the Cretaceous. Because survival of even a few individuals from a species can allow the species to persist, it is likely that almost all of the living things on the planet were killed. It was a catastrophic event, one of the most catastrophic in the history of the Earth.
The solution to this puzzle—how the dinosaurs and others were killed—was not found until fairly recently. At the K/T boundary, sedimentary rocks around much of the world contain a thin clay layer. This layer is rich in iridium, an element rare on Earth but common in meteorites. This clay layer contains bits of rock that were melted and refrozen rapidly to form glass, such as are produced by meteorite impacts. Quartz grains in the layer contain shock features, which are caused by very high pressures applied very rapidly, but by no other known mechanisms such as volcanic eruptions. The layer is rich in soot (black carbon) from fires. The layer is thicker in and near the Americas than elsewhere. Around the Caribbean, the layer includes a deposit of broken-up rock such as would be produced by a huge wave. And on the Yucatan Peninsula is a large crater, the Chicxulub Structure, that is dated to the K/T boundary. The crater is partially buried by younger rocks, but easily detected using geophysical techniques, drilling, etc. The crater is at least 110 miles (180 km) across, and perhaps as much as 180 miles (300 km) across.
This evidence indicates that a large meteorite, perhaps 6 miles (10 km) across, hit the Earth (the hole or crater made by an energetic projectile is usually a whole lot bigger than the projectile). Such a collision would have released more energy than all of the nuclear bombs that were on Earth when the U.S. and Soviet arsenals were at their largest.
Debate continues on exactly how such a meteorite would kill things, but it looks like fire and ice, and maybe acid. The impact would have blasted huge amounts of rock, from the meteorite and the Earth, into the stratosphere or above. As this rock fell back to Earth, friction with the air would have generated heat in the same way that a re-entering space capsule or a “shooting star” is heated. For a little while, the air would have been a toaster-broiler oven, cooking and burning everything beneath.
Following that, cold probably descended. The impact site included sulfur-containing rocks. The heat of the impact would have vaporized those rocks, and that vapor would have cooled later to form clouds in the stratosphere. The small particles of these clouds wouldn’t fall fast enough to heat up much; modern space capsules and meteorites fall fast enough to get hot, but raindrops and dust particles do not. But many, many small particles would block part of the sunlight and cool the Earth. We know that such cooling occurs with modern volcanic eruptions—big ones such as Mt. Pinatubo in 1992 cool the Earth by a degree or two for a year or two. A nuclear war might do much more, creating nuclear winter or at least nuclear fall. Even more of the sunlight would have been blocked after a huge meteorite impact, and the world may have frozen for a few years.
The sulfur particles, when they fell, would have made sulfuric acid, giving much stronger acid rain than the recent human-produced pollution. The incoming meteorite may have heated the surrounding air enough to burn the nitrogen in it, forming nitric acid that also would have produced acid rain.
The meteorite impact was not nearly big enough to roll the Earth over, notably move the orbit, rearrange the continents, or anything similarly cataclysmic for the physical behavior of the planet beyond the few years or decades of the heat, cold and acid, but the event was cataclysmic for life—almost all of the living things on Earth would have died. Who would have survived? Plants with long-lasting seeds, hibernators, things that live in ocean sediment or along spreading ridges, scavengers, probably some others with appropriate characteristics. The general pattern is that the surviving animals were small, and mammals did better than dinosaurs. (Although, yes, birds are a branch of the dinosaurs, and the birds are still with us.)
After the fire and ice were over, the “jobs” (ecological niches) of many of the dinosaurs were left open. There were no big plant eaters, or meat eaters, left. Over tens of millions of years, the mammals, freed of the competition from the dinosaurs, slowly came to take over the jobs of the dinosaurs. Some of the larger offspring of some species were successful, although most mammals remained small (mice, voles, etc.). Lining up the fossils over time, we see an evolutionary shrubbery—lots of branches, many extinctions where those branches were cut off, persistence of small creatures, but appearance of some large creatures, with those leading to modern lions and tigers and bears—and people. Biodiversity was hugely reduced by the meteorite, but over the next millions of years new species appeared a little more often than existing species went extinct, so that diversity increased back to more-or-less what it was before the meteorite, just with different species doing the jobs.
If today, Coke and Pepsi and all other soft-drink companies suddenly magically disappeared, new soft-drink companies would be started fairly soon, not because of a magical affinity for soft-drink companies, not because soft-drink companies must exist, but because we usually figure out how to take advantage of opportunities. In the same way, wiping out dinosaurs opened up a space for mammals.
Meteorite impacts have been happening throughout Earth history, and a Mars-sized body colliding with the Earth and blasting things into space is the best explanation for the formation of the moon. Big impacts were common early on. Evan Pugh Professor Jim Kasting of Penn State helped show that the heat from many of the early, huge impacts would have been enough to evaporate the whole ocean during the first few hundred million years of the planet, and that it was probably only about 3.8 billion years ago when the last impactor hit that was large enough to evaporate the sunlit upper layer of the ocean. Since then, collisions have been much smaller. The dinosaur-killer was larger than any that happened since, and than any for a long time before, but was not nearly big enough to evaporate much of the ocean.
There still are many large rocks out in space that go whizzing by the planet, and large impacts remain possible. One scientific estimate found that your chances of being killed by a meteorite impact are about the same as being killed in the crash of a commercial airliner. Commercial-airliner crashes kill a few hundred people per decade, and a meteorite might wait ten million years and then kill hundreds of millions of people, so the statistics are not exactly comparable, but the number is interesting. (In comparison, car crashes kill waaaaay more people, and far out-do tornadoes and hurricanes and earthquakes and food poisoning and bee stings and airliner crashes as killers of people in the developed world. In the short term, the most dangerous thing you do is probably driving a car. In the long term, smoking, over-eating, under-exercising, and other poor health habits are much more important.)
Scientists are coming up with ways to divert asteroids, and are looking for the asteroids, to help avoid such collisions. If we see an asteroid coming from far enough away, then we need to turn its path only a tiny bit to miss the Earth. One idea is to hit the asteroid with a bag of dust that would spread across one side, changing the reflectivity of the surface; the difference between reflecting and absorbing the sunlight would cause a tiny push that would steer the asteroid. Another idea is to send a spacecraft to sit next to the asteroid for a year; the tiny gravity of the spacecraft, tugging on the asteroid, would turn it a tiny bit. The fate of the dinosaurs has stimulated much of the interest in this, so some day we may be saved by paying attention to the big beasts of the past.
Meteorites are not the only ways to cause extinctions. A meteorite ended the Mesozoic age of dinosaurs and ushered in the Cenozoic age of mammals, but there doesn’t seem to have been a meteorite at the even bigger extinction that ended the Paleozoic age of shellfish and brought in the Mesozoic age of dinosaurs. The leading hypothesis for that extinction, still “hotly” debated, is that immense volcanic outpourings (the biggest known hot-spot head, if you remember far enough back to flood basalts and hot spots) over a million years or so released enough CO2 to make the climate really hot, and also produced easily weathered rocks that broke down and supplied fertilizer to the ocean. The warmer ocean held less oxygen because heating drives gases out of water; plants growing in the ocean released oxygen near the surface to go into the air just like always, but when these abundant plants died and sank into the low-oxygen deep layers of the ocean, decay used up the oxygen there and made "dead zones." The volcanic outpouring would have supplied sulfur as well, and with the oxygen gone that would have led to an ocean containing hydrogen sulfide, which causes the smell in rotten eggs and which is highly poisonous to most things, and some of that hydrogen sulfide would have escaped to the air. Special molecules from photosynthetic bacteria that use hydrogen sulfide are found from rocks of that age. We will discuss human-caused global warming next week, but don't panic, we do NOT expect it to cause the ocean to belch out deadly gases that kill most of the life on Earth!
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 11.
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Dinosaur National Monument
(Provided by UCGS)
This week's lone GeoMation features Dr. Alley's visualization of the evolutionary process. Two additional GeoClips discuss "Ancient Bees" and "Extracting Fossils." Hope these add some insight into your Unit 11!
The fossil record includes some amazing things. If there were trees and insects far back in time, wouldn't you expect that the insects would have burrowed into the trees then as they do now? And wouldn't you expect that a tree with some of those burrows would be fossilized? Well, here is one example. Park Paleontologist William Parker of the Petrified Forest National Park explains fossil burrows to the CAUSE team.
Petrified Forest National Park is best-known for trees turned to stone, but also has an immense wealth of fossils of various types from the late Triassic (in the Mesozoic, about 210 million years ago). Here, Park Paleontologist William Parker and assistant Randall Irmis explain to the CAUSE class the paleontological excavation of plates of the armored amphibian known either as Koskinonodon or Buettneria.
Optional Enrichment (no, these won't be on the quiz!):
Agate Chunks in Sand - The formation of fossils is both rare and normal—most dead things are eaten, burned, or otherwise recycled before they are turned to stone, but over the diverse environments of the planet, conditions favoring fossilization are bound to occur in some places at some times. Here, Irene Meglis and Dr. Alley use a little geological sleuthing to understand why the fossil trees of Petrified Forest National Park were preserved. (Transcript)
Deep Time film clips - What do beauty, saving money at Las Vegas, religion, oil exploration, emerging new diseases, and the planet’s recovery from global warming have in common? All in some way involve deep time, the immense age of the Earth. Eric Spielvogel filmed a discussion of these and other issues with Dr. Alley, for a special “time” issue of Research! Penn State. These "Deep Time film clips" will give you something to think about, and may even help with the course. Enjoy!
During three weeks in May 2004, two hardy Penn State geoscientists traveled through 12 stunning National Parks of the southwestern United States with 13 lucky students. The trip was sponsored by CAUSE (Collaborative Active Undergraduate Student Experience), an annual course offered by the College of Earth and Mineral Sciences. Richard Alley, Evan Pugh professor of geosciences, led the expedition. CAUSE 2004 was an extension of his course, "Geology of National Parks," and allowed students to interact with and learn from the rocks and landscapes of Arizona, Utah, and Colorado. In the following videos, Alley explains the concept of deep time, how it tells the history of our planet, and how it affects our lives.
Richard Alley, Ph.D., is Evan Pugh professor of geosciences in the College of Earth and Mineral Sciences, rba6@psu.edu.
—Emily Rowlands
The Unit 11 lecture features Dr. Richard Alley and is 50 minutes long.
Check out the Unit 11 Presentation used in the online lecture.
Many very good sources are available on evolution. The interested reader may wish to start with Teaching About Evolution and the Nature of Science (1998), National Academy of Sciences, National Academies Press, Washington, DC.
What follows, in question-and-answer format, is a synopsis of some of the objections that the author, Dr. Alley, has heard or read against evolution, together with brief answers. The author expresses some opinions toward the end on teaching of science, but they are quite in line with the broader scientific view and with materials already discussed in class. The author really believes that science is a tremendously useful way for humans to find out how the world works to help us stay fed and clothed and housed and healthy so that we can address big questions. The author also includes quotes from two noted people (Pope John Paul II and US President James Earl Carter) that tend to promote religion as well as science.
Answer: They don’t have to be. The author is religious, and is convinced of the overwhelming scientific evidence for evolution. When the author wrote a commentary on the subject for Pennsylvania newspapers, the pastors at the author’s church (a mainline Protestant denomination) approved of the piece. Most of the religious people in the U.S. belong to groups that have accepted evolution. Pope John Paul II added the Catholic Church to those groups accepting evolution (“Truth cannot contradict truth;” Address of Pope John Paul II to the Pontifical Academy of Sciences, October 22, 1996).
Perhaps the most famous Sunday-School teacher ever in the U.S., former president James Earl (“Jimmy”) Carter, said in January, 2004 that “he was embarrassed by the Georgia Department of Education proposal to eliminate the word ‘evolution’ from the state’s curriculum” (CNN story). He went on to say, “The existing and long-standing use of the word ‘evolution’ in our state’s textbooks has not adversely affected Georgians’ belief in the omnipotence of God as creator of the universe. There can be no incompatibility between Christian faith and proven facts concerning geology, biology, and astronomy. There is no need to teach that stars can fall out of the sky and land on a flat Earth in order to defend our religious faith.”
There surely are people who believe in evolution and who dislike or even attack religion, and there are many religious people who dislike or attack evolution. But, evolution is not anti-religious in any way. The author is of the opinion that most leaders of evolutionary research wish to coexist with religion, and that most religious leaders wish to coexist with evolution.
Answer: Very often, “what is” becomes entangled with “what ought to be”—“Letters to the Editor” on the subject frequently include the worry that science displacing religion inevitably leads to a lack of divine authority for moral codes, which leads to lack of morality. As noted above, this just doesn’t make sense—the morality of evolution-accepting Jimmy Carter or of the late Pope John Paul II has not been seriously in question. (But, the author suspects that questions of morality are more important than questions of science to many of the critics of evolution.)
In regard to racial purity or some similar such nonsense, consider for a moment the case of the author. He peers out from behind thick glasses, and his daughters both wear corrective contact lenses. It is likely that he has a genetic predisposition causing near-sightedness in individuals who, while still young, use their eyes for much close-up work such as reading. This genetic predisposition seems to be hereditable—he has passed it on. (There is still medical debate about genetic roots of bad eyesight, but the interpretation here is probably correct.) Leaving the author in the gene pool may “weaken” it a little bit. How should the author be dealt with in a world that recognizes evolution? Should he have been sterilized as a youth, or killed, or forbidden to mate? Or, should he be recognized as suffering from a handicap for which he should receive affirmative action? The government could have subsidized lessons for him during his youth on how to be attractive to a potential mate while peering through thick glasses. (Fortunately, he was successful in marrying a wonderful woman, but maybe there are other downtrodden thick-glasses wearers who should have been helped by outreach efforts.) Or, should we just recognize that glasses (and now, contacts) work just fine, and why worry about it? The reality of evolution in no way dictates one’s morality! Evolution is what is, not what ought to be—we have to decide how to use the scientific information about evolution.
Answer: No. We covered this one in the text at some length; the fossil record is beautifully consistent with evolution. The gaps present are the gaps expected based on the nature of speciation and the incompleteness of the fossil record, and the gaps are filled by transitional forms in those groups that are commonly fossilized and for which you would expect to find transitional fossils. Even a little consideration shows that not every creature is fossilized, and that big and relatively rare land creatures will have somewhat sketchy fossil records whereas small and relatively common shelly shallow-sea creatures will have rather complete fossil records. This is observed. One can look at the likelihood of fossilization, and then generate predictions on how complete the fossil record will become as more fossils are collected, and these work.
Answer: Again, this was discussed a lot in the text. There are commentators who make arguments against science that might seem sensible to those who don’t know the field, but those arguments can be shown to be completely wrong with just a little care.
Consider one that the author has been shown several times. The author has helped count over 100,000 annual layers in a Greenland ice core, and to do all of the careful testing using fallout of historically dated volcanoes and other time markers to show that the results are reliable. The “counter-argument” from the young-Earth supporters was that a flight of World War II planes that was forced to land on the ice sheet had been buried a couple of hundred feet in only 50 years, so a couple of thousand feet would be 500 years, and the 10,000 feet of ice thickness in central Greenland would be about 2,500 years, so the ice sheet started safely after Noah’s flood. Seems perfectly reasonable, doesn’t it?
But, when we discussed glacier flow back a few chapters, we noted that a glacier is a bit like pancake batter, spreading across a griddle under the influence of gravity. If you put a dollop of pancake batter in the middle of a griddle and watch, the layer thins as it spreads. Put another dollop on top, and both spread and thin. Keep putting dollops on top, and letting the batter drip off the end of the griddle, and eventually you’ll have a whole pile of layers. The one at the bottom will have been spreading and thinning the longest, and will be the thinnest.
An ice sheet is similar, spreading and thinning as more snow is piled on top. Very crudely, an annual layer will become half as thick and twice as long while it is moving halfway from wherever it is toward the bed. Friends of the author have directly measured the motion and spreading of the ice, and it fits this pattern beautifully. (These measurements show a little extra downward motion associated with squeezing snow to ice, but glaciologists usually speak in terms of ice-equivalent thickness—the air already mathematically squeezed. And very deep, the halfway-to-the-bed-thins-by-half breaks down, because the ice sheet had to form sometime so the first layers weren’t thinned while flowing through the ice sheet, and because the flow over bumps in waffle-iron fashion also complicates things a little). The fact that Greenland makes icebergs, hence is spreading, means that deeper layers are thinner, and the simple airplane burial calculation is completely wrong.
At the time the author was writing this, a quick search of the Web found numerous sites that slightly improved the young-Greenland calculation while still getting it completely wrong. These sites noted the thinning of layers with increasing depth, picked a place near the edge of the ice sheet where the ice was relatively thin, picked a thickness of annual layers at the surface, picked an erroneously thick annual value at the bottom, and suggested using the numerical average of the surface and bottom thickness in the calculation to get the age of the ice sheet. Fourteen inches thick at the top, less than two inches thick at the bottom, call it two inches, take the average of 14 and 2 inches and get 8 inches per year, and a few thousand feet of ice in the thin margin of the ice sheet still squeaks in after Noah’s flood—the scientists must be confused. Seems reasonable, right?
Absolutely not. Try this very simple equivalent, that you can do in your head. Suppose that the ice sheet is two feet thick, or 24 inches, that the top annual layer is 12 inches thick, and the bottom has twelve annual layers each 1 inch thick. A scientist would count the annual layers and find 13, giving a 13-year age for the ice sheet. But the technique advocated by the young-Earth websites would average the top and bottom thicknesses (6.5 inches per year), and then calculate that fewer than 4 years were required to build up the two feet of ice, not the actual 13 years. The mathematical error made on the websites is a very simple one, and one that most students will have learned to avoid while still in middle school. (The author doesn’t know whether the mistake on these young-Earth websites represents stupidity or deliberate misrepresentation, but the mistake is so flagrant that it is not easy to think of a third option.) Do the calculation right for Greenland, and you end up with a very old ice sheet. If you count the annual layers, or calculate the age using the flow of the ice sheet, or estimate the age by identifying abrupt climate changes and using the ages from tree-ring or other layer counting of those abrupt changes elsewhere, or use any of the other dating techniques, you’ll get the same Answer: Greenland’s ice is much older than written history.
The scientific community is continually improving age-dating techniques, arguing about them, working on them, and thus far, no serious problems have been found with the old age of the Earth. The arguments presented against the old age, such as the buried planes or the 5,000-year-old living clam (see the Enrichment from the Grand Canyon) prove not to be problems after all.
Answer: No. The second law of thermodynamics says that entropy increases in a closed system; the law of conservation of information is something promoted by the intelligent-design supporter William Dembski in his writings arguing against evolution, and appears to be basically a special case of the second law of thermodynamics, although there is no scientifically recognized “law of conservation of information.” In fact, order can emerge out of chaos, and does so all the time—a snowflake does form from randomly oriented water molecules, for example. The second law summarizes a great number of observations showing that, in growing the snowflake, heat must be removed, which causes an increase in disorder of the surroundings. Some creationist websites have suggested that followers not use this second-law argument because it is completely wrong.
An interesting note is that a whole field of Evolutionary Computing now exists (linked to genetic algorithms, artificial intelligence, etc.). In trying to “teach” computers to solve complex and difficult problems, computer scientists have found it useful to mimic evolution and natural selection—have the computer start with a possible answer, see if it works, then tweak the answer and see if that works better, throwing away ones that work worse and keeping ones that work better. If, for example, you’re trying to improve the routing of airplanes to fly the shortest distance while carrying the most passengers, you could start with the current route map, and then randomly “perturb” it a little, to see if that works better. You need to define “better” (how many more miles of flying is it worth to allow you to sell another passenger ticket?), but then the technique is successful. Usually, many small perturbations are used in sequence, but occasionally a slightly larger one may be useful; huge ones usually fail. The approach obtains order—a useful optimization of the flight plan—from the chaos of reality by random adjustments followed by selection of those that work better. Selection doesn’t require intelligence, but simply telling the computer to save the coordinates if the cost is lower. In biology, this selection is achieved by survival—if you survive to have kids, you have been selected. The lessons of biological evolution thus have been used to help computer scientists solve hard problems—science works. And the idea that somehow the second law of thermodynamics prevents evolution is just silliness.
Note that, in the absence of the “selection” step, randomly generating new options is almost guaranteed to fail in finding an optimal answer. Many of the anti-evolution websites and other anti-evolution materials point to how incredibly unlikely it is for random processes to generate something useful. These sites are completely correct, but completely misrepresenting evolutionary theory. The step of randomly generating lots and lots of new “experiments” is followed by the step of picking “successful” ones—in evolutionary computing, the successful ones meet some criterion such as saving the airline money; in evolution by natural selection, the successful ones promote survival to have kids.
Answer: This one takes a bit of discussion; it seems common-sensical, but turns out not to make sense.
To many biologists, evolution is defined as something like “change in gene frequencies over time,” with genes being the basic inherited instructions for making and running living things. There is no “micro” or “macro” in this; the distinction is simply not meaningful in the modern theory of evolution. “Macro” is just more of “micro”; they are not separate things.
A different way to view this question is that, biologically, there is a division between species (either things interbreed, or they don’t, or they do somewhat, as discussed next). All of the other divisions that we draw between different living types (kingdom, phylum, class, order, family, genus) are human constructs to help us understand the world. If different people with well-developed science had named all of the creatures, those people likely would have picked the same species, but may have picked different ways to group the species. “Macro” in this sense presumably is used by the evolution-skeptics to refer to changes between the larger groupings (there are wolves and coyotes and domestic dogs, but these people claim that there is a fundamental difference between “dogness” and “catness”). But larger groupings are primarily conveniences for us; evolutionary biology addresses whether creatures interbreed and exchange genes, or not.
“No, no, no” a skeptic might say, “We don’t care about what biologists think; we care about the reality, and we know that there is ‘dogness’ and ‘catness,’ these are different types or sorts.” This is harder to answer; such people presumably are postulating some unknown barrier that somehow prevents genetic “experiments” beyond pre-defined boundaries—evolution can bring about new types of dogs, or new types of cats, but only to some boundary and not beyond. But what forms such a boundary? Where is the ‘police officer’—biochemical or otherwise—that would check after two dogs had sex to make sure that the potential offspring did not include a genetic experiment that went one DNA base pair outside of the defined limits of ‘dogness’? An omnipotent deity could of course do such a thing, but there exists no scientific evidence for this—make another base substitution in DNA, and you have another experiment. Let nature choose the “good” experiments (those that lead to lots of surviving kids) and evolution happens. “Macro” evolution really means “evolution that takes long enough to occur that humans in their lifetimes won’t see much change in large animals (although plenty of such changes are happening to disease organisms).”
Dogs and cats really are different, because many successful evolutionary experiments have accumulated over tens of millions of years. Evolution really is gradual, and “hopeful monsters” (a dog gives birth to a cat, for example) do not happen. But, given the observed rates at which variability is produced by reproduction, and the rates at which natural selection is observed to function, mathematical modeling shows that there has been more than enough time in geological history for all of evolution to have occurred. There is no problem, for example, in going from the known damage that happens to cells in the bright sun (sunburn), to cells that are a bit more sensitive to light to help a creature know where the light is and avoid it, to groupings of those cells, and on to an eye. The question is not whether evolution could have happened in that much time, but why evolution ran as slowly as it did—most of the time, not a lot of change has been occurring, probably because creatures had found pretty good evolutionary solutions to problems and tend to stick with those solutions.
An additional note is that the species with us today are really not as separate and distinct as some people might have you believe, but have blurry edges, overlaps, etc., as you would expect from the scientific understanding of evolution. For example, “ring complexes” of animals are observed—species A interbreeds with B, which interbreeds with C, but A and C don’t interbreed. Are A and C the same species? Would they be if B became extinct? (Imagine what would happen if we had nothing but miniature poodles and Great Danes—could they ever “get it on”?) The messy world of biology shows that the world is not populated by “types” but by evolving populations with greater or lesser degrees of gene exchange.
Answer: “Intelligent design” is a resurrection of a very old idea. Proponents of “intelligent design” argue that there exist “irreducibly complex” parts of plants and animals, such that these parts would not have been useful while they were evolving but only after they evolved, and thus that they could not have evolved. If evolution made a useless something that later became part of an eye, that something wouldn’t help the creature and might hurt it, and so wouldn’t be saved and experimented upon to generate the rest of the eye—the intermediate steps usually should be useful for evolution to work.
In decades gone by, the eye was often cited as an irreducibly complex structure—without all of the parts of your eye, you wouldn’t be reading this. But it is very easy to find successful creatures on Earth with no eyes, and others with rudimentary eye spots, and others with slightly more complex eyes, and so on—gradual improvements on the slight sensitivity of all cells to light can lead to an eye. Because each step in the evolution of an eye has utility, the eye is not “irreducibly complex”. (Ask someone whether they would prefer to be completely blind or to be able to discern light and darkness, or vague shapes, and that person is not likely to opt for blindness—less-than-perfect eyes are still valuable.) As their eye-argument failed, intelligent-design advocates have switched to other arguments, such as the flagellum or the clotting of blood, but scientists working on these topics are not finding these structures to be irreducibly complex, either.
Scientists working in many fields related to evolution have been vocal through their leading organizations, such as the National Academy of Sciences, and the American Association for the Advancement of Science, in noting that “intelligent design” is not science, even if it happens to be stated by people with scientific training. Scientists after all can say all sorts of things that are not science. (And yes, you can find a few scientists saying anti-religion things, but those are not science either.) The leading hypothesis of “intelligent design” seems to be that there are some things that evolutionists cannot explain. This hypothesis does not lead to useful predictions that can be tested and falsified, so it just isn’t science. Notice, by the way, that successful scientific explanation of one so-called irreducibly complex item such as the eye has not falsified “intelligent design” to its supporters, who can always propose that something else is irreducibly complex. In a widely watched court case in Dover, Pennsylvania in 2005, a federal court stated these results very clearly: “intelligent design” is not science. (See the Kitzmiller case (.pdf) at uscourts.gov if you’d like 139 very interesting pages on this topic.)
After the author wrote a newspaper column advocating the teaching of science in science classes, he received communications (e-mail, phone, letter) from numerous interested people, including many intelligent-design supporters. Aside from a couple of unpleasant “You’re going to burn in Hell” e-mails from the fringe, the exchanges were respectful, interesting, and informative, from a broad spectrum of beliefs. Notice that these are not the leaders of the “intelligent design movement”, but mostly-Pennsylvanian newspaper readers responding to a column. Two observations are:
1) None of the “intelligent design” supporters seemed to seriously consider the possibility that the “intelligent designer” was a flying spaghetti monster or a space alien; where it could be determined, the correspondent identified the “intelligent designer” as God as worshipped in the Judeo-Christian tradition, and the correspondent came from a fundamentalist, conservative, or traditional Christian background.
2) None of the “intelligent design” supporters seem to have come to their religious beliefs based on the perceived difficulty of evolutionary biochemists in explaining blood clotting, flagellae, or eyes. Awe and reverence for the glory of nature did seem to figure in some religious beliefs, but supposedly irreducibly complex items were not in the forefront.
This leads to some additional, important points. The broad umbrella of “intelligent design” allows an old Earth and allows evolution to have occurred, except at those moments when an unspecified intelligence tinkered with the process, so if you are a true sacred-book literalist, “intelligent design” may give you scant comfort. And many religious people object to “intelligent design” based on the argument that it is lousy theology—if an unspecified “intelligent designer” is introduced to high-school biology students based on a claimed difficulty that scientists are having in explaining intermediate steps in blood clotting, might future success in explaining blood clotting raise questions in the minds of those students about the validity of belief in the “intelligent designer”? The pastors at the author’s church seemed remarkably uninterested in getting high school biology teachers to take over religious instruction based on explanations of flagellae.
Answer: This is a hard one, because we so strongly believe in fairness and in hearing a diversity of ideas. But, if you present “both sides” in science class, it isn’t science class any more. Science is the human search for ways to make accurate predictions, and that means setting aside the ideas that don’t work (the Earth is flat, the Earth is the center of the universe) and keeping the ones that do. Teaching the controversy over evolution in science class would be akin to teaching the controversy over whether the Earth or the Sun is more nearly the center of the solar system. Students are not really equipped to answer such a question (it took bright people, telescopes, calculations, and observations over centuries to figure out that the Earth does most of the moving in the Earth-Sun system). Teaching “intelligent design” is not good science and does not stimulate good science. “You biology students will fail in explaining and predicting some things” is not an especially motivational approach to teaching. Imagine if you did the same thing for math homework—“Do problems 1-24, but a few of them are impossible, so if you hit a hard one, just skip it”—what kind of motivation would that give to students?
The scientific controversy over evolution by natural selection is similar to the controversy over gravity. We don’t have a complete understanding of gravity yet. The grand unified theories of physics have not succeeded in explaining the quantum world and gravity through one set of equations, gravitational waves have not been observed, and other research frontiers await. But we have a pretty good idea that if you knock your pencil off your desk, the pencil will fall down. There are limits to predictability—although gravity causes things to fall down, and gravity keeps the atmosphere near the Earth, gravitational attraction is weak enough compared to thermal energy that you can’t easily keep track of the position of an air molecule, for example. But, scientists can hit a tiny meteorite with a spacecraft from across the solar system, and can deliver a small exploding device to a particular window in the palace of a particular former dictator, using gravitationally based calculations. In the same way, there are lots of evolutionary questions, things we don’t know, and fascinating research frontiers, but the basic idea that kids are mostly like parents, differences affect success, and this leads to evolution, is very well established with very little uncertainty.
Discussions about “intelligent design” surely can be included in school—a school that does not recognize the reality of religion in the world is not preparing its students for that world. But “intelligent design” is not science, and the author believes that science classes should teach science.
No magic is involved in petrification of bone, or wood or other materials. The chemical environment inside organic materials is very different from the chemical environment outside. All groundwaters carry minerals, and when those mineral-carrying groundwaters encounter a change in chemical conditions, some chemicals usually are picked up while others are put down. Remember that water can poison people by picking up lead and other chemicals from old pipes and that water can clog old pipes by putting down scum and hard-water deposits—picking up some chemicals and putting down others is the usual behavior for water. Water softeners work by pulling calcium out of water and putting sodium back in.
When organic material is buried in mud, if enough adding and subtracting of chemicals occurs before the material is eaten by worms or fungi or bacteria, then the organic material can be “turned to stone”. This petrification is rare—most organics are recycled—but occurs often enough to give us plenty of fossils to study.
For example, silica is very common, and is relatively insoluble in acidic solutions but very soluble in basic solutions. Groundwater in dry environments often is basic and so carries much silica. When this silica-carrying groundwater meets the organic acids in a buried tree or bone or other formerly living thing, the silica may precipitate. Actual chemical processes are typically a little more complicated than explained here, but the principle is the same. Materials scientists are even succeeding now in using wood as a template for “growing” ceramics, replacing the wood by zeolites, silicon carbide, or other materials to make useful things by human-accelerated petrification.
If you go to a casino to gamble, you are likely to lose. This is partly because the games are stacked against you—the odds on all casino games favor the house. (The odds on state lotteries typically are even more in favor of the house, by the way.)
But, you also lose at a casino because you are poor and the casino is rich. Suppose that you start off with $10 and the casino starts off with $999,990. Together, you have $1,000,000. Suppose further that this is a bizarre casino with a perfectly fair game—you and the casino are equally likely to win. If you win $10, then you have $20 and the casino has $999,980. But, there is slight catch—if you lose $10, you are at $0 and the casino will show you the door. If you stay and gamble for a while, the outcome is almost guaranteed; you will hit $0 before the casino does, the casino will have all of your money, and you will go home broke. The $0 mark is an absorbing boundary—you can’t bounce off of it (the casino won’t give you your money back), nor can you go negative and then return (well, you might go take out a loan, but your ability to get loans to cover gambling losses is much smaller than the casino’s ability to get loans, so you’ll eventually reach the point of no return—your money will have been absorbed by the casino).
Extinction works the same way. Once a species is gone, it is gone. You can’t have a negative number of tigers and then later have a positive number. So random fluctuations eventually kill off species. But, at certain times, such as when the meteorite hit at the end of the Mesozoic, or now as we humans spread across the surface of the planet and squeeze others out (more on this coming soon!), extinctions go a whole lot faster than normal.
Column by R.B. Alley, published in Harrisburg Patriot-News (2004) and republished in
Pittsburgh Post-Gazette and Centre Daily Times early in 2005.
The School Board of the Dover (PA) Area School District recently (November, 2004) mandated the teaching of so-called “intelligent design” alongside Darwinian evolution in science classes, and similar actions are at least being considered elsewhere. As a religious person and a scientist, I hope that school boards will avoid mixing apples and angels in the classroom.
Like many scientists, I am fortunate to teach. We know that our students will soon discover things we missed, often correcting our mistakes in the process. Thus, a scientist would be foolish to claim that science gives absolute knowledge of Truth. If I successfully predict the outcome of an experiment, I’m never sure whether my understanding of the world is True, whether I’m pretty close but not quite right, or whether I’m really confused and was just lucky this time.
But, our society has agreed to act as if science is at least close to being true about some things, and this makes us very successful doing those things. Carefully crafted bits of silicon really are computers, airplanes designed on those computers using principles of physics really do fly, and medicines from biological laboratories really do cure diseases. The military has investigated psychics as well as physicists, but continues to rely on the physicists because they are so much more successful. Science, tightly wedded to engineering and technology, really does work, and is the best way we humans have invented to learn to do many things.
Science asks a high price for the value it gives, however, and that price is real dedication to science. The cartoonist Sidney Harris once drew a panel showing two long strings of blackboard equations connected by “Then a miracle occurs,” with one scientific-looking character saying to the other “I think you should be more explicit here in step two.” For a plane to fly, for a medicine to cure disease, every step must be tested, and everyone else must be able to follow those steps. Science students are welcome to rely on divine inspiration, but they cannot rely on divine intervention in their experiments. Scientists, like athletes, must follow the rules of the game while they’re playing.
What, then, are the rules? First, scientists search for a new idea, by talking to people, or exploring traditional knowledge, or in the library or other places. We look for an idea that explains what we see around us, but that also disagrees with an old idea by predicting different outcomes of experiments or observations. Then we test the new idea against the old one by doing the experiments or making the observations. An idea that repeatedly makes better predictions is kept; an idea that repeatedly does more poorly is set aside. An idea that can’t be tested also is set aside; it isn’t scientific. Even if I really love an idea, or really believe it is True, but I can’t think how to give it a fair test, I have to set it aside for now. Some people find this limiting and avoid science; others find it exhilarating and are drawn to science. Doing this well gives us good things from good science.
Who decides what is or isn’t science? Scientists--with other thinkers watching over our shoulders--do the day-to-day deciding, but ultimately the whole bill-paying, newspaper reading community is checking on us to make sure that we really are producing useful insights.
Does science have limits? Will we run out of new ideas? Will we hit problems that we can’t solve? Perhaps. But when I come out of a classroom of bright young students, I am convinced that we’re nowhere near any limits that might exist, and that there is much to discover yet.
So, what about Intelligent Design, or even Young-Earth Creationism, and teaching them in science class? They’re interesting ideas. But, some parts we don’t know how to test. Even if they are said by scientists, they aren’t science. And the testable parts have been tested and found wanting--they don’t do as well as the “scientific” view in explaining what we see around us, or in predicting what we find as we collect new tree-ring records and ice-core samples, or as we search for oil and valuable minerals, or as we watch dangerous new diseases appear faster than our bodies can respond to them. We spend a few hours discussing the main pieces of evidence in class, and a lifetime isn’t enough to cover all the details, but scientists have been working on these questions for centuries and have a pretty good idea of what works now. Evolution “in the dark backward and abysm of time” is scientific theory, not Truth, but it is very good science.
How does this fit into the bigger picture? Although some people are happy to view science as merely a tool, others do believe that the remarkable success of science means that we are getting closer to truth. But even these people disagree about that truth: a mechanistic universe, a benevolent and omnipotent deity, or something else? Fascinating as they are, such questions are for now outside of science. Many scientists and religious people are thinking about such questions, but no experimenter knows how to guarantee the cooperation of an omnipotent deity.
By all means, students should ask deep questions, think and discuss and probe. Science does not tell us what we ought to do, and students will have to join us in addressing what ought to be as well as what is. But if we want to face the big questions with better medicines, with computers that function and planes that fly, with clean water and buildings that don’t fall down, I believe that we should teach science in science class.
You have reached the end of Unit 11! Double-check the list of requirements on the Unit 11 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Following are some supplementary materials for Unit 11. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
It's a cold night. You wake up, curled in a little ball, shivering, and you remember that there is a nice, warm blanket in the hall closet. What do you do?
If you picked D, you probably should be thinking about wise responses to global warming, because the physical basis for expecting that the ongoing human addition of carbon dioxide to the atmosphere will raise the Earth's average temperature is about as strong as the physical basis for expecting warming from putting another blanket on the bed. Nature might do something bizarre and offset what we're doing—a huge number of volcanoes might erupt and throw things into the stratosphere to block the sun, or space aliens might come and get in the sun's way—but we're putting a blanket on the planet, and the planet is likely to get warmer.
However, that doesn't tell us what, if anything, to do, so let's go take a look at the options. There's some money to be made here, for you business majors, and engineers, and everyone else.
You will have one week to complete Unit 12. See the course calendar for specific due dates.
As you work your way through the online materials for Unit 12, you will encounter a video lecture, several vTrips, some animated diagrams (called GeoMations and GeoClips), additional reading assignments, a practice quiz, a "RockOn" quiz, and a "StudentsSpeak" Survey. The chart below provides an overview of the requirements for this unit.
REQUIREMENTS | SUBMITTED FOR GRADING? |
---|---|
Read/view all of the Instructional Materials | No, but you will be tested on all of the material found in the Unit 12 Instructional Materials. |
Submit Exercise #6: Fossil Fuels (& Driving Hazards) | Yes, this is the last of 6 Exercises and is worth 5% of your total grade. |
Take the Unit 12 "RockOn" quiz | Yes, this is the last of 12 end-of-unit RockOn quizzes and is worth 4.5% of your total grade. |
If you have any questions, please feel free to email "All Teachers" and "All Teaching Assistants" through Canvas conversations.
On the following pages, you will find all of the information you need to successfully complete Unit 12, including the online textbook, a video lecture, several vTrips and animations, and an overview presentation.
Students who register for this Penn State course gain access to assignment and instructor feedback, and earn academic credit. Information about registering for this course is available from the Office of the University Registrar.
The difficulty lies, not in the new ideas, but in escaping from the old ones, which ramify, for those brought up as most of us have been, into every corner of our minds. —John Maynard Keynes, economist, Preface to The General Theory of Employment, Interest and Money (1935)
Prediction is very difficult—especially about the future. — Attributed to physicist Niels Bohr
The Arctic National Wildlife Refuge (ANWR) sprawls across the North Slope of Alaska, from the Brooks Range to the coast of the Arctic Ocean, and is nearly as large as the state of Maine. ANWR is home to grizzly and polar bears, wolves loping across the tundra, moose and caribou, uncounted waterfowl, and snowy owls ghosting on white wings.
And beneath it, there probably is oil. The nearby Alaska Pipeline has pumped billions and billions of dollars worth of petroleum south from regions near the North Slope. But as that Prudhoe Bay oil runs out, the pipeline may soon be left empty--a very expensive conduit with nothing to carry. Similarity of geology suggests that ANWR also has oil to fill the pipeline, and to fuel automobiles in the U.S., or China, or somewhere. Not a lot of oil—maybe 10 billion barrels, according to the USGS. That is just over two years of U.S. oil imports at recent rates, not much more than one year of total U.S. oil use, and not exactly “energy independence,” but at $100 per barrel, it represents something like $1 trillion. The argument between wilderness and development has been going on for years, and is not ending soon; the recent emphasis on "fracking" for oil and gas in the lower-48 has temporarily shifted press attention away from the Arctic, industry interest in the north remains strong.
Plants have an amazing ability. They take carbon dioxide (CO2) and water (H2O) and use energy from the sun to turn them into more plant material (which has an average chemical composition fairly similar to CH2O), releasing oxygen (O2) in the process. An approximate formula for photosynthesis is:
CO2+H2O+energy→CH2O+O2
Most of the rest of us—animals, fungi, many bacteria, as well as forest fires—run this reaction backwards, combining plant material with oxygen to release energy, carbon dioxide, and water. Done rapidly, this is “burning” in a fire; done slowly, it is still burning of a sort, which you might call "respiration." Plants usually include a bit of nitrogen that we didn't write in the simplified formula above, and animals often use the plant material with its nitrogen to make proteins that make animals, but after the animals die, they are almost always "burned" by bacteria or fungi or other animals that eat them to release the energy.
But, what happens if dead materials end up in a place without oxygen? It turns out that a lot of burning can be done by replacing oxygen with other things such as sulfate, but if these run out too, burning is no longer possible. Then, fossil fuels—coal, oil, and natural gas—become possible.
Coal is formed when bacteria break down dead plants. When there is no free oxygen in the air or water, bacteria remove the oxygen and hydrogen that are included in the plant material, leaving mostly carbon and forming a brown material called peat. When peat is buried by more sediment, heating and pressing drive off more and more of the little remaining oxygen and hydrogen, thus forming coal. Brown coal (lignite) has not been cooked much; it is common in the western United States. “Normal” or bituminous coal, formed from lignite by heat and pressure, is common in many places including western Pennsylvania. (This is a Penn State class, so we do keep track of Pennsylvania!) In a few places including eastern Pennsylvania, closer to the center of the old Appalachian Mountains, the bituminous coal was cooked to metamorphic anthracite coal. Peat is found with loose sediment, lignite with not-too-hard sedimentary rock, bituminous with harder sedimentary rock, and anthracite with metamorphic rock.
Oil is something like coal, but it is formed from dead algae buried in mud, usually from marine settings but sometimes from lakes. Algae start with more hydrogen and less oxygen than wood, so they produce a different fossil fuel. Heat breaks down the algae to release liquid oil. More heat breaks down the oil and makes natural gas, which is primarily methane (CH4). Some natural gas also is made at low temperatures by bacteria. While the heat is making oil and gas, the mud is being squeezed to make shale. Pennsylvania contains some oil and natural gas, and the first oil well in the world was drilled in western Pennsylvania. Humans had used petroleum before, but from natural seeps rather than wells. The push for petroleum that led to the first oil well was fueled by a looming shortage of whale oil, as the demand for that product far exceeded the ability of the oceans to grow the whales that produced the whale oil that was used to light homes on dark nights.
Oil and gas are low in density, float on water, and so tend to rise through water-filled pores in rocks and escape at natural seeps on land or beneath the sea. In some places such as off the U.S. Gulf of Mexico coast, biological communities have been found living on oil seeps on the sea floor. (Oil is natural, and nature uses oil in small quantities. But if a supertanker wrecks or an oil well blows out, nature cannot use all that oil at once, and problems result.) Because of the tendency for oil to escape, a large accumulation of oil can form only if there is a trap of some sort. Many different types of traps exist. For example, fluids do not pass through shale easily because it has only tiny spaces; if shale lying above sandstone is folded a little, oil and gas may be trapped in the sandstone layer, as shown in the figure.
At present rates of use, and at costs vaguely similar to what we see today, the oil and gas will last for most of a century, and the coal for a few centuries. If we were willing to pay more for gasoline, say $50 per gallon, more fossil fuels would be available. Although the store of fossil fuels is certainly limited, there is a rather large quantity. Until very recently, U.S. production had been dropping, causing us to import most of our oil from other countries, although as noted in the next section, "fracking" is affecting this. World production will probably peak in the next few decades (and some experts have suggested a peak within a few years). Demand for fossil fuels is rising rapidly, particularly in China and India as they industrialize. Shortages of fossil fuels, and worries about such shortages, already cause political problems; some observers believe that such worries are one reason the U.S. government spends so much on military preparedness.
In Pennsylvania and some nearby states, there is much interest now in the natural gas being produced by "fracking" the Marcellus Shale, while Texas is fracking the Barnett Shale, North Dakota is fracking the Bakken Shale, and others are coming into play or being considered, in the U.S. and overseas. When a deep ocean ran out of oxygen back in the geologic past, mud and organic material accumulated on the sea floor, and these became black shales over time. Burial and heating broke down some of the organic material to make oil and gas, and some of that escaped from the shale, either leaking out at the sea floor or being trapped in special places that we have drilled into to get the fossil fuels.
But, a lot of organic material remains in the black shales, stuck between the very tiny clay particles that make up the shale, and not able to get out. People have learned how to drill into such layers, and then turn the drill to go along the layer. After a long hole is made, water and sand and various chemicals are pumped into the hole at very high pressure, breaking the rocks and then squirting into the new fractures to hold them open, in a process called “fracking.” Gas then leaks out along the fractures and up the hole, where it is collected and used by people.
In the big picture, this doesn’t make a lot of difference—the estimates of the total amount of fossil fuel that might be available have included the Marcellus gas since before the technology for recovering the gas became commercial, under the assumption that eventually we would figure out how to recover that gas. But, right now, there is much more gas on the market than there was a few years ago, and some oil is coming this way, some people are getting jobs or checks from gas companies for access to the land, other people are worried about the effects of the gas on the environment, and there is a lot of heated discussion going on. If you want to learn more about this particular issue, see the Enrichment.
Let’s be honest. We use fossil fuels for good reasons. Most of our energy is obtained from fossil fuels. We run washing machines, rather than hand-scrubbing our clothes, primarily with fossil-fuel energy. Most of us are freed from the manual labor of hoeing and shoveling to grow our food because fossil-fueled tractors plow and plant and cultivate and harvest. Many of us have been freed from freezing to death in the winter, or perishing of heat stroke in the summer, or dying because we can’t get to the hospital in an emergency, because of fossil fuels. Humans have so far largely avoided the Malthusian trap of having lots of kids who have lots of kids until we exceed our food supply and then starve, primarily through being clever and through using cheap energy. Our energy use in the U.S. is equivalent to each of us having more than 100 people working to take care of us.
For humans, and for the few types of domesticated animals that have benefited from our use of fossil fuels (pigs, rice, chickens and soybeans, for example), there is little question that fossil fuels have been good. For other species on Earth, our use of fossil fuels to tame much of the planet has been less beneficial. How much easier did trains make it for humans to travel west to shoot bison? How much easier is it to cut a tree with a chain saw than with a stone hatchet? However, this is complicated by the fact that we have let some trees grow back, and we quit burning whales (or whale oil) to light evenings because we switched to burning the long-dead algae and trees that are fossil fuel.
There clearly are other costs of fossil-fuel use. Marks of oil exploration on the tundra may last for decades or longer. Acid rain from coal-fired power plants has killed the trout in headwaters streams on the Laurel Ridge of central Pennsylvania, and in some other places. Smog is not good for us, and shortens our lives. However, our lifespan continues to lengthen, so the dangers of smog and other “modern” hazards, such as industrial chemicals and radiation, are less than the advantages of the technologies that gave us all of these things. In fact, our biggest dangers are probably infectious diseases (evolution of new nasties), each other (automobile accidents, murders, and wars) and ourselves (smoking and drinking and eating too much, while exercising too little) rather than the technological things or natural disasters so many people worry about.
If not for the greenhouse effect, we humans probably would not be here. The Earth’s atmosphere allows the shortwave radiation (what we usually call "sunlight") coming from the sun to pass through to the Earth’s surface, without much interference. (There is a little interference. Among other things, blue light is scattered off air molecules a bit more than red light is, so the blue light bounces around in the atmosphere and reaches our eyes from all directions in the sky, whereas the red comes more directly from the sun, which is why the sky is blue.) The windows in a greenhouse similarly allow sunlight to enter easily.
But, the sunlight heats the Earth or the inside of a greenhouse, which then radiate longwave radiation back upwards. As we saw way back at the Redwoods, the total energy reaching the planet on average exactly equals the total energy leaving, but the arriving energy is mostly shortwave electromagnetic radiation (light that we can see) while the leaving energy is mostly longwave electromagnetic radiation (infrared radiation that we cannot see without special sensors). The windows of a greenhouse do not allow longwave radiation to pass through easily; some gets through, but some is trapped.
When the sun rises after a cold night, energy enters a greenhouse but has trouble leaving, and the extra energy warms the greenhouse. Warming makes the floor of the greenhouse emit more longwave radiation, forcing some through the glass until a new balance is reached between incoming and outgoing radiation, but with the greenhouse at a higher temperature than would occur without the windows of the greenhouse. Some gases in the atmosphere act in the same way as windows on the greenhouse, intercepting some of the outgoing longwave radiation and keeping that energy in the Earth system. Water vapor contributes the most to the greenhouse effect, but carbon dioxide, methane, and others also matter. Without these greenhouse gases in the atmosphere, the Earth would be mostly frozen. (For a brief meteorological perspective on greenhouses, see the Enrichment, which also explains why water vapor is a “slave” to the other greenhouse gases, so that carbon dioxide is more important for changing the climate.)
Human activities are increasing the concentrations of several greenhouse gases in the atmosphere. Carbon dioxide, from burning of fossil fuels, burning of forests, and a few other sources, is the greenhouse gas we hear the most about. Refrigerants (chlorofluorocarbons and related compounds) also are potent greenhouse gases, and are increasing in the atmosphere. Those refrigerants that are especially damaging to the ozone layer that protects us from harmful very-shortwave (ultraviolet) radiation are being phased out, but the less-ozone-damaging replacements are still greenhouse gases. Methane, produced in cow guts, landfills, rice paddies, and other places where carbon breaks down in the absence of abundant oxygen, also has been increasing. Eventually, the huge carbon-dioxide source and the long time that carbon dioxide survives in the atmosphere mean that carbon dioxide will dominate global warming.
The world is warming. Of this there is almost no question—thermometers, including thermometers far from cities and in weather balloons and on satellites, in the ground and in the ocean, analyzed by government and university scientists in many different countries, including with industrial funding, show that warming is occurring. So do the temperature-sensitive types of snow and ice (NOT the top of Antarctica at -40, which won't melt even with a fairly large warming, but we see reductions in the seasonal river and lake ice, seasonally and "permanently" frozen ground, springtime snow, mountain glaciers and more). The great majority of significant changes in where plants and animals live, and when during the year they do things, are in the direction expected from warming.
Furthermore, there is high confidence that the warming is from the carbon dioxide. The physics of warming from rising carbon dioxide is unavoidable. Must of this was actually worked out by the Air Force after World War II. They were not doing global warming, but instead were worrying about such things as the appropriate sensors on heat-seeking missiles to shoot down enemy bombers. Use the wrong sensor, and you can't "see" the target because carbon dioxide is in the way. And carbon dioxide interacts with infrared radiation going from Earth to space in the same way. Satellites confirm this every day. Scientists have worked very hard to find some other explanation of the warming, but if anything, the sun has dimmed a little over the last 30 years, while volcanic eruptions have thrown up particles that blocked the sun a bit, we have put up particles from smokestacks that help the volcanoes in blocking the sun and causing cooling, we have replaced dark forest with more-reflective crops to cause a little cooling, and nature has not done anything with cosmic rays or space dust or anything else that would explain what is happening. Indeed, we see the warming from rising greenhouse gases despite the other influences primarily pushing towards cooling; really, the best answer to "How much of the warming did human-released greenhouse-gases cause?" is "More than all of it, because other things have reduced the warming." Finally, the "fingerprint" of the warming (for example, warming down here but cooling high in the stratosphere) is just what is expected from the effects of greenhouse gases, and completely unlike the pattern expected from changes in the sun, volcanoes, El Niño, or other natural fluctuations. Computer models of the climate system, when forced with the known natural causes of climate change such as changes in the sun and volcanoes, do a pretty good job of simulating the climate changes that happened before greenhouse gases had risen much but do a lousy job of simulating recent changes; adding greenhouse-gas effects causes the models to simulate what happened quite accurately.
Even prominent “skeptics” have now publicly admitted the high probability that humans are warming the world. (One of the well-known, often-seen-on-Fox-News skeptics directly stated this during a public debate with Dr. Alley a few years ago.) If you follow the news, you know that there is a lot of argument about how large the warming will be and whether it will be good or bad, although scientific consensus is very strong that the warming will be large compared to recent natural changes, and will be more bad than good for humans and other living things.
Some of the argument involves feedbacks. Feedbacks are “extra” processes in a “system” such as the Earth’s climate. If you force a system to change by doing something to it, many other things may then change. Some of these will amplify what you just did, making the changes bigger than you could have accomplished by yourself; these are positive feedbacks. Others will oppose what you just did, making the changes smaller than what you initially forced; these are negative feedbacks.
You, for example, have all sorts of negative feedbacks built in. If you are placed in a warmer room (the forcing), your body will begin to warm up. But then a negative feedback kicks in—you start to sweat, and that cools you off. Your body temperature doesn’t change nearly as much as the temperature outside of you changed. But, if you have certain diseases, they may fool your body so that its negative feedbacks are reduced and may even become positive feedbacks. Fever is usually a good thing, helping the body fight invading germs more effectively, but people die of fever when the feedbacks become positive and the body “burns itself up.” If you’re in a canoe with a really enthusiastic golden retriever, you may try to lean as the dog leaps about in such a way as to stabilize the canoe—you are providing a negative feedback on the tipping. But if the dog tips the canoe, and the ice chest falls to the low side, the ice chest is acting as a positive feedback to amplify the dog’s motion and tip the canoe further. If you lose your balance when the dog lunges to the side, you may suddenly fall toward the dog, providing another positive feedback and perhaps flipping the canoe.
The Earth certainly has positive and negative feedbacks. The easiest stabilizing one is the very fast change in radiation—a warmer place glows more brightly and sends more heat toward space, tending to cool the hotter places faster. Other than this almost-instantaneous change, most of the feedbacks over times that matter to us (years through millennia) are positive, amplifying changes; over still-longer times (say, millions of years), the feedbacks tend to be negative, stabilizing the climate. Climate changes over years or centuries thus can be almost as large as climate changes over millions or billions of years.
Put some extra carbon dioxide in the air and the climate will become warmer, speeding the rock-weathering chemical reactions that remove the carbon dioxide to produce dissolved ions that are used to make shells, removing the extra carbon dioxide. This negative feedback, which was explained by Penn State’s famous professor Jim Kasting and coworkers, stabilizes the Earth’s climate. Given too much atmospheric carbon dioxide, the excess is removed through warming-enhanced weathering; if too little, low temperatures slow weathering and allow carbon dioxide released by volcanoes from subducted shells to build up in the atmosphere. However, this takes hundreds of thousands of years or longer to act; over mere centuries, increased weathering will have little effect on the carbon dioxide we release, much of which will stay up for centuries, millennia or longer.
If we put carbon dioxide into the air and warm the Earth a little, several positive feedbacks begin to function. Warmer air can “hold” more water vapor (the saturation vapor pressure roughly doubles for a 10oC or 18oF warming), and water vapor is an important greenhouse gas, so warming causes more warming. Some of the shortwave radiation from the sun that hits the Earth bounces right back to space without first warming the Earth. This occurs especially over snow and ice, which have very high albedo or reflectivity. But, warming the Earth removes some snow and ice, which then allows more of the shortwave radiation to be absorbed, which warms the Earth more—a positive feedback.
Clouds reflect some sunlight (so cloudy days are cool), but clouds also interfere with outgoing longwave radiation (so cloudy nights are warm). The largest uncertainties in predicting how much warming will result from a given amount of fossil-fuel burning are probably related to how clouds will change, and whether these changes will produce positive or negative feedbacks. However, these uncertainties are not nearly large enough to affect the conclusion that future warming from fossil-fuel burning is highly likely.
We clearly wish to predict the future. If burning of fossil fuels, combined with bovine belches and leaky refrigerators, are going to cause too many problems, we might want to change our ways now. To predict the future, we need to do experiments. But, we have only one world. We cannot look at many different futures of one world, nor do we wish to wait many decades for the experiments to end. The solution we use is to build little worlds in computers, and run the experiments on those.
Geologists are important in this effort in two ways: we help find out how the world works, so that the right things can be put in the computers. The computer models always will be simpler than the real world, so careful choices must be made about what to put in. And, we provide data against which the models can be tested. You wouldn’t trust a model that had never been tested, but you wouldn’t want to wait a whole lifetime for a test. If the models can successfully simulate very different, warmer and colder climates of the past, then the models are probably pretty good. So we need to know about climates of the past.
The computer models of today actually are doing very well at “retrodicting” climate, predicting things that already happened. Modelers set up the configuration of ice sheets and ocean and continental positions and orbits and solar brightness, then model the climate and see if the computer results can match the climate that is recorded by the fossils and other climatic indicators in the rock record without “cheating” (so you can’t go in and tweak a lot of things to make the model match the data and then claim that the model is great—the models actually do work on past climates without such cheating). The models are also doing quite well at predicting the patterns of change we have observed with instruments over the last decades. Predictions made by modelers over the last decades are really occurring now. (For a little more on this, see the Enrichment, which also gives a bit more on why each doubling of CO2 has about the same effect on temperature, as covered next.)
Models predict that the world will warm about 3oC or just over 5oF for a doubling of CO2. The full warming will be delayed a few decades behind the rise in CO2, because the air can’t warm all the way until the ocean and ground have warmed and some ice has melted, which takes a while. The global warming to date, somewhere between 1oF and 1oC, is similar to the typical temperature variability at most places, so you really have to be paying attention to know that anything is changing. If we proceed to burn all the fossil fuels, something like an 8-fold increase above “natural” levels is possible over the next few centuries, or a warming of about 9oC or over 16oF, large enough that no one would have any doubt about the change. As noted in the Enrichment, uncertainty in the models is about 50%—the models project that the warming may be only about half this much, or more than one-and-a-half times this much, if we continue with business as usual and octuple CO2. However, because data on past climates are very difficult to explain if the models showing smaller changes are correct but are easy to explain if the models showing mid-range or large changes are correct, the low-end changes are less likely than the high-end changes.
Few people care about the temperature, and many care about what the temperature means for humans and other living things. Impacts are harder to estimate, both because of the additional uncertainties involved in turning temperature change into economic change, and because of the involvement of human values. For example, if warming causes species to become extinct, but those species were not economically important, how bad is that? You will find really strong differences of opinion across the political spectrum in the U.S. and the world.
Some disagreement is expected. Many places are already too hot for comfortable human habitation; making them hotter will not be good. Other places are too cold for comfortable human habitation today, so warming them might be good for humans there, unless you melt their ice and flood the coasts. Warming may remove one of the barriers that help keep malaria and other tropical diseases from spreading into today’s temperate zones, which is not good. Sea-level rise from warming is almost inevitable, as glaciers melt and the warming causes the ocean water to expand. Drying of the great grain-growing regions of the world during summers is likely. As a crude summary, most models indicate that human-induced warming will make our lives harder. People in the developed world are likely to cope; some in the developing world will cope, and others may fail (and die). A little warming is likely to hurt most of the world’s people (who live in too-warm places) but help most of the world’s economy (which is concentrated in fewer people in colder climates); too much warming then hurts everyone. A lot of loud radio commentators focus on the slight possibility of good outcomes, but ignore the possibility of disaster. (For a bit more on this, see the Enrichment.) Many professionals in the field note the benefits of caution—slow the ship in the fog until you have time to see how many icebergs are ahead, while starting to turn now to avoid the icebergs we already see.
Much of the argument focuses on economic concepts such as discount rates. Economically, something in the future is worth less than something today because of uncertainty (an apple in 30 years may do me little good, because I may be dead). If one uses this sort of reasoning, then any environmental change that takes more than a few decades to occur is of little concern—we humans will just deal with it when the time comes, using the great economic engine we’ve built.
But some people argue that intergenerational transfers should be treated differently—“we” humans who will “deal with” the changes will be our grandchildren, not us. Do we have the right to give them a greatly altered (probably degraded) Earth? Thus far, classical economic reasoning is “winning” in the real world, but the best science says that humanity will be better off if the reality of climate change is included in the decision-making.
Serious scholarship shows that, with a few decades of real investment in science and engineering, abundant energy can be provided while controlling greenhouse gases for a cost of roughly 1% of the world economy. That is in line with the costs of other clean-ups—sewers and garbage and catalytic converters and others. That is much smaller than most recent estimates of the coming damages from global warming if we do nothing. That also is a whole lot of money—roughly $400 billion per year now. Depending on one’s politics, the clean-up of greenhouse gases can be presented as a money-making prospect, as a low-cost obligation to the future, or as frighteningly expensive. Eventually, the switch away from fossil fuels must be made, because the fossil fuels are finite. And one can argue that $400 billion per year is a potential future industry that merits a real investment in research now. If we move away from fossil fuels, good jobs will be lost in fossil-fuel industries, but most studies point to a gain of more jobs in other fields. Note the problem, though—people in the industry who are worried about losing their jobs know who they are, and vote. People who will get the new jobs are students now, and don't know about the new jobs. So, there is some learning to be done.
Most of the national parks were established to preserve geological features. A few parks, such as Sequoia and Redwood, were established for biological reasons. Increasingly, however, the national parks are visited, used, preserved, and managed for biodiversity. We humans continue to spread. More and more land is brought under cultivation. More of the produce of the sea is netted and served to humans. Some estimates are that we and our immediate friends—cats and corn and cows—now use about half of everything that the world makes available for us and everything else. And, with about 7 billion of us here now, heading for 9 or 10 billion, and with a couple of billion of us not using much but hoping to use more, all humans taken together may double what we're using. Ultimately, this leads to extinctions and loss of biodiversity.
This is a geology class, and biodiversity is a bit far-afield, but we have time for a quick detour. We saw that there have been mass extinctions in the past—times when many living types became extinct in a short interval. There is a real chance that a geologist far in the future would place our current time as the latest mass extinction, the end of the Cenozoic, and the start of the Anthropocene.
Early humans were surprisingly hard on biodiversity. Wherever humans arrived with their efficient tool kits—in Australia, New Zealand, other islands, the Americas—extinctions of large animals followed. Direct human hunting, or competition from the rats, pigs, dogs, and others that arrived with the humans, likely contributed.
Some people don’t like the idea that early humans were hard on biodiversity. Many people, including good scientists, have argued that the extinctions of large animals in the Americas were caused by climate change, which happened to occur at about the same time as human arrivals in some places. Dr. Alley has listened to talks in which data he helped produce were used to argue that the climate changes were so large and rapid that they must have been responsible. But, the work by Dr. Alley and others showed that dozens of such abrupt climate changes occurred, and the extinctions did not occur until the humans arrived. It is hard to imagine that a couple of dozen abrupt climate changes happened without killing off many species, and that just when fluted-point spears showed up in the rib-cages of mammoths, the next abrupt climate change was solely responsible for killing the mammoths, and humans did not play a role.
But, the earlier extinctions were mostly of large creatures. Since the industrial revolution, “modern” humans have contributed to extinctions of various creatures. And, the rate of extinction may pick up soon as we increasingly occupy the planet. To see why, let’s take a little detour into island biogeography.
If you were to visit a lot of islands that are more-or-less the same distance from the mainland, and count the number of species on each island and measure its area, you would find that the bigger islands have more species. Roughly, an island with ten times the area as another will have twice as many species. If you visited islands of about the same size at different distances from the mainland, you would find that those closer to the mainland have more species.
At least some of what controls these observations is not too difficult to understand. If you have a small island, it can hold only a few individuals of a species. From year to year, populations go up and down depending on food supply, predators, and other things. With a small population, a small drop can hit the absorbing boundary of zero individuals and cause extinction, but a large population can survive a small drop. So, extinction is more likely on a smaller island, and smaller islands have fewer species. The mainland is there to supply new individuals to islands to replace those that die, and repopulation is easier for islands closer to the mainland, so those islands closer to the mainland have more species.
These patterns of island biogeography are well-established. Studies of repopulation of islands sterilized by volcanic explosions, and even of very tiny islands that were deliberately depopulated and then allowed to come “back to life,” have shown that this is the way the world works naturally.
Now, think about Yellowstone. Originally, the boundaries drawn for the park separated wilderness inside from wilderness outside. Today, as shown in the satellite photo, some of the park boundaries are easy to see from space because loggers outside the park work right up to the boundaries. Yellowstone remains connected to other wilderness regions in other directions; it is not an island (yet).
But what if Yellowstone were an “island,” as some other parks are or soon will be? Suppose a park becomes surrounded by farmland, which is used to feed humans and keep us alive. Farmland does not support a lot of wild orchids or wolves. Farmland is impoverished in biodiversity, with just a few species, carefully selected to feed us. A park surrounded by farmland is an island, because many species have great difficulty crossing the farmland, just as many species have difficulty crossing the ocean. And, from the well-established principles of island biogeography, isolation of a parkland from other wilderness will cause extinctions in the park. Perhaps more worrisome, if the only remaining wilderness is in parks, there is no longer a “mainland” to replace species lost to local extinction on the island—extinction in the park is then extinction from the world.
We know that as climate changed in the past, plants, and animals migrated long distances to stay with their preferred climate. As the climate changes in the future, migration will be required, but may be impossible if the parks become isolated.
One can ask whether biodiversity is worth preserving. This is proving to be a difficult topic, and one that will be discussed much in the future. Certainly, many of our medicines have come from plants, and if many plants become extinct before we can study them carefully, we are likely to lose many possible medicines. Engineers and designers are increasingly using “biomimetic” techniques—mimicking nature. Evolution has worked over vast times to select the most successful biological patterns, and we can learn from them, if they are here to be learned from.
More-diverse ecosystems seem to be more productive (if you have hot-loving and cold-loving and wet-loving and dry-loving types in a region, then something will grow well no matter what weather arrives; if you have only one type, and the weather is bad, so is the crop), so if producing more is good, biodiversity seems good. But, the difference is not huge.
Living things have frequently served as “canaries in coal mines”. Miners would take a canary along in the mine, not only for companionship, but because the birds were more sensitive to bad air than were people, and a sick or dead bird would warn miners to get out before the miners became sick or dead. Birds of prey served that function for us with DDT. This chemical was going to free us of pests, increase crops, wipe out diseases—until the falcons, hawks, eagles, and other predatory birds started disappearing. A little DDT on a plant led to more DDT in an insect that ate lots of plants, and still more DDT in a bird that ate a lot of those insects, and became so concentrated in a falcon that ate the birds that the falcon’s eggs broke, and young ones couldn’t be raised. It became clear that such “bioconcentration” threatened us with problems as well—the other living things gave us a warning. Loss of biodiversity means loss of warning sensors.
And, many people like diversity (look at the money spent on zoos, and the interest in charismatic macrofauna in parks). Further, some people see a moral issue—do we really have the right to terminate the existence of other living things?
Some planners today are trying to establish corridors connecting wilderness areas, so that the parks do not become islands and lose species. How successful this plan will be remains to be seen. The “simple” answer is that, to maintain many species on Earth, we have to maintain much wilderness. And that in turn has implications for how we humans choose to behave.
Join Dr. Alley and his team as they take you on "virtual tours" of National Parks and other locations that illustrate some of the key ideas and concepts being covered in Unit 12.
Join Dr. Alley in this optional video (no, it won’t be on the quiz) to learn about formation of Fossil Fuels, in Barataria Reserve, Jean Lafitte National Historical Park and Preserve, Louisiana, from the PBS television special Earth: The Operators’ Manual.
This last GeoMation of the semester is a short piece about people, environmental and habitat change, and their effects on biodiversity. We hope this adds a bit of insight to Unit 12 for you, and we wish you success as you wrap up the end of the semester!
The Unit 12 lecture features Dr. Richard Alley.
Check out the Unit 12 Presentation (the PowerPoint presentation used in the online lecture).
As an aside, some of our friends over in meteorology are not happy that the effect of CO2 on climate is called the “greenhouse effect.” They fully understand that CO2 does warm the planet, and they know that the glass of a greenhouse affects radiation in much the same way that CO2 in the atmosphere does—the shortwave radiation from the sun comes through glass or CO2 more easily than the longwave radiation from the Earth goes out through glass or CO2. But, the meteorologists note that this effect of glass on radiation is not the only reason why a greenhouse is warm, nor is this the major reason. Greenhouses also are warmer than their surroundings because the glass blocks the convection currents that take much of the sun’s heat away from the ground outside of greenhouses. Some meteorologists have even suggested renaming the atmospheric phenomenon to avoid possible confusion. But, the “greenhouse effect” is catchier than “the effect that warms the Earth through modulation of radiation balance, akin to the radiative effect that contributes to but does not dominate daytime warming of greenhouses.” Maybe our meteorological friends would be wise to “chill out” on this one. Notice that this little discussion about terminology in no way affects the reality that more CO2 in the atmosphere warms the planet—nature works, regardless of what words we use to describe it.
Many different models have been constructed of the Earth’s climate system, ranging from attempts by large teams to include essentially all Earth-system processes into models that tax the world’s largest computers, to small-group or individual-scientist efforts to build fast and flexible models that allow exploration of uncertainties in many parameters. Across a range of models, the equilibrium warming from a doubling of CO2 is often stated to be between about 2oC (maybe as low as 1.5oC) and 4.5oC, with a central value near 3oC (and with the most recent results pointing to a bit above 3oC). Comparisons to the past, for both the last century and for much longer times, largely exclude the low end of that range—models that change global average temperature near 1.5oC for a doubling of CO2 are not able to accurately simulate the changes of the past, whereas models with larger temperature change in response to CO2 do better in simulating past changes. Based on the paleoclimatic record, warming of near 3oC or more for a doubling of CO2 seems reasonable, and values above 4.5oC cannot be totally excluded.
Water vapor is the most important greenhouse gas in terms of the warming being provided now, but we usually don’t talk about water vapor with global warming. Why not? Simply, water vapor is almost entirely a slave to other things. Put some more CO2 up in the atmosphere, and the atmospheric concentration of CO2 remains high for centuries or millennia or longer before chemical processes remove it. Put more water vapor up, and in just over a week, on average, that water has rained out. The burning of fossil fuels makes equal numbers of water-vapor and CO2 molecules, but because the water vapor stays up less than two weeks and the CO2 perhaps 2000 years, our effect on the atmospheric concentration is more than 100,000 times larger for CO2 than for water vapor. We can change CO2 fairly easily (and are doing so!), but we can’t put up water vapor fast enough to make much of a difference, nor can most other natural processes affect global water-vapor loading very much. However, changes in the atmosphere’s water-vapor content are easily caused by changes in temperature.
Remember from back at the Redwoods that cooling reduces the equilibrium water-vapor pressure or “water-holding capacity of the air” (by about 7% per degree Celsius of cooling). Remember that as full-of-water air came in from the Pacific and was forced up over Redwood National Park and then Yosemite and Sequoia National Parks, the air cooled by about 0.6oC/100 m, raining on the way. The temperature at the top of the Sierra was controlled by the height of the Sierra and the temperature of the air before the rise began, and the amount of water left in the air at the top was controlled by the temperature at the top. The air then goes down over Death Valley, and the water-vapor content of the air there depends on the temperature at the top of the Sierra.
So, if the temperature is increased over the Pacific by an increase in CO2, the water-vapor content and its greenhouse effect are increased over the Pacific, going up the Sierra, going back down over Death Valley, and on to the Atlantic or Gulf of Mexico. Water vapor acts as a positive feedback—warming increases the water-vapor content of the atmosphere, causing more warming.
You can find lots of climate-change skeptics or contrarians or denialists who love to point out that water vapor is the big greenhouse gas, CO2 less so, so the scientists must be out to lunch by focusing on the small one and not the big one. Sounds sensible, right? But, it is totally stupid or deliberately misleading, or somewhere in-between. If we pulled all the water vapor out of the air, more would evaporate in a week or so. Pull all the CO2 out of the air, and the cooling would remove a lot of water vapor, with a rather high chance that the whole Earth would freeze over into a snowball. Thus, although water vapor gives us more warming than CO2, you can argue that the CO2 is more important overall.
Environmental problems seem to follow a fairly predictable path. First, someone has a good idea. Refrigerators and air conditioners and freezers are useful, but if you use ammonia in the pipes and you’re in the way when a pipe breaks, you might die, so chlorofluorocarbons were a great idea. Then, scientists discover an unintended consequence—the chlorofluorocarbons might break down ozone and allow harmful ultraviolet rays to give living things “super-sunburns,” causing cancer and other problems. There follows a period when the scientists work to improve their understanding.
But, there also follows a lot of yelling and not-entirely-scientific discussion. Some people fear that they are going to lose their jobs, or lose a lot of money, and these people respond to the scientists by arguing that there is no problem, that the problem that does not exist must be caused by nature rather than humans, and that this natural problem that does not exist would cost way too much to clean up. A very common approach is to attempt to convince the public, or policymakers, that scientists are still having a big debate, even if they are not. It is fairly easy to find a few skeptics, fund them and promote their statements, and to “cherry-pick” favorable results from the scientific literature and present them out of context.
Politics often feeds into this. Usually, if a problem is identified that affects a lot of people, the government ends up dealing with the problem. You are not allowed to tear out your sewer or septic system, poop in a pot, and dump it over the fence into my yard. Nor are you allowed to smoke in many public places now, or a number of other things, and these are laws that are passed by and enforced by the government. So, if you don’t much like government, you may not want the government trying to clean up a problem. And, if you can keep the argument focused on whether or not the science is good, rather than on possible wise responses to the problem, there is little danger that the government will do anything—we usually don't do much about a problem until we agree that there is a problem.
The press makes all of this worse, attempting to maintain "balance" by presenting both sides of a “scientific dispute,” even if one side is being manufactured and does not have much scientific basis of its own. Recent scholarship has demonstrated clearly that a reader of the mainstream press in the U.S. would have a very skewed view of the degree of scientific agreement over global warming, for example—many press outlets present a conflict that really doesn’t exist.
But, some forward-looking people also see the problem as a possibility—a new invention may make a lot of money. And, history indicates that problems usually are followed fairly quickly by new inventions, the cost of dealing with the problem typically is much less than previously stated (often about 10% of the previously stated cost), the cleanup becomes part of the economy, and life goes on. (Imagine life without toilets and sewers…)
The twin energy problems—finding replacements for the finite fossil fuels, and doing, so before the world is changed too much in bad ways—are arguably the biggest environmental problems we have ever faced, but they can be solved. Because of the huge size, the solutions will take longer, and more inventions will be required than for the ozone hole or DDT or lead in gasoline. The scholarship is clear that the sooner we start, the better off humanity will be.
In Pennsylvania, as this is being rewritten (autumn of the year 2017), the Marcellus Shale is the center of a “fracking” industry to extract gas and sell it. Some jobs have been created, and some economic activity. Some people have received money because gas companies paid for the right to extract the gas. Other people who did not own the mineral rights of their land are rather unhappy that the money from the gas on their land is going to someone else.
The issues related to the Marcellus have been contentious for a decade, and it remains virtually impossible to write anything that is not: i) quickly out of date, and ii) likely to make a lot of people mad. But, we cannot ignore the big issues, so here is a little more background.
The estimates of available fossil fuels that you see in various publications may be very different things. Textbooks often list the total amount that is considered to be available if we continue inventing new technologies and pushing the price slowly higher. This is a fuzzy number—how high can the price go before alternatives are cheaper and we quit trying for fossil fuels? But, this sort of calculation has long assumed that we would go after the gas in the Marcellus and other shales, and the oil in oil shales, and tar sands, and more.
Other numbers on fossil-fuel reserves are much smaller because they count only the nearly-ready-to-pump-out-of-the-ground fuels or some other similar definition. Over the last few years, a lot of gas in the Marcellus and other such black shales has moved from the “available in theory” to “available in practice” column, which has changed a lot of business balance sheets. You will see estimates, often communicated by industry, that the Marcellus and other such deposits contain "100 years of gas," although other estimates (including from the government) have been lower, including 25 years of gas. But, gas is supplying a little over one-quarter of U.S. energy use, so an all-gas system might last 6 to 25 years based on those estimates. That is lots of energy, lots of money, but not even vaguely close to a long-term, sustainable energy solution.
When conditions are right, organic materials accumulate in sea-floor muds. Gases dissolve more easily in colder waters (heating a carbonated beverage drives off the fizz). So, when the climate has been hotter in geologic history, there has been a tendency to have less oxygen in the ocean and lakes. Extra fertilizer at the surface may favor the growth of so many plants that the oxygen in the deeper water is used up in “burning” the plants after they die and sink, before the dead plants are used up. Low oxygen in deep water is also favored by restricted ocean basins that prevent vigorous currents to supply newly oxygenated waters. The still conditions of a deep, restricted ocean basin mean that mud isn’t washed away, but that no big chunks such as sand or gravel are washed in. So, an organic-rich mud forms in such places, especially during hot-climate, well-fertilized times.
Eventually, this is a stabilizing feedback on Earth’s climate—when high CO2 makes the climate hot, CO2 is removed from the air by being converted to plants that are buried in these muds. In just a few hundred thousand years, this can make a big difference to how much CO2 is in the air, especially because a hot climate can also remove a good bit of atmospheric CO2 by rock weathering in a few hundred thousand years. Nature thus will remove the CO2 that we are putting up now, and if you have a few hundred thousand years to wait, our CO2 is no big deal. If you care about your great-great-great-great grandchildren, though, these natural processes just aren’t fast enough to help much.
Anyway, the muds in the low-oxygen ocean or lake basins are often black, partly from the organic material, and a lot from having iron sulfide and other black stuff that forms in low-oxygen environments (add oxygen, and the iron in iron sulfide rusts and turns red). As more mud accumulates on top, the black muds get hotter from the heat of the Earth, and the mud slowly is squeezed and recrystallized to make a rock called shale. Shale is the commonest sedimentary rock, and there is a lot of black shale on Earth.
The heating also changes the organic material, making oil and gas. Most oil, and a lot of gas are formed in this way.
However, the spaces between clay particles in the mud/shale are tiny, and even tiny gas molecules and the smallest of oil molecules have trouble moving rapidly through. As more oil and gas are produced, the pressure inside rises (splitting molecules off dead things tends to increase the total volume taken up by the organic materials). Eventually, this opens cracks and allows the oil and gas to leak out. As noted in the main text, most of the leaking oil and gas escape, slowly, to the Earth’s surface, where they are broken down by bacteria or other living things that use the stored chemical energy in the fossil fuels. A little of the leaking oil and gas are trapped in special geologic places on the way, and oil companies have gotten very good at finding those places, drilling into them, and recovering the oil and gas.
Much of the organic material remains in the black shales, though, unable to make its own fractures and escape.
Oil companies have been “fracking” for a long time. Suppose that oil and gas leaked from a black shale into one of the special reservoirs that oil companies like to drill into, arriving in just a million years. That is fast compared to many geologic processes, but if it took the oil company another million years to get the oil out of the rock, the oil company would be very unhappy. So, for rocks that allow oil and gas to move slowly, but that contain a lot of oil and gas, the oil companies learned to make fractures that let the fossil fuels move faster.
one of history’s fascinating stories, the actor John Wilkes Booth, before he shot President Lincoln, was an investor in a Pennsylvania oil well. With much of his money committed, and the well yielding slowly, Booth and his co-investors decided to “shoot” the well, exploding gunpowder to fracture the rocks and yield more oil. Instead, they destroyed the well. Now much poorer, Booth switched his focus and soon thereafter shot President Lincoln. But, a new invention, the Roberts Torpedo, from a Union soldier who saw an exploding Confederate shell fracture dirt, soon made shooting wells much safer and more widespread.
Imagine you’re working for the gas company. If you blow up a balloon too vigorously, it breaks. Pump too much air into the bicycle tire, and it may rupture. This is the basis for fracking rocks. Drill a hole to the rocks you want to break. Put in some sort of plug with a tube or pipe going through (so that when you “blow up the balloon” on the other side of the plug, the pressure is pushing against the rock down there and not squirting out of the hole up here). Then, pump water through the pipe into the space below the seal, and pump hard enough that you really raise the pressure, “break the balloon” by cracking or fracturing the rocks. Release the pressure, and oil and gas can leak out of the new cracks to your hole, and up the hole to the surface, perhaps aided by a well pump that you use. (Companies that generate geothermal power also may “frack” to make spaces for hot water to move easily.)
Notice, though, that if you pry open a crack, and then quit prying, nature may squeeze it back closed. But, if you put some sand in below the plug, when the rocks break, the sand can squirt into the cracks and keep them from closing. A crack that is twice as wide carries eight times more fluid, all else equal, so this can make a big difference.
In addition to the sand, you might add something to keep microbes from growing in your new cracks, eating your valuable methane and clogging the cracks with dead-bug “gunk.” And, you might want to add some surfactants. For example, many antacids include a chemical (simethicone) that reduces surface tension so that small bubbles in the stomach can combine into larger ones that can be moved along more easily; chemicals to do the same job may be added to fracking fluids, for the same reason. You might even think of other things to add.
Once you have this fluid, some of it will stay in the cracks you make. And, the fracking is generally way down, so it is very unlikely that the cracks will reach all the way to the surface, or to the shallow level of water wells. But, some of the fracking fluids must come back up the hole to get out of the way so your gas can come out. You may recycle this “flowback,” or dump it in the creek or at the sewage treatment plant, or pump it into even deeper wells somewhere just to get it out of the way. Deep disposal and recycling have largely replaced earlier techniques. But, in Oklahoma and elsewhere, when some of this fluid was pumped into deep wells, it seems to have triggered earthquakes. (The quakes happened where and when the pumping was going on, not huge quakes but big enough to crack plaster and otherwise damage houses.)
A gas well used in fracking must be drilled through shallow and intermediate depths on the way to the shale to be fracked. The plan is to seal the well so that there are no leaks of gas or other fluids into those shallow and intermediate depths where people may have water wells. But, if the sealing isn’t done properly, it could leak. Leaks of that sort from gas and oil wells have been reported, but there aren’t a lot of public data on just how often, and just how large the leaks are. The mix of chemicals may be a trade secret, so people worry may not even know exactly what is there to leak.
Regardless of what your politician or pundit may have said, the science really does give us high confidence that a long-term shift away from fossil fuels is economically beneficial to avoid the damages of changing climate. Natural gas produces more energy from the same number of carbon molecules than coal (or oil from tar sands), so if natural gas is used in place of coal, this can reduce warming. However, if natural gas is used in addition to coal, then it increases warming.
And, there is a qualification. If you let natural gas (which is almost entirely methane) leak into the air without burning it, the methane is a greenhouse gas. Per molecule at current levels, methane causes more warming than CO2, and after a decade or so the methane is oxidized to CO2 and then stays up typically for additional centuries. So, if natural-gas production is too leaky, that may be even worse than coal in terms of causing warming; leakage also wastes money for the gas companies.
In Pennsylvania over the last few years, the public discussion included everything above, but especially focused on jobs and the economy and taxation. Every major gas-producing state in the US except Pennsylvania had taxes or fees that provided notable funds to the public coffers, with Pennsylvania having only a relatively quite small local impact fee. Those who favored the no-taxes approach in Pennsylvania typically have argued that taxes might drive away the gas companies. Supporters of taxing have often noted that no gas company would be stupid enough to come into Pennsylvania assuming the no-tax position was guaranteed for the long term. The companies can look across the nation, see that every gas-producing state except Pennsylvania had levied taxes, and that Pennsylvania had budget problems that could be eased by levying taxes, and it seems reasonable that a planner for a company would run the numbers assuming that Pennsylvania would follow all the other states.
However, once a company decided to enter the state, a good business person would note that production achieved before any taxes were levied would be more valuable. This might justify hiring more workers and otherwise working faster, which would stimulate the economy but provide less long-term stability. It also might justify cutting some corners, allowing more leakage or otherwise adopting less-efficient but faster approaches. Reliable studies balancing these issues have been lagging behind the pace of development.
Lurking behind all of this discussion is something that economists argue about a lot, called the “Resource Curse.” With certain notable exceptions, countries that rely heavily on the extraction of concentrated valuable resources (diamonds or oil, for example) have often had less economic growth than might be expected based on the size of the economic input from the resource, and many residents of western democracies do not like the social conditions in such resource-producing states. (You can think of the leading oil exporters in recent years, including Saudi Arabia, Russia, Norway, Iran, United Arab Emirates, Venezuela, Kuwait, Nigeria, and Algeria, and decide whether this “academic” view of the resource-reliant economies seems accurate, nearly accurate with maybe a few exceptions, or really wrong.) Mere production of fossil fuel is not seen as an economically bad thing by the economists who have recognized the resource curse; the worry is attached to reliance on the money from the valuable resource (the US is a major oil producer, but oil production is not a dominant part of our economy). There are many ideas on why the resource curse exists (and a few experts who still deny that it exists); perhaps the most common idea is that if people put effort into controlling the temporary resource rather than building a sustainable society, they end up with a poorer place to live.
The observation of the author is that there has been rather little public discussion in Pennsylvania on whether reliance on a soon-to-be-depleted natural resource is the best way forward for the Commonwealth.
The author will also state that he isn’t coming out for or against Marcellus Shale gas development, that it may ultimately turn out to be more “good” than “bad,” or the other way around. Furthermore, even if the author did take a side, almost no one would care. The development has been going fast, many people love it, and many people don’t.
We really are throwing wads of money at people to get energy from oil wells that really are starting to run dry, and burning oil really does make carbon dioxide that really does have a warming influence on the climate. If we keep doing what we're doing, we have high scientific confidence that the fossil fuels will run out, but not before the exhaust changes a lot of things on the planet. The short-term solutions may involve conservation, or nuclear energy, or capturing carbon dioxide and putting it back in the ground, or other ideas. Or we could just barge ahead and hope for the best. The long-term solution is highly likely to involve the sun, through photovoltaics, or wind, or biomass, or other things. Fossil fuels are just stored energy from the sun—think of them as a great battery, charged up over a few hundred millions of years, that we are discharging over a few hundred years, and we can already see the bulb on the flashlight starting to dim. Creedence Clearwater Revival watched the "big wheels keep on turning" in the John Fogerty song Proud Mary; in this parody, we review the ways to keep the big wheels turn without making our lives a lot harder.
Climate changes driven by fossil-fuel burning may trigger extinctions and reduce biodiversity. But, if we cut down forests and burn trees rather than burning fossil fuels, extinctions may occur that way, too. Most of the logging in the Pacific Northwest is for lumber rather than for fuel, but the trees are cut down just the same. Hoyt Axton wrote Joy to the World (Jeremiah was a Bullfrog), Three Dog Night made it famous, and countless DJs have used it to get the wedding party lurching about the dance floor at the reception. In this parody, we visit the raptor center at Penn State's Shaver's Creek Environmental Center, to discuss the impacts of deforestation on spotted owls, and the general issues of biodiversity.
Many people, including US senators, have offered the opinion that our inability to forecast weather more than a week or two in advance means that we cannot forecast climate years ahead. This seems sensible, but is actually really wrong. And, anyone who understands the game of Wheel of Fortune knows why. Here’s a musical explanation, prepared for Groundhog Day 2015, when Dr. Alley was inducted into the Punxsutawney Weather Discovery Center Hall of Fame.
You have reached the end of Unit 12! Double-check the list of requirements on the Unit 12 Introduction page and the Course Calendar to make sure you have completed all of the activities listed there.
Reminder - Exercise #6 is due this week. Check the Course Calendar for the specific due date.
Following are some supplementary materials for Unit 12. While you are not required to review these, you may find them interesting and possibly even helpful in preparing for the quiz!
Please feel free to send an email to ALL of the teachers and TA's through Canvas conversations with any questions. Failure to email ALL teachers and TA's may result in a delayed or missed response. See "How to send email in GEOSC 10" for instructions
We’ve been through a lot—a tour of some of the finest places in the world, and a “tour” of the science of the Earth. What should you have learned from all of this?
First, science is a successful human activity that allows us to do things we want by first learning how things work. Scientists come up with ideas, and then try very hard to prove those ideas wrong. Situations are found in which a new idea makes predictions that differ from those of old ideas. These situations are then created (experiments), and the outcomes show which idea predicts better. An idea that makes wrong predictions many times is discarded; an idea that makes correct predictions many times is accepted tentatively. It may be right, or just a good approximation, or just lucky. If we behave as if the ideas that make accurate predictions are in fact true, we are able to do things successfully. Many people believe that this shows that the successful ideas are in fact true, or at least close to being true, but there is no way to prove this.
Geology is the science of the Earth. Geologists study the materials that make up the Earth, and the processes that happen to them or have happened to them in the past. Much of geology is highly practical: finding oil, clean water, metals, and other valuable things in the Earth; warning people of geological hazards such as landslides, earthquakes, and volcanoes. Some of geology will become important; geologists are helping in writing the “operator’s manual for the Earth” to help us and the Earth live in harmony in the future. Some of geology is just plain fun—learning about the dinosaurs, for example. Much of geology is fundamental, and contributes to all of these.
The Earth formed about 4.6 billion years ago. It has a layered structure, with a mostly-iron core (which is solid in the middle but liquid outside of that), with a silicate mantle (but, as silicates go, with relatively much iron, magnesium, and calcium and relatively low silica), and a thin silicate crust on top that is rich in silica, sodium, potassium, and aluminum.
Crudely, the heat of the Earth drives processes that build mountains, and the heat of the sun drives processes that tear the mountains down. Inside the Earth, radioactive decay makes heat, with a little help from other processes. Heating rocks softens them and reduces their density. The heating of rocks from below causes convection cells. Columns of hot rock rise from deep in the Earth, perhaps from the core-mantle boundary, to feed hot spots that erupt basaltic lava (higher in silica than the mantle) and produce volcanic chains such as Hawaii. Where a hot spot exists under a continent, the basaltic lava may melt some continental rock to become silica-rich and explosive, as at Yellowstone.
Most of the heat loss from the Earth is through roll-shaped convection cells higher in the Earth’s mantle. These reach into the upper mantle; the very upper mantle and the crust make the cold and brittle lithosphere, which floats on the softer rocks deeper in the mantle. Where the convection cells rise and spread, the rocks above are raised high and split apart. Basaltic lava leaks up through the cracks and hardens, making sea floor at spreading ridges in the oceans. Where a spreading ridge passes under a continent, the ridge splits the continent to form fault-block, pull-apart mountains separated by deep valleys, including the Sierra Nevada and Death Valley.
Ocean lithosphere initially is warm, but cools and becomes denser with time. Eventually, it sinks back into the mantle at a subduction zone. The downbending at a subduction zone makes a deep trench in the ocean. Sediments may almost completely fill a trench near a continent, but trenches lacking abundant sediments are the deepest places in the oceans. Sediments scraped off a downgoing slab may pile up to make coastal mountains, such as those of Olympic National Park. Some sediments go down with the subducting slab, but are heated as they go down, and eventually melt to make andesitic lavas, higher in silica and lower in density than the basaltic sea floor. These feed explosive, steep stratovolcanoes such as Mt. St. Helens and Mt. Rainier. Sometimes, such stratovolcanoes form offshore rather than on the land, giving island arcs. When such an island arc, or another continent, reaches a subduction zone, the silica-rich rocks are too low-density to sink. The collision of an island arc or a continent with another continent is called obduction, and forms folded and thrust-faulted mountains such as the Appalachians, including the Smoky Mountains and Mt. Nittany.
Above all of this activity, the Earth is wrapped in a thin layer of air and water and ice that erupted from volcanoes over time. The sun heats the air, causing convection cells. Convecting air cools by rising and expanding, or by radiation, and this cooling causes rain and snow.
Rocks at the surface of the Earth often formed in the Earth under different conditions and are unstable at the Earth’s surface. Physical processes break these rocks into smaller pieces, and chemical processes change the minerals in the rocks to other types. Silicate rocks release ions that wash to the sea, and clays, quartz sand and rust that mix with organic materials (worm poop, etc.) to form soil. If loose soil and rock are too steep and wet, they slide or creep downhill in mass movements.
Rain that reaches the surface of the Earth is mostly returned to the atmosphere after being used by plants, with most of the remainder soaking into spaces in the ground to form groundwater. Groundwater eventually returns to the surface to feed streams; we humans intercept this flow in the ground with wells. Pollution of this groundwater is a common problem, and very difficult to clean up.
The rocks and water that are moved across or beneath the surface of the Earth feed streams. The characteristics of streams—whether wide and braided or deep and meandering, for example—depend on the debris supply as well as the water supply. If humans interrupt the usual pattern of water and sediment transport, we usually cause other changes (erosion, sedimentation, subsidence, etc.) that cause problems for us.
Glaciers are really cold-weather streams. Glaciers move water and rock from place to place on the Earth, and typically are more effective at moving rocks than are other mechanisms of sediment transport such as mass movement, streams, and wind. Glaciers have grown and shrunk many times over the last million years, covering as much as 1/3 of the modern land surface. The Earth’s surface in those formerly-glaciated regions is still dominated by features left by the glaciers. Many of our most beautiful landscapes, including Glacier and Yosemite, center on the works of glaciers.
Mass movement, streams, and glaciers wear down the mountains. These processes also deposit sediments. The types of sediments, their fossils, and other features contain a history of events at the Earth’s surface. Reading this history, we find that the Earth is very old. Annual layers in the very youngest deposits, including trees and ice sheets, record more than 100,000 years. Rocks beneath the trees and ice sheets contain evidence of old seas and deserts and mountains, which early geologists realized had required more than about 100 million years to form. Radioactive clocks tell how long the time gaps were in the sedimentary record and how much time was required to make the oldest rocks, revealing that the Earth is about 4.6 billion years old.
The fossils in the rocks show that offspring differ from their parents slightly; these differences affect ability to survive to have offspring, and the genetic basis of these differences is passed on to offspring in greater or lesser degree. Over long times, accumulation of these differences has led to changes in the types of living things on Earth (evolution by natural selection). At some times, catastrophic events such as meteorite impacts have caused mass extinctions; at other times extinction has been a gradual process.
In the geologic record, recent times would look like a mass extinction. Humans are using more and more of the Earth’s resources. Because these resources are not available to other species, individuals and entire species have disappeared. Humans are also changing the climate, and spreading pollutants across the globe. These changes are not yet catastrophic, and may never be, but humans will need to reduce our impact in the future if we wish to prevent more extinctions and possible catastrophic harm to ourselves. This must be accomplished through some combination of controlling our numbers, or controlling our impact per person.
Science has proven, over and over again, to be the best tool humans have ever developed for learning enough about the world that we can keep ourselves sufficiently fed, clothed, free of disease, and otherwise healthy so that we can worry about the big questions of what we are here for and how we get along with each other and stay happy. Geology is an important piece of the big picture of science. National parks have proven, over and over again, to be beautiful, enjoyable preserves of our shared natural and human heritage. It is our fondest hope that you will have a happy and prosperous future, that you will appreciate and perhaps even contribute to the benefits of science, including geology, that you will enjoy using the benefits of science to contemplate big questions, and that you will do some of your contemplation in your national parks. Pleasant journeys!!!
Be sure to check out the video review on the next page as well!
The final lecture is a course review and it features Dr. Richard Alley.
Check out the Course Review - Unit 13 Presentation (the PowerPoint presentation used in the online lecture.)
You can also review with Dr. Alley's final Rocking Review.
Building a real online course takes a community. The primary author of GEOSC 10 is Richard B. Alley, and I’d like to thank a whole lot of other people who have worked very hard to make the course a reality. In no special order, here are a few of the key folks.
The transition from all-paper to online content was made in concert with the College of Earth and Mineral Sciences' CAUSE trip to the Grand Canyon and surrounding national parks during 2004. Eric Spielvogel was the primary architect and guiding force of that trip, and was aided by Anna Brendle with the enthusiastic support of Dutton e-Education Institute director David DiBiase and help from Annie Taylor. WPSU contributed a lot of resources. Topher Yorks filmed the whole thing and kept us in batteries, good music, and goodwill, Tom Keiter and Joe Myers really bolstered filming during part of the trip, and the WPSU management (including Ted Krichels, Jeff Preston, and Tracy Vosburgh) supported the effort. The CAUSE students were Sam Ascah, Risa Goren, Raya Guruswami, Ryan Hanrahan, Dave Janesko, Kym Kline, Lainie McGuiney, Irene Mckenna, Amish Shah, Steph Shepherd, Sameer Safaya, Sheri Shannon, and Dave Witmer, TV stars and stars in and out of class. Dave Witmer really worked to get the “GeoClips” in order after the course ended.
Sridhar Anandakrishnan was the other prof on the trip, is the other prof in the course, and is the best rubber-duck herder ever. His contributions are woven seamlessly through the course in so many ways, and it would not be what it is without him.
At the Dutton e-Education Institute, Eric Spielvogel has been responsible for filming, course planning, course building, HTML programming, and a great range of additional things from the very beginning—you’ll find his handiwork throughout. Annie Taylor and Jennifer Babb have made the course work, solved the myriad technical problems, planned, and upgraded. David DiBiase has been a continuing guide and supporter. Marty Gutowski and Dave Babb have helped out, too.
Cindy Alley has done the textbook publishing, the figures in the textbook, the geovisualization of the rock videos, a lot of the filming of rock videos and various other parts of the course, as well as putting up with a lot of complaining over the years. Eric Spielvogel started the rock-video filming, including bringing his truly artistic eye to the effort. Additional help for the rock videos has come from Janet Alley, Karen Alley, Adam Jones, Amanda Jones, Carl Pillot, Will Ross, and WPSU. Flash animations in the exercises were supported by the Penn State Blended Learning Initiative, with Dean Blackstock’s animation and help from Mark DeLuca.
Much of the work of the course is done by the Teaching Assistants. We have had great ones. The course dates back to 1996 and began the transition to electronic in the fall of 2005. There have literally been dozens of TA's that have contributed to this course. I wish I could list them all, but the list is just too long.
The externally funded research that Sridhar and I conduct informs the teaching we do. Although we have not been funded directly to build the course, our funding agencies encourage education and outreach, and we are grateful for that. Our research has especially been funded by the US National Science Foundation, primarily through their Office of Polar Programs. We have also received support from NASA, and from the Comer Science and Education Foundation. Supplying this material to so many students, and making this material available to the public, is a way to pay back our funding agencies for support that ultimately originated with the public.
Some “Greatest Hits”:
The CAUSE effort and the development of GEOSC 10 have been rather prominently featured in various Penn State efforts and more broadly. At the Outreach Conference “Engaging Faculty: Public Service Media as Service Learning” in 2004, the effort was described in talks by Sridhar Anandakrishnan, David DiBiase and Eric Spielvogel as well as by me. CAUSE student Stephanie Shepherd and I described our efforts to the Partners of the Penn State Trustees. The work was featured prominently in my College of Earth and Mineral Sciences (EMS) G. Montgomery and Marion Mitchell Innovative Teaching Award (“Rocking the Parks: Big Canyons, Big Craziness, and a Big Team Innovating in General Education”). Similarly, GEOSC 10 and CAUSE figured prominently in my Eisenhower Teaching Award from Penn State. The film of the CAUSE trip won an Emmy Award, has been shown on WPSU many times, and in the EMS museum many more times. Shorter versions were shown to the New Student Convocation at University Park for at least two years, to incoming Schreyer Honors Scholars, in the University President's State of the University address, and in the Road Scholars orientation for new faculty at Penn State. We were also featured on the cover of the EMS undergraduate recruiting packet. In addition to offering GEOSC 10 to students across the Penn State system, our materials have been incorporated in locally taught courses at Penn State’s Dubois, Delaware County, and Lehigh Valley campuses, and are being used in university education outside of Pennsylvania. Dispatches from the CAUSE trip were highlighted on Penn State’s website, and a discussion on Deep Time with Eric Spielvogel was featured at Research Penn State. We’re happy that thousands of students have learned lots from GEOSC 10 while providing high ratings on Student Educational Experience Questionnaires. With many more people studying our materials outside of the normal classroom (including more than 5000 views of the YouTube spinoff of our Rock Videos as of the end of 2008), we’re even happier.
Want to see more?
Here are some optional vTrips you might also want to explore! (No, these won't be on the quiz!)
Canyonlands National Park
(Provided by UCGS)
Mammoth Cave National Park
(Provided by UCGS)
Jean Lafitte National Historical Park and Preserve
(Provided by National Park Service)