We are happy you have joined us in our exploration of the fundamental principles of biology. Biology is the scientific study of life. Biologists are scientists who study living organisms.
Life is everywhere. Biologists study the human body and figure out how it works when it is healthy, and compare that to the way it works when things go wrong. They study the bacteria that live in your digestive system and explore the reasons why these bacteria are critical to your health. They study mice and flies and yeast and plants and learn about how living organisms are alike, and how they are different.
Different kinds of biologists study different things. Cell biologists are currently researching the causes and symptoms of various types of cancer, enabling doctors to find cures and treatments. Ecologists are biologists who specialize in studying the way that living organisms interact with their environments. They examine important issues such as acid rain and observe what happens to the health of living systems when acid rain falls. There are even specialized biologists, called astrobiologists, who study conditions on other planets and design experiments to determine whether or not life could exist in other places in our solar system.
Understanding biology can help us understand ourselves and the environment around us. For example, understanding how climate change affects ecosystems will help identify the ways in which sustainable food production can keep up with a growing human population. Biological research is constantly adding to our understanding of human health and disease, and often results in treatments that lead to longer, healthier, and more productive lives. Biological research can also help address the growing demand for renewable energy and sustainable resource use.
Every day we are bombarded with information about health issues, new vaccines and medicines, ecological problems, and other issues related to biological research. This information comes at us from a myriad of diverse sources, including friends, healthcare providers, the media, your neighbors, and the Internet. As you gain a deeper understanding of life and the processes that sustain it, you will be able to more effectively sort through and interpret this deluge of information, and make informed decisions about your health and how you interact with the environment.
After completing this course, you will have a basic understanding of biological principles and how they relate to your life.
Portions of this course are built on materials developed and generously provided by University of Maryland University College, made available with permission under a CC-BY-NC license. Direct use of specific activities and media elements are noted throughout the course. |
The purpose of this course introduction is to prepare you conceptually and technically for the course. Since you may not have experienced an online course like this before, we will start with a short section describing the course and offer some learning strategies that will help you use the materials most efficiently. Finally, we will discuss what biology is all about — the "big picture" — our framework for exploring the relationship between the themes in biological research and the fundamental concepts and principles included in this course.
Information in this course is organized into units. Each unit begins with an introduction that orients you toward the major themes you’ll explore in that section. The unit introduction will also show you how the content fits into the course as a whole. Each unit consists of several modules. Modules are like chapters in a book, and when you start a new module, you will see the list of learning outcomes you will achieve after completing that section of the course. Each module consists of several pages designed to help you achieve the learning outcomes. The introduction highlights what you will learn and how it relates to the big picture. The following pages make up the informational “meat” of the module. This explanatory content consists of short passages of text with information, examples, images, and explanations. As you work through the content, you will have many opportunities to practice what you are learning. The practice usually takes one of two forms:
Directed feedback during these learning activities will help you stay on track as you assimilate new information. Most modules will also include an “Application Spotlight” that gives you the opportunity to apply your understanding of the module’s content to a specific case study or real-life example.
After completing a module, you will have a chance to demonstrate what you learned by taking a graded quiz. The module quizzes will assess and reinforce your learning as you progress through the unit. When you complete all the modules in a unit, you can participate in a “My Response” activity, where you will assess your own understanding of the unit content. After reflecting on your ability to achieve the learning objectives, you will have a chance to submit questions to your instructor, and then you will conclude your activities with a unit quiz.
You can navigate through this course using the navigation bar at the top of the screen, the course syllabus, the course outline, and the page number box. All are accessible on any page in the course.
Also provided is an Appendix, which includes information such as a glossary of key terms and their definitions, a table illustrating the various ways biologists represent chemical structures, and an interactive image of the cell (the smallest unit of life).
This course is divided into 10 units. This first unit, "Biology: The Science of Life," lays the foundation upon which the rest of the course is built. This unit begins by defining biology and characterizing its relationship to other fields of scientific study. It then elaborates on a few of the themes that emerge over and over throughout this course. Finally, in this unit you will explore the nature of science. This introduction represents the toolkit you’ll need to understand the rest of the course, and will enable you to create context for all the new information you will learn.
The diagram above represents the "Big Picture" of biology. It illustrates how the course is organized and depicts the relationships between the main topics you will learn about as you progress through the material. You will find that as you go, the material is telling a story. It is a story based on the research and collaboration of centuries of scientific thinkers. The story begins with the smallest particles of matter: atoms. Atoms are the building blocks of all matter and they can be put together, like Legos, to form molecules. A special class of molecules called biomolecules are the “legos” used to build the structures required for life. The fundamental unit of life is the cell, and all cells are made of biomolecules. The cell is the first level of organization to display all the characteristics of life. Cells, and all living organisms, require energy to function. Organisms capture and use energy through a chain of chemical reactions called metabolic processes. The energy that cells capture through metabolism can be used to do work, including the work required for reproduction. All living organisms reproduce, passing their genetic information from generation to generation. This results in a unity and diversity of all living organisms, because we all share a common ancestor. Some organisms are more likely to survive than others, and this leads to evolutionary processes that result in the incredible diversity of life you see around you. The final chapter in the story reminds us that the diversity of life exists within a rich context of interactions, causes, and effects. Life does not take place in a vacuum; instead, living organisms rely intimately on other living organisms, as well as in the nonliving environment. All matter, living and nonliving, exists within a delicate interconnected balance, and humans are just one part of this biosphere.
This is a fascinating story and one that inspires a deep understanding and appreciation of the living world within us and around us. As we study the fundamental principles of biology, we will systematically explore the following characteristics of life.
Atoms are tiny particles that are put together to build living organisms. While atoms do not possess life, they are required for life. The atom is the smallest organizational level that we will explore in this course, and you will learn more about atoms in Unit 2: Introduction to Chemistry. You will also learn how atoms can be linked together by chemical bonds to form molecules. In Unit 3: Biological Macromolecules, you will focus on the four classes of carbon-based macromolecules that comprise living systems: carbohydrates, fats, proteins and nucleic acids.
Atoms and molecules are not alive. After exploring the world of atoms and molecules, you will learn how these building blocks are put together to construct the fundamental unit of life: the cell. The cellular level of organization is the first time all the characteristics of life emerge, resulting in a structure that is able to maintain its own internal constancy. In Unit 4: The Cell, you will learn how the unique structures that make up a cell enable it to maintain homeostasis and carry out the varied functions of life.
Energy is required to sustain life and all living organisms need energy to fuel their metabolic activities. Some organisms get this energy directly from the sun through the process of photosynthesis. Other organisms harvest their energy from the food they eat. Humans are an example of organisms that obtain energy from food. In Unit 5: Metabolism, you will take a closer look at what energy is and explore how different cells acquire and use energy.
All living things grow, develop and reproduce. We will focus on the ways different types of cells reproduce themselves in Unit 6: Cell Division. Genetic information in cells provides the instructions for carrying out life processes. In order for life to continue, organisms must pass information to the next generation. How genetic information is passed from one generation to the next is discussed in Unit 7: Classical Genetics. The physical traits an organism displays are ultimately determined by the DNA found within the organism’s cells. In Unit 8: Molecular Genetics, we will connect heredity to DNA, the genetic material of the cell. We will specifically explore DNA function and figure out how DNA determines the heritable traits that individual organisms can pass to their offspring.
Life changes and evolves over time. In Unit 9: Evolution, you will examine the process of evolution taking place within populations of organisms. This unit will explicitly link evolutionary change with the heritable characteristics you will learn about in Classical Genetics and Molecular Genetics.
You will conclude your exploration of biology by learning about the interactions between groups of living organisms and their environments. Unit 10: Ecology discusses how humans are part of a living biosphere and how our actions, both intentional and inadvertent, have widespread consequences.
Biology is the scientific study of life and is the branch of science that studies living organisms and the way organisms interact with their environments. The subject is vast and includes topics as diverse as acid rain, evolution, and genetically modified foods. In this module, you will investigate the definition of life and explore some of the characteristics of living systems.
![]() |
![]() |
There are five distinct qualities used to determine whether or not something is (or was) alive. A living organism is something that displays all these qualities. To be considered alive, something must:
The cell is the smallest unit that displays all of these characteristics. Because of this, living organisms are often identified based on whether or not they are made of cells. Nonliving things can show several of these characteristics. For example, a rock crystal can “grow” in a simple fashion. However, if even one of these conditions is not met (rock crystals do not reproduce with help from DNA), the object in question cannot be considered alive.
Matter is any substance that has mass and takes up space. All matter can be classified in one of two categories: biotic (living) or abiotic (nonliving). Matter is considered biotic if it was ever alive at some point in time. In this sense, a dead human buried underground in a wooden coffin is still biotic, as is the wood used to make the coffin (the wood came from a tree that was once alive). However, not everything within, or made by, a biotic organism is biotic. For example, urea, a chemical component of urine, is an abiotic substance.
To be classified as biotic, all of the required conditions for life must be met or have been met in the past. Otherwise, the matter being classified is considered to be abiotic, or nonliving.
Observe the following diagram.
All living things consist of smaller parts that are organized in a hierarchical way.
Living things are highly organized and structured, following a hierarchy that can be examined on a scale from small to large. In this course, the smallest level we will examine is the atom, which is the basic unit of matter. The atom consists of a dense nucleus surrounded by electrons. Atoms join together to form molecules. A molecule is a chemical structure consisting of at least two atoms held together by a chemical bond. Macromolecules are biologically important molecules, and they are technically polymers. Polymers are made by combining smaller units called monomers, which are simpler macromolecules. An example of a macromolecule is the genetic molecule deoxyribonucleic acid (DNA) that contains the instructions for the development of all living organisms. DNA is built of four kinds of monomers (nucleotides). They are strung together, or polymerized, in a sequence that codes for the structure of proteins and other biological molecules. The DNA in one of your microscopic cells contains a sequence of almost three billion nucleotides.
When macromolecules are used as building blocks to form a membrane-bound sphere, you have a cell, which is the fundamental unit of life. A cell is essentially a tiny droplet of water and other molecules enclosed by a fluid “skin” or membrane. The cell is the smallest and simplest entity that possesses all the characteristics of life. There are two main types of cells: prokaryotes and eukaryotes. The cells of prokaryotes are relatively small and simple; they do not have any clearly defined compartments inside of them. The cells of eukaryotes, by contrast, include membrane-bound organelles: compartments inside the cell that contain specific groups of macromolecules and carry out specific cellular functions. One of these organelles is the nucleus; it encloses the DNA within the cell.
Some organisms consist of just one cell and include unicellular organisms such as bacteria and protists. Single-celled life forms are typically referred to as microorganisms. Other organisms consist of many cells working together. These multicellular organisms include animals, land plants, and most fungi. Most multicellular organisms have cells that are specialized to carry out specific functions. Tissues are formed when many different kinds of cells work together to fulfill the same detailed function. Organs are collections of tissues that work together to carry out a common general function. Organs are present not only in higher level animals but also in plants. An organ system is a higher level of organization that consists of functionally related organs. Mammals have many organ systems. For example, the circulatory system transports blood through the body and includes organs such as the heart and blood vessels. Organisms are individual living entities that survive and reproduce as a unit. For example, each tree in a forest is usually an individual organism.
Consider this example to help clarify the nature of the levels between a cell and an organism. A human is an organism which has a circulatory system (organ system) that transports blood through the body. It is made up of organs such as the heart and blood vessels. Each of the organs, in turn, is made of more specific tissues. Your heart, for example, has muscle tissue for pumping and nerve tissue that helps coordinate each heartbeat.
The hierarchical organization of living systems continues beyond single organisms. A population consists of all the individuals of a species living within a specific area. For example, a forest may include many pine trees. All those pine trees represent the population of pine trees in that forest. As you know, many different populations can live in any specific area. All of these populations can interact with each other in positive and negative ways, and together they form a community. Continuing with our example, the forest with pine trees includes populations of flowers, mammals, birds, insects, fungi, and bacteria, all of which can interact. These interacting populations make up a community. An ecosystem consists of all the living things in a particular area together with the abiotic, or nonliving, parts of that environment. The pine forest ecosystem includes not just plants, animals, and microbes but also rocks, water, temperature changes, air chemistry, and other abiotic factors that interact with living organisms in the area. Finally, the highest level of organization in living systems is the biosphere, which is the collection of all ecosystems on Earth. The biosphere includes all habitable zones on the planet, including land, soil, and rocks to a great depth in the Earth’s crust; water and ice; and the atmosphere to a great height.
Life is organized into hierarchical levels of increasing complexity. The study of biology involves all these levels, from single atoms or molecules up to global relationships among organisms and the environment. As we ascend through this hierarchy to more and more complex levels, emergent properties appear. These are characteristics of a system that are not present in any of its component parts. Take, for example, an automobile. The separated parts of the automobile amount to a heap of junk. Only when properly assembled, with gas, the right key, and a human driver, does the car fulfill its function, which is to transport us from place to place.
Life is an emergent property, and one that appears at the cellular level of organization. Molecules are not alive, but they are the components of life. More than 2,000 years ago, Aristotle said “The whole is much greater than the sum of its parts.” In biology, this is a constant theme: we can learn much about a system by looking at its details, but we also must step back and look at the big picture to truly understand the workings of life.
Have you ever gotten sick with the flu and said something like, “I’ve come down with a bug,” or “I’ve got the flu bug”? In fact, the flu (short for influenza) is not really caused by a “bug” at all. It is caused by a virus, and there is some controversy in the scientific community regarding whether or not viruses are alive. But what are viruses, really? And how would you determine whether or not they are alive? In this lesson, you’ll learn more about what viruses are, so you can figure out if they are actually alive.
Viruses contain genetic information in the form of either DNA or RNA. The genetic information is surrounded by a protein coat called a capsid. Some viruses also have a membrane structure surrounding their genetic information. Viruses are tiny particles and they lack the machinery necessary for growth and reproduction. In order to reproduce, a virus must infect a host cell and hijack the host cell’s machinery. In other words, a virus cannot reproduce without using the tools of another cell.
So are viruses alive? There is much debate on the subject. Complete the following activity and draw your own conclusions.
Throughout the study of biology, several themes emerge. These recurrent concepts will continually appear as we delve deeper into the scientific study of life. As we progress through the course, we will highlight the lessons that help to illustrate these six themes:
In this module, we will explore an overview of each of these themes separately.
Biologists describe how organisms are put together (structure) and explain how organisms stay alive, move, grow, reproduce, and do other activities (function). In biological systems, the structure of a cell or body part is tightly linked to its function within the life of the organism.
The structure of something is determined by two factors: its three-dimensional shape, and the materials from which it is made. The structure that something takes directly influences its possible functions. Consider a piece of wood. Depending on how it is shaped, it could be used as a spear for hunting, a cup for drinking, a pipe for smoking, or even a flute for playing music. The wood also possesses unique properties that influence how it can be used. For example, the piece of wood could never function as a hot air balloon, no matter what kind of shape it took. It would always be too heavy to lift off the ground.
At the smallest levels of biological organization, the structure of a molecule determines its function within a cell. In this course, you might investigate the structure of a molecule and then learn what it does. It will help you to remind yourself: the structure and the function of this molecule are connected.
Enzymes are molecules found within a cell that speed up the rate of the chemical reactions necessary to support life. Enzymes break down food molecules, help build muscle proteins, can destroy toxins, and much more. An enzyme’s ability to function depends directly on its three-dimensional shape. If exposed to excessive heat or harsh chemical conditions, enzymes unravel and change shape. When this occurs, the enzymes stop working and the chemical reactions of life slow down or cease.
At a larger scale, the structure of a cell is directly linked to its function within the body of a multicellular organism. Compare the following images of some different cell types and explore how structure relates to function at the cellular level.
![]() |
![]() |
![]() |
![]() |
Structure also determines function at higher levels of biological organization. Your body’s organs function as they do because of how tissues are put together, forming filters, pumps, levers, surfaces for gas exchange, and pipes for air and fluid flow. And, of course, the entire body of an organism is suited to its lifestyle and environment. For example, the earthworm is well-suited for burrowing: it is slimy, muscular, flexible, and cylindrical. Yet there are many surprising details in the worm’s structure. The next time you get the chance, feel an earthworm. Pinch it carefully near the head with one hand, then rub your free forefinger against its belly. You’ll notice prickly hairs on the worm’s underside. And you’ll find that you only feel them when you move your finger toward the worm’s head: the hairs are set at an angle, enabling the worm to grip the soil to and move it forward through the earth.
You can probably imagine other ways to use different shapes and materials to build a functional body that enables Structo to overcome many different challenges. Environmental pressures have led (and continue to lead) to the evolution of organisms with a virtually infinite range of such combinations.
In fact, organisms are so well adapted to their environments that humans are now looking to them for inspiration. Today, a fast-growing field called biomimicry brings biologists together with engineers to develop new products based on solutions found in nature. One of the earliest products of this approach was the invention of Velcro (the hook and loop attachment) by Swiss engineer Georges de Mestral. He was inspired by the burrs that got stuck in his dog’s fur. To learn more, visit the Biomimicry Institute.
Living organisms detect and respond to changes in the conditions of their external environment. In response to threats and opportunities, organisms may move or change their activities. For example, plant stems can grow toward a source of light. Over time, the trunks of trees strengthen when they are flexed by the wind. Animals, of course, respond to stimuli with a huge range of behaviors from the hibernation of bears to your own development of “goosebumps” on a chilly day.
In contrast to the extreme variability of the outside world, the interior of a cell is a remarkably constant environment. For example, cells typically keep their internal pH (acidity) within a narrow range. This is important because changes in pH can cause molecules like enzymes to change their shapes. As you already know, a molecule’s shape (or structure) determines its function. If its shape changes, it may lose function. Because of this, many of the chemical reactions of life will not work properly if cell conditions change too much.
The ability or tendency of organisms and cells to maintain stable internal conditions is called homeostasis. The term homeostasis comes from the Greek words homeo (same, alike) and stasis (standing). It describes how life stands in one place despite many changes in the surrounding world.
Environmental conditions can also affect cellular homeostasis. If the cell’s surroundings change quickly, conditions inside the cell may be temporarily disturbed. Organisms react to these changes in a corrective way: they detect changes and do something to oppose them. For example, when you get cold, your muscles start to contract in rapid bursts, causing you to shiver. The process of shivering then generates heat that warms you back up again. As a result of this and other adjustments, your temperature is maintained close to 99 degrees F (Fahrenheit) [37 degrees C (Celsius)]. At this temperature, your enzymes work very efficiently.
Homeostasis is an important theme in biology. All living systems use resources to maintain homeostasis. When cells fail to maintain homeostasis, disease results. Ultimately, if homeostasis is not restored, an organism will die.
Matter is traditionally defined as anything that has mass and takes up space. Matter is made of atoms. Matter is reused and recycled in living systems. To live and grow, organisms and cells must take in (or absorb, or ingest) certain forms of matter. Any matter an organism needs but cannot make for itself is considered a nutrient for that organism. Not all matter can be used by an organism, which is why all living systems release other forms of matter. When an organism or cell releases (or excretes) matter, the excreted matter is considered waste for that organism.
Both nutrients and wastes are made of atoms; they occupy space and have mass. The atoms retain their identity through the processes of life, even though they can be combined with other atoms in different ways. Consider the picture above. When a wolf is eaten by a bear, the atoms that made up the wolf become part of the bear. The matter in the wolf that is not absorbed or used for growth by the bear becomes waste. The bear excretes this waste in urine, in feces, and in its exhaled breath. Atom for atom, all the matter that was the wolf can be accounted for, and it is recycled through the ecosystem. The waste produced by one organism can provide nutrients to another.
Energy can be defined as the capacity to do work or to make a change in the location, temperature, or structure of matter. Energy does not have mass and it does not take up space, but it can be measured in terms of what it does. Energy comes in many forms, including heat, chemical potential energy, kinetic energy of motion, and light. Energy is required for all organisms to maintain homeostasis, grow, and reproduce.
It is important at this point to clarify the difference between matter and energy and to highlight a form of energy that is very important in biology. Food contains nutrients that are useful to humans. These nutrients take up space and can be weighed. Imagine a serving of breakfast cereal. You can weigh and measure the volume of its nutrients, and on food labels nutrient content (the amount of fat, carbohydrates, protein, etc.) is reported in grams. But the same serving of cereal also has a property called chemical potential energy. This is energy stored in the structure of molecules that can be converted to other forms of energy by a chemical reaction. If the cereal is burned (a chemical reaction), its molecules will react with oxygen gas. Atoms will change partners and the molecules will become much simpler. In the process, energy will be released as heat and light. We can determine the amount of energy in the cereal by burning it and measuring how much heat is released in the process.
In biology, chemical potential energy is a key form of energy. It is measured in units called calories. When we’re talking about energy, we’ll say “the calories stored in ...” or “the chemical potential energy of ...” a substance. We are talking about what the substance can do if it is put through a chemical reaction, like burning, that will release its stored energy. Returning to the breakfast cereal example, nutrition labels indicate the chemical potential energy of food by listing the calories per serving.
Energy and matter move through living systems in different ways. Energy flows through living systems, changing forms as it goes. For example, the energy in sunlight is captured by green plants, which use this energy to build sugar molecules. The energy from the sun is now stored in the sugar and when an organism eats the sugar, the stored energy can be harvested and used to do work. The energy flows through the system; it is never recycled. Matter, on the other hand, cycles within living systems. For example, the atoms in the sugar molecule start out as nutrients, and will ultimately become waste. The waste might become nutrients for something else. Those same atoms will be used over and over again. Energy flow and nutrient cycling are themes in biology at every level.
Have you ever played a game of Jenga, where you try to remove one block at a time from a tower structure without causing the entire tower to collapse? In this game, each block depends on the other blocks for stability, and if you’re not careful, the removal of a single block can cause the destruction of the entire structure. Life is much the same. At levels from the individual to the biosphere, the various parts of living systems are interdependent. Your trillions of cells are intimately dependent on each other for survival. Your cells are specialized, and you exist because of a massive team effort from all these different cells. If any one of your major organs were to suddenly fail, you might die very quickly. Moreover, your health also depends on trillions upon trillions of bacterial cells that live in and on your body. They provide your cells with vitamins, help keep out invaders, and may even influence your mood. So your body itself is a community of interdependent cells. |
Jenga Game![]() |
An ecosystem consists of all the organisms living within a defined area along with the abiotic components of that particular environment. Within an ecosystem, organisms interact with each other in helpful and harmful ways. Some types of organisms play overlapping roles with other species (e.g., six species of oak tree may have similar, if not identical, roles in a forest). Others, however, play unique and essential roles that other organisms depend upon for their survival. For example, in a Colorado forest, researchers found that a species of sapsucker (a bird similar to a woodpecker) played a vital role in providing nesting holes and food for many other species.
To learn more: Gretchen Daily and others, 1993. Double keystone bird in a keystone species complex.
All organisms grow and develop. Growth is just an increase in size. In development, structure and function change in an orderly way as an organism passes through its life cycle. An individual’s pattern of development is partly determined by genetic instructions. DNA, the molecule of inheritance, encodes proteins and other molecules that build cells and make them work. You can think of genes as “recipes” for proteins. Each cell has a huge library of thousands of recipes in its DNA. But most of the recipes don’t get used in any given cell. Chemical controls tell the “cooks” whether to make a given protein, or ignore its recipe and make something else. Therefore, even though your cells share the same DNA, their activities change radically in different parts of your body and at different stages of development.
In this time-lapse video, a few days of development are sped up for your viewing pleasure. In the earliest frames you can see large cells dividing. With each division the individual cells get smaller. Chemical signals in each cell are passed to its descendants and to nearby cells. As a result, DNA is turned “on” or “off” in each cell, leading to different shapes and structures. Toward the middle and end of the video, individual cells are much too tiny to see. Specialized cell types form. Cells move from place to place, grow, die to create gaps, and divide to create bulges. In this way, body structures begin to emerge.
Reproduction occurs when an individual organism passes on its genetic information to a newly independent organism, or offspring. Individual organisms have a limited life span. All are subject to extreme events that bring death, such as being eaten by a predator. Even if they avoid such a fate, most organisms decline in health and die toward the end of their typical life span. Reproduction maintains genetic information by passing it to new individuals. Offspring resemble their parent(s) and reproduction maintains the continuity of life over time.
Evolution is a scientific theory that explains how and why life changes over time. Evolution provides the explanation for why all living organisms share profound similarities, and yet, the life forms on our planet are so incredibly diverse. The two fundamental tenets of evolution are shared ancestry and natural selection.
As you learned previously in this module, reproduction results in the passing of characteristics from parents to offspring and maintains the continuity of life. You can see this in a family photograph: offspring resemble their parents and siblings resemble each other. Biologists look at all life on the planet and see the same patterns.
According to evolutionary thinking, life forms are similar because of their shared ancestry. All cells have DNA as the molecule of inheritance and they also share many other detailed features. Organisms and cells always come from parents, and this is where they get their genetic information. Trace this process back in time, and it is reasonable to suppose that all life forms inherited DNA and many other features from a long-ago common ancestor. The unity of life is very striking when we examine molecules, which can be almost identical in species as different as bacteria and whales. This unity of structure and function is also evident when we compare skeletons, kidneys, or hearts among animals.
The second major idea of evolution is that of natural selection. Natural selection helps explain how groups of organisms become well-suited, or adapted, to their surroundings. Individuals are always a bit different from their parents and from each other, partly because of changes to their genes. These differences may be helpful or harmful to the individuals that inherit them. In nature, individuals often have very low odds of surviving to reproduce. Individuals with slightly harmful or even average characteristics might be less likely to make it, and those with traits that fit in very well with the local habitat will have the greatest chance to survive and reproduce. This sorting process goes on generation after generation. Each time only a tiny fraction of those born are well-suited and lucky enough to pass their genes along; the others die, leaving no descendants. As generations of time pass, inherited features that help organisms produce offspring become more common within a population. Harmful features that reduce reproduction become less common. Many experiments have demonstrated the effects of natural selection on populations in the laboratory and in nature. The process of natural selection is one way that scientists have explained the vast diversity of life on our planet.
Science is a powerful tool used to understand the natural world. It is the process that has resulted in the knowledge we are exploring in this course. However, the process of science does have limitations. In other words, there are some questions that science cannot be used to answer. For example, science cannot answer questions about supernatural things. It also cannot be used to make moral or aesthetic judgments. While questions of these types may be important for humans to explore, the process of science cannot be used to study them.
In this module, the process of science will be discussed in more detail. Then you will learn about how to design a quality scientific experiment. Finally, you will have the opportunity to practice identifying examples of science and to compare those to examples of pseudoscience, or things that claim to be scientific, but are not.
Science is a process that helps us to understand how the natural world works. When we use the term “natural” in this context, we mean all that can be observed with our senses or with instruments that extend our senses. The “natural” world studied by science can be reliably observed and measured, from the far reaches of outer space to the man-made chemicals in our air and water. This module will examine the process of science and explore how science has led to an increased understanding of living organisms.
Biology is an example of a scientific field that deals with understanding living organisms. Chemistry is a scientific field that deals with understanding atoms and molecules. Overall, scientific fields can be broken down into two main branches: natural sciences, and social and behavioral sciences. Social and behavioral sciences focus on human cultures and the behaviors of humans in groups and as individuals. Biology is a natural science that includes many related sub-branches, such as ecology, biochemistry, and microbiology. Each sub-branch of biology examines different aspects of living organisms. Other natural sciences are physical sciences like chemistry and physics. It is important to remember that even though scientific knowledge is broken into branches and sub-branches of study, the knowledge gained from the various scientific disciplines is interconnected. For instance, to understand how a cell responds to a certain environmental signal, it is necessary to know about the chemical composition of that signal, so the science of chemistry is important to understand the science of biology. Likewise, understanding how the brain functions (neuroscience) can aid in understanding the human behaviors studied in the social and behavioral sciences.
Science deals with testable knowledge about physical phenomena in the universe. The goal of science is to understand how the universe works. Biology focuses on understanding living things. To gain knowledge about nature and physical phenomena, scientists use a particular approach called “scientific inquiry.”
Scientific inquiry is the best approach we have to understanding the natural world and predicting natural phenomena. Evidence for this claim can be found in the successes of science-based technologies. Take medicine, for example. Prior to the 1700s, most medical practices were based on folk traditions or on ideas promoted by religious leaders. Some of these prescientific remedies worked, but the process for discovering new treatments was a slow and haphazard system of trial and error. Ineffective treatments were often accepted simply because there was no clear procedure for evaluating them. Today, with science-based medicine and public health practices, we have gained unprecedented control over threats to our health. According to the Centers for Disease Control, the average life expectancy in the United States has increased by more than 30 years since 1900.
Scientific inquiry has not displaced faith, intuition, and dreams. These traditions and ways of knowing have emotional value and provide moral guidance to many people. But hunches, feelings, deep convictions, old traditions, or dreams cannot be accepted directly as scientifically valid. Instead, science limits itself to ideas that can be tested through verifiable observations. Supernatural claims that events are caused by ghosts, devils, God, or other spiritual entities cannot be tested in this way.
The rest of this module will focus on the methods of scientific inquiry. Science often involves making observations and developing hypotheses. Experiments and/or further observations are often used to test the hypotheses, and the data gathered are carefully interpreted. Methods and results are then communicated to other scientists within peer-reviewed scientific journals.
A scientific experiment is a carefully organized procedure in which the scientist intervenes in a system to change something, then observes the result of the change. Scientific inquiry often involves doing experiments, though not always. For example, a scientist studying the mating behaviors of ladybugs might begin with detailed observations of ladybugs mating in their natural habitats. While this research may not be experimental, it is scientific: it involves careful and verifiable observation of the natural world. The same scientist might then treat some of the ladybugs with a hormone hypothesized to trigger mating and observe whether these ladybugs mated sooner or more often than untreated ones. This would qualify as an experiment because the scientist is now making a change in the system and observing the effects.
When conducting scientific experiments, researchers develop hypotheses to guide experimental design. A hypothesis offers a testable and falsifiable explanation of observations. For example, a scientist might observe that maple trees lose their leaves in the fall. She might then propose a possible explanation for this observation: “cold weather causes maple trees to lose their leaves in the fall.” This statement is testable. The scientist could grow maple trees in a warm enclosed environment such as a greenhouse and see if their leaves still dropped in the fall. The hypothesis is also falsifiable. If the leaves still dropped in the warm environment, then clearly temperature was not the main factor in causing maple leaves to drop in autumn.
In the activity below, you can practice recognizing scientific hypotheses. As you consider each statement, try to think as a scientist would: can I test this hypothesis with observations or experiments? Is the statement falsifiable? In other words, is it possible to gather evidence that clearly indicates that the statement is not true? If the answer is “no,” the statement is not a valid scientific hypothesis.
Throughout the rest of this module, we examine the scientific process by discussing an actual scientific experiment conducted by researchers at the University of Washington to investigate whether a vaccine may reduce the incidence of the human papillomavirus (HPV). The experimental process and results were published in an article titled, "A controlled trial of a human papillomavirus type 16 vaccine" [1] , available for viewing online at the National Institutes of Health's online database of publications: PubMed.gov
Preliminary observations made by the researchers who conducted the HPV experiment are listed below:
You’ve successfully identified a hypothesis for the University of Washington’s study on HPV: People who get the HPV vaccine will not get HPV.
The next step is to design an experiment that will test this hypothesis. There are several important factors to consider when designing a scientific experiment. First, scientific experiments must have an experimental group. This is the group that receives the experimental treatment necessary to address the hypothesis.
The experimental group receives the vaccine, but how can we know if the vaccine made a difference? Many things may change HPV infection rates in a group of people over time. To clearly show that the vaccine was effective in helping the experimental group, we need to include in our study an otherwise similar control group that does not get the treatment. We can then compare the two groups and determine if the vaccine made a difference. The control group shows us what happens in the absence of the factor under study.
However, the control group cannot get “nothing.” Instead, the control group often receives a placebo. A placebo is a procedure that has no expected therapeutic effect — such as giving a person a sugar pill or a shot containing only plain saline solution with no drug. Scientific studies have shown that the “placebo effect” can alter experimental results because when individuals are told that they are or are not being treated, this knowledge can alter their actions or their emotions, which can then alter the results of the experiment.
Moreover, if the doctor knows which group a patient is in, this can also influence the results of the experiment. Without saying so directly, the doctor may show — through body language or other subtle cues — his or her views about whether the patient is likely to get well. These errors can then alter the patient’s experience and change the results of the experiment. Therefore, many clinical studies are “double blind.” In these studies, neither the doctor nor the patient knows which group the patient is in until all experimental results have been collected.
Both placebo treatments and double-blind procedures are designed to prevent bias. Bias is any systematic error that makes a particular experimental outcome more or less likely. Errors can happen in any experiment: people make mistakes in measurement, instruments fail, computer glitches can alter data. But most such errors are random and don’t favor one outcome over another. Patients’ belief in a treatment can make it more likely to appear to “work.” Placebos and double-blind procedures are used to level the playing field so that both groups of study subjects are treated equally and share similar beliefs about their treatment.
A variable is a characteristic of a subject (in this case, of a person in the study) that can vary over time or among individuals. Sometimes a variable takes the form of a category, such as male or female; often a variable can be measured precisely, such as body height. Ideally, only one variable is different between the control group and the experimental group in a scientific experiment. Otherwise, the researchers will not be able to determine which variable caused any differences seen in the results. For example, imagine that the people in the control group were, on average, much more sexually active than the people in the experimental group. If, at the end of the experiment, the control group had a higher rate of HPV infection, could you confidently determine why? Maybe the experimental subjects were protected by the vaccine, but maybe they were protected by their low level of sexual contact.
To avoid this situation, experimenters make sure that their subject groups are as similar as possible in all variables except for the variable that is being tested in the experiment. This variable, or factor, will be deliberately changed in the experimental group. The one variable that is different between the two groups is called the independent variable. An independent variable is known or hypothesized to cause some outcome. Imagine an educational researcher investigating the effectiveness of a new teaching strategy in a classroom. The experimental group receives the new teaching strategy, while the control group receives the traditional strategy. It is the teaching strategy that is the independent variable in this scenario. In an experiment, the independent variable is the variable that the scientist deliberately changes or imposes on the subjects.
Dependent variables are known or hypothesized consequences; they are the effects that result from changes or differences in an independent variable. In an experiment, the dependent variables are those that the scientist measures before, during, and particularly at the end of the experiment to see if they have changed as expected. The dependent variable must be stated so that it is clear how it will be observed or measured. Rather than comparing “learning” among students (which is a vague and difficult to measure concept), an educational researcher might choose to compare test scores, which are very specific and easy to measure.
In any real-world example, many, many variables MIGHT affect the outcome of an experiment, yet only one or a few independent variables can be tested. Other variables must be kept as similar as possible between the study groups and are called control variables. For our educational research example, if the control group consisted only of people between the ages of 18 and 20 and the experimental group contained people between the ages of 30 and 35, we would not know if it was the teaching strategy or the students' ages that played a larger role in the results. To avoid this problem, a good study will be set up so that each group contains students with a similar age profile. In a well-designed educational research study, student age will be a controlled variable, along with other possibly-important factors like gender, past educational achievement, and pre-existing knowledge of the subject area.
After the experiment is completed, results are compiled and interpreted. This involves the measurement of the dependent variable. In the case of our HPV experiment, remember, the dependent variable is the rate of HPV infection.
Although the HPV study suggests that the vaccine protects against infection by HPV, is the finding significant? In science, as in life, things can happen for many different reasons. A convincing study will rule out “luck” (random chance) as an explanation for the results. Strong results are said to be significant: very unlikely to occur by chance or random events.
Whether the outcome is significant often depends on the size of study; the larger the number of individuals enrolled, the more convincing the results are likely to be. For example, imagine only 10 women were enrolled in the study. In the control group, 2 in 5 of the women became infected. In the experimental group, 0 in 5 were infected. At first you might think this proves the vaccine’s effectiveness, but it is NOT a convincing or significant result. Why not? Random events could easily explain the difference between the groups. For example, perhaps none of the five women in the experimental group were sexually active over the study period. They therefore stood no chance of acquiring HPV. The vaccine might appear to work, but a skeptical reader could account for the results by proposing many other scenarios.
However, imagine if the same study were done with 10,000 women, and the infection rates were 2,000 of 5,000 in the control group and zero of 5,000 in the experimental group. Random events would be spread out among a very large group of people in this study; on average, the two big groups should have similar sexual behavior and other factors influencing infection rates. If there is a big difference at the end of the study, it is very unlikely that this result occurred by random chance.
Statistical analyses did support the significance of the HPV vaccine result.
After the results are interpreted and conclusions are drawn, researchers often return to their work and begin asking further questions. In this way, scientific inquiry is a powerful tool for exploration.
Now that you have a pretty good idea about the process of science, you’ll have a chance to identify examples of science and compare those to examples of pseudoscience.
Pseudoscience is any claim that purports or pretends to be scientific in nature, but does not actually have the characteristics of true scientific inquiry. Pseudoscientific ideas often involve the supernatural. Sometimes they involve claims about forces or processes that cannot be measured using traditional tools or instruments employed by scientists. Finally, pseudoscientific claims are often quite dramatic. They are “amazing” ideas that would seem strange or unlikely to most scientists working in a related field.
The clearest line separating pseudoscience from “real” science is publication in peer-reviewed scientific journals. These are publications, usually run by scientific societies or academic publishing companies, in which scientists publish their findings according to a well-established system of oversight. Before it gets published in such a journal, a piece of research is carefully reviewed by two or more researchers in the same field of study. The methods and logic of the paper are evaluated carefully, and if it makes bold or unusual claims, the study is subject to especially close scrutiny. Reviewers and editors may demand not just rewrites, but also additional evidence if a claim is weakly supported.
Truly scientific research is published in reputable peer-reviewed journals. These journals exclude pseudoscience rigorously. If you hear that some idea is “scientifically proven,” check the source to see if there’s any reference to a scientific journal. Track down the article to see if it really supports the claim. Finally, double check to make sure that the publication is a respected peer-reviewed journal. One way to get this information is to ask a scientist at a research university, consult with a librarian, or do a careful Internet search. Your ability to distinguish between science and pseudoscience will help you be a scientifically literate citizen.
Matter is anything that has mass and takes up space. The chair you are sitting in is made of atoms. The food you ate for lunch was built from atoms. Even the air you breathe is made of atoms. In this unit, we will learn more about atoms, the fundamental unit of matter.
The cell is the fundamental unit of life; therefore, you might think that your study of biology should begin with the cell. In fact, before you can truly understand how a cell functions, you must understand the building blocks from which cells are made. In the Introduction to Biology module, you learned that all life exhibits a hierarchical organization. While cells are the first level of organization that displays all the properties of life, their component parts (atoms and molecules) are not alive, as illustrated in the inverted pyramid below.
To make sense of how cells work, you must constantly return to their parts (atoms and molecules) and the interactions among them. This is why a clear understanding of some chemistry is essential to understanding how life functions. This unit will focus on the atom. You will learn how atoms combine to form different molecules, making up all the diverse matter you see around you. In this unit, you will explore only nonliving components of the hierarchy of life.
In this module, you will learn about the basic structure of the atom, the fundamental unit of matter. Living things are made up of atoms arranged in a complex and nonrandom way. This is one of the common features of all living organisms. The atom is the smallest level in the hierarchy of life that we will explore in this course. Understanding the properties of atoms allows us to predict and understand how atoms interact with one another to build molecules, such as hormones and DNA. In addition, an understanding of atoms allows us to predict how cells will react to different therapeutic drugs and toxins in the environment.
All living and nonliving things are composed of matter. Matter can be defined as anything that occupies space and has mass. The mass of an object is a measure of how much matter it has (that is, how much “stuff” is in it). Mass is not exactly the same as weight, but here on Earth we can measure and compare the masses of different objects by weighing them.
When we explore matter scientifically, we find that it takes on many different structures as we zoom in at smaller and smaller scales. Everyday objects are mostly mixtures of molecules, which in turn are made up of atoms. Even air is a mixture of molecules. The atom is a good focal point for understanding matter; all matter is composed of atoms.
Atoms are unimaginably small. Even within a single microscopic cell, there is room for not just billions, but trillions or even hundreds of trillions of atoms. Amazingly, however, physicists now understand that most of the volume of an atom is actually made up of empty space. The atoms themselves are made of even smaller subatomic particles called protons, neutrons, and electrons. We cannot look at the parts of atoms with a microscope; they are simply too small. However, physicists have learned a great deal about atoms and subatomic particles through indirect methods. One model of the atom is shown in the diagram below.
There are 92 different kinds of atoms that are naturally occurring on Earth. These different types of atoms are called elements. Each element has its own set of properties that are unique to atoms of its kind. Atoms differ from one another in their number of protons, electrons, and neutrons. Atoms are the smallest unit of an element that retains all of the properties of that element. So, an element is a substance that is composed of a single type of atom. For example, a hydrogen atom contains one proton and one electron. In contrast, a carbon atom contains six protons, six electrons, and six neutrons. Elements are designated by either one- or two-letter abbreviations and they can be organized into a chart called the periodic table. We will take a closer look at the periodic table (also called the periodic chart) later in this unit.
In biological systems, the major elements are carbon (C), hydrogen (H), nitrogen (N), oxygen (O), phosphorus (P), and sulfur (S). These elements represent more than 95% of the mass of a cell. Carbon is a major component of nearly all biological molecules. Some elements are found in relatively small amounts and are called “trace elements.” Examples include sodium (Na), potassium (K), chlorine (Cl), manganese (Mn), and Zinc (Zn). Throughout the course you will see how atoms of these elements are very important to the functioning of a cell.
Elements (atoms) are characterized by their atomic structure, which is made up of subatomic particles: protons, neutrons, and electrons. Protons and neutrons reside in the nucleus (center) of the atom and have a mass of one atomic mass unit (amu) each. Electrons are found outside of the nucleus, in zones that are called “shells.” Electrons have almost no mass.
The mass of an atom is called the atomic mass. When calculating atomic mass, we pay attention only to the protons and neutrons; the electrons have almost no mass. The atomic mass is the sum of the number of protons and the number of neutrons. By summing the atomic mass of all the atoms in a molecule, one can estimate the molecular mass of the molecule, which is expressed in atomic mass units (called Daltons). Each of the heavy particles (neutron, proton) weighs one atomic mass unit, so a Helium (He) atom, which has two protons, two neutrons, and two electrons, weighs about four atomic mass units; that is, two protons plus two neutrons.
Particle | Charge | Location in Atom | Relative Mass |
proton | positive | nucleus | 1.0 |
neutron | no charge | nucleus | 1.0 |
electron | negative | orbit outside nucleus | negligible |
In addition to location and mass, each subatomic particle has a property called “charge.” Charge can be “positive” or “negative.” Items with the same charge tend to repel each other and items with opposite charges tend to attract each other. Protons have a positive charge and neutrons have no charge, giving the nucleus a positive overall charge. Each electron has a negative charge that is equal in strength to the positive charge of a proton. Electrons and the protons of the nucleus attract each other, and this is the force that keeps the atom together, much like the force of gravity keeps the moon in orbit around Earth.
Atoms of different elements have many different features, including different sizes and levels of reactivity. For example, sodium is an incredibly reactive element, especially when it combines with water. Lead, on the other hand, is relatively inert. Yet all elements share some regular patterns that make it easier for us to categorize them and understand why they behave as they do.
The first key characteristic that truly identifies an element is its atomic number, which is the number of protons in each atom. The atomic number is constant and identical for all atoms of an element. For example, hydrogen (H) atoms always have only one proton and have an atomic number of one. If an atom has two protons, it has an atomic number of two, and it is helium (He). For this reason, if you started with one hydrogen atom and added a proton, the original hydrogen would now be helium. If an atom has three protons in its nucleus, it is a lithium (Li) atom. Beryllium (Be) has four protons in its nucleus, and so on. Based on their atomic numbers, elements can be organized into a periodic table of elements like the one pictured below. The number at the top of each box in the table is the atomic number of the element.
In addition to providing information about each element’s atomic number, most periodic tables also provide an atomic mass for each element. The atomic mass, also known as the atomic weight, reports the mass of the nucleus. The mass of the nucleus is the sum of all protons and neutrons found in the nucleus. A hydrogen atom has one proton and would therefore have an atomic mass of one. A carbon atom, which has six protons and six neutrons, would have an atomic mass of 12.
The units of atomic mass are Daltons, abbreviated Da, named after the British scientist John Dalton. One Da is equal to 1.66 x 10-27 kg, which is the mass of a single hydrogen (H) atom. Helium (He) has a mass of 4 Da or 6.67 x 10-27 kg. Because the mass of a single atom is so small, it is not a convenient unit for everyday use. Typically, the atomic mass is multiplied by the number of atoms in a mole to give the atomic weight in grams. A mole contains 6.022 x 1023 particles and is defined as the number of particles of carbon (C) that give an atomic weight of 12 grams.
Notice that the atomic mass or atomic weight on the periodic table is not generally a whole number. For example, the atomic mass of carbon (C) is 12.01. How is that possible? Atoms cannot have fractions of protons or neutrons. The reason for this is that not all atoms of an element have the same number of neutrons. For example, carbon atoms always have six protons. But some carbon atoms have six neutrons, while others have seven or even eight neutrons. A carbon atom that has six neutrons would have an atomic mass of 12; a carbon atom with eight neutrons would have an atomic mass of 14. Two atoms with the same atomic number but a different atomic mass are called isotopes. For example, most atoms of carbon have an atomic mass of 12. There are some atoms of carbon with an atomic mass of 13 (carbon 13), and some with 14 (carbon 14). These are isotopes of the carbon atom. Carbon 12 is the most abundant or prevalent isotopic form of carbon. Because carbon 12 is the most abundant form and taking into account all of carbon's different isotopes, the average value for the atomic mass for carbon is near 12: it's listed at 12.01.
Scientists have learned that isotopes have many useful properties. Some isotopes can be used to trace chemical reactions. If a special isotope is placed in fertilizer, for instance, that same isotope will later show up in the tissues of plants, then animals and decomposers as the nutrient moves through the food web. Some isotopes break down into others at a constant rate, allowing us to determine the age of fossils or artifacts. When an isotope is radioactive, this means that it emits some energy (radiation) when it breaks down to a different isotope. Radioactive isotopes can be used to visualize certain tissues for medical diagnosis. For example, radioactive iodine tends to concentrate in the thyroid gland. It emits radiation as it breaks down to another isotope of iodine. Special cameras can pick up this radiation and use it to create an image of the thyroid gland. This can help determine if the patient’s thyroid is normally active or to detect thyroid cancer.
The electron is the reactive part of the atom. The number and location of electrons determines the interactions between atoms. As you recall, negatively charged electrons are pulled toward the positively charged nucleus. But electrons are kept in orbit by kinetic energy. Electrons that are close to the nucleus are less energetic than those electrons farther from the nucleus.
Electrons reside in the space outside the nucleus in regions called shells. There are specific rules for filling up the shells of the atom. The first shell can hold two electrons; the second and third shells can hold up to eight electrons.
Electrons fill up lower shells before moving to the next higher shell. The electrons that occupy the outermost shell are called valence electrons and are the electrons that are involved in chemical bonding. The chemical properties of an element depend mostly on the number of valence electrons.
A key characteristic of atoms is that they are most stable when they have full outer electron shells. Helium (He) has a full outer shell with two electrons; directly below it in the periodic table is Neon. Neon (Ne) has an inner shell with two electrons and a full second shell with eight electrons. In fact, all of the elements in that column of the periodic table have full outer shells, are stable, and do not react with other elements.
Opposite charges cancel each other out when they are close together. So what would you predict about the overall charge of an atom with an equal number of protons and electrons? That’s right — it is neutral. There is no charge; the atom will not be attracted to or repelled by charged objects. All atoms, in their elemental state, have equal numbers of protons and electrons, and for this reason they have no net charge.
As you recall, an atom is most stable when it has a full outer shell. Some atoms achieve this endpoint by taking on extra electrons, or giving them up. When this happens, the atom becomes an ion and it takes on an overall charge.
Some elements readily take on one or more extra electrons. They become anions that carry a negative charge, because they have more electrons than protons. On the periodic table, elements to the right (with the exception of the last column) are likely to take on extra electrons. For example, chlorine (Cl) has seven electrons in its valence shell. It is likely to take on one extra electron to fill its outer shell, becoming an anion with a charge of -1. The same can be said for other elements in the same column with chlorine. Conversely, elements on the left side of the periodic table tend to give up one or more electrons and become cations with a positive net charge. For example, sodium (Na) has one electron in its valence shell. It readily gives up this electron, becoming a cation with a charge of +1. The same can be said for lithium (Li) and other elements in this column.
In the next section, we will explore how these electron transfers occur, and will look at some of the other ways that atoms interact to achieve stable, full outer electron shells.
Scientists have learned to use isotopes in many different ways. For example, stable isotopes can be used to trace chemical reactions through a food web. For example, if a stable isotope is used in fertilizer, that same isotope will later show up in the tissues of plants, then in animals, and finally in decomposers as the nutrient moves through the food web. Other isotopes are unstable and they tend to decay (or break down into other isotopes) at a constant rate. Scientists have learned how to use unstable isotopes to determine the age of fossils or artifacts. Finally, some isotopes are radioactive. Radioactive isotopes emit energy in the form of radiation when they decay into different isotopes.
Radioactive isotopes can be used to visualize certain tissues for medical diagnosis. One example of this is radioactive iodine, which tends to concentrate in the thyroid gland. It emits radiation as it breaks down to another isotope of iodine. Special cameras can pick up this radiation and use it to create an image of the thyroid gland to determine if the patient's thyroid is functioning normally. If there is abnormal activity in the thyroid gland, this can indicate cancer.
Another example of radioactive isotopes is positron-emission tomography (PET). In PET, a patient ingests sugar that is marked with a radioactive isotope. Cells that are using a lot of sugar (or need a lot of energy) will absorb more of the radioactive isotope. The patient is then moved into a PET scanner, where the radioactive isotopes are visualized and the machine translates the data into an image. This can help in a variety of applications, from cancer and Alzheimer’s disease diagnosis to addiction research (see image).
Atoms are the smallest units of an element that retain all the properties of the element. Atoms combine to form a larger and more complex entity called a molecule. Molecules are composed of two or more atoms held together by covalent bonds.
Chemical bonds are attractions between atoms that hold atoms and molecules together. There are three types of chemical bonds that are important in biology: covalent bonds, ionic bonds, and hydrogen bonds. Covalent bonds are very stable, while ionic and hydrogen bonds are less stable. The relatively weaker ionic bonds and hydrogen bonds allow reversible interactions between different molecules in biological systems.
Joining atoms together builds molecules that are more complex and larger than the individual atoms that compose them. The table below illustrates some similarities and differences among ionic, covalent and hydrogen bonds.
Covalent bond | Ionic bond | Hydrogen bond |
strongest | moderate strength | weak |
between atoms | between ions | between different molecules or parts of molecules |
sharing of electron pair, creating a strong bond between atoms | moderately strong electrostatic attraction between oppositely charged ions | weak electrostatic attraction between areas of molecules with opposite partial charges |
The most stable situation for an atom is to have its outer shell completely filled with electrons. It is not easy to explain why this is true, but it is a rule of thumb that predicts how atoms will react with each other. Recall from the discussion on electrons, in the Atoms module, that the first electron shell is full with two electrons, and the second and third shells are full with eight electrons. Atoms tend to bond to other atoms in such a way that both atoms have filled outer shells as a result of the interaction.
Ionic bonds are the interactions between ions of opposite charges. Atoms form ions when they either take up or release electrons in order to fill their outer shell. Elements in the outer columns of the periodic table often react in this manner; elements like sodium tend to lose electrons and become cations (positively charged ions), whereas elements like chlorine tend to gain electrons and become anions (negatively charged ions).
In fact, the reaction between sodium and chlorine is a great example of ionic bonding, and produces a compound you have in your kitchen — table salt. Chlorine (Cl) has an atomic number of 17, so it has 7 electrons in its outermost shell. Chlorine needs one more electron to have a full outer shell.
Sodium (Na) has an atomic number of 11, with one electron in its outermost shell. Sodium needs to get rid of an electron, and then it will have a full outer shell.
By losing an electron, sodium becomes a cation with a positive charge (+1). By gaining an electron, chlorine becomes an anion with a negative charge (-1). Now each ion has a net charge and the two charges are opposite. The ions attract one another. The electrostatic interaction between the sodium ion and the chlorine ion is an ionic bond .
Instead of transferring their electrons completely, atoms may remain in very close contact and share electrons so that their outer shells are filled. In essence, a shared electron is counted “twice” and participates in a larger shell that joins the two atoms. A single pair of shared electrons makes a single covalent bond. Atoms can share two pairs of electrons (in a double bond), or even three pairs of electrons (in a triple bond).
This sharing of electrons is called a covalent bond.
Carbon, for example, has an atomic number of 6 (see the figure below). The outer shell of carbon has 4 electrons. Carbon can share an electron with four other atoms. Hydrogen has an atomic number of 1. It has a single electron in its outermost shell, and can share this electron with one other atom. A carbon atom can form a covalent bond with four hydrogen atoms to form a molecule called methane, CH4 . In a methane molecule, carbon effectively has a “full” second shell (8 electrons) and each hydrogen has a “full” second shell (two electrons).
Nitrogen, another example, has an atomic number of 7 (see the figure above). The outer shell of carbon has 5 electrons. Nitrogen can share an electron with three other atoms. Hydrogen has an atomic number of 1. It has a single electron in its outermost shell, and can share this electron with one other atom. A nitrogen atom can form a covalent bond with three hydrogen atoms to form a molecule called ammonia, NH3 . In an ammonia molecule, nitrogen effectively has a “full” second shell (8 electrons) and each hydrogen has a “full” second shell (two electrons).
When two atoms are joined by a covalent bond, the new structure that forms is called a molecule. This is in contrast to the structure formed when two atoms are joined by an ionic bond, which is called an ionic compound. When drawing a molecule on paper, covalent bonds are often drawn as lines between atoms. A single covalent bond is drawn as one line; a double covalent bond is drawn as 2 lines. Each atom is represented by its element symbol. Thus a structural formula of methane (CH4) would be drawn like this:
Notice that the carbon in methane has formed four covalent bonds with four hydrogen atoms. Because carbon has four electrons in its outer shell, it can always form four covalent bonds with other elements.
The molecular mass or weight of a molecule is the sum of the individual atomic weights. As with atomic weights, it is convenient to use the weight of a mole of molecules. For example, the molecular mass of a single methane molecule is 16 Da (12 from the carbon, 4 from the hydrogens). The weight of a mole of methane molecules is 16 grams.
Structural Formulas: There are many different ways to draw molecules. For example, look at the glucose molecule shown on the right in a structural formula. Notice that the structure consists of a carbon backbone running down the middle of the molecule. Each carbon has formed a covalent bond with four other atoms. (The first carbon actually formed two bonds with the oxygen atom; this is called a double bond. This carbon still formed a total of four bonds, even though two of the bonds are with the same oxygen atom.) Skeletal Formulas: Chains of carbon atoms are very common in carbon-based molecules. Sometimes such molecules are drawn using skeletal formulas, also known as shorthand formulas. In these diagrams, the carbon atom, typically indicated with the letter “C,” is not shown. Instead, each C atom is drawn simply as a corner on the diagram. |
![]() |
Below are skeletal diagrams of glucose. In these examples, the carbons are located at each angle in the diagram.
You may have another question about the skeletal diagrams: what happened to the hydrogen atoms attached to the carbon atoms? It looks like some of the carbons only have two or three covalent bonds! The skeletal diagrams are often drawn to be simple, not to show every atom. It is understood that carbon must have four bonds, and in any carbon-based molecule there are often one, two, or even three hydrogen atoms bonded to each carbon. To clean up the diagram, the H atoms are not drawn. It is assumed you will understand that any “missing” bond is actually provided by a hydrogen atom.
These conventions simplify diagrams of molecular structures. The cholesterol molecule, shown below, looks much cleaner than if we had to draw each carbon and hydrogen atom. You may also notice that this image has different-looking bonds. These solid and striped triangle-shaped bonds describe the three-dimensional shape of the molecule.
Recall that in a covalent bond, the electrons are shared between two atoms. However, the two atoms don’t necessarily share the electrons equally. Some atoms are more likely to draw the shared electrons closer to themselves. Atoms have a high electronegativity if they tend to draw the electrons towards them. Electronegativity increases as one moves from the left to the right across the periodic table. Hydrogen is moderately electronegative, with a value of 2.1, carbon is somewhat more electronegative, with a value of 2.5, and fluorine is the most electronegative atom, with a value of 4.0.
The degree of unequal electron sharing depends on the difference in electronegativity of the atoms involved in the bond. For example, in carbon-carbon bonds, both atoms have the same electronegativity, so the electrons are equally shared between the two carbons. In contrast, a carbon-oxygen bond involves two atoms that have different electronegativities. The less electronegative atom (carbon) will donate part of its electrons to the more electronegative atom (oxygen), resulting in a partial positive charge on the carbon (indicated by δ+) and a partial negative charge on the oxygen (indicated by δ-). The size of the partial charges is proportional to the difference between the electronegativities of the two bonded atoms. Bonds in which electrons are unequally shared between the atoms are called polar covalent bonds.
Throughout this course, you will be studying many different molecules. Many of the important molecules of life, like DNA, proteins, and even ordinary water, share a key characteristic: they all form hydrogen bonds.
Hydrogen bonds are not like the covalent bonds you just learned about. They do not join atoms into molecules. Instead, they are the attraction of an electronegative atom to a hydrogen that is covalently bonded to another electronegative atom. This involves the attraction of the hydrogen with a partial positive charge to the electronegative atom with a partial negative charge. Only hydrogen covalently bonded to an electronegative atom can participate in hydrogen bonding.
Hydrogen bonding has a significant effect on the properties of molecules. For example, the structure of the DNA molecule is in part held together by hydrogen bonds. Hydrogen bonds also create coils and other structures within the complex protein molecules that are essential to life’s diversity. Finally, water behaves the way it does because of the hydrogen bonds that attract water molecules to each other.
The water molecule is a very important example of a molecule that can form hydrogen bonds. It is the hydrogen bonding of the water molecules that gives water many of its unique and life-sustaining properties. One water molecule consists of two hydrogen atoms covalently bonded to a oxygen atom.
Recall that in a covalent bond, the electrons are shared between two atoms. However, the two atoms don’t necessarily share the electrons equally. Some atoms are more likely to draw the shared electrons closer to themselves. These atoms have a high electronegativity. Both oxygen and nitrogen are molecules with a high electronegativity.
In a water molecule, oxygen is more electronegative than hydrogen. This means that oxygen has a greater affinity or attraction for the shared electron pair than hydrogen does. Because of this, the electrons tend to spend more time close to the oxygen atom and they spend less time close to the hydrogen atom. As a result, the oxygen becomes slightly negative and each hydrogen atom develops a slightly positive charge. Whenever two atoms with different electronegativities form a covalent bond, we say that the bond between the atoms is polar, with each atom carrying a partial charge. Oxygen and nitrogen are the two highly electronegative elements that are often bonded to hydrogen in biological molecules. O-H and N-H bonds are strongly polar, and this sets the stage for hydrogen bonding.
Hydrogen bonding is the attraction of an electronegative atom for a hydrogen that is covalently bonded to another electronegative atom. This involves the attraction of a hydrogen with a partial positive charge to an atom with a partial negative charge. However, only hydrogen covalently bonded to an electronegative atom can participate in hydrogen bonding.
In water, the partially positive hydrogen atoms are attracted to the partially negative oxygen atoms of neighboring water molecules. Hydrogen bonds are responsible for many of the unique properties of water. We’ll explore this in more detail in the next module. Watch the following animation that illustrates hydrogen bonding.
Biologists have different ways of representing chemical structures. Each type conveys different information.
Have you ever wondered why scientists spend time looking for water on other planets? It is because, at least here on Earth, water is essential to life. Evidence of water on Mars, and some of the moons of Saturn and Jupiter, increases the odds that life may exist there. Here on Earth, water is one of the most abundant molecules. The water content of our bodies and our cells ranges from 70 to 95 percent. All of life’s chemical reactions take place in watery fluid. Without water, life as we know it simply would not exist.
There are four properties of water that make it such a unique and important molecule:
These properties of water are due to the hydrogen bonds formed between water molecules. Hydrogen bonds are discussed in the Chemical Bonds module.
Solutions are homogeneous mixtures. This means that you can sample any part of the solution and the composition of the solution will be the same.
Solutions have two components:
If you add some salt to water, you are making a solution. Water is the solvent and the salt is the solute. Solutions can have more than one solute. If you add sugar to your salt solution, both sugar and salt would be the solutes of the solution. Solutions in which water is the solvent are called aqueous solutions. The inside of our cells and our body fluids are examples of aqueous solutions. When a solute is added to a solvent and a solution is formed; the solute is described as “dissolving” in the solvent.
Substances that will dissolve in water are hydrophilic or water-loving. What kinds of molecules are hydrophilic? Ionic substances like table salt (NaCl) are hydrophilic. They split into positive and negative ions and dissolve in water. Polar water molecules surround the charged particles, breaking them away and pulling them into the fluid.
Polar molecules also are hydrophilic. Polar water molecules readily surround and dissolve polar molecules or molecules with polar functional groups. Examples include sugars and alcohols, which have hydroxyl groups. Molecules that do not dissolve in water are hydrophobic (from the Greek word meaning water-fearing or water-hating). Nonpolar molecules are hydrophobic. Examples include hydrocarbons and fatty acids with their abundant nonpolar C-H bonds. Have you ever tried to mix water and oil? Oils and fats are nonpolar and will not form a solution with water.
Next we will examine hydrophobic molecules more closely. Substances with hydrophobic molecules will not dissolve in water and instead will tend to separate from water.
Some molecules are hydrophobic, which literally means fear (phobia) of water (hydro). These molecules do not like to dissolve in water. When hydrophobic molecules are mixed with water, they will tend to separate into distinct phases (or layers), one containing water and the other containing the hydrophobic molecules. This makes it very difficult, for example, to “wash” oil off of waterfowl after an oil spill.
Things composed of hydrophobic molecules generally include oils, fats, and greasy substances. Materials containing hydrophobic substances are often used for removal of oil from water and management of oil spills. Other molecules are hydrophilic, which means love (philic) of water, and readily dissolve in water.
Now that you know some common hydrophobic and hydrophilic molecules, let’s look at their structures to understand why they interact with water in different ways.
The additional atoms that are found only on glucose are electronegative, meaning that they will withdraw some of the electrons from the atoms that form bonds with them. The electronegative atom will have a slight negative charge, and the atom that it is bonded to it will have a slight positive charge. The unequal distribution of electrons makes the bond a polar bond.
The interaction between a polar bond on water and polar bonds on hydrophilic molecules is given a special name — the hydrogen bond, because it involves a hydrogen that forms a bridge between the two molecules. Hydrogen bonds are fairly stable, so quite a bit of energy is released when they are formed. The release of energy helps hydrophilic molecules like glucose dissolve in water. Since hydrophilic molecules contain polar bonds, they are often referred to as polar molecules.
Hydrophobic molecules lack the electronegative atoms that are required to generate a polar bond. Thus, hydrophobic molecules are also referred to as nonpolar.
The fact that nonpolar molecules cannot form hydrogen bonds in one reason why they have low solubility in water; however, there is another far more important force that drives nonpolar molecules from water, called the hydrophobic effect.
When a non-polar molecule dissolves in water it becomes completely surrounded by water molecules. The water molecules cannot form hydrogen bond with the non-polar molecule, however they do form hydrogen bonds with the other water molecules that surround the non-polar molecule, forming a layer of hydrogen bonded water molecules that cover, or form a cage, around the dissolved non-polar molecule. A representation of these water molecules organized around dissolved butane is shown in the following Jmol:
The dissolved butane is highlighted in yellow. It is surrounded by a cage of water molecules, forming hydrogen bonds (purple) with each other. Place your cursor over the structure left-click, hold, and move your pointer to rotate the structure. Right-click to display a pop-up menu and select zoom-in or zoom-out to examine the chemical bonds. (Source: pdb file modified from Udachin et al., J. Chem. Phys, 134, 121104 (2011).
The formation of the water cage is unfavorable, because it requires water molecules to become ordered. Ordering anything requires energy, whether it is a cage of water molecules surrounding a nonpolar molecule, or your messy dorm room. The energy cost of ordering water molecules around nonpolar compounds is so high that they are forced out of the water, leading to a separation of phases; oil and water don’t mix because of the hydrophobic effect. Most biological molecules contain both polar and nonpolar regions. Their behavior in water depends on the relative number and type of polar groups versus nonpolar groups. For example, consider this series of simple alcohols.
Molecules that contain a very polar part and a very nonpolar part are called amphipathic. The polar part interacts with water, forming strong interactions, while the nonpolar part doesn’t interact with water because of the hydrophobic effect. In order to accomplish this, amphipathic molecules usually undergo spontaneous self-assembly into structures that expose the polar part to water and keep the nonpolar section away from water. One example of amphipathic molecules are fatty acids.
The carboxylate group (COOH) on the left interacts strongly with water, while the remaining part of the molecule is very hydrophobic and is forced away from water. Fatty acids form structures in solution called micelles. These spherical structures have the hydrophilic part exposed to water on the surface of the sphere and bury the hydrophobic part in the center of the sphere.
You might be wondering why triglycerides are not amphipathic, given that they are similar to fatty acids, with a polar group of atoms and a nonpolar group. The reason is that the polar region of triglycerides interacts weakly with water because there is no free -OH group.
Have you ever filled up a glass of water to the very top and then slowly added a few more drops? Before it overflows, the water forms a dome-like shape above the rim of the glass. This water can stay above the edges of the glass because of the property of cohesion. In cohesion, water molecules are attracted to each other (because of hydrogen bonding), keeping the molecules together. Cohesion allows surface tension, the capacity of a liquid’s surface to resist being ruptured when placed under tension or stress. When water is sprinkled onto a solid surface, surface tension causes the water to remain in compact droplets instead of spreading out into a thin film.
When you drop a small scrap of paper onto a droplet of water, the paper floats on top of the water droplet, even though the object is denser (heavier) than the water. This occurs because of the surface tension that is created by the water molecules. Cohesion and surface tension keep the water molecules intact and the item floating on the top. Many insects are specially adapted to exploit surface tension, which allows them to literally “walk on water.”
These cohesive forces are related to water’s property of adhesion, or the attraction between water molecules and other molecules. This is observed when water “climbs” up a straw placed in a glass of water. You will notice that the water appears to be higher on the sides of the straw than in the middle. This is because the water molecules are attracted to the straw and therefore adhere to it.
Cohesive and adhesive forces are important for sustaining life. For example, because of these forces, water can flow up tubes within plants. These tubes connect the roots of plants to their leaves, and carry water to plant tissues. Adhesion helps water cling to the walls of the tubes, and cohesion keeps the water in an intact column. Without these properties of water, tall plants would be unable to receive the water and nutrients they require, and they would die.
Imagine you have placed a metal pan containing water on the stove. You transfer heat energy to the pan. If you touch the metal pan after a few minutes, it is warm or even hot to the touch, but the water does not yet feel warm. The metal’s temperature changes much more quickly than that of the water, even though both substances are receiving heat from the same source. Temperature is a measure of the vibrational energy of molecules within a substance; it reflects how fast the molecules are jiggling around. If you add heat to metal, the molecules respond quickly with an increase in their “jiggling” and the temperature increases rapidly. If you add the same amount of heat to water, the molecules respond much less quickly. Why is water’s temperature so resistant to change?
When water is heated, a large part of the heat energy goes into breaking the hydrogen bonds between the water molecules. Only a small part actually increases the kinetic energy of the water molecules. Imagine a rack of billiard (pool) balls. When the cue ball strikes them, energy is transferred and the balls “break,” rolling in every direction. What would happen if the balls were connected by rubber bands? They would not move nearly as much. Similarly, the hydrogen bonds among water molecules make it hard to increase their “jiggling.” As heat is added, hydrogen bonds are broken and water warms up, but due to the hydrogen bonding, the water’s temperature changes only slowly. When water cools down, it releases a great deal of heat as hydrogen bonds are re-established. This effect is particularly strong when water freezes; water molecules are joined in a regular crystal structure by stable hydrogen bonds. As these bonds form, heat is released to the surrounding environment. You may have noticed that snowy days are often not quite as cold as the dry clear days of winter. This is yet another example of the way that water moderates temperature.
When water freezes, the water molecules become arranged into a regular crystal. Each water molecule is bound by four hydrogen bonds to its neighbors. In this arrangement, the water molecules are further apart than they are in liquid water. Thus, the density of ice is less than the density of liquid water. Ice, the solid state of water, can float in the liquid state. This is important for lakes and ponds that contain fish and plants. When the temperature goes below freezing, ice is formed. The ice is less dense than the liquid, so it floats on the top of the lake or pond. The surface ice insulates the pond, helping it to retain heat, and keeps the pond from freezing solid.
Water-Ice |
|
The three dimensional structure of ice is shown above. Place your cursor over the structure, left-click, hold, and move your pointer to rotate the molecule. Click on “Highlight Void” to show the empty volume generated due to the hydrogen bonds in ice. Click on the "Spin on/off" checkbox to automatically spin or stop spinning the molecule.
Solid water is less dense than liquid water because in ice, the regular pattern of hydrogen bonds pushes the molecules further apart.
We use detergents to remove oily dirt from things. We use dish detergent to remove the grease on dishes, laundry detergent for oils on our clothes, and shampoo for oils in our hair. Detergents make the hydrophobic oils soluble so that the dirt can be rinsed away. How does this happen?
All detergents are amphipathic molecules; they have a polar, water-loving (hydrophilic) part and a nonpolar, water-hating (hydrophobic) part, which allows them to dissolve in both water and oils. The chemical structure of a common detergent, sodium dodecyl sulfate (SDS), is shown below. The hydrophilic part of DS interacts favorably with water by hydrogen bonding and electrostatic (charge-charge) interactions between the negative charge on the DS and the partial positive charges on the polar water molecules. The hydrophobic part has no polar atoms to form favorable interactions with the water.
When placed in water, detergents spontaneously form organized spherical structures called micelles. In a micelle, the hydrophilic part of the detergent is on the surface of the sphere and interacts with the water. The center of the sphere contains the hydrophobic part of the detergent, hidden from the water. A cross section image and a three-dimensional structure of a micelle are shown below. When a micelle comes in contact with oily (hydrophobic) compounds, the oils dissolve into the interior of the micelle and can be washed away with the micelle.
This course has looked closely at the water molecule because it is vital to life. Water covers 70 percent of the Earth’s surface and the cell, the fundamental unit of life, consists of 70 to 95 percent water. In this module, we will examine one very small (and simple) substance that, when dissolved in water, can have an enormous impact on life: the hydrogen ion (H+). Hydrogen ions are hydrogen atoms that have had their electrons removed. What remains is only a single proton. The simplicity of this ion, however, is quite misleading. In this module, we will investigate the dynamics of H+ and learn more about the effects H+ can have on solutions.
The story of acids and bases begins with water (H2O). Water in the liquid state is highly dynamic. Not only are hydrogen bonds constantly forming and breaking between water molecules, but individual water molecules are breaking apart and then reforming again to make water. When they break apart, a hydrogen ion (H+ ) is transferred to another water molecule to make a hydronium ion (H3O+), leaving a hydroxide ion (OH-) from the original water molecule. This dynamic reaction can be represented by this chemical equation: H2O + H2O [double arrow] H3O+ + OH-. To simply things we will represent the hydronium ion as a hydrogen ion (H3O+=H+) in the following discussion.
To simply things we will represent the hydronium ion as a hydrogen ion (H3O+ = H+) in the following discussion.
Notice that the two arrows point in opposite directions. This simply means that the equation can move in either direction. In other words, hydrogen ions and hydroxide ions can combine to form water, or water can break down into hydrogen ions and hydroxide ions. These two reactions are in equilibrium, so at any instant in a sample of pure water there are equal concentrations of H+ and OH-.
Pure water is said to be “neutral.” This means it is neither acidic nor basic because it has equal concentrations of H+ and OH-. Some substances can dissolve in water and they will not disturb the balance between H+ and OH-. However, other compounds disturb this balance. Some compounds, when added to water, cause an increase of hydrogen ions (H+) in the solution. These compounds are called acids. Other compounds, when added to water, cause a decrease in hydrogen ions (H+) in the solution. These compounds are called bases. Some bases accept H+ directly, but others release hydroxide ions (OH-) that then combine with H+ to produce water. The end result is the same. When the base is dissolved in water, the concentration of H+ ions goes down. A compound that does not result in a net gain or loss of hydrogen ions (H+) when in aqueous solution is said to be neutral; it does not alter the acidity of the water.
Acids | Bases |
Readily give hydrogen ions (H+) when dissolved in water | Readily take hydrogen ions (H+) when dissolved in water; may donate hydroxide ions (OH-) |
Acidic foods often taste sour | Basic foods often taste bitter |
Many foods and beverages are acidic | Bases often make your skin feel slimy or slippery because they release soap-like molecules from your lipids. |
Strong acids can cause burns and are caustic | Strong bases can cause burns and are caustic |
Acids commonly induce damage to living organisms by denaturing proteins: changing their shape so they do not function normally | Bases commonly induce damage to living organisms by breaking down fats and denaturing proteins |
Understanding acids and bases is important to biology because living organisms must maintain homeostasis (balance) and cannot tolerate drastic changes in the acidity of their body fluids or cytoplasm. Cellular functions are disrupted if these fluids become too acidic or basic.
The acidity of a solution is measured on the pH scale. pH is a measurement of the concentration of hydrogen ions (H+) in a solution. You can think of pH as “parts Hydrogen ion,” but remember that the pH scale is “backwards.” The pH scale ranges from 0 to 14, with zero being the most acidic (highest concentration of H+) and 14 being the most basic. Neutral solutions (like pure water) have a pH of 7. Solutions with pH measurements below 7 are acidic, while solutions with pH measurement above 7 are basic. A solution at pH 7 is neutral because at that pH the number of hydrogen ions (H+) equals the number of hydroxide ions (OH-), just as they are balanced in pure water. Hydrogen ion concentration can vary over many orders of magnitude. To represent very large differences in concentration levels, a logarithmic mathematical function is used to denote the pH scale. |
Acids and Bases in Solution![]() |
The pH scale is the -log of the hydrogen ion concentration; the minus sign changes the negative numbers that would be obtained from log[H+] to positive ones:
pH = -log[H+]. |
The log conversion reduces a tenfold change in hydrogen ion concentration to a one-unit change in pH. Thus, starting at pH 7, a solution at pH 6 would have 10 times more H+ than a neutral solution. On the other hand, a solution at pH 8 would have 10 times fewer H+ than the solution with pH 7. If you skip a few units, the differences between solutions become very great indeed. Lemon juice (pH 2) has one million times more H+ ions per unit volume than seawater, which has a pH of about 8.
One of the ways that scientists can measure the pH of a solution is to use indicator paper. This is special paper that is infused with dyes that change color when placed in an acid or a base. The most simple indicator paper is called litmus paper. Litmus paper comes in two colors — red and blue.
In the presence of a neutral solution, red litmus paper remains red, and blue litmus paper remains blue.
If an acid gives a hydrogen ion (H+) in solution and a base accepts a hydrogen ion by giving a hydroxide ion (OH-), you may be wondering what will happen when an acid and a base are combined. Let’s consider the combination of equal amounts of hydrochloric acid (HCl), a strong acid, and sodium hydroxide (NaOH), a strong base. When the hydrogen ion (H+) is released from hydrochloric acid, and the hydroxide ion (OH-) is released from sodium hydroxide, the two ions combine to form water (HOH). Water, as you know, has a neutral pH, so the process of combining an acid and a base is called neutralization. Remember, in a neutral solution (like water) the number of hydrogen ions equals the number of hydroxide ions. The remaining sodium ion (Na+) from the sodium hydroxide and the remaining chloride ion (Cl-) from the hydrochloric acid also stay in solution as dissolved table salt (NaCl). This process can be written as it is below:
HCl + NaOH ➨ HOH + NaCl |
Notice that the same four atoms (H, Cl, Na, and OH) are found on each side of the equation. Using this example we can generalize the process of neutralization as follows:
acid + base ➨ water + salt (in solution) |
The process of neutralization also gives off heat, indicating that chemical potential energy is being released during the course of the reaction. Neutralization has many useful applications. One such application in the laboratory is to neutralize strong acids or bases so that they can safely be disposed of. As you will learn later in this course, the addition of acids and bases to the environment leads to ecosystem damage, so before a strong acid or base can be disposed of, it must be neutralized (to make water and salt) to prevent injuries or environmental damage. In your everyday life you may be familiar with acid and base neutralization through your lawn care and gardening activities. Grass grows best in a very mildly acidic soil (pH 6.5 to 7). So if your soil is too acidic (less than pH 6.5), you would need to add a compound like limestone (a basic mineral) to your soil to neutralize the excess acidity and correct the pH level. Likewise, if your soil is too basic (pH above 7), you would need to add sulfur, which interacts with water in the soil to form sulfuric acid, to help neutralize the excess base in the soil.
Another common application of neutralization is the standard volcano science fair project, where a mixture of vinegar (an acid) and baking soda (a base) are combined to make the volcano "erupt." In this case, the reaction releases some carbon dioxide gas, generating fizz and fun.
If you ever come into contact with a strong acid or base and receive a chemical burn, DO NOT attempt your own science experiment to try and neutralize the compound. The damage to your skin is already occurring and can be exacerbated by adding more chemicals; a vigorous neutralization reaction could cause further harm through the release of heat, causing a physical burn on top of the chemical burn already suffered. Your safest option is to wash the affected area under a heavy stream of running water, and to seek medical attention for your burn. The goal is to remove the agent causing the burn to prevent further damage.
Living organisms are quite sensitive to even small changes in pH. The chemical reactions of life are tuned to a specific and often narrow pH range; outside this range, the reactions will not proceed normally. Therefore, organisms have mechanisms that work to maintain a constant pH level. This is true both at the cellular level and at the level of the whole organism. Maintaining a stable pH in a living organism is an important part of homeostasis. Homeostasis refers to the ability of an organism (or cell) to maintain stable internal conditions despite constantly changing environmental conditions.
One way in which changes in pH are moderated is through the use of buffers. Buffers are aqueous solutions that can resist changes in pH. When an acid or a base is added to a solution containing a buffer, the pH of the solution will only exhibit a very minor change. Blood is a buffered solution. The liquid portion of blood contains carbonic acid (H2CO3). It dissociates (breaks apart) to produce bicarbonate ions (HCO3-) and hydrogen ions (H+). This reaction is easily reversible. If acid enters the blood (H+ concentration goes up), most of the excess hydrogen ions will bond with bicarbonate, restoring carbonic acid. If a base enters the blood, the H+ concentration will initially go down. But then carbonic acid will dissociate to replace the “missing” hydrogen ions. Thus the easily reversible reaction works to oppose any pH change in the blood. A buffered solution cannot absorb an unlimited amount of acid or base. As acid or base is added, any buffered solution will eventually reach a limit called the “buffering capacity.” At this point, no additional hydrogen ions can be absorbed or produced, and the buffer stops working. The solution’s pH will change rapidly if further acid or base is added.
While maintaining a pH close to neutral is usually the goal of homeostasis, there are certain conditions and parts of organisms that are maintained at a lower (more acidic) pH. For instance, the process of digestion is aided by the presence of stomach acid (pH 2). Not only does stomach acid help to break down the compounds in our food, it also helps to prevent disease by killing many of the bacteria and viruses found in food and water. Lysosomes, digestive organelles found in eukaryotic cells, have an internal pH of about 5. Again this acidic environment is used to break down cellular food, waste, and even invading bacteria. Likewise, the surface of our skin is acidic (pH of 4.5 to 6), which helps to support the growth of our normal bacterial flora, as well as to help prevent the growth of pathogenic bacteria.
When the normal pH levels of the body are disrupted, there are detriments to the health of the organism. For instance, the female vaginal tract is moderately acidic, with a pH between 3.8 and 4.5. When the pH increases in this area of the body (becomes more basic), women are more prone to bacterial infections of the vagina. Conversely, if the pH decreases (becomes more acidic), women are more prone to yeast infections.
A higher-stakes example of loss of pH regulation occurs in diabetic ketoacidosis. In people with diabetes, cells cannot take up glucose (sugar) from the blood efficiently. As a result, cells switch from using glucose (sugar) as a fuel source and use fatty acids (fats) instead. When fatty acids are broken down, ketone acids are produced as a by-product. In healthy individuals, a small amount of ketone acids are released by normal metabolic processes; for example, if you are losing weight, fat breakdown will add some ketone acids to your bloodstream. This poses no problem because your blood is buffered, and your liver and kidneys are able to remove the excess acids before they cause problems. However, in people with diabetes, chronic release of ketone acids can overwhelm these systems. Ketone acids build up in the blood and urine, leading to a condition called ketoacidosis. The presence of excess ketone acids lowers the body’s pH, leading to several health problems as detailed on the figure below. If the pH balance is not restored, diabetics in ketoacidosis can face serious harm and even death. Treatment consists of restoring proper insulin levels in the body (so that carbohydrates can be absorbed by cells and used as a fuel source) and ensuring proper hydration (to help the kidneys flush out the excess ketones).
While it is clear that the regulation of pH in living organisms is crucial to their survival, the pH in the environment is just as important to maintaining life on this planet. Different environments have different pH levels, and organisms that live in those locations have adapted to those conditions. When environmental pH levels are disrupted, the health and vitality of the ecosystem becomes threatened.
As you have learned, seawater is slightly basic at pH 8. The oceans and the organisms that live in them play a critical role in maintaining the balance of oxygen and carbon dioxide on this planet. You will learn more about this in the Photosynthesis / Cellular Respiration module in the Metabolism unit. With the increased burning of fossil fuels in the modern industrial society, more man-made carbon dioxide (CO2) is being released into the environment than ever before. Much of this carbon dioxide is absorbed by the oceans. When carbon dioxide is absorbed into water, carbonic acid is produced. This then dissociates to produce bicarbonate ions and hydrogen ions, and the pH of the ocean is lowered. Acidification of the ocean threatens the health of the organisms that call the ocean home. Many marine life forms, including reef-building corals, are harmed or killed when ocean water is acidified. As carbon dioxide levels in the atmosphere continue to climb, there is great concern about the worsening impact of ocean acidification.
Another environmental problem involving pH balance is acid rain. When nitrogen oxides and sulfur dioxide, two common by-products of burning fossil fuels, combine with water in the Earth’s atmosphere, acid rain is produced. In some regions, acid rain reduces the pH of fresh water sources like streams, rivers, and lakes. It also may change soil chemistry and harm plants. Some regions have rocks and soils that contain buffers, and they are less susceptible to damage from acid rain. However, when soil buffering capacity is exhausted (buffers cannot accept any more hydrogen ions), pH may decline very quickly. Tighter regulations on coal-burning power plants have helped reduce the acid rain problem in recent decades, but it remains a threat to some forest and aquatic ecosystems and the organisms living within them. We will learn more about acid rain in the Ecology unit.
In the Introduction to Chemistry unit, you learned that atoms are the building blocks of all matter. You also learned that atoms can be linked together with chemical bonds, to form molecules that have their own unique properties. In this unit, you will take a closer look at a special group of molecules that form the backbone of living structures: the biological macromolecules. Biological macromolecules are special molecules that contain carbon atoms covalently bonded with hydrogen atoms.
There are four classes of biological macromolecules that we will study in this course: carbohydrates, lipids, proteins and nucleic acids. These macromolecules are probably already familiar to you, because they make up the nutrients you ingest every time you eat. In this way, you provide your cells with the building materials and energy necessary to sustain life. In the next unit, we will use these fundamental building blocks to form the cell, the first level of organization that shows all the characteristics of life.
In chemistry, the word “organic” has a special meaning. It is different from the grocery-store meaning, where “organic” refers to foods that are grown in a particular way and are free of pesticides. In fact, many of the chemical pesticides that an organic farmer would never use are actually organic molecules.
An organic molecule is any molecule that contains a carbon to hydrogen (C-H) covalent bond. They are often complex and many store a lot of chemical potential energy. Examples of organic molecules include glucose, methane, DNA, protein, and fat.
All the molecules discussed in the "Introduction to Chemistry" unit were inorganic molecules. Molecules like water, oxygen gas, carbon dioxide, and ionic salts like sodium chloride fall into this category. Most inorganic substances are relatively stable, simple, and store little chemical potential energy. What exactly are organic molecules? All organic molecules have several common properties that help distinguish them from inorganic molecules:
In this module, you will learn more about how to recognize organic molecules. You’ll also explore in more detail the important characteristics of organic molecules.
All organic molecules contain carbon atoms covalently bonded to hydrogen (C-H bonds). They can also contain oxygen, nitrogen, or sulfur. In the next activity, you will practice determining whether or not a chemical structure is an organic molecule.
Why are organic molecules so diverse and flexible in their structure? The characteristics of carbon are essential for building the complex and diverse structures needed for life. A single carbon atom is able to bond with up to four other atoms, allowing the formation of chains, branched chains, and even rings.
Carbon forms covalent bonds with many other elements; these strong bonds make organic molecules durable.
Organic molecules come in all shapes and sizes. One of the key features of organic molecules is their modularity: a limited set of monomers can be connected in different ways to form a vast array of polymers. It may help to consider an analogy: with a few dozen different kinds of Lego bricks, you can make an almost limitless diversity of structures. In organic molecules, monomers can be linked together by bonds between different atoms, different monomers can be used, and they can be strung together in many different arrangements. Therefore, organic molecules are virtually limitless in their diversity.
The four major classes of biological macromolecules are carbohydrates, proteins, nucleic acids, and lipids. These molecules carry out a diverse set of essential functions. Carbohydrates are essential for energy storage and cellular communication. Nucleic acids (such as DNA) are essential for information storage within a cell and passing on this information to the next generation. Lipids are essential for energy storage and maintaining a boundary between the living organism and its environment. Proteins are essential for carrying out most of the necessary functions of life.
Each major class of organic molecules represents a diverse assortment of polymers that are built from a handful of possible monomers. The monomers, smaller molecules bound together, are also called subunits. The table below lists the subunits for each class of organic molecule. Different carbohydrates, for example, are polymers made up of sugars (the monomers). Proteins are polymers made up of amino acids (the monomers). Nucleic acids are polymers made up of nucleotides (subunits or monomers). Even lipids are generated by combining separate fatty acid chemical components.
Macromolecules and Their Subunits | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|
|
There are many different types of proteins, carbohydrates, lipids, and nucleic acids. Each has a different function and purpose in living organisms. There are also many different types of subunit amino acids, sugars, fatty acids, and nucleotides. The combination of these different types of subunits is what determines the properties and function of the macromolecule. For example, the function of a particular protein is determined by the amino acids in it.
Perform the following activity to review the relationship between polymers (macromolecules) and monomers (smaller subunits). Indicate the macromolecule composed of each subunit.
In the remainder of this unit, you will be exploring the chemical structures of the four major types of biological macromolecules. You will see molecules represented in a variety of ways. The following activity will help you practice interpreting the different ways molecules can be illustrated.
Different representations of methane, ammonia, and albuterol are shown in the table below. Formula gives the chemical formula, Molecular indicates how the atoms are bonded together, 2D-Structural provides some information about the three-dimensional structure using a two-dimensional drawing. The solid wedges indicate that the atoms are above the plane of the page and the dashed wedges indicate that the atoms are below the page. 3D-Structural images are seen in the Jmol images at the right of each row. You can manipulate the 3D-structures by placing the cursor in the window and moving the mouse. You should understand the relationship between the 2D-structural representation and the 3D-structural representation. You should also take note of the standard color coding in 3D representations: usually carbon is gray, hydrogen is white, oxygen is red, nitrogen is blue, and sulfur is yellow.
|
|
|
|
|
|
Functional groups are parts of organic molecules that have specific properties or functions. Because organic molecules can contain more than one type of functional group, a particular molecule may have multiple properties. Identifying functional groups in molecules is an important skill, because once you identify a molecule’s functional groups, you can predict many aspects of its biological behavior.
Functional groups can be divided into three groups based on their physical properties. There are the nonpolar groups, the polar groups, and the charged groups.
Functional groups in this category are also referred to as hydrophobic (water-hating) groups. They only contain carbon and hydrogen, and lack electronegative atoms such as nitrogen, oxygen, and sulfur. Nonpolar functional groups are often found on amino acid side chains of proteins, and they also make up a major part of most lipid molecules. The hydrophobic nature of nonpolar functional groups often affects the shape of molecules containing these groups. For example, proteins will fold so that nonpolar groups are clustered together and are not in contact with water. Some lipids bury the nonpolar section of the molecule by forming lipid bilayers, which shape the boundaries of all cells.
The key nonpolar functional groups are:
Also, long chain alkanes, (CH2)n, found in lipids, are among the nonpolar functional groups.
Polar functional groups contain electronegative atoms like nitrogen (N), oxygen (O), and sulfur (S). The presence of electronegative atoms in a functional group results in an unequal distribution of charges on the atoms, causing the bonds to become polar. Because polar bonds interact favorably with water, compounds with polar functional groups also interact favorably with water, making them hydrophilic, or water-loving. In addition to interacting favorably with water, the polar atoms can also participate in chemical reactions and polar functional groups are usually responsible for the catalytic properties of enzymes.
The key polar functional groups are:
Some compounds contain both nonpolar and polar functional groups; molecules of this type are referred to as amphipathic molecules. Phospholipids and soaps are examples of amphipathic molecules.
Charged functional groups are acids, meaning that they form ions by the release of hydrogen ions (H+), which are also called protons. Depending on the specific functional group, these groups of molecules lose a hydrogen ion (or deprotonate) and become either charged or neutral as a result. Charged functional groups play key roles in biological systems. For example, many proteins bind to DNA (or RNA) by utilizing the electrostatic interaction between positive charges on the protein and negative charges on the DNA or RNA. Note that when these groups are in their uncharged state, they can be considered to be polar functional groups.
The key charged functional groups are:
A special case of the phosphate group is a phosphate diester, which links together nucleotides in DNA and RNA. The phosphate diester is always negatively charged at neutral pH.
If you need a reminder of chemical structures or of how to view molecules in 3D, review Viewing Chemical Structures in the Introduction to Chemistry unit.
Carbohydrates are organic molecules that consist of carbon, hydrogen, and oxygen atoms in a 1:2:1 ratio. The most abundant class of macromolecules found in living systems, carbohydrates are the primary source of energy in living systems. Large complex carbohydrates can be used to store energy. For example, a potato is full of starch, which is a complex carbohydrate that the potato plant uses to store energy. When you eat the complex carbohydrate, your body breaks the large molecules into their smaller subunits (sugars), which you will then use to fuel your own body’s energy needs.
Carbohydrates can also be used as structural building materials. Cellulose is an example of a complex carbohydrate used in plants for structural support. Within plant cell walls, cellulose molecules form chains that provide high tensile strength. In some plants, cellulose is also included in a secondary cell wall, which adds rigidity and waterproofing. This helps tree bark, plant leaf stalks, and other structures resist wind and other physical forces in the environment.
![]() |
![]() |
Cellulose, while indigestible to humans, is an important part of our diets because it makes up dietary fiber, which has been linked to lowered risk of diabetes and heart disease. In this module, you will take a closer look at the structure and function of these important macromolecules.
The paper you write on, the bowl of cereal you eat for breakfast, and the energy you use to walk up a flight of stairs all come from carbohydrates. Rice, wheat, and corn are some of humanity’s most important crops; these foods are the primary source of energy for much of our population. All three of these foods are high in carbohydrates.
A carbohydrate is an organic molecule that contains carbon, hydrogen, and oxygen. Carbohydrates are either simple (often referred to as sugars) or complex. Simple carbohydrates (sugars) are made up of only one or two sugar monomers. Each monomer has the proportion of carbon to hydrogen to oxygen in the ratio of 1:2:1, or (CH2O). You can see why these compounds are called carbohydrates; in a monomer, each carbon is “hydrated.” Complex carbohydrates are made up of more than two sugar monomers linked together. Carbohydrates can be further subclassified, again on the basis of structure, depending on the number of monomers in each molecule: monosaccharides and disaccharides (simple carbohydrates) and polysaccharides (complex carbohydrates).
Simple carbohydrates are quickly and easily accessed to generate energy the cell can use, while complex carbohydrates are used to store energy for a longer period of time. Some complex carbohydrates are also used as structural components. Carbohydrates also play a role in cell signaling and recognition within multicellular organisms.
Monosaccharides are the simplest sugars. The root “mono” means "one," and “saccharide” refers to an organic molecule where the ratio of carbon:hydrogen:oxygen is 1:2:1 (i.e., CNH2NON). The more common monosaccharides contain six carbons and have a molecular formula of C6H12O6. Glucose, fructose (fruit sugar), and galactose are examples of six-carbon monosaccharides. A common five-carbon monosaccharide is ribose (the sugar component of RNA), which has a molecular formula of C5H10O5. DNA contains a modified ribose where one of the oxygen atoms has been removed, hence the name deoxyribonucleic acid. Shorter monosaccharides exist as linear molecules while five- and six-carbon monosaccharides form ring-like structures.
Disaccharides are made of two monosaccharides linked together by a covalent bond (the root “di” means "two").
Polysaccharides are long chains of monosaccharides (“poly” means "many").
Glycogen, starch (amylose and amylopectin), and cellulose are all made of many linked glucose monomers. Glycogen is mainly used for energy storage in animals. Starch is mainly used for energy storage in plants. Cellulose is mainly used for maintaining plant structure. These three molecules are all made of glucose, so why do they have such different functions? The main difference is how the glucose monomers are linked together. This is a prime example of one of the central themes of biology: form affects function.
These differences are explained in more detail below.
Chitin, a modified structural polysaccharide, is best known as a major component of the exoskeleton (hard outer shell) of arthropods (e.g., beetles, crabs, lobsters) and mollusks (e.g., snails, clams, scallops). However, chitin also plays a minor role in the cell structure of some fungi, algae, and yeasts. Chitin (C8H13NO5)n is comprised of slightly-modified glucose monomers.
Digestible carbohydrates provide energy in a form that is easily accessible. Glucose, with the chemical formula C6H12O6, is a simple sugar and a monosaccharide. It is one of the most abundant and important energy molecules for living things. The energy stored in covalent bonds between atoms (within a glucose molecule, in this case) is released during cellular respiration, when glucose is broken down and converted to simpler, more stable molecules. This energy can be captured to make ATP, which can then be used to power cellular processes. This will be covered in greater detail in the metabolism unit.
In animals, excess glucose molecules are linked together (through an anabolic pathway) to make a long branching polysaccharide called glycogen. Glycogen is primarily stored in the liver, with small amounts being stored in the muscles. Once glycogen stores are filled, the body begins storing excess food calories as fat. Conversely, when food intake is insufficient to keep up with energy requirements, the body will break down its glycogen stores to release individual glucose monosaccharides.
Starch, another branching polysaccharide, is the primary form in which excess glucose is stored in plants (and other producers). Starch, just like glycogen, is composed of individual glucose monomers joined together. And, like glycogen, starch is broken down to release individual glucose molecules that can be used to build ATP. The difference between these two energy storage polysaccharides lies in the organization of the glucose monosaccharides.
Cellulose, like glycogen and starch, is a polysaccharide that is made up of many glucose monosaccharides linked together. Cellulose is an important structural carbohydrate that provides support within plant cell walls. This carbohydrate is indigestible by humans and most animals. Some animals (e.g. cows, horses, and termites) can digest this carbohydrate. These animals have symbiotic (“living together”) bacteria that inhabit their digestive tract. The bacteria produce enzymes that break down cellulose, making sugar available to their hosts.
Deoxyribonucleic acid (DNA) and ribonucleic acid (RNA) are the molecules involved with storing our genetic information (DNA) or the conversion of that information into proteins (RNA). Sugars (deoxyribose in DNA, ribose in RNA) play a part in the structure of these nucleic acid molecules. They alternate with phosphate groups to make up the long backbone of nucleic acids, as shown below.
Although we will not delve into the specifics of this topic, carbohydrates play a variety of roles in identification and signaling. One of the ways your immune system is able to identify foreign invaders is through detection of the different carbohydrates displayed on the surface of cells. Additionally, the first step in fertilization between a sperm and egg is generally believed to involve carbohydrate-to-carbohydrate interactions. As a final example, the specific carbohydrates attached to the surface of your red blood cells determine your “blood group” (A,B,O) and predict what kind of blood you can safely receive in a transfusion.
Indigestible carbohydrates, which cannot be degraded by human digestive enzymes, are referred to as dietary fiber. Cellulose, mentioned above as a major component of plant cell walls, is one of the largest contributors of dietary fiber. Pectin is another structural carbohydrate found in most plants, especially apples and citrus fruits, and is commonly used as a gelling agent to make jams and jellies. A diet high in fiber has several benefits. Dietary fiber plays a role in maintaining regular digestive functioning, slowing the absorption of sugar into the bloodstream, and reducing bad cholesterol. So the old saying “an apple a day keeps the doctor away” does have some truth to it, since there are proven health benefits associated with the consumption of foods that are high in fiber, like apples. Other foods that are high in fiber are oatmeal, popcorn, raspberries, black beans, lentils, and peas, to name a few.
For a quick boost of energy, simple sugars are the carbohydrate of choice; diabetics will oftentimes carry a sugary food or drink with them in case their blood sugar level drops too low. However, excessively high blood sugar levels can be equally dangerous. Accordingly, doctors routinely suggest that diabetics eat a diet high in fiber to help manage their diabetes, since fiber helps to slow the absorption of sugar into the bloodstream, which helps diabetics to prevent spikes in blood sugar. In contrast with simple sugars, complex carbohydrates are broken down by the body more slowly and energy is released in a stepwise fashion.
Lipids are a diverse group of macromolecules united by their hydrophobic, nonpolar nature. We are most familiar with the lipids known as fats. Fats are used to store energy for later use. They also provide structural support and cushioning for many animals. Some dietary fats are healthier than others. In this module, you will learn about the differences between saturated, unsaturated, and trans fats.
People are most familiar with fats, but there are other types of molecules that are not fats, but are lipids. Cholesterol is a familiar example of a lipid that is not actually a fat. You will learn more about cholesterol in this module as well.
Lipids include a diverse group of compounds that are united by two common features. First, they are largely nonpolar in nature. This is because they are hydrocarbons that include many nonpolar carbon-carbon or carbon-hydrogen bonds. Second, because lipids are nonpolar, they are also hydrophobic (water-hating), or insoluble in water.
Lipids perform many different functions in a cell. Cells store energy for long-term use in the form of fats. Lipids also provide insulation from the environment for plants and animals. For example, they help keep aquatic birds and mammals dry because of their hydrophobic nature. Lipids are also the building blocks of many hormones and are an important constituent of the cell’s plasma membrane. Lipids include fats, oils, waxes, phospholipids, and steroids.
Because lipids are such a diverse group of biomolecules, we will study them in four categories. Examine the following table for a general understanding of lipid structure and function.
Waxes are nonpolar lipids that form protective layers on plants and animals. For example, waxes cover the feathers of some aquatic birds to keep the feathers dry. Plants, on the other hand, often have leaves coated with waxes to prevent water from evaporating off the plant surface. Because of their hydrophobic nature, waxes prevent water from sticking on the surface of these structures.
The chemical structure of a wax is linked to its function. Waxes are the simplest lipids in nature, consisting of two long hydrocarbon chains linked in the middle by an ester group. It is this structure that enables waxes to function the way they do in living systems.
Steroids are a family of lipids based on a molecular structure with four fused carbon rings. This family of lipids includes many hormones and cholesterol. Hormones (such as estrogen and testosterone) are used by some animals as long-distance messengers. This means that they are produced by cells in one part of the body and affect cells in a different part of the body.
Cholesterol is an example of a familiar steroid that plays many important roles in your body. While most people are familiar with cholesterol, they often don’t realize how important it is for healthy function. In fact, most of the cholesterol in your body is synthesized by your liver. Cholesterol has many functions. Cholesterol is used to build steroid hormones, including testosterone and estradiol, which are secreted by the sex organs. Cholesterol is also the precursor to vitamins D and K. Some cholesterol is converted to bile salts, which help in the absorption of fats from the digestive system into the body. Cholesterol is also used by animals to maintain the proper consistency of the cell membrane, which is the structure that surrounds every cell. Other organisms use different steroids for this function. For example, plants use a steroid called stigmasterol, while fungi use ergosterol.
Fats and oils are called triglycerides because they are made of three (“tri”) fatty acids attached to one glycerol. A fatty acid is a carboxylic acid that contains a long hydrocarbon chain. The fatty acids are usually 16 to 22 carbons long, but can range from 4 to 36 carbons in length.
Fats are mainly energy storage and insulating molecules. Per gram, fats contain twice as much energy as carbohydrates. Layers of fat also surround the vital organs of animals to help cushion and protect them. In some animals, layers of fat under the skin provide insulation. This is particularly true of marine endotherms (i.e., organisms that maintain internal body temperature using their own metabolism) that live in cold polar regions, like penguins, whales, and seals. These are all examples of animals with thick insulating “blubber” beneath the skin.
In one triglyceride there are three fatty acids. All of these can be the same, or there can be two or three different types of fatty acids in a single triglyceride. The fat’s properties depend on the types of fatty acids that are present.
In saturated fatty acids, all of the carbon-carbon bonds are single bonds and each carbon is bonded to two or three H atoms; each carbon is “saturated” with hydrogen. Because all the bonds are single, the entire fatty acid can adopt a linear shape. Unsaturated fatty acids have one, two, or even three double bonds along the carbon “backbone.” Notice in the diagram below that at each double bond, the carbons involved are bonded to only one instead of two hydrogen atoms. Thus these fatty acids have less hydrogen; they are “unsaturated” with respect to hydrogen. The double bonds found in the unsaturated fatty acids have an effect on the overall shape of a triglyceride: unsaturated fatty acids are normally “kinked,” because the double bond prevents free rotation between the two carbons involved.
Saturated fats have three saturated fatty acid “tails” — all of which are linear in shape. With their compact shape, saturated fat molecules pack together efficiently and form a solid at room temperature. These are usually produced by animals; everyday examples of saturated fats include lard and real butter. Unsaturated fats have at least one unsaturated fatty acid. With their “kinky” fatty acid(s), unsaturated fats will not pack into a regular structure and thus remain fluid at room temperature. They are called oils, and are commonly produced by plants; everyday examples include olive oil, corn oil, and canola oil.
A diet rich in saturated fats may contribute to formation of plaques in the arteries (atherosclerosis) and increase your risk of heart disease. To reduce this risk, it is recommended that you replace some foods rich in saturated fats (e.g. fatty cuts of pork, beef, and high-fat dairy products like butter) with foods that provide unsaturated fats (available in foods like olive oil, canola oil, seafood, and walnuts). Such a shift is believed to help improve blood cholesterol levels and may reduce your risk of cardiovascular disease.
In the diagram above, you see the physical behavior of two different triglycerides (with saturated and unsaturated fatty acids) at three distinct temperatures. Figure 1 shows that at a temperature of 0oC (freezing), both types of fatty acids stack neatly. At 16oC (Fig. 2), the unsaturated fatty acids lose the tightly stacked formation and become more mobile. At 70oC, the saturated fatty acids also become mobile (Fig.3). At 27oC (room temperature) and at 37oC (body temperature), triglycerides with unsaturated fatty acids are liquid oils, whereas saturated fats remain solid.
Phospholipids are an important class of lipids found in the membranes of all cells and organelles. A phospholipid contains only two fatty acids attached to a glycerol, which in turn is bound to a phosphate group. Together, the phosphate and the glycerol make up the “head” of the phospholipid. The fatty acids make up the molecule’s “tail.” The phosphate group carries a negative charge and is therefore hydrophilic. The tail, on the other hand, is made up of nonpolar fatty acids and, like other lipids, is hydrophobic. Thus phospholipids, like fatty acids, are amphipathic. An amphipathic molecule has both a hydrophilic end and a hydrophobic end.
Remember that hydrophilic molecules dissolve in water and hydrophobic ones do not. But phospholipids have a “split personality” in this regard. How do they behave in water? When small numbers of phospholipids are in an aqueous solution they will self-assemble into micelles, structures that exclude water molecules from the hydrophobic tails while keeping the hydrophilic head in contact with the aqueous solution. If enough phospholipids are present, they will form a bilayer. This is the favored formation because it is the most stable orientation for phospholipids in a water solution. View the animation that demonstrates the formation of micelles and bilayers.
With their “split personality,” phospholipids are able to play a very important role in biology. Phospholipids, together with other molecules in smaller quantities, form membranes that surround the cell and intracellular organelles such as mitochondria. The cell membrane is a fluid, semipermeable bilayer that separates the cell's contents from the environment. See animation below.
Although some cholesterol is essential for healthy physiological functioning, excess cholesterol can lead to health problems. Because cholesterol is hydrophobic, it cannot be easily transported through the body in the blood stream. A carrier is required to transport cholesterol around the body to the cells that need it. These carriers are called lipoproteins. You can think of lipoproteins as cholesterol boats that carry cholesterol to the body’s cells (where it can be used to build or repair cell membranes or synthesize steroid hormones). Lipoproteins that carry cholesterol to the cell for use are called low-density lipoproteins (or LDL). LDL is often referred to as “bad” cholesterol, because excess LDL levels have been linked to plaque deposition in the arteries, which can result in heart disease. Another type of lipoprotein, known as high-density lipoprotein (or HDL) transports cholesterol to the liver, where it is often turned into bile salts and excreted. HDL is often referred to as “good” cholesterol, because high blood concentrations of HDL have been linked to a decreased risk of heart disease. Exercise and proper diet can increase HDL levels. Increasing intake of unsaturated fats and decreasing intake of saturated fats is also a good strategy for increasing HDL levels.
Proteins are the most functionally diverse group of biomolecules we will examine in this unit. Critical to our diet, protein can be found in animal products like meats and cheeses, as well as in plant products like beans and grains. Kwashiorkor, which causes a distinct swelling of the abdomen, is often seen in malnourished children who lack sufficient protein in their diets. Proteins are also found in many toxins, such as the incredibly poisonous toxin produced by the bacterial species Clostridium botulinum. Studies have shown that just one teaspoon of this poison, which is a protein, would be enough to kill 20 percent of the world’s population! This module will take a closer look at the structure of a protein and examine how protein structure enables such a wide range of diverse functions.
Proteins are macromolecules built from amino acids. There are 20 different amino acids that can be joined in a myriad of ways to produce molecules with an enormous variety of possible structures that can perform a huge number of critical cellular functions. The following list represents just a sampling of the many things proteins can do in living systems.
Take a look at this diagram to see even more ways that proteins are used in your body.
Amino acids are the building blocks of proteins. A protein is composed of a series of amino acids attached end-to-end via covalent bonds. Consequently, a protein looks like a string of beads, with each bead representing an amino acid. The bond between the amino acids is referred to as the peptide bond.
An amino acid is a small organic chemical that is made up of four parts. There is a nitrogen-containing amino group on one end of an amino acid. The other end of an amino acid has a carboxylic acid group. These two groups (the amino group and the carboxylic acid group) are the reason for the name “amino acid.” The amino and carboxylic acid groups are linked by a single carbon atom called the alpha carbon. Finally, there is a variable “R group” also attached to the alpha carbon. The amino acids differ in the nature of the R group that is attached to the central alpha carbon. There are 20 different R groups commonly found in nature. In this way, all amino acids are identical except for the different R groups (also called side chains) attached to the alpha carbon. The properties of different proteins depend entirely on the sequence, or arrangement, of the amino acids.
There are many ways to represent an amino acid. Take a look at some of the representations below.
![]() |
![]() |
Proteins are built when amino acids are linked end-to-end by covalent bonds. The covalent bonds that link amino acids are called peptide bonds. A peptide bond is formed when the amino group of one amino acid reacts with the carboxylic acid of another amino acid. After a peptide bond forms, the three conserved parts of the two amino acids (the nitrogen group, the alpha carbon and the carboxylic acid) are all linked, forming the main chain, or backbone, of the protein. The variable side chains of each amino acid (the R groups) project out from the main chain.
Two amino acids join to form a dipeptide. Longer chains of connected amino acids are often called polypeptides. When polypeptides are synthesized, new amino acids are added only to the carboxylic acid end of the existing chain. A completed polypeptide is directional; it has two different ends. One end has a free amino group, while the other end has a free carboxylic acid group. When amino acids in a protein are counted, the numbering starts with the amino acid that has the free amino group. The last amino acid is also the last one that was added when the chain was built: it has a free carboxylic acid group.
The order of the amino acids in a particular protein is determined by information encoded in the cell's genes, and the order of amino acids is referred to as the sequence of the protein. An example of a protein sequence is shown below where the one-letter abbreviations are used for each of the 20 amino acids used in cellular protein synthesis. By convention, the first amino acid is at the end with the free amino group and the last amino acid is at the end with the carboxylic acid group.
The linear chain of amino acids will spontaneously fold into a three-dimensional shape, which is usually the active form of the protein. For example, look at the drawing of the three-dimensional structure of myoglobin (shown on the left). You can manipulate the 3D structure using the Jmol image on the right.
![]() |
|
Many folded proteins bind small organic molecules that assist the protein in performing its function. For example, the heme group in myoglobin (colored gray in the right image) is not part of the protein chain, but fits within a pocket in the myoglobin protein. Myoglobin is responsible for binding oxygen in muscles. Heme groups also are present in the oxygen transport protein, hemoglobin.
Proteins spontaneously fold to form their functional three-dimensional shape, reaching the lowest energy state, because this is the most stable state. The folding process can be reversible, meaning that a folded protein can unfold and then refold. However, unfolding can also be irreversible, where the unfolded protein chains tangle with each other, forming an aggregate and precipitating out of solution. This is often referred to as a denatured state. Conditions that cause proteins to unfold and denature are extremes of temperature and pH. Cooking denatures proteins by heating so they unfold and denature. The same effect occurs when proteins are exposed to the low pH in your stomach. This helps your body digest the proteins found in your food, and it helps kill ingested bacteria and viruses.
Although many factors stabilize the final shape of a protein, the most important factor is the hydrophobic effect. Amino acids with nonpolar (hydrophobic) side chains are driven into the central core of the protein because they are repelled by the water. This is known as the hydrophobic effect, and it results in proteins that have amino acids with hydrophilic side chains on the surface of the protein. The folded protein reaches the lowest energy state, which is the most stable, by folding in a way that will optimize the burial of nonpolar amino acid side chains while still leaving the polar groups on the surface. Therefore, the shape of the folded form of the protein depends on the order of the hydrophobic amino acids in the protein.
Consider a short 12 amino acid protein consisting of polar (white) and nonpolar (black) amino acids. This polypeptide will fold to bury most of the nonpolar amino acids, removing them from contact with water, giving the lowest energy state. Other structures are possible, but these will be higher energy (and consequently less stable) because they expose more nonpolar amino acid side chains to water.
The biological function of most proteins involves binding to something else. The binding event may be the first step in transport, signaling, regulation, or an enzyme-catalyzed reaction. Molecules that bind to proteins without being modified are called ligands. Although ligands are usually small molecules, they can also be larger than the protein to which they are binding. For example, when proteins bind to DNA, it is the DNA that is the ligand. Oxygen is a ligand when it binds to hemoglobin (a protein) during oxygen transport to the tissues. Molecules found on the surface of bacterial cells are ligands when antibodies (proteins) in your blood bind to them during an immune response. Sometimes ligands are given special names to remind us of their function. For example: enzyme inhibitors are ligands that bind to and inhibit enzymes; antigens are ligands that are recognized by antibodies (which are proteins).
The ligand binds to the protein by interacting with amino acids in the protein’s
binding site.Binding proteins have binding sites for their ligand (L).
Binding proteins have amino acids in their binding site that are complementary to the ligand. Generally, a higher degree of complementarity leads to tighter binding and a more specific interaction. The molecules can be complementary in these different ways:
The interactions between proteins and their ligands are incredibly diverse, but they all share some basic ground rules. Protein-ligand binding is:
Enzymes are proteins that have the ability to bind substrate in their active site and then chemically modify the bound substrate, converting it to a different molecule — the product of the reaction. Substrates bind to enzymes just like ligands bind to proteins. However, when substrates bind to enzymes, they undergo an enzyme-induced chemical change, and are converted to products.
The substrate binds to the enzyme by interacting with amino acids in the binding site. The binding site on enzymes is often referred to as the active site because it contains amino acids that both bind the substrate and aid in its conversion to product.
You can often recognize that a protein is an enzyme by its name. Many enzyme names end with “-ase.” For example, the enzyme lactase is used to break down the sugar lactose, found in mammalian milk. Other enzymes are known by a common name, such as pepsin, which is an enzyme that aids in the digestion of proteins in your stomach by breaking the peptide bonds in the proteins.
Enzymes are catalysts, meaning that they make a reaction go faster, but the enzymes themselves are not altered by the overall reaction. Examine this image to see how enzymes work.
The amino acids in the active site of enzymes play two roles, and sometimes those roles overlap. Some of the amino acids in the active site are responsible for binding of the substrate and others are responsible for facilitating the chemical reaction. Enzymes are generally quite specific for their substrates. Although lactase and pepsin both catalyze the same type of reaction, breaking a bond using water (hydrolysis: "hydro" means "water" and "lysis" means "to break"), lactase only functions when lactose is its substrate and pepsin can only break peptide bonds.
Enzymes catalyze many different types of chemical reactions. Some of the reactions are synthetic and result in products that are more complex than the original substrates. This is often done by binding two substrates together. Synthetic metabolic pathways like this are called anabolic pathways.
Other enzymes catalyze reactions that reduce the substrate to simpler products. These enzymes catalyze reactions in catabolic pathways.
An example of an enzyme that is involved in catabolism is lactase. Lactase is produced in the intestinal tract of mammals and breaks the sugar in milk, lactose, down into the monosaccharides glucose and galactose so that they can be metabolized to produce energy.
Unfortunately, in many adults the ability to produce lactase is diminished. Without adequate lactase, ingested lactose from milk and cheese is not broken down and absorbed in the small intestine. Instead, the lactose goes to the large intestine, where it is converted into large volumes of carbon dioxide (CO2) gas by the bacteria in the large intestine. These individuals are lactose intolerant and should avoid eating foods with high milk content.
In all chemical reactions, there is an initial input of energy that is required before the reaction can occur. If this initial energy requirement (called the activation energy or energy barrier) is small, then the reaction will happen quickly and easily. If the activation energy is large, then the reaction will take longer to occur. Enzymes function to reduce the activation energy required for a chemical reaction to occur.
First, the enzyme binds to the substrate and slightly distorts its shape. The change in shape activates the substrate molecule and decreases the total activation energy required for the substrate to be turned into product. As the number of activated substrate molecules increases, so does the conversion of substrate to product. An analogy for this effect is a ski hill, with skiers at the bottom of one side of the hill representing substrates, skiers on the top of the hill representing activated substrates, and the products being the number of skiers that ski down the other side. If the height of the hill is lowered (due to the presence of the enzyme), then more skiers can make it to the top, increasing the number that ski down to become products.
The activity of an enzyme (its ability to convert substrate to product), depends on a number of parameters that are listed below. Many of these parameters can help control the activities of enzymes to optimize the utilization of cellular resources.
Lactase is an example of an enzyme that breaks large molecules down into smaller ones. Lactase is produced in the intestinal tract of mammals and breaks the sugar in milk, lactose, into the monosaccharides glucose and galactose. The monosaccharides can then be metabolized to produce energy.
Lactase breaks lactose into galactose and glucose by adding a water molecule to the bond between the two sugars. Without adequate lactase, ingested lactose from milk and cheese is not broken down, and therefore cannot be absorbed in the small intestine. Instead, the lactose goes to the large intestine, where it is eaten by bacteria that live there. After metabolizing the lactose, the bacteria produce large volumes of carbon dioxide (CO2) gas, which results in significant discomfort for the individual lacking lactase.
Unfortunately, as many individuals age, their ability to produce lactase is diminished. This is especially true for individuals from Asian, African, Native American, or Mediterranean descent. The incidence of lactose intolerance is lower for those of European descent. Approximately one in seven American adults suffer from lactose intolerance, and African Americans can show signs of lactose intolerance as early as two years of age.
These individuals are lactose intolerant and should avoid eating foods with high lactose content, such as milk. Milk products, such as buttermilk, yogurt, and cheese, cause less of a problem because the amount of lactose in these products is reduced by the microorganisms that are used to make them. Lactose intolerant individuals can consume milk under certain conditions.
In 1953, a group of scientists contributed to one of the most significant scientific discoveries in history when they determined the structure of DNA, a nucleic acid that is the hereditary material in a cell.
Nucleic acids are macromolecules that carry out two main functions in the cell: storage of genetic information and synthesis of proteins. Two types of nucleic acids specialize in these functions: deoxyribonucleic acid (DNA) and ribonucleic acid (RNA). DNA is the genetic material that stores information for making proteins in all living organisms. Some viruses store their genetic information in RNA instead of DNA. This may seem like an exception to the universal use of DNA as genetic material; however, recall that viruses are not cellular, and are not considered living organisms.
A nucleic acid consists of a chain of nucleotides. There are different kinds of nucleotides that can be linked to build the different nucleic acids. DNA is probably the most familiar nucleic acid, but in this module, you will also learn about RNA. The nucleotides that are used to build RNA molecules are different from the nucleotides that are used to build DNA. You'll also take a look at a very special nucleotide — ATP, which acts as the energy currency of the cell.
Like any polymer, nucleic acids are made up of repeating subunits or monomers. The subunits of nucleic acids are called nucleotides. Each nucleotide contains three building blocks: a phosphate, a sugar, and a nitrogenous base. The nitrogenous base is a type of organic molecule that consists of one or two ring structures. Note that the term “base” is used here to refer to a specific part of the nucleic acid structure; the meaning is very different from the use of the term “base” in the discussion of pH and acid-base chemistry.
There are five different nitrogenous bases that can be found in various nucleotides. In DNA, you will find these nitrogenous bases: adenine (A), guanine (G), cytosine (C), and thymine (T). In RNA, thymine is replaced by uracil (U).
Nucleotides can also differ in the type of sugar they contain. All sugars found in nucleotides are pentose sugars, which means they have five carbons; however, they can be different pentose sugars. For example, DNA contains the sugar deoxyribose. RNA, on the other hand, contains ribose.
Finally, nucleotides can differ in the number of phosphate groups they contain. DNA and RNA nucleotides contain just one phosphate. But a very common nucleotide used for energy (adenosine triphosphate) contains three phosphates. We will look at this molecule in greater detail in the Metabolism unit.
Let’s examine the structure of a nucleic acid polymer. How are nucleotide building blocks connected to each other? The alternating "phosphate - sugar - phosphate" building blocks form a backbone of the linear polymer. Bases are attached to the backbone via the sugars and are equivalent to the “side chains” of amino acids in proteins.
Just as with proteins, a DNA (or RNA) strand has directionality: one end is different than the other. In the case of proteins, an amino group is found on one end and a carboxylate group is found on the other end. Since it is the order of bases that carries out the function of DNA (information storage), we often represent DNA as just a sequence of bases (GAGGCT) and do not bother representing the backbone. By convention, the first letter is the base at the 5’ (five prime) end of the DNA strand. Thus, the sequence GAGGCT is shorthand for 5’-GAGGCT-3’. Determine how this directionally occurs in DNA in the following learn by doing.
All DNA found in living organisms comes in sets of two polymers or strands, usually referred to as double-stranded DNA. Some viruses are an exception to this rule and contain single-stranded DNA as their genetic material. The discovery that DNA structure is a double helix (two strands wound around each other) is considered one of the greatest discoveries in science. Understanding of the structure of DNA opened up a whole new realm of biological and biomedical research (molecular biology) and led to the development of new technologies in science and medicine. Let’s examine the structure of DNA using the activity below.
DNA structure is an antiparallel double helix: the two DNA strands run in opposite directions. The sugar-phosphate backbone is on the outside, and the bases are on the inside of the helix. The two strands are held together by base pairing: hydrogen bonding between specific bases. Adenine (a purine) always pairs with thymine (a pyrimidine); guanine (a purine) always pairs with cytosine (a pyrimidine). These base pairing rules (A-T, G-C) are very important to the structure and function of DNA.
Two of the most important nucleic acids are DNA and RNA. DNA contains the instructions for building proteins. RNA, on the other hand, is involved in actually building the proteins. Because of these different functions, the structures of the two molecules are very different.
Compare DNA and RNA and answer the questions about these structures.
In this module you learned that nucleotides are the monomers that make up the nucleic acid polymers. Adenosine triphosphate (ATP) is a nucleotide. It consists of a single adenosine (the base adenine and the sugar ribose), linked to three phosphate ions. However, ATP has another essential function: it acts as a general energy source for most cellular activities. You will learn in detail about ATP in the unit dedicated to metabolism.
ATP is a relatively unstable molecule; consequently, it is never used for the long-term storage of energy in the cell. This job goes to other more stable compounds, like fats and sugars. However, ATP is specialized for direct and rapid transfers of energy. The bond between the second and the third phosphates can be broken in a reaction producing ADP (adenosine diphosphate). The energy released in this reaction can be used for chemical reactions or cellular work. Conversely, if there is a surplus of energy, specialized reactions can produce ATP from ADP and phosphate, storing the energy temporarily before being used for other processes.
The ADP generated when energy is released is recycled back to ATP using energy gained from the metabolic process. Large amounts of ATP are consumed while providing energy for biological functions.
The cell is the first level of organization that exhibits all the properties of life. Made up of biological macromolecules, the cell’s unique structure enables it to carry out the functions of life.
In the "Biological Macromolecules" unit, you explored matter and learned how atoms can be combined to form molecules that, through their own unique structures, are able to carry out specific functions. You learned that proteins, such as enzymes, are able to function like little molecular machines. However, enzymes and other biological macromolecules made of matter do not possess the properties of life. In fact, it is only after we combine these molecular building blocks to form a cell that we finally see the emergent property of life. Take a moment to review examples of each level of organization forming these building blocks using the activity below.
One of the hallmarks of living systems is the ability to maintain homeostasis, or a relatively constant internal state. The cell is the first level of complexity able to maintain homeostasis, and it is the unique structure of the cell that enables this critical function.
In this unit, you will learn about the cell and all the parts that make it functional. You will also focus on the cell membrane, which is the structure that surrounds the cell and separates its internal environment from the external environment. This separation enables the cell to maintain homeostasis and exhibit the emergent property of life.
The microscopes we use today are far more advanced than those used in the 1600s by Antony van Leeuwenhoek, a Dutch shopkeeper who had great skill in crafting lenses. Despite the limitations of his now-ancient lenses, van Leeuwenhoek observed the movements of bacteria, other single-celled organisms, and sperm. He labeled these moving microbes “animalcules.”
In a 1665 publication called Micrographia, experimental scientist Robert Hooke coined the term “cell” for the box-like structures he observed when viewing cork tissue through a lens. Later, he confirmed van Leeuwenhoek’s 1678 discovery of bacteria and protozoa. Later advances in lenses and microscope construction enabled other scientists to see some components inside cells.
By the late 1830s, scientists had closely examined many plant and animal tissues under the microscope. Comparing notes, botanist Matthias Schleiden and zoologist Theodor Schwann realized that cells were found in every tissue they had studied. They proposed the Unified Cell Theory, which states that all living things are composed of one or more cells, and that the cell is the basic unit of life. You, for instance, are made of approximately 60 trillion cells, all of which originated from one single cell, the fertilized cell produced when an egg from your mother was fertilized by the sperm cell from your father.
However, there were still controversies between those who believed in the existence of a vital force able to “create” life (spontaneous generation) and those who supported biogenesis, which claims that living cells can arise only from pre-existing cells. A series of experiments by Louis Pasteur and Rudolf Virchow showed that living organisms could only come from other living organisms. The currently accepted tenets of Cell Theory are:
Modern cell theorists assert that all functions essential to life occur within the cell and that during cell division the cell contains and transmits to the next generation the information necessary to conduct and regulate cell functioning.
In spite of the fact that all cells share certain characteristics, there is incredible diversity in structure and function among different cells. The human body alone contains trillions of cells of more than 200 different types, each with a unique structure and function. Cells are categorized into two types: Prokaryotes are small, simple, single-cell organisms; bacteria are the most prevalent kind. Eukaryotes are larger, and most often they are multicellular organisms, including plants, animals, and fungi. Eukaryotic cells are about 15 times wider than the typical prokaryotic cell, and up to 1,000 times greater in volume. The images below show examples of a prokaryote and single-celled eukaryotic organisms.
Prokaryotic MRSA Bacteria Cells![]() |
Eukaryotic Single-Celled Organisms![]() |
You have already learned that cells are made up of organelles, molecules, atoms, and subatomic particles. These parts assemble into various structures to perform specific functions within the cell. Organelles are specialized structures formed when a specific set of molecules bonds, providing a subunit within a membrane-like enclosure that performs particular functions within the cell. Some structures are the same for both prokaryotic and eukaryotic cell types and some structures are different. The remainder of this page will focus on a handful of structures that all cells have in common. Then, later in this module, you will learn about the structures and characteristics of each cell type in more detail.
There are several parts (referred to as structures) common to all cells regardless of the cell type. All cells are surrounded by a cell membrane. The cell membrane provides a barrier between the interior and exterior of the cell and it regulates the flow of substances in and out of the cell. All cells also have cytoplasm, which is the fluid that occupies the space inside the cell. Cytoplasm is the space in which the chemical reactions that enable life take place. All cells also contain DNA, which is often called the “master molecule” of the cell because it contains the instructions for synthesizing all of the cell’s proteins. As you already know, proteins are the raw materials used to build many important structures in living systems. Finally, both prokaryotes and eukaryotes contain ribosomes. Ribosomes are the molecular machines that use the instructions contained in the DNA to build all the proteins needed by the cell.
The following table summarizes the major structures found in cells and the primary function of each structure. In addition, it lists the types of cells in which each structure is found. (There are many other important structures and features of cells. The table below shows only the most important structures.)
The essential differences and similarities among prokaryotic and eukaryotic cells are:
Check your understanding of cell structures with the following activity and then learn more about the structures and characteristics of each cell type on the next page.
Every cell on Earth belongs to one of two categories:
Prokaryotic cells were the first cells to appear on our planet. All prokaryotes alive today are unicellular (one-celled), and include bacteria (singular form is "bacterium") and archaea (singular form is "archaean"). Prokaryotes are small cells that don't have a nucleus or membrane-bound organelles.
Eukaryotic cells appeared 1.5 billion years after prokaryotes. The main difference between the two is that eukaryotes have a central control structure, called the nucleus (plural form is "nuclei"), where DNA is housed. In prokaryotes, the main DNA molecule (bacterial chromosome) is present in a region called the nucleoid, but the nucleoid lacks a surrounding membrane. Smaller DNA molecules called plasmids can be also found in prokaryotes. Prokaryotic DNA is circular, in contrast to the linear structure of eukaryotic DNA.
Both eukaryotic and prokaryotic cells have a cell or plasma membrane, which surrounds and defines the inner environment of the cell. The cell membrane is made of a phospholipid bilayer containing a variety of proteins and additional components. The cell membrane is responsible for mediating interactions between the cell and its environment. The Cell Membrane module contains a more detailed discussion of the structure and function of the cell membrane.
Typical Animal Cell (Eukaryotic)![]() |
Bacteria Cell (Prokaryotic)![]() |
Prokaryotic cells have a simpler structure than eukaryotic cells, and they range in diameter from 0.1 to 5.0 µm (micrometers). Most prokaryotes have a protective layer called the cell wall that is made of peptidoglycan, which is a combination of polysaccharides and amino acids. Prokaryotes also have a cell membrane and cytoplasm. Many prokaryotes also have external appendages such as a flagellum. The cytoplasm contains the DNA and the ribosomes, where protein synthesis takes place. Several types of RNA are involved in the process of protein synthesis, and ribosomal RNA (rRNA) is the main component of ribosomes. Both prokaryotes and eukaryotes have ribosomes, but they are different. Prokaryotic ribosomes are smaller and lighter than their eukaryotic counterparts.
Note that certain abiotic entities such as viruses and prions (infectious proteins causing diseases such as mad cow disease) are also studied in biology. However, they are not cells, and while they may exhibit certain characteristics of life, they do so only in certain conditions.
All eukaryotes contain a nucleus. Animals and plants are familiar eukaryotes; in fact, all large complex organisms are eukaryotes. Fungi (the singular form is "fungus"), which include molds and mushrooms, are also eukaryotes. There are even single-celled eukaryotes, called protists. All eukaryotes have a nucleus, but they can also have other cell structures in common.
One simplified but useful analogy of a cell is that of a factory. Just as in a factory, cells have a wall or membrane providing a protective enclosure, a planning center where the product blueprints are stored, a source of energy, an assembly line for production, packaging and shipping facilities, storage facilities, etc.
Eukaryotic cells contain compartments with specialized functions called organelles. Organelles are surrounded by membranes. Organelles are similar to the specialized work areas in the factory above. Because a cell is a protein-producing factory, we can take a closer look at the different organelles and see how they function toward this goal.
Nucleus: Control Center![]() |
The nucleus is the control center of the cell and it stores the DNA, which contains the instructions for how to build all the protein products required by the cell. The nucleus is like the factory command center, which stores the instructions needed to build its product. A single molecule of DNA is called a chromosome. The chromosomes are like the different books in the factory’s control center. The nucleus is surrounded by a double-layered membrane called the nuclear envelope. The nuclear envelope is studded with pores that allow information from inside the nucleus to enter the cytoplasm. You can imagine that a factory command center would not be very effective if it did not have doors or windows through which to pass information to the rest of the factory. |
Ribosome: Protein Production![]() |
There are several important structures found within the nucleus. The most visible of these is the nucleolus. In contrast to other organelles, the nucleolus is not bound by a membrane. Instead, it is an aggregate of molecules where ribosomes, another type of nonmembranous organelle, are assembled. Most organelles, including ribosomes (after they are built in the nucleus), are found in the cytoplasm, which is the substance found between the nucleus and the cell membrane (number eight in the cell factory diagram). The cytoplasm is analogous to the factory floor, where all the work takes place. |
Mitochondria: Power Plant![]() |
All this work requires energy. Most factories need some sort of power plant that converts fuel into a form of energy that can be captured to do work. In the cell, this job is accomplished by organelles called mitochondria (singular form is "mitochondrion"), which take fuel in the form of sugar (glucose) and convert it to usable energy — ATP. |
Chloroplasts: Sugar Production![]() |
Plant cells have an additional type of organelle — chloroplasts, which exist only in plant cells — involved in energy transfer. The chloroplasts provide sugar. Chloroplasts capture energy from the sun and use that energy to build sugar molecules. Mitochondria then harvest the energy stored in the sugar molecules and use it to do work. The chloroplasts and mitochondria are the organelles responsible for providing energy for all cellular functions. |
Endoplasmic Reticulum: Product Assembly![]() |
Factories often have assembly lines that put together the company’s product. In the cell, proteins and other cellular components are put together, or assembled, by the endoplasmic reticulum (ER), a series of sacs and tubes. In eukaryotes, ribosomes are associated with the "rough ER," which gets its name from the beaded appearance that the ribosomes give it. The smooth ER (without ribosomes) can have different functions depending on the cell type, but it is often the site for the synthesis of lipids. |
Vacuoles: Storage![]() |
The Golgi apparatus is the packaging and shipping center of the cell, where the proteins that were built by the ER assembly line are delivered to different parts of the cell, or in multicellular organisms, to different parts of the body. Often the Golgi apparatus packages proteins in vesicles and vacuoles, which are membrane-bound sacs that function in storage and transport. Vesicles are specialized for transport and some other functions. Their membranes can fuse with the plasma membrane, allowing them to empty their contents into the extracellular space. Vesicles also may fuse with the membranes of the endoplasmic reticulum and Golgi apparatus, allowing them to empty their contents into those organelles. Lysosomes are specialized vesicles found only in animal cells. Lysosomes contain powerful digestive enzymes that can recycle cellular parts or destroy external invaders. Vacuoles are specialized mainly for storage. Their membranes do not fuse with the membranes of other cellular components. |
Eukaryotic cells include other important structures not illustrated in our factory diagram; they are the cytoskeleton, actin, centrioles, microtubules, flagellum, and cilia.
The cytoskeleton is formed by a series of protein filaments, and is both a scaffold for the cell structure and a framework for many cellular activities, including movement and cell division. In our factory analogy, the cytoskeleton is represented by structural beams, as well as any transport infrastructure, like hallways, elevator tracks, or even small railroads. In the cell, the cytoskeleton is made up of many different protein filaments. One important filament is actin, which has a prominent role in cell movement, maintaining cell shape, and connection to other cells, as well as intracellular transport. A barrel-shaped structure only present in animal cells is the centriole, which plays a role in the spatial organization of the cell and cell division. Centrioles are formed by microtubules, another filament type of the cytoskeleton. Microtubules also form appendages such as the flagellum of the sperm cell and the cilia of the cells of the respiratory system.
Both animal and plant cells are eukaryotic cells. However, they both contain some specialized structures. Plants have chloroplasts and a rigid cell wall. Chloroplasts are unique organelles able to harvest solar energy to make sugars from carbon dioxide. This process, called photosynthesis, is the basis of life as we know it on our planet. This process is possible due to the presence of special pigments, called chlorophylls, which are found in the chloroplasts.
The cell wall is a porous structure that protects, supports, and gives shape to the cell. The plant cell wall is different from that of prokaryotes, and it is mainly formed by polysaccharides, particularly cellulose. Plant cells can also have a central vacuole, which can be a place of storage, degradation, defense, and even physical support for the cell. Even though plants have a cell wall, this does not mean they don’t have a cell membrane. All cells have a cell membrane. Some, like plant cells, just have a cell wall as well.
The structures shown above are parts of a “generic” eukaryotic cell. While most eukaryotic cells will present these structures, variations occur depending on the function of the cell. Think about cars — while most cars have the same parts, there will be differences between a sedan and an off-road vehicle. Among animal cells, extreme examples of specialization include red blood cells (containers of hemoglobin to transport oxygen; see picture on the left) and nerve cells (dedicated to transmission and integration of signals; picture on the right).
Blood Cells![]() |
Nerve Cell![]() |
Until the late 1600s, only large life forms such as plants and animals were known to science. Humans were completely ignorant of the teeming masses of microscopic life forms that inhabit our world. In fact, they didn’t even know that plants and animals are composed of cells.
The invention and refinement of the microscope changed all that. Today we are aware of a vast diversity of single-celled, microscopic life forms that interact with us every day. In fact, in and on your body, bacterial cells outnumber your own human cells 10 to one. The bacteria that live on and in you are often critically important to your own health and well being. Bacteria have mutually beneficial relationships with every large organism on Earth and they also cleanse the water we drink and renew the air we breathe. While bacteria can be invaluable partners in the quest for healthy living, it is also well known that bacteria cause disease. Interestingly, this fact was not fully demonstrated until the late 1800s. The earliest measures taken to control bacterial diseases involved improving sanitation using procedures such as sterile surgical techniques, water purification, and sewage treatment. While these strategies helped reduce infections caused by bacteria, it didn’t change the fact that once bacterial infections took hold, they were often fatal.
In 1928, Alexander Fleming serendipitously discovered the first antibiotic, Penicillin. Fleming was studying bacteria in his lab and was growing different cultures on Petri dishes. He wasn’t very tidy and had accidentally forgotten to clean some Petri dishes before leaving his lab for a monthlong vacation. When Fleming returned to the lab, he noticed that a mold had grown on one of the plates. Looking closely, Fleming saw that his bacteria had not grown well in a zone surrounding the moldy spots. Following up on a hunch, he deliberately added mold to some Petri dishes seeded with bacteria and found that all bacteria were killed near the mold colonies. By the mid-1940s, just in time to help wounded soldiers in World War II, scientists were able to mass-produce a drug derived from the mold discovered by Fleming.
Fleming had discovered the first antibiotic. Antibiotics are substances that are produced by one type of organism (usually bacteria or mold) and used to kill or suppress other organisms (usually bacteria). Antibiotics have a very specific way of working, which is called their “mode of action.” Penicillin works by interfering with cell wall construction in bacterial cells. The cell wall is a somewhat rigid “shell” that encloses a bacterial cell. In the presence of penicillin, a bacterium’s cell wall cannot be built (or rebuilt) and the cell wall eventually collapses, killing the cell. Cell walls are not present around human cells, or those of any other animal. Therefore, penicillin does not harm our cells — only those of bacteria. This is a general feature of antibiotics. When taken in prescribed doses, antibiotics are usually harmless to humans.
During and just after World War II, penicillin was touted as a wonder drug, and many believed the battle against pathogenic (harmful) bacteria would be won in a matter of years. However, the initial success of this antibiotic was relatively short-lived, as bacteria began to evolve resistance to the drug. In any group of bacteria, individuals can have variations in their genetic makeup that help them survive in the presence of specific antibiotic compounds. Over time, if the antibiotic is used, vulnerable types of bacteria die out and the “resistant” types of bacteria reproduce much better than average. Eventually most or all bacteria in the environment are resistant, making the antibiotic useless.
Presently, we are facing a growing threat to human health; the overuse and misuse of antibiotics has led to the evolution of “superbugs” — bacterial strains that are resistant to many different types of antibiotics. Such infections are very hard to treat, and can be deadly. Perhaps you have heard about methycillin-resistant Staphylococcus aureus (MRSA), which can be picked up in hospitals and places of close body contact (locker rooms, retirement homes). It has evolved a resistance to the antibiotic methycillin. If it penetrates the body, it may dissolve the skin and underlying muscle.
One way to reduce the development of antibiotic resistance is to use antibiotics only when necessary, so that there are many antibiotic-free people and places in the environment. This makes it easier for vulnerable strains of bacteria to reproduce, and keeps these strains more common. Within a sick patient, however, the goal of antibiotic therapy is to keep reducing the numbers of harmful bacteria until your immune system can kill off the last (and probably most resistant) holdouts. Therefore it is important to take antibiotics at full dosage and to complete the round of treatment as directed, continuing until it is certain that your bacterial foes have been eliminated.
We are not helpless against the rising tide of antibiotic resistance. Today, with our great knowledge of microscopic life and nonliving — but equally dangerous — viruses, we have an unprecedented ability to understand and address health challenges.
The cell membrane, also called the plasma membrane, is the boundary of the cell; it determines what enters and exits the cell, and it is how the cell interacts with its environment. Have you ever looked closely at the colors swirling in a soap bubble? This may give you a feel for the fluid nature of a cell's membrane. A membrane's components are in constant motion, as if they were flowing in a river. A variety of different proteins, carbohydrates, sterols, and other molecules are embedded in the phospholipid bilayer. This gives the impression of a tile mosaic, which has variously shaped and colored tiles embedded in grout. Because of its fluidity and its variety, biologists currently describe the cell membrane as a fluid mosaic.
The phospholipid bilayer is the main fabric of the membrane. The bilayer's structure causes the membrane to be semipermeable. The hydrophobic core blocks the diffusion of hydrophilic ions and polar molecules. Small hydrophobic molecules and gases, which can dissolve in the membrane's core, cross it with ease.
Other molecules require proteins to transport them across the membrane. Proteins determine most of the membrane's specific functions. The plasma membrane and the membranes of the various organelles each have unique collections of proteins. For example, to date more than 50 kinds of proteins have been found in the plasma membrane of red blood cells.
What is important about the structure of a phospholipid membrane? First, it is fluid. This allows cells to change shape, permitting growth and movement. The fluidity of the membrane is regulated by the types of phospholipids and the presence of cholesterol. Second, the phospholipid membrane is selectively permeable.
The ability of a molecule to pass through the membrane depends on its polarity and to some extent its size. Many nonpolar molecules such as oxygen, carbon dioxide, and small hydrocarbons can flow easily through cell membranes. This feature of membranes is very important because hemoglobin, the protein that carries oxygen in your blood, is contained within red blood cells. Oxygen must be able to freely cross the membrane so that hemoglobin can get fully loaded with oxygen in your lungs, and deliver it effectively to your tissues. Most polar substances are stopped by a cell membrane, except perhaps for small polar compounds like one-carbon alcohol, methanol, and water. Glucose is too large to pass through the membrane unassisted and a special transporter protein ferries it across. Certain types of diabetes are caused by misregulation of the glucose transporter. Charged ions, such as sodium (Na+) or potassium (K+) ions seldom go through a membrane; consequently, they also need special transporter molecules to pass through the membrane. The inability of Na+ and K+ to pass through the membrane allows the cell to regulate the concentrations of these ions on the inside or outside of the cell. The conduction of electrical signals in your nerves is based on the ability of cells to control Na+ and K+ levels.
Selectively permeable membranes allow cells to keep the chemistry of the cytoplasm different from that of the external environment. It also allows them to maintain chemically unique conditions inside their organelles.
The cell membrane must be a dynamic structure if the cell is to grow and respond to environmental changes. The fluidity of the membrane is demonstrated in the following animation.
To keep their membranes fluid across a range of temperatures, cells alter the composition of their membranes. Phospholipids with differing fatty acid tails have different levels of mobility in the membrane. The right ratio of saturated to unsaturated fatty acids keeps the membrane fluid at any temperature conducive to life. For example, winter wheat responds to decreasing temperatures by increasing the amount of unsaturated fatty acids in cell membranes. The unsaturated fatty acid tails keep membranes fluid because they are kinked and resistant to packing. In animal cells, the membrane is made up of mostly saturated fatty acids, so it is relatively stable (not too fluid). Cholesterol plays a key role in keeping animal cell membranes fluid across a range of temperatures. At high temperatures, cholesterol molecules interfere with phospholipid movement and reduce membrane fluidity. At low temperatures, cholesterol keeps the saturated fatty acid tails from packing and maintains adequate fluidity. Other sterols play a similar role in the cell membranes of plants, fungi, and even some prokaryotes. Fungi, for example, use ergosterol.
Structure of Cholesterol |
Cholesterol in the membrane maintains the correct
fluidity.
|
The cellular or plasma membrane is a lipid bilayer composed of phospholipids and associated proteins, specialized lipids, and carbohydrates. The composition of the plasma membrane varies according to the type and function of the cell.
Sterols are used by some cells to maintain proper membrane fluidity.
The extracellular surface of the cell membrane is decorated with carbohydrate groups attached to lipids and proteins. These short carbohydrates play a role in giving a cell its identity (i.e., distinguishing self from nonself) and are the distinguishing factor in human blood types.
Membranes also contain proteins, which carry out many of the functions of the membrane. Some functions of membrane proteins are:
In order to carry out their functions, some proteins need to be embedded in the membrane. These proteins are called integral proteins (“integral” because they are “integrated” into the membrane). For example, a membrane protein that transports materials into or out of the cell needs to completely span the membrane. On the other hand, a protein that binds the cell to another can simply be attached to the outer surface of the membrane. Proteins that are attached to the inner or outer surface of a membrane are called peripheral proteins.
Prokaryotes are an enormously diverse group of organisms. However, they can be organized into two primary groups: bacteria and archaea. While both bacteria and archaea are prokaryotes (both groups lack nuclei), the archaeans have some interesting structural characteristics in the cell membrane that have unique functional significance.
The cell membranes found in both eukaryotic and bacterial cells consist of a phospholipid bilayer that separates the internal environment of the cell from its external environment and regulates the materials that pass in and out of the cell. This enables the cell to maintain internal homeostasis, regardless of changing external conditions. You already know that cell membranes can be structurally different, to enable different functions. For example, some bacteria live in the frozen Arctic Circle. They are able to maintain membrane fluidity because of an increased concentration of unsaturated fatty acid tails found within the membrane phospholipids.
Archaeans are often found living in extreme environments such as high temperature, pH, or salinity, and as such, are often called “extremophiles” (loving extreme conditions). How can their cell membranes stay intact in such conditions? The cell membrane must have a unique structure to withstand these conditions.
You will explore characteristics of the archaean cell membrane in the following activities. First, you need to orient yourself to the phospholipid found in a bacterial cell. You might want to review the structure of phospholipids.
You figured out that the structure of the archaean phospholipid is significantly different from the structure of the bacterial phospholipid. These structural differences lead to functional differences that enable archaeans to survive in extreme environments. This is an excellent example of how structure determines function at the cellular level of organization.
The cell membrane determines what enters and exits the cell, and it is how the cell interacts with its environment. As you have already learned, the cell membrane is a phospholipid bilayer structure that provides a semipermeable barrier between the inside and the outside of the cell. In this module, you will explore the following three ways molecular substances enter and exit the cell:
This module will discuss how the cell membrane controls the transport of nutrients, ions and signals between the highly variable outside environment and the relatively well-defined interior of the cell.
Molecular substances enter and exit the cell so that: nutrients and waste can be exchanged; the cell can perform its function; and, when required, the cell can send appropriate signals to other cells. Transporting some substances may require the cell to exert energy and other molecules may cross through the membrane passively, requiring no use of energy by the cell. One way molecules passively move through the cell membrane is a process called simple diffusion.
In simple diffusion, both large and small molecules spontaneously move from areas of high concentration to areas of low concentration following random movements, referred to as Brownian motion. The classic example is the diffusion of a drop of ink placed in a beaker of water. The concentrated drop of color slowly disperses (diffuses) until at some point equilibrium is reached and the water in the beaker appears to have a uniform color. The following animation depicts this simple diffusion process. Add ink to the beaker and watch the diffusion process. After a period of time, how has the distribution of the ink changed in the beaker? Follow the yellow ink for some time. Does its behavior change as time passes?
The cell membrane provides a semipermeable barrier between the inside and the outside of the cell. The phospholipid bilayer structure of the membrane allows selected ions and organic molecules to pass through the plasma membrane and regulates the movement of molecular substances. This characteristic of the membrane, known as selective permeability, acts as a filter that allows only selected substances that are needed for the survival and functioning of the cell in and out. The movement of molecular substances across the membrane can occur passively without the cell exerting energy or it can occur through the cell's use of energy to transport substances. Selective permeability affects the energy required to transport substances in and out of the cell. This page will explore the conditions under which molecules can pass across the membrane passively through diffusion.
Molecules can be divided into four categories with regard to their ability to cross the plasma membrane. The first category is nonpolar molecules. These hydrophobic molecules can easily cross the membrane because they interact favorably with the nonpolar lipids. Note that these molecules can accumulate in the membrane because they interact so well with the lipids. The second category is small polar molecules. Although they don’t interact with the lipids, their small size allows them to pass through small temporary holes in the membrane. The third category is large polar molecules. These have difficulty crossing the membrane because of their size and poor interaction with the lipids. The last category is ionic compounds. Their charge interacts very unfavorably with the lipids, making it very difficult for them to cross the membrane.
The size, polarity, and charge of a substance will determine whether or not the substance can cross the cell membrane by diffusion. The cholesterol was an example of a lipid, and is highly soluble in the nonpolar environment of the lipid bilayer. You saw, in the animation above, the cholesterol freely passing into the hydrophobic environment of the membrane. Cholesterol distributes freely in the membrane and then some fraction will dissolve in the aqueous environment of the cytoplasm. Water, on the other hand, while polar, is small enough to cross the membrane at a slow rate. Note that specialized transport proteins in certain cell membranes can provide a channel for the water, greatly increasing its rate of crossing the membrane. The lipid bilayer is much less permeable to the ion, because of its charge and larger size. As a general rule, charged molecules are much less permeable to the lipid bilayer.
Cells must be able to move large polar and charged molecules across the lipid bilayer of the membrane in order to carry out life processes. To allow these molecules, which are not soluble in the lipid bilayer, to pass across the hydrophobic barrier, it is necessary to provide ports, channels, or holes through the membrane. The molecules will still move spontaneously down a concentration gradient from high to low concentration. These channels can either remain open at all times, allowing the molecules to move freely according to the concentration gradient, or they can be gated channels that open and close in response to the needs of the cell. In most cases these channels are very discriminatory and will only allow specific molecules to pass. The process of moving impermeable molecules across a membrane (down their concentration gradients) using channels or pores is referred to as facilitated diffusion. Because the molecules are moving down a concentration gradient, the process is driven by simple diffusion and does not require the input of additional energy from the cell. The following simulation depicts the facilitated diffusion of glucose across the membrane using the glucose permease transporter.
![]() |
|
In some cases it is necessary to move molecules against their concentration gradient. The eukaryotic cell has many compartments within it, each surrounded by a membrane. In most cases, the environment within the compartment is different from the environment in the cytoplasm. An example is the lysosome, an organelle whose function is to digest macromolecules delivered either from outside the cell or from other compartments within the cell. To carry out this function, the lysosome maintains a very low internal pH compared to that of the cytoplasm. Thus, there is a steep pH gradient across the lysosome's membrane. This contrasts with the equilibrium state, in which the concentration of hydrogen ions would be the same inside and outside the lysosome.
To decrease the pH inside the lysosome, the concentration of protons will need to be greater inside the lysosome than in the cytoplasm. To accomplish this, protons will need to move from a low concentration to a high concentration. This is a nonspontaneous process and requires the cell to do work to move the ions "uphill" against the concentration gradient. To do work, the cell must expend energy and actively move (pump) the ions. This process is referred to as active transport. The source of energy for this process in most biological systems is the hydrolysis of ATP.
The animation that follows illustrates an example of an active transporter that uses energy from the molecule ATP to transport sodium (Na+) ions out of the cell and potassium (K+) ions into the cell. Both ions are moved against their concentration gradients. The main stages of the process are:
Observe the active transport of sodium and potassium molecules using ATP.
Facilitated diffusion and active transport both require transport proteins that act as channels or ports in the membrane. Transport proteins are classified both by structure and by function. The structure (shape) of each channel helps determine what materials can pass through. Even within classes of transport proteins that carry out similar functions (e.g. ion channels) there are many different structures that can discriminate between specific substances. Dozens of different transporters swirl around within the fluid membranes that enclose your cells and organelles, contributing greatly to their "mosaic" quality.
Molecules can also be moved across the membrane in bulk. Additionally, larger items (such as entire cells) can also be taken into a given cell. These processes are types of endocytosis and exoctyosis and will be discussed next.
Facilitated diffusion and active transport are not the only ways that materials can enter or leave cells. Through the processes of endocytosis and exocytosis, materials can be taken up or ejected in bulk, without passing through the cell's plasma membrane.
In endocytosis, material is engulfed within an infolding of the plasma membrane and then brought into the cell within a cytoplasmic vesicle. To begin endocytosis, a particle encounters the cell surface and produces a dimple or pit in the membrane. The pit deepens, invaginates further, and finally pinches off to form a vesicle in the cytoplasm of the cell. Note that during the process the inside surface of the newly formed vesicle is the same as the exterior surface of the cell. Thus the integrity of the cytoplasm and the orientation of the plasma membrane are preserved. Once internalized, a new vesicle containing solid materials may fuse with a lysosome so that its solid contents are digested. The resulting molecules may be released to the cytoplasm for use within the cell.
There are two general forms of endocytosis: phagocytosis and pinocytosis. Phagocytosis is the uptake of large solid particles such as bacteria or cellular debris. Pinocytosis is the uptake of fluid and any small molecules dissolved within it. Cells are also capable of recognizing specific particles and engulfing them in a more targeted way, a process called receptor-mediated endocytosis. In this case, the particle first binds to a membrane protein receptor on the surface of the cell. Binding of the target particle induces the cell to engulf it.
Exocytosis is just the reverse of endocytosis. In exocytosis, an internal vesicle fuses with the plasma membrane and releases its contents to the outside. The balance of exocytosis and endocytosis preserves the size of the plasma membrane and keeps the cell's size constant. The following animation depicts endocytosis.
How are endocytosis and exocytosis important to everyday life? Immune cells protect animals by recognizing and destroying foreign objects such as bacteria. Disease-causing bacteria are recognized by proteins called receptors on the surface of the immune cell. The phagocytic immune cell will then engulf the bacterial cells (phagocytosis). The vesicle that contains these bacterial cells is called a phagosome ("phago" means "eating" and "-some" refers to "body"). The phagosome next fuses with lysosomes. Finally, the digested bacterial products are excreted through the process of exocytosis.
Cells continually encounter changes in their external environment. Most cells have a similar blend of solutes within them, but external fluids can vary dramatically, from pure water to brine or syrup. What will happen if there is a strong concentration gradient between a cell's interior and the fluid outside? As you know, molecules will tend to move down their concentration gradients until equilibrium is reached. You might think that solutes will flow into our out of the cell until the solute concentrations are equal across the membrane. However, not all molecules can pass through the cell membrane. The plasma membrane (lipid bilayer) is significantly less permeable to most solutes than it is to water. Therefore the WATER tends to flow in a way that establishes an equal concentration of solutes on either side of the membrane. The water flows down its own concentration gradient, with a net movement toward the region that has a higher concentration of solutes. This movement of water across a semipermeable membrane in response to an imbalance of solute is called osmosis.
Cells may find themselves in three different sorts of solutions. The terms isotonic, hypertonic, and hypotonic refer to the concentration of solutes outside the cell relative to the solute concentration inside the cell. In an isotonic solution, solutes and water are equally concentrated within and outside the cell. The cell is bathed in a solution with a solute concentration that is similar to its own cytoplasm. Many medical preparations (saline solutions for nasal sprays, eye drops, and intravenous drugs) are designed to be isotonic to our cells. A hypotonic solution has a low solute concentration and a high concentration of water compared to the cell's cytoplasm. Distilled (pure) water is the ultimate hypotonic solution. If a cell is placed in a hypotonic solution, it will tend to gain water. The solutes will "stay put" within the cell, but water molecules will diffuse such that their net flow is toward the area with a higher concentration of solutes. A hypertonic solution has a high solute concentration (lower water concentration) compared to the cell cytoplasm. Very salty or sugary solutions (brines or syrups) are hypertonic to living cells. If a cell is placed in such a solution, water tends to flow spontaneously out of the cell.
Cystic fibrosis is an inherited condition that affects the respiratory and digestive systems in children and adults. While there is no cure for the condition, in the 1950’s, people diagnosed with cystic fibrosis were lucky if they lived long enough to go to elementary school. Medical research has made an enormous difference in the life expectancy of people with cystic fibrosis, and today, they routinely live into their thirties and forties.
A diagnosis of cystic fibrosis (CF) is suspected if a patient has chronic weight loss accompanied by problems with the respiratory system. Most people with the condition are diagnosed by the time they are two years old. A century ago, cystic fibrosis was diagnosed when a doctor licked the forehead of a patient and found it to be unusually salty. While the medical equipment used to diagnose the condition today is much more professional, the same symptoms are identified. People with CF exhibit a variety of symptoms [1] that may include:
Cystic fibrosis is caused by a malfunctioning transport protein called Cystic Fibrosis Transmembrane Conductance Regulator (CFTR). CFTR is a channel that allows for the passage of chloride ions (Cl-) into and out of the cell. CFTR is located in the cells that line the respiratory tract, the pancreas, and sweat glands. It allows for the movement of Cl- into or out of a cell. CFTR is an active transporter and requires energy in the form of ATP to function.
People with cystic fibrosis have respiratory difficulties, usually caused by unusually thick mucus in the respiratory tubes in the lungs. In the lungs, CFTR transporters secrete chloride ions into the center of the respiratory tubes. This increases the concentration of solutes in the mucus lining the tubes, which in turn causes water to flow into the space inside the tubes (yellow). Under normal conditions, this process allows for a runny mucus to line the tubes.
Most of the treatments for cystic fibrosis involve making sure the lungs remain clear of the thick, sticky mucus. This can be done using drugs that thin the respiratory mucus, like Pulmozyme. A more recently discovered treatment involves the inhalation of hypertonic saline solution. This salty solution causes water to move from the cells and into the airways (via osmosis), making the mucus more watery and easier to cough up.
The cell is the fundamental unit of life. It is the first level of organization that presents all the properties of life. In "The Cell" unit, you learned about the different structures that make up a cell. Implicit in that knowledge was the notion of cells as continuously changing and dynamic entities; cells: grow, rebuild, repair, reproduce, and react to the external environment. All of these functions require energy.
All living organisms — from bacteria to trees to humans — are composed of molecules. To make a cell, many molecules must be combined, and it takes energy to build these complex structures. Covalent bonds hold these molecules together, and the chemical bonds contain energy associated with the force that holds two atoms together. As you know, some bonds are relatively weak and unstable. When they are replaced by stronger, more stable bonds, chemical potential energy is released that can be captured by the cell and used to do work. In cells, certain high-energy molecules contain many unstable bonds and are used for fuel. When these molecules are broken down into simpler molecules with more stable bonds, chemical potential energy is released. In this unit, you will learn how cells manage to capture this energy and use it to do work (e.g., power movement or build new macromolecules).
Energy and carbon flow through the biosphere in a tightly coupled manner. The coupling between energy and carbon is accomplished within an organism through the process known as metabolism. Metabolism is the sum of all of the metabolic pathways within an organism. A metabolic pathway is a series of chemical conversions, each of which are catalyzed by an enzyme (protein catalyst). The compounds that are internal to the pathway are referred to as intermediates in the pathway. A number of pathways, such as those related to energy production, are essential to life and are found in all cells. Although specialized cells may possess different pathways, there are common characteristics that apply to most pathways. These will be covered in the discussion on "Common Characteristics of Metabolic Pathways." In the Energy module, you will learn about ATP, the universal energy carrier in cells, and about how energy is transformed and transferred during metabolic reactions.
Metabolic pathways are either anabolic or catabolic. Anabolic pathways involve the use of energy and simple organic building blocks to create more complex molecules. Catabolic pathways convert complex molecules to simpler ones, releasing energy for use by the organism. In the Pathways and Regulation module, you will learn the differences between catabolic and anabolic pathways.
The principle source of energy for all organisms on earth is the sun. This energy is absorbed by photosynthetic organisms such as plants and algae and used to convert carbon dioxide (CO2) to glucose and oxygen during photosynthesis. The end result is that the energy from the sun has been stored in glucose, a small six-carbon carbohydrate. The oxygen and glucose are then available to other living organisms. Humans (and all other organisms, including plants) use glucose as food, and release the energy in glucose through a process called respiration. The figure below illustrates this process of photosynthesis and respiration. During this process, the carbon atoms in glucose combine with oxygen to produce CO2; the glucose is oxidized, releasing energy. The energy from the oxidation of glucose is stored in a small chemical called ATP (adenosine triphosphate) for immediate use by the cell. ATP is used for almost all of the energy needs of the cell, from the synthesis of other complex molecules to doing mechanical work, such as running. You will learn more about these two pathways in the section about “Anabolic and Catabolic Reactions.”
Many microorganisms, such as yeast, can utilize the energy stored in glucose in the absence of oxygen (anaerobic growth); the most common form of this process is fermentation. In the absence of oxygen, yeasts still produce ATP by the partial oxidation of glucose, releasing smaller amounts of CO2. The remaining carbon atoms in the glucose are converted to ethanol. The energy stored in ethanol can be used as fuel for our internal combustion engines. Humans also undergo anaerobic metabolism during vigorous exercise, but in humans, lactic acid is produced instead of ethanol. When sufficient oxygen becomes available, the carbon atoms in lactate are converted to CO2. Other organisms that perform fermentation produce a variety of other end products. Alternative pathways such as fermentation are addressed in a separate module.
All metabolic pathways are regulated such that they are only active when the products of the pathway are required by the cell. In this way, the organism is able to optimize the use of its carbon and energy resources.
The first part of the process of respiration is called glycolysis. It consists basically of the breakdown of glucose into a smaller compound called pyruvate. Pyruvate can then go on to either respiration or fermentation, depending on the availability of oxygen. Glycolysis will occur if the cell requires energy.
The glycolysis pathway is regulated to turn glycolysis on or off, depending on the energy needs of the cell. There would be two dire consequences if the energy that is stored in food were to be released immediately after consumption.
First, the organism would run out of energy quickly and would therefore have to consume food all the time, which isn’t possible for most organisms, especially mammals, which need to sleep.
Secondly, the heat that would be released by the production of energy would raise the temperature of the organism to intolerable levels. Misregulation of metabolic pathways can cause disease. For example, in diabetes, the metabolism of glucose is incorrectly regulated.
All living things need energy to run the processes required for life. Energy is defined as the capacity to do work. You can think of “work” as any change that won’t happen on its own. Work can involve movement, building larger molecules, increasing the concentration of chemicals, increasing the temperature of an object, making a sound, or even glowing (bioluminescence).
Energy is recognized by what it DOES: it makes “work” happen. It makes things change. Energy exists in many forms. Some biologically important forms of energy are light energy, electrical energy, sound energy, and thermal or heat energy. Kinetic energy is the energy of motion. Potential energy is stored energy. Energy can change forms and often transforms from potential to kinetic energy, and back. In biology, a key form of potential energy is chemical energy: the energy stored in chemical bonds. It can be released when certain bonds are broken. Chemical energy is released when wood is burned: the bonds in the wood are broken, releasing light and heat energy.
The first example below shows a single-celled organism — an amoeba — using energy to move. The second example illustrates the chemical reaction resulting in urea, a molecule found in urine. The third example shows bioluminescence — the emission of light by living organisms; in this case, the aurelia aurita jelly fish.
Example 1: Cell movement. Double-click on the above image to observe an
amoeba using energy to move.
|
|
Energy can be converted from one form to another. In the image below, the chemical energy in the wood is converted to light and heat. After the fire dies, the heat and light have dissipated and are no longer available for living organisms. Energy in our biosphere is a one-way system: It comes from the sun, is captured by living organisms, and eventually is released back out as heat.
Energy cannot be created or destroyed. Instead, energy is transformed from one form to another. Imagine that you climb a mountain trail ending in a steep cliff. Your body’s movement (kinetic energy) is fueled by chemical energy from your food. As you move uphill, your body’s chemical potential energy decreases (you break down molecules and use up stored food energy). At the same time, your body’s gravitational potential energy increases. This kind of potential energy can be released by letting an object drop from a height. A fall from the top of the cliff would do a great deal of “work,” so watch your step!
Where does chemical energy in food come from? The sun provides solar energy, the energy source for most living things. Plants and other photosynthetic organisms can use solar energy to build sugar molecules from simpler raw materials (carbon dioxide and water). A sugar molecule stores chemical energy in the bonds that join its atoms. The energy stored in sugar can be released and used to build other types of biological molecules (proteins, lipids, etc.), which also contain stored chemical energy in their complex structures.
To us, it seems that the sun has burned forever and will always keep shining. However, the sun is fueled by nuclear energy generated by atoms colliding in its incredibly hot core. Eventually, the supply of fuel in the sun will be depleted and our sun’s long life will end. This illustrates an inescapable fact about energy: it tends to lose quality over time. It dissipates, spreading out in space, and loses its intensity.
Heat is the energy possessed by any substance because of the random jiggling of its molecules and atoms. It is a low form of energy, disorderly, and spreads out rapidly through most materials. Every kind of energy conversion generates at least some heat energy as a byproduct. For example, the breakdown and use of chemical energy in food also releases heat. Heat can be helpful in keeping our bodies warm, but for the most part heat energy is a loss: an energy waste that cannot be recovered. No organism can capture heat and use it to do work or to make food. Instead, heat seeps out of organisms and eventually into space, leaving the biosphere forever.
Adenosine triphosphate, or ATP, is an organic molecule that acts as the direct energy source for almost all cellular activities. ATP consists of a single adenosine molecule (composed of an adenine bound to a ribose sugar) linked to three phosphate ions.
ATP is a relatively unstable molecule. Consequently it is never used for the long-term storage of energy in the cell. This job goes to other more stable compounds, like fats and sugars. ATP is specialized for direct and rapid transfers of energy. It is because of this instability that ATP is such a good energy currency for the cell (it’s easy to “spend” this energy). ATP is like cash, while long-term energy storage molecules (such as fats and sugars) are more like a savings account.
ATP is an important molecule in biological systems because the chemical bond between the last two phosphates is unstable, and reactions that involve the removal of the terminal (last) phosphate release energy. This bond can be broken in a reaction involving water. This reaction is called a hydrolysis reaction ("hydro" means "water" and "lysis" means "separation"; together they mean "using water to separate"). The products of this reaction are inorganic phosphate (P) and adenosine diphosphate (ADP), which has only two phosphate groups. Taken together, ADP and P are more stable than ATP. The net result of ATP hydrolysis is the release of energy. This process is illustrated in the animation below.
ATP, the molecule with the higher energy content, can react to form ADP + P. This chemical reaction can also be reversed. ADP + P can combine to form ATP, a reaction referred to as a condensation reaction since two molecules are condensed into one. We know the first reaction (ATP → ADP + P) releases energy. Therefore, the second reaction (ADP + P → ATP) requires energy. The opposing reactions are like using and charging a rechargeable battery. When you use the energy stored in the rechargeable battery, it is like using up ATP (ATP hydrolysis). When you recharge the battery, you must put energy in (by plugging it into the electrical socket). This is like ATP formation, which is usually fueled by chemical energy extracted from sugar or fats.
Virtually all the activities of the cell are powered by the energy released when ATP loses its terminal phosphate. For example, ATP hydrolysis provides the energy needed to form complex molecules from simpler ones (synthesis). ATP can also transfer energy by making proteins change shape when ATP is converted to ADP. When proteins change shape, their microscopic movements can add up to macroscopic motion, such as occurs when entire muscles contract. Similarly, ATP can transfer energy to specialized transport proteins embedded in the cell membrane and enable materials to be pumped in or out of the cell, against their concentration gradients (active transport).
The energy difference between ATP and ADP is largely due to the interaction between the negatively charged phosphate groups in each molecule. You may remember that unlike charges (+ and -) attract each other, while like charges (+ and +; or - and -) repel each other. Energy is therefore required to move two negative charges close to each other, which is what happens when ATP is made from ADP and phosphate. This is similar to the way in which energy is required to bring the two north poles of a bar magnet together.
Examine the following figure, which illustrates the change in energy found within a chemical system as a phosphate ion is slowly brought closer to an ADP molecule. As the negatively charged phosphate gets closer to the negatively charged ADP molecule, more energy is required to overcome the particles' natural tendency to repel each other.
Most life on Earth is fueled by solar energy through the process of photosynthesis. Photosynthetic organisms use sunlight to make organic molecules from simple raw materials (carbon dioxide and water). Cellular respiration and other processes break down organic molecules and use the energy released to build ATP molecules. ATP then transfers energy to drive many important processes in cells including movement, transport, and synthesis. In this way, the energy from the sun is transformed to chemical energy in molecules such as glucose and ATP. Eventually energy is released in forms like motion, heat, sound, or light.
Metabolism is the sum of all chemical activity within a living organism. Reactions that release energy are called exergonic (“exer” refers to “out” and “gonic” refers to “energy,” so these are “energy out” reactions). Reactions that require energy are called endergonic (“energy in”). Exergonic reactions can be used to power endergonic reactions. For example, building ATP from ADP and phosphate requires energy (it is an endergonic reaction). Exergonic reactions can provide the necessary energy to build ATP.
Catabolic reactions (also called “catabolism”) break down larger, more complex molecules into smaller molecules and release energy in the process. The smaller end products of a catabolic reaction may be released as waste or they may be fed into other reactions. The energy that is released by catabolic reactions can be captured and used in many ways. Some of the energy is released as heat and increases the temperature of the cell. Sometimes the energy is stored in the chemical bonds of another molecule. And sometimes it can be used to do work, such as movement of cellular machinery to power the active transport of materials across cell membranes. Catabolic reactions are central to biological processes such as cellular respiration and the digestion of food molecules.
Cellular respiration is catabolic because it takes glucose (a complex molecule with 6 carbons) and reduces it to carbon dioxide (each with only one carbon) and water (oxygen joined to two hydrogens). Some of the energy released by this process is captured in ATP, which can be stored for later use.
Digestion of protein is also a catabolic reaction. It takes a very large protein molecule and breaks it down into several smaller polypeptides, then dismantles the polypeptides to yield individual amino acids.
Anabolic reactions use energy to build more complex molecules from relatively simple raw materials. “Anabolic” and “catabolic” sound similar but are opposites. To remember the difference, it may help to think about how “anabolic steroids” promote the buildup of muscle mass. All of the complex molecules of life — carbohydrates, lipids, proteins, nucleic acids — are generated by anabolic reactions. Anabolic reactions are central to processes like photosynthesis, protein synthesis, and DNA replication.
At this point, you may wonder why the cell has to break down complex molecules in order to build them up again. Different cells use and require different biomolecules for their specialized functions. For example, carbohydrates are good sources of energy. In plants, carbohydrates are usually stored as starch. In animals, they are stored as glycogen. Both are polymers of glucose, but they have different structures, and an animal cell cannot readily store or access starch. When an athlete “carbo-loads” the evening before a race, starches from foods like pasta are digested to simple sugars in a catabolic reaction. They are then rebuilt as glycogen (in an anabolic reaction) in the muscles and liver. That way, there is some extra energy stored and easily accessible for the muscle cells to use for the big race. Biomolecules are broken down to raw materials, then rebuilt in a different form to meet the needs of the organism. One molecule that is unique to your cells is DNA. No other organism (except an identical twin) has the same DNA as you, so your cells need to build it anew in order to pass it on to the next generation
Photosynthesis is an anabolic process you’ll learn more about in this unit. Protein synthesis is an amazing feat of anabolism. Protein chains (polypeptides) are produced by linking together relatively simple amino acid monomers. In one polypeptide, thousands of amino acids may be strung together, and several different polypeptides may fold and bind to each other to form a large and complex protein molecule.
Life requires energy to accomplish its anabolic reactions. This energy cannot be “made” from nothing, because energy cannot be created or destroyed. But energy can be transferred and transformed. Energy that is available to the cell must somehow be packaged in useful form and delivered where it is needed. This problem is solved through chemical transfers of energy from catabolic reactions to energy carriers and then from the energy carriers to the substrates that are joined in anabolic reactions.
Life’s universal energy carrier is adenosine triphosphate (ATP). ATP acts as a shuttle that transfers energy from catabolic reactions that release energy from complex molecules (such as cellular respiration) to the anabolic reactions that consume energy (such as protein synthesis). The energy released from catabolic reactions can be stored for later use through the formation of ATP by combining ADP plus a phosphate group. Then, when energy is required for anabolic reactions, ATP can be broken down into ADP and a phosphate group, releasing energy that was stored in the phosphate bond.
Plants, algae, and some bacteria can use light energy to turn carbon dioxide and water into simple sugars through the process of photosynthesis. These sugars can act as “building blocks” for the synthesis of other organic molecules, or they can store energy for later use. While the process of photosynthesis stores chemical energy for later use, most photosynthetic organisms access the stored energy the same way consumers do: they break down the sugars to build ATP using the process of cellular respiration, which occurs in subcellular organelles called mitochondria.
Photosynthetic producers have evolved an effective system for absorbing light and converting it into chemical energy. This biochemical solar energy system begins with a diverse array of (mostly green) pigments that are embedded in membranes. In plants and algae, these pigments are contained within specialized subcellular organelles called chloroplasts.
You may want to review the Cell Structures page.
The main purpose of photosynthesis is to convert energy from sunlight and store it as bond energy within sugars. Photosynthesis uses pigments (such as chlorophyll) in the chloroplasts to absorb light energy. This energy is used to produce sugars from carbon dioxide, while oxygen is given off as a byproduct. You will examine the function of pigments in more detail later.
Some photosynthetic organisms (such as most algae and cyanobacteria) live in water. Their cells are equipped with pigments to absorb light, and they absorb water and other nutrients directly across the cell membrane. They are faced with a challenge involving the attenuation (loss) of light energy with increasing depth. As a result, many algae and cyanobacteria have adaptations that keep them floating near the water’s surface.
Land plants face a different set of challenges. Plant leaves are the main location of photosynthesis. They are often large, flat structures that are adapted for maximum light absorption. In many plants, leaves also are held high above the ground by tall stems. Growing tall is an adaptation that helps a plant get high-intensity sunlight, because short plants are more likely to be shaded by taller ones.
Land plants also face the risk of drying out. Plants have evolved traits that protect their above-ground portions from excess water loss. For example, plant leaves are sealed with a layer of waxy material called cuticle. Protected in this way, leaves cannot efficiently absorb water when it rains. Instead, land plants absorb water from the ground (through their roots).
Plants must then transport water upward to leaves and other tissues. This is quite a challenge for a tall plant. Plants have no moving parts that could power a pump, so their circulatory systems must work on different principles. Plants use specialized cells to move water. These cells are essentially long, thin tubes that transport water from the roots all the way up to hollow cavities in the leaves. Evaporation of water from leaves provides negative pressure that sucks water up the tubes. The thick cell wall that allows the tube cells to function without collapsing also provides the plant with structural stability. “Wood” is mainly the structural material that remains after these cells have died. Ultimately, an adaptation for moving water helps woody plants grow even taller than their less sturdy cousins, further enhancing their access to light.
The adaptations of plants have long been a subject of biological study. One of the first questions asked by early scientists was, “How do plants get their resources for growth?”
We know that plants require light. But “light” is a complex form of energy. Visible light is just a small part of a much broader spectrum including many forms of electromagnetic radiation. In many ways, light behaves as a wave. Different colors of light have different wavelengths. In a given period of time, short wavelengths deliver more energy to a surface than longer wavelengths. Thus blue light is higher in energy than red light.
Are all wavelengths of light equally useful for photosynthesis? In 1883, a scientist named George Engelmann designed an experiment to answer this question. He shined a light through a prism to separate the different wavelengths so that they landed on different zones along the length of filaments of Spirogyra, a type of green algae. Examine the following image depicting Engelmann’s results.
Pigments are colored substances that absorb light energy. Why does a pigment look red, green, purple, or blue? A given pigment absorbs only a limited set of light wavelengths. For example, a red pigment absorbs green and blue wavelengths. Red light, by contrast, is NOT absorbed: it is either reflected or transmitted through a solution containing the pigment.
What happens to light energy that is absorbed by a pigment? In keeping with the law of conservation of energy, this energy does not disappear. Instead, it excites electrons in the pigment, boosting them to higher energy levels. In plants, the excited electrons from some pigments are transferred to other molecules, where they start the transformation of light energy into chemical potential energy.
If excited electrons are not removed from a pigment molecule, they will eventually “fall” back to their starting point. At this point, the pigment will release energy. In some pigments, a portion of the released energy is emitted as light. If you expose these fluorescent pigments to light and then place them in a dark place, they will glow. All pigments release at least some of the energy they absorbed from light as heat.
O2 is a product of photosynthesis. This product is essential for many living organisms, including humans and any other organisms that go through cellular respiration. When photosynthesis became widespread in the early Earth’s oceans, it generated excess oxygen gas that changed the Earth’s atmosphere. This increase in atmospheric oxygen concentration eventually changed the face of our planet as organisms that couldn’t adapt died off and new organisms evolved.
As we have emphasized thus far, cells juggle energy through opposing reactions. Photosynthesis is an anabolic pathway fueled by solar energy that allows the synthesis of organic molecules from simple raw materials. Catabolic pathways reverse this process. They break molecules down into simpler components. The centrally important catabolic pathway that is complementary to photosynthesis is cellular respiration. Although there are may other catabolic reactions occurring in any cell, cellular respiration supplies the vast majority of ATP molecules required for cellular work in aerobic organisms.
If you compare the summary reactions of photosynthesis and cellular respiration, you will see that cellular respiration is photosynthesis “running in reverse”:
sugar + O2 → CO2 + H20 + energy |
Cellular respiration is a key pathway in the metabolism of all aerobic organisms. On the macro scale, respiration refers to breathing: taking in oxygen and removing CO2. But ultimately, the reason we need to breathe is to provide the oxygen needed to carry out cellular respiration in our cells and to remove the carbon dioxide produced as a byproduct.
Three important things happen during cellular respiration:
In eukaryotes, the breakdown of sugar begins in the cytoplasm (cytosol), while the consumption of oxygen and the generation of CO2 and ATP synthesis happen inside specialized membrane-bound organelles called mitochondria.
Energy is released in small portions, so that some of the released energy can be captured in chemical bonds, ultimately as ATP. The process can be regulated at different steps, allowing cells more control over the process.
Cellular respiration is a complex, multistep process. Dozens of different enzymes and intermediates are involved. Initial steps are accomplished in the cytoplasm, and additional steps occur in the mitochondria. At various points along the way, intermediates may be removed and used to synthesize fats and other molecules. Similarly, different types of food molecules may “feed in” to cell respiration at specific junctions in the process.
Even though glucose is the starting substance used in cellular respiration, organisms do not consume pure glucose as an energy source. Instead, many different kinds of fuel molecules must be partly broken down and then fed into various stages of the cellular respiration pathway. For example, complex carbohydrates are readily converted to glucose or similar sugars. Fats and proteins can also be used in cellular respiration, but they must be modified before they can feed into the process.
As you may recall, photosynthesis occurs in some bacteria; in eukaryotes it occurs in cells that contain chloroplasts. Such cells are found in algae and in the leaves and stems of plants. Photosynthetic organisms use light energy and simple building blocks (carbon dioxide and water) to make their own food. Cellular respiration is very widespread. It is completed in mitochondria, bacteria-like organelles that are found in protists, fungi, plants, and animals. Although the details of the two pathways are different, the overall reaction of cellular respiration is photosynthesis running in reverse. Cellular respiration extracts the stored energy from food molecules, uses it to “charge up” ATP, and releases carbon dioxide and water as wastes.
In the last two sections, the topics of photosynthesis and cellular respiration have been introduced and discussed as independent processes. The focus of this module is on how the two processes are part of one larger cycle. There is a yin-yang relationship between the two processes — the product of one process is the starting material for the other.
In the image below, you see a summary of photosynthesis. But photosynthesis does not occur in a vacuum. In fact, it is inevitably PAIRED WITH cellular respiration in most producers. They take most of the glucose they have produced and break it right down again in their own cells.
A full representation joins photosynthesis to cellular respiration in a cycle. The products of one reaction are the starting materials for the other.
This is how the cycling between photosynthesis and cellular respiration occurs: in photosynthesis, carbon dioxide and water, in the presence of light energy, are converted to make glucose and oxygen; while in cellular respiration, the products of photosynthesis (glucose and oxygen) are metabolized to make energy in the form of ATP and heat, releasing carbon dioxide and water. Because each process starts where the other ends, they form a cycle.
The photosynthesis / cellular respiration cycle isn’t restricted to the cells of an individual plant. To the contrary, it is a global cycle that makes all life on Earth interdependent. A growing photosynthetic producer carries out more photosynthesis than cellular respiration. Thus, it releases excess oxygen into the atmosphere and stores carbohydrates and other molecules in its body as growth. These materials are essential to virtually all other life forms on Earth. This interdependence is depicted in the diagram below.
The cycling that occurs between photosynthesis and cellular respiration is vital to the health of planet Earth. If there was no way for the carbon dioxide produced through cellular respiration to be utilized, respiring organisms (like humans, dogs, and even grass) would soon die of asphyxiation. Additionally, photosynthetic organisms are the base of almost every food chain on the planet, so without these organisms, mass starvation would ensue. Luckily, this planet is full of photosynthetic organisms. All of the breathable oxygen on Earth comes from photosynthesis. Just over half of the oxygen is produced by phytoplankton — drifting photosynthetic algae and bacteria — in the oceans. The rest comes from plants (trees, grass, etc.) on the land. Without this vital connection between photosynthesis and cellular respiration, life as we know it would cease to exist.
An understanding of this balance is important in the modern world, with global warming in the headlines. Since the late 1700's, the amount of carbon dioxide in the Earth’s atmosphere has been rising. This indicates that the natural recycling of carbon has been upset. The main source of excess carbon dioxide in our atmosphere is the burning of fossil fuels, which were created by photosynthesis millions of years ago. Meanwhile, deforestation removes photosynthetic organisms (trees) from the surface of the planet. Taken together, these human activities have changed Earth’s atmosphere. The carbon dioxide concentration of Earth’s air has increased by more than 35 percent since the 1700's. Such changes are very likely changing Earth’s climate, and pollution and deforestation also have other detrimental effects on environmental health. We will return to explore these topics further in the Ecology unit. You will need to use your understanding of photosynthesis and respiration to fully grasp these issues.
Metabolic pathways constitute a complex network of chemical reactions, with some pathways functioning like central “freeways” to which many other pathways connect. Pathways related to the harvesting of energy (chemical or solar) and biosynthesis of the major cellular components are conserved in many types of cells and organisms.
Many organisms use energy harvesting pathways other than photosynthesis and cellular respiration. Bacteria, in particular, have an astonishing array of metabolic options. The variety of metabolic options in the bacterial domain helps scientists classify evolutionary relationships between these organisms.
Many of these metabolic options do not require oxygen and are called anaerobic (“an-” means without and “aerobic” refers to oxygen). Fermentation breaks down organic molecules and stores energy as ATP. Although fermentation is the only ATP-producing pathway for some microbes, it is also commonly used by many organisms as a special-purpose pathway. For example, yeasts that decompose fruit can switch between aerobic cellular respiration (used when oxygen is abundant) to a form of fermentation that produces ethanol (used when oxygen is depleted). This allows yeast to continue to produce ATP even in rotten, oxygen-poor fruit.
Some human tissues also use fermentation under certain conditions. Sprinters and weightlifters have well-developed fast-twitch muscles. Fast-twitch fibers have fewer mitochondria and are not able to renew their supply of oxygen rapidly. Therefore, they have relatively poor endurance. However, fast-twitch fibers excel at producing strong contractions very quickly. They do this using the ATP created through a form of fermentation. This pathway does not break sugar all the way down to CO2, but stops at an intermediate (lactic acid).
Oxygen is not used in fermentation, and although some CO2 may be produced, glucose is not fully broken down. Instead, fermentation ends with complex products (lactate, ethanol) that still contain quite a bit of chemical potential energy. The breakdown process is incomplete, and relatively little energy is available to drive the production of ATP. In addition, the byproducts of fermentation can be toxic. For example, a buildup of lactic acid may cause stiffness in an overworked muscle.
Fermentation is a very common metabolic pathway in microorganisms such as bacteria and yeast, resulting in alcoholic beverages, fermented milk products, and bread. In fermentation of alcoholic beverages, yeast is added to sugary juices or malt syrups. The container is sealed to exclude oxygen. Under these conditions, the yeast switches to fermentation and generates ethanol, the intoxicating “active ingredient” of alcoholic beverages. Eventually, however, ethanol concentrations reach a point where the yeasts themselves are inhibited and fermentation grinds to a halt. In this way, wine and beer are naturally limited in their alcohol content. High-alcohol liquors can be produced only by distillation, a chemical process that concentrates the alcohol.
*Some bacteria can carry out forms of respiration that do not require oxygen. | |||||||||||||||
|
Cellular respiration and fermentation both break down organic molecules to build ATP. However, cellular respiration is aerobic while fermentation is anaerobic; cellular respiration produces more ATP. The ATP yield in the table originates from one glucose molecule.
In the previous modules, you learned about metabolic pathways in cells — those “freeways” that shuttle energy and carbon in different directions. Metabolic pathways can be either anabolic or catabolic. Anabolic pathways build things up and catabolic pathways break them back down again, sometimes as a source of energy, other times to obtain building materials. Among the main pathways of the cell are photosynthesis and cellular respiration, although there are a variety of alternative pathways such as fermentation.
Metabolic pathways consist of an ordered set of chemical reactions that are catalyzed by enzymes. At the heart of metabolism are enzymes. These proteins embrace substrates, quickly generating products. One or more substrates bind to an enzyme’s active site. As a result, the activation energy is reduced and the reaction proceeds much more quickly than it otherwise would. Without enzymes, metabolism would grind to a halt. Furthermore, the structural characteristics of enzymes allow the regulation of their function, thus allowing the regulation of the metabolic pathways they participate in. A huge diversity of enzymes is required to run the metabolic reactions necessary to maintain even a single cell.
Regardless of their purpose, all metabolic pathways share four important features:
Many pathways are universal among living organisms. When you look at life, you may see a great deal of diversity. Those who study metabolism, by contrast, see a huge degree of overlap among the cells of all living things. Many of the fundamental enzymes that keep our cells working are found in only slightly different forms in creatures as varied as zebrafish, fruit flies, yeast, and nematode worms. This is partly why studies on model organisms can teach us so much about our own biology.
Metabolic pathways (such as cellular respiration) are nearly universal. These organisms all use similar pathways.
![]() |
![]() |
![]() |
Many metabolic pathways are reversible, which means that certain chemical reactions can go “either way,” depending on the needs of the cell. For example, some of the reactions that are part of cellular respiration (which is a catabolic pathway) can become anabolic when there is a surplus of energy. A small number of steps utilize different enzymes in the forward versus the reverse direction. These enzymes are regulated in a coordinated fashion such that a pathway operates in only direction at time.
At the heart of metabolism are enzymes. These proteins embrace substrates, quickly generating products. One or more substrates bind to an enzyme’s active site. As a result, activation energy is reduced and the reaction proceeds much more quickly than it otherwise would. Without enzymes, metabolism would grind to a halt. Furthermore, the structural characteristics of enzymes allow the regulation of their function, thus allowing the regulation of the metabolic pathways they participate in. You may want to review the functions of enzyme proteins.
Metabolic pathways are regulated in order to maintain maximum efficiency or in response to environmental changes.
There are two major methods of pathway regulation. The first method involves changing the amount of the enzymes in the pathway. Since the reactions are catalyzed by the enzymes, if the enzyme is not present, the pathway is effectively shut off. Enzyme levels are usually altered by controlling the amount of mRNA produced from the gene that encodes the enzyme. You will explore how enzyme levels are regulated in the section on transcription.
Regulation of pathways by altering the amount of enzymes in the pathway is usually too slow to meet the second-by-second regulation needs of the organism. Consequently, many pathways are regulated by increasing or decreasing the activity of an enzyme due to the binding of a small regulatory compound. Typically, only a few steps in the pathway are regulated, and the regulated steps typically occur near the beginning of the pathway to prevent the accumulation of intermediates, which are seldom of use to the cell. Think of an assembly line: the speed of production will change depending on how much of the starting material is available, how quickly each of the workers completes the task, and how much of the final product is needed. Similarly, pathways are influenced by the availability of the starting materials, the activity of each enzyme in the pathway, and the concentration of product already present in the cell.
Let’s look at an example of a metabolic pathway and explore one mechanism that regulates it. You may remember that cellular respiration is an essential catabolic pathway. Cellular respiration starts with a pathway called glycolysis, which degrades three-carbon glucose into a three-carbon molecule called pyruvate. Glucose can come from ingested carbohydrates, or from degradation of glycogen, a polysaccharide. When there is an excess of carbohydrates, they are stored as glycogen in an anabolic pathway.
Six general mechanisms can regulate the rate of each step in a pathway. The mechanisms differ in how rapidly they can respond to changes in the environment. Each of them is discussed below, with the more rapid forms of regulation at the top of the list.
In summary, metabolic pathways are dynamic processes, continuously changing in response to environmental changes and cellular (or organismal) needs, while avoiding waste of resources. How much substrate or enzyme is available will determine the rate of reactions. Moreover, extreme “fine-tuning” of metabolism is possible through changes in enzyme activity due to the binding of inhibitors. This is a prime example of the relationship between structure and function.
To this point, you have learned much about how pathways work. They proceed as a series of steps catalyzed by enzymes. They are regulated by factors that modify enzymes or alter the expression of genes.
In living organisms, metabolic processes are tightly regulated under normal conditions. The cell, and the organism as a whole (e.g., the human body), tend to respond to external influences in ways that maintain a relatively stable internal environment. This ability or tendency is called homeostasis and depends on a complex set of stabilizing responses.
Cells live in dynamic environments. Therefore, it takes work to keep conditions stable in a cell. Precise regulation enables the cell (or body) to establish consistent internal conditions. Factors like pH, ion concentrations, and levels of various substrates and products are held within narrow bounds. This unique environment allows a diverse set of reactions to take place. Many of these chemical reactions are essential for life and all of them help maintain overall health.
You had to work at staying upright because your body was in motion. In the cell or inside a body, countless dynamic processes are running at once. Many of them could easily disturb internal chemistry and throw things out of whack. So could changes in the external environment. As a result, many levels of regulation are constantly at work within your body, keeping internal processes balanced and compensating for external changes.
These concepts are not just academic, textbook issues. To the contrary! Many of today’s most challenging health problems center on the regulation of metabolism. The loss of proper regulation can result in a number of diseases or disorders.
Substances that inhibit enzymes in a central pathway can have deadly effects for the cell or organism. For instance, the potent poisons cyanide and arsenic are inhibitors of cellular respiration. Lack of regulation, or dysregulation, of metabolic pathways can lead to serious diseases. Examples include phenylketonuria (PKU), Tay-Sachs disease, galactosemia, and maple syrup urine disease. Each of these is an inherited condition in which an altered gene produces an enzyme that does not function normally. As a result, metabolism is disrupted and health problems occur.
On the other hand, drugs often target metabolic pathways to correct diseases. For instance, angiotensin-converting enzyme (ACE) inhibitors are drugs that are commonly used to control high blood pressure. These drugs inhibit ACE; the result is reduced production of angiotensin, a molecule that increases blood pressure.
Many humans are able to maintain a stable adult body weight over a large number of years. Although it may seem perfectly normal, this is an amazing feat of metabolic regulation. As you eat, foods are broken down and metabolized. But the energy stored in the food’s chemical bonds cannot be destroyed. Simply put, if more calories are consumed than are required for growth, activity, and maintenance, the excess energy is stored within the body. This energy is stored when enzymes catalyze the formation of new chemical bonds to create specific kinds of macromolecules (i.e., glycogen, fat, and protein), which tends to increase overall body mass. On the other hand, if not enough food calories are consumed, the body will begin to break down its own tissues as a source of energy, and will decline in mass as a result.
Note that when “calories” are being discussed in reference to food consumption, the word "calorie" actually refers to a kilocalorie. One kilocalorie is the amount of energy required to raise the temperature of one kilogram of water by one degree Celsius. It is the unit typically used when discussing human nutrition and energy balance. For your reference, a typical adult male requires about 2,500 calories per day; a female of average size might require 2,000 calories daily. The body’s calorie requirement declines with age, and depends on many factors, including body size, sex, and physical activity.
As you have seen, the balance between caloric intake and output must be held within narrow limits to maintain a steady body weight. Yet some humans do this without conscious effort. How does this remarkable balancing act work?
First, consider what happens over the course of a typical day. After a meal, extra calories are initially stored in the liver and muscles as glycogen (a complex carbohydrate). Glycogen is broken down and used to supply glucose for intense exercise, and to provide glucose while we sleep. Breakfast provides fuel to rebuild some of the body’s glycogen stores that were depleted overnight. Thus, over short time scales, glycogen storage is an efficient way to provide glucose when it’s needed between meals.
Over longer time periods, body fat is used to store excess energy. When glycogen stores are full, the body switches over to storing the excess energy as fat. Remember that energy is stored within chemical bonds, and so it is the chemical bonds within glycogen and fat molecules that contain energy. Fat is an efficient way for the body to store extra calories, because it is able to pack more calories per unit mass when compared with carbohydrates or proteins (see Figure 8.2 below). You can think of body fat as a long-term, highly compacted energy reserve.
How are caloric intake and energy output balanced? Your levels of hunger and overall metabolic rate change in response to your nutritional status. When your stomach is empty, a hormone called ghrelin is secreted. This acts on receptors in the brain to stimulate feelings of hunger. When the stomach is full, ghrelin secretion stops and so do your cravings for food.
If a person takes in less calories than required for maintenance and activity, glycogen reserves are broken down first. This leads to rapid weight loss early on, because each gram of glycogen associates with up to four grams of water. As glycogen is broken down, water is released in urine. Most of the initial weight loss is actually “water weight.”
Later weight loss comes more slowly, because body fat stores a great deal of energy and is associated with much less water. Importantly, body fat is not just a storage bin. It also has a role as an endocrine (hormone-secreting) organ. For example, body fat secretes a hormone called leptin. This acts on target cells in the brain to produce feelings of “fullness.” Loss of body fat leads to decreases in leptin and increasing feelings of hunger. Second, loss of weight also has effects on a person’s energy output. As a person loses weight, total metabolism goes down simply because the body is getting smaller. More significantly, the per-pound rate of metabolism is turned down to conserve energy and resist what the body perceives as “starvation.” A person in this state will feel sluggish and reluctant to exercise. Taken together, these and other mechanisms tend to automatically adjust our hunger and activity to maintain our weight.
Obesity is a condition in which body fat is excessive and interferes with normal function or impairs health. Obesity can make many physical activities difficult or impossible. It is associated with many increased health risks; type II diabetes (see below), hypertension, and heart disease are three serious conditions that are clearly more common among obese people than their normal-weight peers. Obese people also suffer socially and professionally because of weight-related bias and discrimination.
Why don’t the normal regulatory mechanisms work to limit weight in people who become obese? This is a question that has fueled considerable research and debate. Findings suggest that there are many contributing factors. For example, many (but not all) obese individuals have very high circulating levels of leptin in keeping with their high body fat reserves. Yet the target cells in their brains do not respond to leptin in the normal way; they do not perceive satisfaction or fullness.
Obesity may result from a mismatch between our evolutionary adaptations and the realities of the modern world. We expend less calories each day on physical tasks than did previous generations of humans. More than 100 years ago, the average person may have expended 30 percent to 50 percent more calories each day than a typical modern American. In addition, today we have continuous access to an unprecedented array of calorie-rich foods. In times past, when food was scarce and could not be preserved easily, it was adaptive to feast on rich foods when they were available. Those who did so could build up fat and glycogen reserves, and thus were more likely to survive famines. Today, in a changed environment, the same behaviors predispose us to obesity.
Our society’s main effort to overcome obesity has come in the form of dieting. Diet fads constantly come and go — Atkins, South Beach, Acai Berry, the Hollywood diet, and pills, drinks, or even machines that purportedly melt away fat. In fact, the total sales of weight loss services and products were estimated at over $55 billion in 2009. Unfortunately, all of this effort and expense does not appear to do much good. Surveys, clinical trials, and literature reviews have repeatedly turned up a depressing pattern: about 80 percent of dieters regain lost weight within one to five years — and very often shoot past their prediet weight with each cycle.
If crash dieting is a bad idea, and we are “hardwired” to overeat, is there any hope for solving the obesity problem? Yes, there are many efforts in this direction.
A variety of information on the scientific and social aspects of obesity can be found at http://www.cdc.gov/obesity.
Information on diabetes can be found at http://www.cdc.gov/diabetes/.
More information and many links can be found through federal websites including the obesity pages of the Centers for Disease Control and the National Heart Lung and Blood Institute.
In this unit [1] , you will learn about another major cell process — cell division, or cell reproduction. The word "reproduction" is used in many different scientific and nonscientific settings. Regardless of the setting, though, reproduction has the same basic meaning. Reproduction is defined as "the act of making a copy of the original." With this in mind, the word reproduction can be applied to many different situations. Some of these include parents having a baby, you photocopying a page in your textbook, and a museum staff member making print copies of a famous oil painting.
In each case, a relatively close copy of the original has been made. The same principle applies in cell reproduction—making new cells that closely resemble the original cell. Cell reproduction is a complex process that involves many intricate, highly regulated steps. Like metabolism, each of these steps must occur in a particular order.
Cell reproduction is important for three reasons:
In this unit, you will learn about how cells carry out these critical processes. Your understanding of the information this unit will lay the foundation for understanding heredity and molecular genetics.
Why do cells divide? Cells undergo different phases, during which they grow, duplicate their genetic material, and then divide. In single-celled organisms, cell division results in two new individuals. Cell division in multicellular organisms is more complex. Most cells in multicellular organisms are somatic cells. Somatic cells consist of all the nonreproductive cells in an organism; for example, tissue cells, nerve cells, and blood cells. When somatic cells divide, they go through the process of mitosis, which is a type of cell division that results in two identical daughter cells.
Meiosis, on the other hand, is a type of cell division that prepares an organism for sexual reproduction. Meiosis begins with germ cells. Be careful not to confuse “germ cells” with the common use of the word “germ,” which refers to any infectious microorganism. In humans, germ cells are found in the ovaries and testes. These germ cells go through the process of meiosis to produce gametes, which in animals are sperm and egg cells. When two gametes from different individuals combine, a unique new cell is produced, called a zygote. This single cell then goes through the process of mitosis to grow into a new individual.
You might be wondering how a single cell can grow into a complex adult organism containing many different cell types simply by mitosis. This happens through a process known as differentiation, in which certain genes found in a cell’s nucleus are turned on or off. Even though every single somatic cell in your body contains identical DNA, these cells were able to differentiate into specific cell types (such as blood cell, liver cells, etc.) simply by turning certain genes on or off. Throughout our lives, we continue to produce more genetically identical cells through the process of mitosis. These somatic cells are essential for growth and replacement of worn-out cells and tissues.
Mitosis and meiosis are the processes by which cells divide and grow; their role in human growth is summarized in the diagram above.
Chromosomes are thread-like structures located inside cells. In eukaryotic cells, the chromosomes are contained inside an organelle called a nucleus . In prokaryotic cells, they are located in a particular area of the cytoplasm. Chromosomes are made up of two major parts: DNA and proteins. Both DNA and proteins are large molecules. Envision the general makeup of a chromosome, as shown in the following diagram.
DNA is the molecule that stores and transmits inherited genetic information. This information includes the directions that tell a cell how to make proteins. Information is stored in DNA in segments called genes. In the diagram above, the ruler represents the entire DNA molecule. The genes are represented as short segments of the ruler. You will learn more about the structure and function of genes in the Molecular Genetics unit.
The blobs of clay in the diagram above represent histone protein molecules. These proteins associate with DNA in a very precise way. First, they couple with DNA at regularly spaced intervals. Notice in the diagram above that the blobs of clay (histones) are located at equally spaced intervals along the ruler. Second, they help to wind the DNA molecule into an organized and compact structure. The following diagram shows how histones wind DNA into a compact structure called a chromosome.
![]() |
![]() |
The winding function of histones plays a very important role in cells. It allows the DNA to be condensed into organized bundles (chromosomes) that can easily be moved around the cell. Because the DNA is wound around the histones in such an organized way, the DNA can be easily wound up and unwound at various points during the cell cycle. For example, chromosomes must unwind slightly when a new copy of DNA is made. Then, the chromosomes become very tightly wound as the cell gets ready to divide during mitosis or meiosis. Without an organized compaction system, the long pieces of DNA in chromosomes could easily become "knotted" and tangled. Think of a long piece of string. Knots can easily form in the string if it is not carefully wound up into a ball. Keeping the DNA organized is critical, especially when you consider that the DNA in one human cell is three feet long when it is stretched out straight. This is quite amazing when you realize that this must fit into a cell so small that it cannot be seen with the naked eye. This feat is achieved by two different winding processes. One of these is the winding of DNA around histone molecules. As a result of this winding, the DNA molecule becomes shorter than before it was wound around the histone molecules.
The videos below are two illustrations of the winding of DNA around histones. The left video shows the DNA strand coiling and condensing to form chromosomes; on the right, a model depicts the structure of DNA.
|
|
Now look at an image of the relationship between DNA and histones in chromosomes that is more realistic than what is shown in our ruler-and-clay diagram above. When the DNA becomes tightly wound around the histone proteins, imagine that the chromosome looks as illustrated below.
![]() |
![]() |
The chromosomes in somatic cells and germ cells are present in pairs. One chromosome in each pair is descended from the organism's father. The other chromosome in the pair is descended from the organism's mother. Each member of a chromosome pair is called a homologue. Together, these two chromosomes make up a pair of homologous chromosomes. The words homologue and homologous both begin with the prefix "homo-," which means "the same or similar." So the two chromosomes in each pair contain the same set of instructions. The following diagram shows the relationship between homologues and homologous chromosomes.
Somatic cells produced through mitosis are diploid. “Di-” means “two,” and diploid cells contain two copies of every chromosome. Chromosomes are the condensed form of chromatin, the combination of DNA and proteins that fill the eukaryotic nucleus. Humans have a total of 46 chromosomes in their diploid cells. One way to write this is "2n = 46." Humans who have more or less than 46 chromosomes sometimes survive, but usually have life-altering conditions, such as Down syndrome. Down syndrome is caused by having three copies of chromosome 21, instead of the usual two copies.
Gametes are haploid cells. They contain only one copy of each chromosome. In human haploid gametes, there are 23 chromosomes. One way to write this is "1n = 23." As you learned in the previous activity, two haploid cells can combine, through the process of fertilization, to form a diploid zygote. The diploid zygote then goes through mitosis to form the diploid somatic cells of the organism.
Remember that germ cells, found in the ovaries and testes, give rise to gametes. In other words, germ cells go through meiosis to produce gametes. Germ cells are diploid cells. When they go through meiosis, the end result is four unique haploid daughter cells. These haploid cells can be fertilized by other haploid cells to produce diploid cells again.
The table below summarizes the characteristics of somatic cells, germ cells, and gametes.
Every living organism goes through various stages of development. For example, human beings go through these stages of development: fetus, infant, child, youth, teenager, young adult, adult. Remember from earlier that cells are living things. They, too, go through a set of programmed stages.
The cell cycle is the "life cycle" of a cell. It is the set of stages that a cell goes through during its lifetime. Cells go through five major stages of development: G1, S, G2, M, and cytokinesis. The events that occur during stages G1, S, and G2 are nearly the same in all types of cells. Together, these three stages of the cell cycle are called "interphase." The events that occur during stages M and cytokinesis are different, depending upon the type of cell. The diagram below shows an overview of the stages in a cell's life cycle:
Notice in the diagram above that we start with one cell on the left side. This is called the parent cell. It is called this because, like a human parent, it reproduces—it makes new cells that are copies of itself. The parent cell goes through the five stages of development in this order—G1, S, G2, M, and cytokinesis.
At the end of the development process, two new cells are produced. These cells represent the next generation of cells in this lineage (just as your parents represent one generation and you represent the next generation). Each of the new cells is called a daughter cell. Do not be confused by the use of the word daughter—it does not mean that these are female cells. Regardless of whether they are male or female cells, the newly produced cells are called daughter cells.
Notice that the parent cell no longer exists—but the parent cell did not die. Rather, the contents of the parent cell have been split into two new cells. Each of the new cells contains a portion of the original parent cell. Previously, we compared the cell cycle to the phases of human development. Notice here a very important difference between the two: The far end of human development is marked by the death of the individual. The end of the cell cycle, however, results in the formation of new cells.
In the diagram "Overview of the Cell Cycle" above, look closely at the parent cell and the two daughter cells. Notice that they look the same. Once again, we revisit the theme of cycles in nature that we discussed above. The cycle generates cells that resemble the cells we started with. In addition, the new cells undergo the same development-and-division process that their parental cell did. For these two reasons, we call this the cell cycle, and draw it like a circle. Similarly, human beings have children, who then have children themselves. (You may have heard this called the circle of life.) Cells do the same thing through the cell cycle. The following diagram shows the cell cycle in circular form.
In cell replication, chromosomes are moved to new locations. Like packing household materials into a box before moving, condensing the chromosomes helps with this process. Chromosomes are condensed by wrapping them more tightly around histones (proteins found in the nucleus).
If a cell is going to reproduce, more DNA is needed. During mitosis, one cell divides into two cells containing identical genetic material. This requires twice as much DNA, because there must be one complete set for each daughter cell. Meiosis also requires the DNA to be replicated. DNA replication (or DNA synthesis) occurs in the synthesis (S) stage of the cell cycle.
After a chromosome has replicated, the two copies remain attached at a point called the centromere. Each copy of the chromosome is called a chromatid.
During cell division, each chromatid moves to one of the two daughter cells. Keeping the two copies of a single chromosome attached helps the sorting process. In human cells, there are a total of 46 chromosomes. After the S stage of the cell cycle, these chromosomes have replicated and there are now 92 total pieces of DNA. The process of mitosis must sort these into 2 identical piles of 46 chromosomes. The cell keeps the matching pairs attached to keep track of which chromosomes need to be sent to opposite poles.
This is rather like sorting a sock drawer. If all the socks are loose in the drawer, it can be challenging to find the identical mate to a sock. But if the sock mates are folded together, each bundle contains both mates in the set.
Both mitosis and meiosis must keep track of the chromatids on a chromosome. In the Meiosis module, you will learn that meiosis must also keep track of homologous chromosomes. A diploid cell has two copies of each chromosome — one copy from each parent. These chromosome pairs are called homologous because they have the same genes, but may have different versions of that gene. For example, a homologous pair of chromosomes may have a gene for hair color. One chromosome in this homologue may contain a “blond” version of that gene and the other homologue may contain the “brown” version. Each gamete that is produced will get either the homologue with the brown version or the homologue with the blond version.
Growing from a single fertilized egg to an adult organism requires many rounds of cell division. Before cell division, a cell must copy its genome so that each new daughter cell receives a full set of DNA. This DNA replication occurs in the synthesis (S) stage of the cell cycle just prior to mitosis. The process of mitosis sorts the DNA equally into two separate nuclei. The process of mitosis is divided into the following steps or phases:
Each new cell gets one of these nuclei during the process of cytokinesis. At the end of mitosis and cytokinesis, each new daughter cell should have the identical genetic makeup of the parent cell. Mitosis of a single diploid cell will result in two diploid cells with identical genomes, unless a mutation or other error occurred. Mitosis of a haploid cell will result in two identical haploid cells. These cells can go through the process again to reproduce.
Let’s begin with a diploid cell that is 2N = 4. This cell has two copies of each of its two chromosomes. During the S phase of interphase, the cell replicates its DNA. This results in two sister chromatids that are attached to each other at the centromere. The chromosomes have not condensed enough at this point to be visible with a microscope.
Centrosomes are also formed during interphase. Centrosomes are used to identify the poles of the cell during the division process; these poles determine where the chromosomes will be sent.
Mitosis begins when DNA condenses into chromosomes visible under a light microscope. This packaging of the chromosomes into condensed bundles makes them easier to sort, but it inhibits protein synthesis. Therefore, it is essential that the cell has produced all necessary proteins prior to the start of this process.
The nucleoli disappear and the nuclear envelope begins to disintegrate. This allows cellular components to act upon the chromosomes (the chromosomes are no longer tucked away inside the nucleus).
Centrioles, contained in the centrosomes formed during interphase, are areas where microtubules originate in order to help sort and organize the sister chromatids. The centrioles begin to move to the opposite poles of the cell. This will determine where the chromosomes go when they are sorted.
Microtubules are disassembled from the cytoplasm and reassembled into the mitotic spindle. This structure will move the chromosomes to the proper location.
This stage is called “prophase,” because it is a preparatory stage. “Pro-” means "before;" “phase” means “stage.” So, this is the stage before the process gets into full swing.
Prometaphase is the stage between prophase and metaphase. During this stage, the nuclear envelope is fully broken down. This allows the microtubules to attach to the centromeres of the chromosomes.
The centromere is the region in the center of the X-shaped chromosomes. Each half of the X is a copy of the same DNA strand. The centromere contains proteins that hold together these two copies and can bind to the microtubules. The proteins on the centromere are called the kinetochore, and the microtubules that attach to them are called kinetochore microtubules. The microtubules that are not attached to the kinetochores are called nonkinetochore microtubules.
Metaphase is so named because the chromosomes line up in the middle of the cell. The root “meta-” means “middle.”
The kinetochore microtubules are used to orient the chromosomes in the center of the cell. Each chromosome will be attached to two kinetochore microtubules. Each of these kinetochore microtubules will be attached to one of the two centrioles.
During anaphase, the kinetochore proteins break down the microtubules attached to them and the connections between each copy of the chromosome will be broken down. This causes the individual chromosomes to move to opposite poles of the cell. The root “ana-” refers to “apart”; the chromosomes are moving apart from each other.
The nonkinetochore microtubules from one pole also push on the nonkinetochore microtubules from the other pole. This causes the cell to elongate. By the end of anaphase, each pole of the cell contains an identical set of chromosomes.
In telophase, the nuclei at each pole form again. The chromosomes are now separated into two identical nuclei. This is the end of mitosis.
However, the two nuclei are still in a single cell. The next step will separate the cytoplasm into two cells. “Telo” comes from the Greek word for “end.”
Cytokinesis is the separation of the cytoplasm into two new daughter cells. Animal cells divide when proteins pinch in the center of the cell until it separates into two. This region is called the cleavage furrow.
Plant cells divide when new cell wall components are laid down in the center of the cell. This is called the cell plate.
Use the activities on this page to review the process of mitosis. Do the activities as many times as you’d like, until you feel comfortable with the whole process.
The cell cycle is highly regulated. There are many checkpoints that are used to determine if the cell should or should not continue through the stages of the cell cycle. Many different environmental and internal signals are involved in determining whether or not a cell will complete division.
A healthy cell has many genes that produce proteins involved in the process of regulating cell division. Some of these genes produce proteins that encourage cell division. These proteins would be produced under circumstances where the cell should divide. For example, if you scrape your knee, the cells in that area need to be replaced. Other genes produce proteins that inhibit cell division. For example, healthy human cells grown on a petri dish would stop growing when they filled the dish. This is also an essential component of regulation. Cells that continue dividing when they should not can cause cancer.
Numerous research projects in the past and present are investigating cell cycle genes. Some of these research projects are focused on cancer treatments and cures and others are focused on understanding the biology of this process. Because much of this research is focused on cancer, the terminology used is generally focused on relationship to the disease.
For example, genes that inhibit cell division are classified as tumor suppressor genes. If these genes are working properly, a cell will not divide uncontrollably because these genes will stop the process. Genes that encourage cell division are called proto-oncogenes. If these genes are working properly, a cell will not receive the message to divide unless it should be dividing.
Mutations could cause these genes to stop working properly. Mutations are errors in DNA that can result from genetic predisposition or environmental factors. Mutagens are chemicals in the environment that produce mutations. Carcinogens are specifically those mutagens that have shown that they can cause cancer. Cigarettes, for example, contain carcinogens that can alter the DNA and disrupt proper regulation of the cell cycle.
The development of cancer requires mutations in BOTH proto-oncogenes and tumor suppressor genes. A malfunctioning proto-oncogene is called an oncogene because it promotes the development of cancer. The root “proto-” refers to "before" and the root “onco-” refers to "cancer." A defective tumor suppressor gene can no longer suppress tumors. Cancer is the result of numerous mutations in multiple genes involved in regulating the cell cycle.
There are two ways for organisms to reproduce: asexual and sexual. Many organisms, such as bacteria, fungi, and even some plants and animals reproduce asexually. During asexual reproduction, all of the genes in a cell are passed to its daughter cells. This means that the resulting cells are clones, or identical copies of the parental cell.
Asexual reproduction is efficient and fast. However, it only provides one set of genes to adapt to the environment. Think about trying to guess a number. Do you have a better chance of guessing the number if you get one guess or if you get two guesses? Sexual reproduction, in which the genes of two parents are combined, results in more genetic variability. This enables offspring to deal more successfully with environmental change. However, sexual reproduction is a more complex process and results in a smaller number of offspring. Interestingly, certain organisms that can reproduce both asexually and sexually will choose sexual reproduction only when environmental conditions are not optimal. That way, a smaller number of more variable, and thus more adaptable, offspring is produced. In optimal conditions, however, asexual reproduction is the favored mechanism.
As you learned, genetic information is contained in DNA, which in eukaryotic cells is bound to special proteins called histones. The combination of DNA and histone proteins is called chromatin. During the S phase of the cell cycle, the chromatin is loose, to facilitate DNA replication. Chromatin can condense to form chromosomes during nuclear division. Chromosomes have a typical appearance and can be organized in a karyotype.
Humans have 23 pairs of chromosomes (2n = 46). Of those 23 pairs, 22 are autosomes, and the 23rd pair are sex chromosomes. Sex chromosomes can be of two types: X and Y. Males have one X and one Y chromosome (XY), while women have two copies of the X chromosome (XX).
In sexual reproduction, two reproductive cells called gametes fuse in the process of fertilization. In order to avoid the duplication of the chromosome number, gametes are formed through a special form of cell division, meiosis, which halves the number of chromosomes. Cells with a single set of chromosomes are haploid (n), in contrast to the somatic (body) cells, which are diploid (2n). The fusion of the gametes during fertilization restores the diploid chromosome number in the first cell of the new individual, the zygote.
In the human life cycle, haploid gametes produced during meiosis fuse to form a diploid zygote, which goes through the process of mitosis to grow into an adult human. When the adult human produces haploid gametes, the process can be repeated.
Germ cells are diploid cells that go through the process of meiosis to form haploid gametes. This process can happen at different times, depending on the organism. For example, in humans, the male gametes (sperm cells) are formed continuously from germ cells in the testes, starting in puberty. Female gametes (eggs or oocytes), on the other hand, are formed in the ovaries during fetal life. They remain in the ovaries, “stuck” in the final stages of meiosis, which is not completed unless fertilization occurs.
Meiosis achieves more than just halving the chromosome number. It also introduces variability in the way the genes are distributed in the gametes. In diploid cells, there are 23 pairs of homologous chromosomes. Each pair of homologues has the same set of genes, though they can have different alleles (or forms) of each gene. One homologue comes from your mother and the other comes from your father. You may have inherited an allele for straight hair from your mother and an allele for wavy hair from your father. When meiosis takes place, the homologous chromosomes will separate, so one gamete will carry the allele for wavy hair and the other for straight hair. Each trait on each chromosome will segregate its two alleles in a completely random way. There are also other additional ways of “shuffling the cards” of genes, which explains why siblings from the same parents can be so different.
Let’s learn more about meiosis. Meiosis consists of two consecutive cell divisions, called meiosis I and meiosis II. Both stages of meiosis consist of prophase, metaphase, anaphase and telophase, just like mitosis. The phases are labeled I in meiosis I, and II in meiosis II.
Meiosis I is often referred to as the reduction division, because it is during this stage that the number of chromosomes in the parent cell are divided in half. Meiosis II is extremely similar to mitosis, except that there are half as many chromosomes.
An important feature of meiosis is the process called crossing over. Crossing over is responsible for generating much of the diversity that exists within a species. Remember that diversity in a species is one of the hallmark characteristics of living organisms. In future modules, you will see that this diversity allows species to survive in changing environmental conditions. So, crossing over is another process that brings variability in the offspring.
Crossing over is a process in which genes swap positions on matching chromosomes. The result is a new combination of genes on each chromosome.
During metaphase I, homologous pairs of chromosomes line up on the metaphase plate. This is very different from mitosis, when sister chromatids line up, completely independent of their homologues.
During anaphase I, homologous pairs are pulled apart, and they move toward the poles of the cell.
In telophase I, cytokinesis occurs and two new daughter cells are formed.
Meiosis II begins with two haploid cells, each containing too much DNA. The events in meiosis II are almost identical to the events of mitosis.
In Meiosis II (which is essentially like mitosis), the sister chromatids separate from each other. Thus, as the result of meiosis, four haploid cells are produced. Note that those four cells are not always viable. In humans, male germ cells will produce four viable sperm cells. However, in the case of females, only one of the four will survive as an egg.
The end result of meiosis is four unique haploid daughter cells. Genetic variability is introduced in various ways during meiosis:
When you add that fertilization is also a random process, you can get an idea of the incredible number of combinations of possible offspring showing up even in one family.
Mitosis | Meiosis | |||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Overall purpose | The purpose of mitosis is to produce new somatic cells and germ cells. These new cells are needed to: (1) increase the number of cells in an organism during growth and development; (2) replace damaged cells during wound healing; and (3) produce replacement cells as other cells undergo naturally preprogrammed death. Mitosis must preserve the same chromosome number in the parent and daughter cells. This is necessary for the parent and daughter cells to be identical to one another. | The purpose of meiosis is to produce new gametes (eggs and sperm). These cells are needed for reproduction. Meiosis must reduce the chromosome number in the daughter cells to half that in the parent cell. This is necessary for reproduction. when an egg and sperm (each with half the full chromosome set) fuse during fertilization, they create a single new cell with the full chromosome set. | ||||||||||||||
Starting cells (parental cells) | Somatic cells | Germ Cells | ||||||||||||||
New cells made (daughter cells) | Somatic cells | Gametes (eggs in female; sperm in male) | ||||||||||||||
Chromosome content of the parent cells | Diploid (2n) | Diploid (2n) | ||||||||||||||
Chromosome content of the daughter cells | Diploid (2n) | Haploid (1n) | ||||||||||||||
Effect on the chromosome number of the daughter cells | Preserves the chromosome number from parent to daughter cells | Cuts the chromosome number in half from parent to daughter cell | ||||||||||||||
Number of cell divisions that occur | 1 | 2 | ||||||||||||||
Number of daughter cells made | 2 | 4 |
Aneuploidy is a chromosomal abnormality characterized by an abnormal number of chromosomes in a cell. It is caused by nondisjunction, which occurs when chromosomes fail to separate properly during anaphase. Genetic problems caused by nondisjunction can be diagnosed prenatally by studying the fetus’ karyotype. A karyotype is a visual representation of the chromosomes of an individual, obtained from a small sample of cells. In fetuses, this is usually obtained from sampling either the amniotic liquid (amniocentesis) or the placenta.
In nondisjunction, chromosomes are not separated correctly during metaphase I or II, resulting in gametes with too few or too many chromosomes. There are many genetic disorders associated with incorrect chromosome numbers. One of the most well known condition caused by nondisjunction is Down syndrome. Down syndrome results from nondisjunction of chromosome 21, resulting in a child with 3 copies of this chromosome (trisomy 21). Down syndrome appears in approximately 1 of every 1,000 children, and older women are at a higher risk of having babies with the condition. Chromosomal abnormalities can include sex chromosomes also. For example, males with Klinefelter syndrome have an extra X chromosome (XXY), and females with Turner syndrome have only one X chromosome (X_). Both these conditions are relatively benign and can be corrected with hormone treatment. Note that nondisjunction due to lack of separation of sister chromatids can occur also during mitosis.
Why do children look like their parents? How do rose growers get so many different colors of roses — pink, red, white, yellow, purple, peach, striped, and many more? How can there be different-colored puppies within one litter?
In the Cell Division unit, you learned that cells pass on a complete set of genetic information (contained in the DNA), which is transmitted from parent cells to daughter cells. This genetic information helps determine many of the characteristics of the offspring. In humans, for instance, DNA is responsible for a child’s height and skin pigmentation. It also helps determine the overall health of the child by regulating many physiological characteristics, such as the condition of the heart and lungs, as well as brain and blood chemistry.
How much of our physiological makeup is determined by genetics, and how much is determined by other factors, such as personal choice and environment? This question is a source of ongoing debate in both scientific and philosophical communities. Environmental factors can influence our DNA, and our DNA can influence how we react to environmental factors. If we are ever able to answer this question, the answer will likely be complex.
In this unit, you will focus on the direct links between genetic information and visible traits. This will help you to learn how genetic information can be passed from one generation to the next and how these genes affect the later generation. We will introduce simple examples of traits that are clearly visible and have a direct correlation to genetics.
The Cell unit discussed the components of the cell. One of these components is the DNA, which is arranged into one or more chromosomes. Each chromosome contains many genes. A gene is a region of a chromosome that has a specific function. For example, genes can contain information required to make a protein. How this happens will be discussed in the Molecular Genetics Unit.
DNA carries the information about how proteins are built, and proteins carry out many of the reactions necessary for living organisms. The Proteins module discussed the variety of roles proteins can play in a living organism. Some proteins produced by living organisms determine the traits we can observe. These visible traits are called the phenotype of an organism.
For example, protein enzymes in plants catalyze production of the pigments, resulting in flowers of different colors. The color of the flower is one of the observable traits, or the phenotype of the plant. The genetic information that produces these traits is called the genotype of an organism. Depending on the pigment production, the flowers can have red, blue, or white color. Flower color is determined by specific genes in the cells. The variations in the same gene are called alleles. In our example, different alleles of the pigment-producing gene can result in more or less pigment, and, consequently, in darker or light flower color.
In this unit, you will learn about the genotype and phenotype of diploid organisms. Diploid cells have two copies of each chromosome; haploid cells have only one copy of each chromosome. Because diploid cells have two copies of each chromosome, they also have two copies of every gene. One set of chromosomes (and, therefore, one copy of each gene) is from the mother (or egg) and one set is from the father (or sperm). Diploid cells can have two of the same alleles of a gene, or they can have two different alleles of a gene. Cells that have the same allele are called homozygous (“homo” meaning "same," and “zygous” referring to the reproductive cells produced from meiosis). These cells are called homozygous because all of their reproductive cells (gametes) will have identical alleles. Cells with different alleles are called heterozygous (“hetero” meaning "different") because they will produce reproductive cells that have different alleles of this gene.
When you studied meiosis, you learned about how gametes were formed. But how are alleles dispersed into gametes through the process of meiosis? This activity will help you answer that question as you review the essential components of meiosis needed to understand inheritance. You may wish to review the Meiosis module prior to continuing with this unit.
![]() |
Gregor Mendel was a nineteenth-century Augustinian monk in Bohemia, a country in Europe that is now part of the Czech Republic. In 1853, Mendel left the monastery for two years to study at the University of Vienna, where he learned about mathematics and botany. When he returned to the monastery, he began investigating how plants passed on their characteristics to their offspring. He was conducting this research only a few years before Charles Darwin published On the Origin of Species. |
Mendel set out to understand inheritance by breeding pea plants. By collecting peas from pods produced after fertilizing two parent pea plants, Mendel could then grow those peas into new pea plants and see how these offspring resemble or differ from the parent plants. The heritable characteristics that Mendel studied included seed color and shape, pod color and shape, flower color and position, and stem length. Each of the traits that Mendel studied occurs only in two forms, or phenotypes. For instance, pea flower color is a trait that can exhibit either the purple phenotype or the white phenotype.
So if we know the phenotype, and we know that those two varieties would always result in same-color flowers, we can deduce that all individuals (plants) in that population (purple or white) have the same alleles for the gene that determines flower color.
Next, Mendel crossed the purple variety with the white variety. He manipulated his plants so he could control which plant pollinated which. In plants, pollen is the male gamete, landing on the flower to fuse with the ovule—the female gamete. The resulting zygote will give rise to the seed, or embryo.
When Mendel crossed purple and white flowers, the first generation of plants, which he designated “F1,” were all purple. Somehow, purple was “stronger,” or dominant over the white. So if the alleles of the purple flowers were symbolized by “B," then the allele corresponding to white flowers was “weaker” and is symbolized by “b.” If the genotype of the parent purple flower is BB, then the genotype of the parent white plant will be bb.
Let’s see what happened during that first cross. You will use a tool called a Punnett square, which graphically shows the possible combinations of the gametes of the parents. In a Punnett square, you place the possible gametes from each parent along the top and left side of the square. What you are representing here is the Law of Segregation: during meiosis I, the chromosome pair separates, resulting in haploid gametes. Each gamete will have only one allele.
Then you combine the possible gametes with each other to see all possible combinations, as shown in this example.
As you can see, 100 percent of the offspring have the Bb genotype, but 100 percent also show the purple phenotype. You might expect the offspring to show a phenotype that somehow blends the parental phenotypes. Sometimes this happens, but in the case of these pea plants' flower color, the purple allele (B) is dominant over the white allele (b). Stated another way, the white allele (b) is recessive to the purple allele (B).
In many, but not all, cases it is possible to pinpoint a phenotype to one allele. Remember, alleles are variants of genes, and genes code for proteins. In this example, the purple color may be due to an enzyme that produces a purple pigment. If the allele codes for a mutated enzyme that does not perform as expected, the resulting flower would be white. A flower with BB or Bb genotype has the normal enzyme — purple pigment is produced — and even the heterozygous flower, which has only one allele for the normal enzyme, produces enough pigment for the flower to be purple. The bb flower, on the other hand, has no functioning enzyme, so the flower remains white.
There are some easily observable human traits that follow the Mendelian inheritance pattern, showing a dominant and a recessive allele. Examples include unattached earlobes (dominant) versus attached earlobes (recessive). The widow’s peak hairline, a cleft chin, rolling tongue, and hitchhiker’s thumb are others. Does the ability to roll your tongue make you more fit for survival? Not really! One common misconception is that dominant traits are more favorable from an evolutionary point of view. Another is that dominant traits are more common. We cannot stress enough that traits, and particularly human traits, are determined by complex interactions between genes and environment. In the early days of genetics, simple observable traits were used to describe inheritance patterns. Once the molecular basis of inheritance was established, scientists could tackle more complex interactions. You will learn more about this in the Molecular Genetics unit.
Mendel’s next experiment involved crossing the heterozygous purple-flowered offspring that resulted from his first cross. The crossing of two individuals heterozygous for a single gene is called a monohybrid cross. Let’s see what happened.
Albinism is a recessive disorder related to the deficiency of an enzyme responsible for the synthesis of the pigment melanin. Only homozygous recessive individuals present the albino phenotype.
After his first experiment with crossing involving one trait, Mendel conducted experiments looking at the inheritance of two traits. For example, he crossed true-breeding plants that had yellow, round seeds with true-breeding plants that had green, wrinkled seeds. The resulting F1 generation plants all produced yellow, round seeds.
The story became much more complicated when Mendel crossed the F1 generation plants with each other. This is called a dihybrid cross, as it is a cross between two individuals, both of whom are heterozygous for two traits.
In the F2 generation, Mendel observed the original phenotypes (yellow/round, green/wrinkled) and two additional phenotypes: green/round and yellow/wrinkled. The four alleles had combined with each other, independently of each other. This outcome demonstrates the law of independent assortment.
Mendel did not know of genes or alleles, he talked only of “units” specifying traits. Let’s look in detail at what happens with the genotypes of a dihybrid cross.
As you remember from the Meiosis module, during meiosis I, after crossing-over, the chromosome pairs separate randomly. This means that Y and y will separate, as will R and r. But which goes to which gamete is random. So the Y allele in a chromosome can be associated with a chromosome containing the R allele or the r allele, and the R allele can be paired with a chromosome containing the Y allele or the y allele. Therefore, there are four possible gametes in a dihybrid cross.
There are exceptions to the law of independent assortment. We know now that peas have seven chromosomes, and Mendel happened to study traits coded by genes on different chromosomes, or on distant positions of the same chromosome, which assort independently due to frequent crossing-over. Genes that are very close together on a chromosome do not assort independently, and are said to be linked. The closer they are, the more probable that the gametes will receive parental combinations of alleles of those genes. All the genes in a chromosome are called a linkage group. Peas have seven linkage groups.
An example of linked traits in humans is human hair and skin color. You may have noticed that red hair and fair skin are often inherited together. This is because the genes for these traits are very close to each other on the same chromosome, and they rarely recombine.
Mendel published his results in 1866, but few read his work, and no one understood it entirely. When he became abbot of the monastery in 1871, he retired from science, and he died in 1884, unaware that his pioneering experiments would form the foundation for the study of modern genetics.
Gregor Mendel’s experiments that you studied earlier involve simple dominance, in which the offspring will show a trait from either one parent or the other, not a blend of those traits. In a case of simple dominance, crossing a red flower and a white one will produce a red flower, not a pink flower. If you think about traits that children inherit from their parents, you will realize that this does not always happen.
Continued genetic research has revealed that inheritance is usually more complicated than simple dominance. Some genes follow Mendelian patterns of inheritance, but many are governed instead by non-Mendelian inheritance. Mendel himself realized this when he looked at species and characteristics beyond the scope of his research. For example, although the cross of a true-bred purple-flowered pea plant with a true-bred white-flowered pea plant will result in purple flowers, the cross of true-bred red flower and true-bred white flower snapdragons will result in offspring with pink flowers. If the pink snapdragons are crossed with each other, pink, red, and white flowers will appear in a 2:1:1 ratio, respectively. How can this happen?
In this inheritance pattern, called incomplete dominance, the heterozygous genotype exhibits an intermediate phenotype. In the snapdragon example, the intermediate phenotype is pink. What is the difference between the purple heterozygotes of the pea flowers and the pink heterozygotes of the snapdragon flowers? Think about what an allele is: a variation of a gene containing the recipe for a protein. In the flower color example, this protein could be an enzyme involved in the synthesis of a pigment. In the case of the pea plant, the presence of a single allele in the heterozygous plant was enough to provide all the pigment needed to give the dominant purple color. However, in the case of the snapdragon, the presence of only one allele does not provide enough pigment to give the flower a fully red hue—but there is enough pigment to tint it partially, thus providing the pink color of the heterozygous flower.
So far, You've only learned about genes with two alleles — that is, two options for a gene coming from either parent — present in a chromosome pair. However, more than two alleles may be present in a population, resulting in what is referred to as a multiple-allele system. Human A, B, and O blood groups exist due to three possible alleles of the ABO gene, which codes for an enzyme that modifies carbohydrates present on red blood cells (Components of the Membrane). The A and B alleles code for two versions of the enzyme, while the O allele is mutated and does not code for a functional protein.
A and B are codominant, which means that if the individual has both A and B alleles, he will be of AB blood group, which expresses both enzymes. On the other hand, O is recessive, so only the homozygous recessive individuals will have an O phenotype (blood group O).
When referring to human blood groups, you may have heard references to “A negative” or “O positive.” The term “positive/negative” here refers to the presence or absence of a membrane glycoprotein called Rh. The Rh glycoprotein was named after Rhesus monkeys because some of the original studies on the Rh factor were done on these animals. The expression of this protein presents a simple dominance pattern, with the presence of the Rh molecule (Rh positive) being dominant over the absence (Rh negative). So a person's blood is said to be A positive (A+) if both A and Rh molecules are present on that person’s red blood cells.
If the immune system of an Rh negative person is exposed to Rh positive red blood cells, it will produce antibodies against the Rh molecule. The next time the immune system meets Rh positive red blood cells, the antibodies will attack and destroy them. This scenario is rather unusual, except when a mother who is Rh negative is carrying an Rh positive baby. After the first pregnancy, her immune system develops anti-Rh antibodies, and in subsequent pregnancies these antibodies can cross the placenta and attack the red blood cells of the fetus. This results in hemolytic anemia, and could seriously affect the fetus. Luckily, Rh negative mothers today receive an anti-Rh antibody injection that prevents development of the anti-Rh response.
Another example of codominance is sickle cell anemia and sickle cell trait. In sickle cell anemia, hemoglobin — the protein that transports oxygen in the blood — is defective, giving the red blood cells a characteristic “sickled” appearance.
A change in one amino acid in the structure of hemoglobin makes it defective: the blood cells cannot transport oxygen effectively, and they accumulate in the blood vessels.
Patients homozygous for the defective allele have sickle cell anemia. Heterozygotes (having one “good” and one “bad” allele for hemoglobin) will present both types of hemoglobin. That condition is called sickle cell trait, where the individuals will present many red blood cells that become sickled under conditions of extreme physical exertion or low atmospheric oxygen. So while the phenotype can be considered intermediate, in fact it is actually a case of codominance. Interestingly, the sickle cell trait protects against malaria infection, and therefore remains prevalent in Sub-Saharan Africa, where malaria is endemic.
Some traits, such as skin color and height, are determined by multiple genes. In humans, skin color is determined by at least three different genes, and probably more. This concept is called polygenic inheritance. Variation along a continuum often indicates that a trait has polygenic inheritance. Most visible traits are polygenic, and their genetics patterns are rather complicated.
In some instances the polygenic effect is rather straightforward. This situation is called epistasis. In epistasis:
For instance, in Labrador retrievers, the allele for black fur color (B) is dominant with regard to the allele for brown (“chocolate”) fur color (b). However, there are also yellow Labs. The coats of yellow labs are created through epistasis. Another gene, E/e, is in play.
The reason for this distinction is an interaction between the two genes. The B gene is involved in the synthesis of the pigment melanin: B results in a black, and b in a brown color. The E gene is related to melanin deposition in the fur: E causes melanin to be deposited, and e suppresses melanin deposits. Being homozygous for e — that is, having the ee genotype (no melanin deposition) — results in fur with neither brown nor black pigment, which appears yellow, independent of the black or chocolate allele (B or b).
Many single genes can have numerous different effects, rather than just two. This capability is called pleiotropy. In the 1930s, for instance, researchers found a gene in chickens that leads most obviously to feathers that curl out, but that also (and less visibly) leads to increased metabolism, blood flow, body temperature, and digestive capacity, as well as fewer eggs laid. This pleiotropic gene is called the “frizzle” gene because of its effect on the appearance of the chickens.
Another gene known to have pleiotropic effects in humans is fibrillin-1, which codes for a connective tissue protein. Marfan syndrome is caused by mutations in the fibrillin-1 gene. People with Marfan syndrome are usually tall and thin with long arms and legs, and they are at risk for heart disease and eye problems. The symptoms can be treated individually, and people with Marfan syndrome may have normal life spans.
Characteristics that are influenced by environmental as well as genetic factors are called multifactorial. The idea of “nature versus nurture” — in other words, the relative influence of genetics versus environmental factors — has been and still is debated. Just looking at the genes of a given organism will not determine how that organism will develop and act. Even identical twins will show different characteristics, depending on the environment in which they live. Everyone is a product of their environment as well as their genetics.
Even when influenced by the environment, phenotypes have a normal range of expression. For instance, human height varies based on nutrition and genetics, but not many people are shorter than 4½ feet or taller than 7 feet. The range of phenotypic possibilities is called the norm of reaction. Hydrangeas, for example, may be blue, pink, or purple, but they are never naturally orange. Hydrangeas are blue in acidic soil with available aluminum, and they are pink in alkaline soil without available aluminum.
Some human characteristics have a narrow norm of reaction, such as blood type. Others have a wide norm of reaction, such as the number of blood cells in humans, which varies depending on factors that include physical fitness, presence or history of infections, and even the altitude at which a person lives.
The environment also affects human genes. Serotonin, a neurotransmitter that acts inside brain cells, lowers anxiety and depression during traumatic times. Mutations in the serotonin transporter gene may cause a reduced ability to cope with stress. That does not mean that the person is always depressed, but if the environment produces stress, the person may become depressed more easily than a person with unmutated serotonin transporter genes.
You already learned about PKU, a pleiotropic disorder caused by defects in a single gene coding for an enzyme that converts the amino acid phenylalanine to tyrosine. Newborns are tested for this defect very early in life, so that if the results are positive, they can be given a diet limiting phenylalanine ingestion. That way, the toxic buildup is prevented and the children can develop normally. PKU is an example in which environmental factors can modify gene expression.
There are around seven billion people on our planet, and everybody looks at least a little different from everyone else. Siblings are different from each other, and even identical twins—who are genetically identical—can show different traits over time. However, when we look at families over several generations, it becomes obvious that certain traits appear generation after generation. How can this extraordinary variability be explained?
In the previous module you learned about the inheritance patterns (Mendelian and non-Mendelian) that were studied using models such as peas, snapdragons, and dogs. Another organism that is often used is the fruit fly Drosophila. Common to these organisms is that they are relatively easy to breed and cross under controlled conditions. They also have a relatively short lifespan, which allows the tracking of traits for many generations. In the case of humans, genetic analysis is much more complicated. We live under a variety of conditions, so the environmental influences are much stronger. Humans choose their mates freely, families are usually smaller, and our lifespan is the same as that of the geneticists. Much information about transmission of human traits comes from the study of pedigrees, a chart of genetic connections similar to a family tree. In this module, you’ll learn how to read pedigrees. You’ll also look at the different patterns and factors that affect human inheritance.
Many human traits are clear-cut and easy to test. These include the ability to bend back the thumb nearly 90 degrees (known as “hitchhiker’s thumb”), to roll the tongue into a U-shape, and to taste a bitter chemical called phenylthiocarbamide. While these traits do not have considerable implications on human health and species survival, many others do. Some of them are related to disorders such as sickle cell anemia, hemophilia, Tay-Sachs disease, and Down syndrome, to name a few.
Some human traits (like the tongue-rolling ability) follow Mendelian patterns, which means they are controlled by a single gene. One of the alleles is dominant, and the other is recessive. Human genetic disorders following a Mendelian pattern are the least common. Most of those tend to be recessive, meaning that the defective allele causing the disorder is recessive. That means only homozygous recessive individuals will show the disorder. The heterozygotes will be carriers, so while they carry the defective allele (and sometimes even express it, as in the sickle cell trait), they phenotypically do not present the full disorder. Other disorders may be dominant, wherein the defective allele causing the disorder is dominant. In this case, the presence of only one allele is enough to provoke the appearance of the disorder, and only homozygous recessive individuals show the healthy phenotype.
As you remember from the Meiosis module, of the 23 pairs of chromosomes found in a human somatic cell, 22 are autosomes, and the 23rd pair are sex chromosomes. Sex chromosomes are of two types: X and Y. Males have one X and one Y chromosome (XY), and females have two copies of the X chromosome (XX). If the gene responsible for the disorder is present on the autosomes, it is called an autosomal disorder. The term “autosomal” refers to chromosomes that are not sex determining. On the other hand, if the responsible gene is present on a sex chromosome, it is called a sex-linked disorder. Due to the fact that the male Y chromosome is very small and contains only genes related to sex determination, sex-linked disorders are due to defective alleles on the X chromosome. The following table shows the inheritance patterns of some human genetic disorders and abnormalities.
Pattern | Disorder | Main symptoms |
---|---|---|
Mendelian: autosomal dominant | Huntington’s disease; Progeria | Degeneration of the nervous system; Premature aging |
Mendelian: autosomal recessive | Cystic fibrosis; Sickle cell anemia | Abnormal glandular secretions provoking lung and digestive dysfunction; Anemia causing effects on the whole body |
Sex-linked | Hemophilia; Red-green color blindness | Inadequate blood clotting; Inability to distinguish red from green |
Changes in chromosome number | Down syndrome; Klinefelter syndrome; Turner syndrome | Mental impairment, heart defects; Sterility, mild mental impairment; Sterility, abnormal ovaries |
Changes in chromosome structure | Chronic myelogenous; Leukemia | Overexpression of white blood cells; Organ dysfunction |
In humans, determination of sex is dependent on a special pair of chromosomes called the sex chromosomes. The other 22 pairs are called autosomes.
There are two types of sex chromosomes — X and Y. The X chromosome is large and carries many genes unrelated to sex. The Y chromosome, on the other hand, is much smaller and carries only the genes containing the instructions for “maleness,” which are molecular signals that instruct the fetal gonads (sex organs) to develop as testes and not ovaries. In the absence of the Y chromosome, a fetus develops as female.
Females have two copies of the X chromosome (XX), while males carry one X and one Y (XY). When gametes are formed, females will have only gametes that contain the X chromosome; males will have some gametes with X and others with Y. For a male baby, a Y gamete from the father has to meet the mother’s gamete. So the determination of a baby’s sex depends on which gamete the father contributes.
About one out of every ten men is color-blind. People who are color-blind are unable to distinguish between certain colors, especially green and red. On the other hand, only about one out of every 200 women is color-blind. Why is there such a drastic difference between the sexes?
In 1910, an American geneticist named Thomas Hunt Morgan made an observation that began to shed some light on this question. One morning, when peering through a hand lens at a male fruit fly, he noticed that it didn’t look right. Instead of having the normally brilliant red eyes of wild-type Drosophila melanogaster, this fly had white eyes. Morgan was particularly interested in how traits were inherited and distributed in developing organisms, and he wondered what caused this fly's eyes to deviate from the norm.
He bred the white-eyed male with several true-bred, red-eyed females and obtained all red-eyed flies in the first generation. (Remember, true-bred organisms have a homozygous genotype.)
Morgan did crosses between the F1 hybrids, and he observed the following: 75 percent of the offspring were red-eyed, and 25 percent were white-eyed. So far, the results supported the Mendelian pattern of 3:1. But then he noticed something odd: all of the white-eyed flies were male.
Morgan wondered why the white-eye trait was associated with the male sex. In Morgan’s time, the idea that an additional pair of chromosomes was responsible for sex determination had just emerged. There were two plausible possibilities: either the trait was somehow linked to the sex chromosome, or the white-eyed trait was lethal for females, meaning that females with the phenotype would not develop past the egg stage. Morgan conducted a series of crosses using the F2 generation, and obtained some white-eyed females also, showing that the white-eyed characteristic was not lethal for females.
In a famous subsequent experiment, Morgan crossed white-eyed females with white-eyed males, obtaining only white-eyed offspring. This cross confirmed that the eye color trait was linked to the female sex chromosome. Because of this, the males are what is called hemizygous for this trait: they have only one allele present, which is in the X chromosome. The Y chromosome does not have the same genes as the X chromosome. Males have no heterozygous option: they either have the dominant allele (red-eyed) or the recessive allele (white-eyed).
The trait of white eyes in fruit flies is a sex-linked trait. It is located on a sex chromosome, so it occurs at different rates in males and females. Because it is located on the X chromosome, it is more specifically called an X-linked gene. In humans, the trait of color blindness results from a gene on the X chromosome. So do some more serious diseases, including Duchenne muscular dystrophy, which causes progressive muscle weakness. People with this disorder rarely live past the age of 25. A gene on the X chromosome codes for a muscle protein that is missing in people with Duchenne muscular dystrophy.
Human chromosomes are numbered from largest to smallest. Each of the 22 pairs of autosomal chromosomes have an identifying number, from 1 to 22; chromosome 1 is the biggest and chromosome 22 is the smallest. As you already know, the pair of sex chromosomes are named either X or Y. The chromosomes of a cell can be visualized using a process that takes a picture of the cell during mitosis, when chromosomes are easily visible. The picture that is produced is called a karyotype.
Karyotypes can be used to identify chromosomal abnormalities in cells or in developing fetuses. Amniocentesis is a medical procedure used to sample fetal cells from a pregnant mother. These cells, which are rapidly dividing, can be used to create a karyotype of the fetus. This karyotype can then help identify potential chromosomal abnormalities in the developing child. Abnormalities could include extra chromosomes, missing chromosomes, and even extra or missing pieces of chromosomes.
This karyotype has an extra chromosome 21. This condition is called “trisomy 21” because the cell has three ("tri") of these chromosomes instead of the usual two. This condition is more commonly called “Down syndrome.” There are relatively few examples of such large-scale chromosomal anomalies. Down syndrome results from an extra copy of one of the smallest chromosomes. Extra or missing copies of sex chromosomes can also result in viable embryos. Embryos with extra or missing chromosomes are often nonviable, and other chromosomal anomalies are not generally seen.
Aneuploidy is the condition of having too many or too few chromosomes, which results from errors in meiosis. If crossing-over does not occur correctly, the chromosomes can have extra pieces or missing pieces. If the chromosomes do not separate properly during anaphase, the resulting cells can have extra or missing chromosomes. This improper separation of chromosomes is called nondisjunction. Nondisjunction can occur in either meiosis I or meiosis II.
The following activity looks at aneuploidy that results from nondisjunction in Meiosis I.
Nondisjunction can also occur at Meiosis II. Complete the following activity to learn more.
Pedigrees are maps that can be used to trace genetic traits through generations of individuals. Pedigrees use the following symbols:
Pedigrees can be used to determine if the trait being studied is dominant, recessive, or X-linked. A trait is recessive if a child anywhere in the family has the trait and both parents do not. This must be the case because if the child has the trait and the parents do not, the only possible genotype option is:
It would not be possible for an affected individual to show a dominant trait that was not expressed in his or her parents, because this sets up an impossible situation where the child would have an allele that could not come from either parent:
If there is no case where a child expresses the trait but neither parent does, then the pedigree is likely to be dominant. In a real-world situation, researchers would need to look at hundreds of individuals to be sure that this conclusion is statistically significant. But for the purposes of this activity, you will look at only a few individuals.
A trait that is significantly more common in males than females is likely to be X-linked. Again, in a real-world situation a much larger sample size would be needed to be sure of these conclusions. The X-linked traits used here are also recessive, so you will notice that they show both characteristics of being X-linked (mostly males) and characteristics of being recessive (child demonstrating a trait that neither parent has).
HGPS is an extremely rare (one case per eight million live births) autosomal dominant disorder characterized by accelerated aging. The disorder is caused in the gene coding for lamin A, a protein involved in chromosome organization. The defective protein accumulates in the nuclear membrane of the cells, affecting many cellular processes. Symptoms start before age two, with thinning skin and weak muscles and bones. Aging is estimated to be eight to 10 times faster than normal. Patients die in their teens from stroke or heart attack. How do people “get” progeria? It is important to remember that because progeria patients die at a very young age, progeria is not inherited, but only appears as a random mutation. This explains its fortunately very low incidence.
In the Classical Genetics unit, you learned about the mechanisms of inheritance, focusing mostly on the organism and population levels in the hierarchy of life. In the Molecular Genetics unit, you will learn about the molecular basis of traits that characterize the organisms. You will develop an understanding on the molecular level of how genetic information is organized and used by living organisms.
If we compare traits of individuals within species, we find that they share certain traits, but they also show many differences among individuals. Fore example, rice (pictured below) shows variation in shape and color of seeds/grains. People also come in different shapes and sizes. What determines physical traits that characterize a person: hair color, eye color, height, weight? How about behavior: some people are risk-takers, others get easily anxious; some people love to be in large noisy crowds, others prefer quiet and solitude. What is more important in determining physical and behavioral traits — genetics or environment? This classic “nature vs. nurture” argument should not be phrased as “either/or,” because the answer is “both.” The interactions between genes and environment determine the traits.
A better question to ask is “To what extent is each trait determined by genetics and to what extent by environment?” The answer will vary by trait and is still being researched. It is not a trivial question to answer, and the numbers below are only rough estimates.
Estimated Genetic Contribution | ||||||
---|---|---|---|---|---|---|
|
In Molecular Genetics, first you will learn about the storage, transmission, and expression of genetic information (central dogma of molecular biology), whereby the information content in DNA sequences, called genes, is ultimately converted to a protein. Then you will look at the structure of genes and genomes and at the interactions between genes and the environment. Finally, you will examine how humans are able to manipulate genes through biotechnology.
If you look at people around you, you will probably notice a fairly large variation in height. Some people are taller than you, while others are shorter than you. Consider this trait as an example that you want to understand on the molecular level. What kinds of molecules may be involved in determining a person’s height? If you and your parents are relatively short (or tall), you may say that you inherited the short (or tall) stature from your parents. What molecule is involved in the transmission of information from one generation to the next? DNA. How does information contained in DNA affect your height? One possible explanation is that your DNA contains a gene encoding a protein called growth hormone. This protein is produced by the pituitary gland, travels to other tissues, and delivers a signal for the cells to grow and divide. Changes in the amount of protein produced or the sequence of amino acids in the protein can affect the function. If bone and muscle cells receive only a small number of signal molecules, the person will end up with a short stature. If they receive a large number of signal molecules, the person may end up being very tall. For a signaling protein (hormone) to deliver the signal, there must be a protein that can bind the hormone and receive the signal (receptor protein). The variation in height will result from any of the following:
Note that a person who inherits “tall” genes may never grow very tall because of illness or inadequate nutrition. Conversely, a person who inherits a faulty growth hormone gene may grow relatively tall if he or she is treated with growth hormone. These are examples of how environment and genes interact to produce traits and phenotypes.
In this module, you will learn about how genetic information is stored in DNA and copied for the next generation. You will also look at a practical application of the process of DNA replication called the polymerase chain reaction (PCR). In the next module, you will learn how the genetic information stored in genes is expressed, resulting in the variety of traits and phenotypes of all organisms.
Nucleic acids are macromolecules that carry out two main functions in the cell: storage of genetic information and synthesis of proteins. Two types of nucleic acids specialize in these functions: deoxyribonucleic acid (DNA) and ribonucleic acid (RNA). DNA is the genetic material that stores information for making proteins in all living organisms. Some viruses store their genetic information in RNA instead of DNA. This may seem as an exception to the universal use of DNA as genetic material; however, viruses are not cellular, and are not considered living organisms. DNA is found in the nucleus of eukaryotes and in two cellular organelles: chloroplasts, and mitochondria.
In prokaryotes, DNA is not enclosed in a separate compartment, but found in the cytoplasm.
Regardless of where DNA is stored in the cell, it contains instructions for building proteins. For protein synthesis to occur, genetic information stored in DNA must first be copied into RNA. In eukaryotic cells, RNA is transported to the cytoplasm, where protein synthesis takes place. Thus, while both RNA and DNA can contain instructions for making proteins, DNA is used for storage of this information, while RNA is directly involved in making proteins.
To understand in more detail how nucleic acids function in transmission of genetic information and in protein synthesis, you must first consider the structures of these molecules and the relationship between structure and function.
The genetic information of an organism is stored in DNA molecules. How can one kind of molecule contain all the instructions for making complicated living beings like ourselves? What component or feature of DNA can contain this information? It has to come from the nitrogen bases, because, as you already know, the backbone of all DNA molecules is the same. But there are only four bases found in DNA : G, A, C, and T. The sequence of these four bases can provide all the instructions needed to build any living organism. It might be hard to imagine that 4 different “letters” can communicate so much information. But think about the English language, which can represent a huge amount of information using just 26 letters. Even more profound is the binary code used to write computer programs. This code contains only ones and zeros, and think of all the things your computer can do. The DNA alphabet can encode very complex instructions using just four letters, though the messages end up being really long. For example, the E. coli bacterium carries its genetic instructions in a DNA molecule that contains more than five million nucleotides. The human genome (all the DNA of an organism) consists of around three billion nucleotides divided up between 23 DNA molecules, or chromosomes.
The information stored in the order of bases is organized into genes: each gene contains information for making a functional product. The genetic information is first copied to another nucleic acid polymer, RNA (ribonucleic acid), preserving the order of the nucleotide bases. Genes that contain instructions for making proteins are converted to messenger RNA (mRNA). Some specialized genes contain instructions for making functional RNA molecules that don’t make proteins. These RNA molecules function by affecting cellular processes directly; for example some of these RNA molecules regulate the expression of mRNA. Other genes produce RNA molecules that are required for protein synthesis, transfer RNA (tRNA) and ribosomal RNA (rRNA).
In order for DNA to function effectively at storing information, two key processes are required. First, information stored in the DNA molecule must be copied, with minimal errors, every time a cell divides. This ensures that both daughter cells inherit the complete set of genetic information from the parent cell. Second, the information stored in the DNA molecule must be translated, or expressed. In order for the stored information to be useful, cells must be able to access the instructions for making specific proteins, so the correct proteins are made in the right place at the right time.
Both copying and reading the information stored in DNA relies on base pairing between two nucleic acid polymer strands. Recall that DNA structure is a double helix (see figure below).
![]() |
DNA Structure (B-helix) |
|
Structure of DNA double helix. On the left: Sugar-phosphate backbone is shown in yellow, specific base pairings via hydrogen bonds (red lines) are colored in green and purple (A-T pair) and red and blue (C-G). On the right: Three-dimensional structure of DNA double helix. Click on the checkboxes to highight the A-T and C-G basepairs. Click on 'Spin On' to automatically rotate the DNA helix structure. Graphic by Madeleine Price Ball (DNA chemical structure) CC-BY-SA-2.0
The sugar deoxyribose with the phosphate group forms the scaffold or backbone of the molecule (highlighted in yellow in the figure above). Bases point inward. Complementary bases form hydrogen bonds with each other within the double helix. See how the bigger bases (purines) pair with the smaller ones (pyrimidines). This keeps the width of the double helix constant. More specifically, A pairs with T and C pairs with G. As we discuss the function of DNA in subsequent sections, keep in mind that there is a chemical reason for specific pairing of bases.
To illustrate the connection between information in DNA and an observable characteristic of an organism, let’s consider a gene that provides the instructions for building the hormone insulin. Insulin is responsible for regulating blood sugar levels. The insulin gene contains instructions for assembling the protein insulin from individual amino acids. Changing the sequence of nucleotides in the DNA molecule can change the amino acids in the final protein, leading to protein malfunction. If insulin does not function correctly, it might be unable to bind to another protein (insulin receptor). On the organismal level of organization, this molecular event (change of DNA sequence) can lead to a disease state — in this case, diabetes.
Before a cell can divide, it must first make a copy or replica of its DNA through a process called DNA replication. This process occurs during the S phase of the cell cycle, so by the time the cell enters the mitotic phase, there are two copies of the DNA molecule. This process of DNA replication takes place in the nucleus of the cell in eukaryotes, where the DNA molecule is found. It also occurs in the mitochondria in animal cells and chloroplasts in plant cells.
Think about this: it is critical that the copies of the DNA are exactly like the original, so the daughter cells are identical to the parent cells. There may be changes to DNA sequence by mutations (as you will see in another section), and meiosis introduces some genetic variability when gametes are produced. But for unicellular organisms or for somatic cells, DNA replication is a process that requires high fidelity.
The process of DNA replication is catalyzed by a type of enzyme called DNA polymerase (poly meaning many, mer meaning pieces, and ase meaning enzyme; so an enzyme that attaches many pieces of DNA). Observe the figure below: the double helix of the original DNA molecule separates (blue) and new strands are made to match the separated strands. The result will be two DNA molecules, each containing an old and a new strand. Therefore, DNA replication is called semiconservative. The term semiconservative refers to the fact that half of the original molecule (one of the two strands in the double helix) is “conserved” in the new molecule. The original strand is referred to as the template strand because it provides the information, or template, for the newly synthesized strand.
DNA replication relies on the double-stranded nature of the molecule. One double stranded DNA molecule, when replicated, will become two double-stranded molecules, each containing one original strand and one newly synthesized strand. You remember that the two strands of DNA run antiparallel: one from the 5’ to the 3’, and the other from the 3’ to the 5’. The synthesis of the new DNA strand can only happen in one direction: from the 5’ to the 3’ end. In other words, the new bases are always added to the 3’ end of the newly synthesized DNA strand. So if the new nucleotide is always added to the 3’ end of an existing nucleotide, where does the first nucleotide come from? In fact, DNA polymerase needs an “anchor” to start adding nucleotides: a short sequence of DNA or RNA that is complementary to the template strand will work to provide a free 3’ end. This sequence is called a primer.
How does DNA polymerase know in what order to add nucleotides? Specific base pairing in DNA is the key to copying the DNA: if you know the sequence of one strand, you can use base pairing rules to build the other strand. Bases form pairs (base pairs) in a very specific way. The figure shows how A (adenine) pairs with T (thymine) and G (guanine) pairs with C (cytosine). It is important to remember that this binding is specific: T pairs with A, but not with C. The molecular recognition occurs because of the ability of bases to form specific hydrogen bonds: atoms align just right to make hydrogen bonds possible. Also note that a larger base (purine, A or G) always pairs with a smaller base (pyrimidine, C or T).
Now that you understand the basics of DNA replication, we can add a bit of complexity. The two strands of DNA have to be temporarily separated from each other; this job is done by a special enzyme, helicase, that helps unwind and separate the DNA helices. Another issue is that the DNA polymerase only works in one direction along the strand (5’ to 3’), but the double-stranded DNA has two strands oriented in opposite directions. This problem is solved by synthesizing the two strands slightly differently: one new strand grows continuously, the other in bits and pieces. Short fragments of RNA are used as primers for the DNA polymerase.
You may remember from the movie Jurassic Park how the DNA of a dinosaur was recreated from a tiny amount of blood from a mosquito trapped in amber. While (so far) this is not possible in real life, the idea is based on one of the most useful tools for the manipulation of DNA: the polymerase chain reaction, or PCR. This technique was invented by Kary Mullis in 1983. PCR uses repeated cycles of DNA polymerase activity to amplify, or make many copies of, a small segment of DNA known as the target DNA. The target DNA resides within a larger DNA molecule that acts as a template. The amplified DNA, or PCR product, can also serve as a template, leading to a “chain reaction” that doubles the amount of PCR product after each cycle. Consequently, PCR can be used to amplify small amounts of DNA from forensic samples or historical artifacts. When scientists have a larger sample of DNA, they can determine the base sequence of the sample and use the data to compare the DNA sequences of different individuals. This process can help forensic scientists figure out the identity of a person who left blood at a crime scene. It can also help decide paternity disputes.
PCR is essentially DNA replication, and just like DNA replication, PCR relies on the enzyme DNA polymerase. Because DNA polymerase requires a primer before it can start building a new strand of DNA, this is the first thing scientists must find (or design). These primers will provide a starting point for the DNA polymerase. In DNA replication, the primers are made of RNA nucleotides. In PCR, the primers are short DNA molecules consisting of a specific sequence that flanks the right and left ends of the target DNA sequence. The PCR product will contain the DNA sequences of the primers and the DNA between the primers.
Step 1: The DNA template and primers must be converted to single-strand DNA. In DNA replication, DNA is separated by helicase. In the lab, we can accomplish the same thing by heating the DNA to 98oC (almost boiling temperature).
Step 2: In order for the DNA to be copied, the primers must bind. 98°C is too hot for the hydrogen bonds between complementary bases to form. Therefore, the reaction is cooled to ~55°C. This allows the hydrogen bonds between the primer and the target DNA sequence to form.
Step 3: In step 3, the temperature is raised to ~78°C, which activates DNA polymerase and stimulates copying of the DNA. The polymerase begins at the primer and copies the DNA to the end of the template.
After this process is complete, the temperature is again raised to 980C and a new cycle begins. At the end of each cycle, the sample contains double the amount of DNA template.
Watch the following video to review the steps of PCR.
You might wonder how the DNA polymerase, as a protein, can resist such a high temperature. In fact, the DNA polymerases used for PCR are derived from organisms that can live at extremely high temperatures. The most common DNA polymerase used for PCR is called Taq, and it is found in the microbe Thermus aquaticus, which was first discovered in the Lower Geyser Basin at Yellowstone National Park.
Much of our DNA contains a number of short DNA segments that are repeated over and over again at different locations in our genome. The number of these tandem repeats at each location in the genome can differ from individual to individual, with 5 to 20 percent of individuals having the same number of tandem repeats at the same location.
Variations in the number of tandem repeats are detected by the polymerase chain reaction, using primers that bind to unique sequences on the edges of the repeat.
If the two DNA samples show a different number of repeats, then the samples must have come from two different individuals. If two DNA samples show the same number of repeats at one location, then there is a chance that the two DNA samples came from the same individual. This is not a very reliable indication that the DNA samples came from the same individual, since a significant fraction of people can have the same number of tandem repeats at any one location in the chromosome. To increase the reliability of a positive match between the two samples, a total of 13 different locations of tandem repeats are amplified, and the number of repeats at each location are compared. If any of the 13 locations differ in the number of repeats, the two DNA samples came from different individuals. However, if all 13 agree, there is a greater than 99.99 percent probability that the two DNA samples came from the same individual. There is still a 0.01 percent chance that the DNA came from two different people who just happened to have the same number of tandem repeats at each of the 13 different locations.
In the previous module, you learned how genetic information is stored in DNA and how it is passed in its totality to the next generation through the process of DNA replication. You also read about PCR, a practical application of the DNA synthesis process. In this module, we will focus on how this information is expressed. Information contained in DNA is organized into genes, and each gene contains instructions for making a functional product: an RNA molecule or a protein. This process of making the functional gene product is gene expression. Different cells can contain the same DNA, but they express different genes and produce different proteins according to their function. For example, red blood cells produce the protein hemoglobin, which is required for delivery of oxygen throughout the body. Nerve cells contain the same DNA, but the hemoglobin gene is not expressed. Instead, nerve cells make RNA copies of the genes that encode proteins required for signaling.
Even cells of the same type may express different genes and therefore produce different proteins, depending on their needs or environmental conditions. For example, pancreatic beta cells produce insulin when blood sugar levels are increased. Weightlifting increases muscle mass because it increases the production of proteins in the muscle cell. Sex hormones are not produced until the beginning of puberty.
Multiple proteins expressed at specific levels at specific times in the cell determine the phenotype. In this module, we will consider each of these processes in more detail, and we will conclude by looking at how changes in DNA sequence can affect the proteins encoded by DNA.
The flow of information in a cell is described by the central dogma of molecular biology.
In eukaryotic cells, DNA stores the genetic information, which is passed from generation to generation through the process of DNA replication. During replication, the whole sequence of DNA is copied. In order for the information stored in DNA to be used by a living cell, the information must be expressed (gene expression). First, DNA is transcribed into one form of RNA called messenger RNA, or mRNA, through the process of transcription; then mRNA can be translated into an amino acid sequence forming a protein through the process of translation. The ribosome translates the RNA sequence into a protein sequence, using three nucleotide bases to encode each amino acid DNA also encodes two other RNA molecules, transfer RNA (tRNA) and ribosomal RNA (rRNA). Although these RNA molecules do not directly encode proteins, they are both required for protein synthesis.
Transcription and translation are often confused. Transcription is copying the information without changing the language. This word comes from the same root as scribe, a word that describes someone who made copies of books in medieval times. DNA and RNA are considered the same chemical “language” because their monomers are nucleotides.
Now, compare transcription to translation. Just as translators of books change information from one language to another, the cellular process of translation changes information from the language of nucleic acids (using nucleotides as the monomers) into the language of proteins (using amino acids as monomers).
Note that there are exceptions to the central dogma. It is thought that during the dawn of life, RNA was the first molecule to store genetic information, and certain viruses still have RNA as their nucleic acid component. In fact, certain viruses (among them the human immunodeficiency virus, or HIV) have an enzyme capable of making DNA using RNA as a template. However, you will recall that viruses are not cells, so it is safe to say that cells use DNA as their genetic material, and the flow of information goes as follows:
Gene expression, a process that is central to all living cells, can be summarized as a list of the following steps:
In eukaryotes, DNA and ribosomes reside in separate compartments, so an additional step is required: the transportation of RNA out of the nucleus and into the cytoplasm.
The animation below illustrates the process of gene expression in eukaryotes. Watch the animation, and then label still frames of the animation in the Learn By Doing activity.
During gene expression, information stored in DNA is first copied (or transcribed) into RNA in a process called transcription. Only one of the two DNA strands is transcribed, and the order of bases in RNA is determined by the base pairing of the ribonucleotides with the DNA. In RNA, the base T is absent and U is present instead. As you know, A pairs with T, but because T is replaced by U in RNA, A pairs with U in RNA. If the transcribed DNA strand has the sequence
5’...AATTGCGC...3’ |
3’...TTAACGCG...5’ |
and the bottom stand is used as a template to produce the RNA copy, the sequence of the RNA will be
5’ ...AAUUGCGC...3’ - transcribed RNA strand |
3’...TTAACGCG...5’ -template DNA strond |
Note that it the transcribed RNA has exactly the same sequence as the upper DNA strand, but with T replaced by U.
The goal of translation is to make proteins. The sequence of amino acids in a protein is specified (or encoded) by the sequence of bases in the nucleic acid. You will recall that originally, the information is stored in DNA, but during transcription, an RNA copy of DNA is made. RNA molecules that contain instructions for making a protein are called messenger RNA (mRNA). The sequence of bases in the mRNA determines which amino acid will be used in the protein. Every three bases encode for a single amino acid. These three bases are called codons or triplets.
The genetic code table below shows all possible combinations of three bases and the corresponding amino acids that the codons encode. For example, the mRNA codon UGG encodes for the amino acid tryptophan. Note that there are 64 codons and only 20 common amino acids. Three of the codons do not specify amino acids, but instead signal the end of translation; these are called stop codons. The remaining 61 codons encode for 20 amino acids. Of these, the codon for the amino acid methionine, AUG, has a dual role. It is used to signal the start of most protein-coding regions in the mRNA; thus, it is the start codon. The same codon is used if methionine occurs at an internal position in the protein sequence.
The genetic code is redundant; some amino acids are specified by multiple codons. This redundancy will become important when we try to predict how changes in nucleotide sequences affect the amino acid sequences of proteins. Remarkably, the genetic code table is almost universal, meaning that most organisms, from bacteria to humans, specify the same amino acids using the same codons. As a result, genes from one type of organism can be transferred to, and expressed in, the cells of another organism. You will learn more about these kinds of technologies later in this unit.
Examine the genetic code table and complete the Learn By Doing activity to make sure you understand how to use the table.
This table summarizes the amino acids that result from each codon on mRNA. The first letter in the codon is on the left. The second is on the top, and the third is on the right.
The genetic code table allows us to convert the information contained in a nucleic acid into an amino acid sequence. How does the cell do this during translation? Translation from one language to another requires a dictionary, a book that connects words in two different languages. In the cell, we need a molecule that connects codons and amino acids. That molecule is transfer RNA (tRNA). Each tRNA contains three bases that can pair with a particular codon in mRNA (complementary base pairing). Each tRNA also carries a specific amino acid. The tRNA binds mRNA via base pairing and brings the correct amino acids to the growing polypeptide chain.
The last type of RNA molecule that is critical for translation is ribosomal RNA (rRNA). rRNAs are components of the ribosome. The ribosome is a large macromolecular complex composed of rRNA and ribosomal proteins. The ribosome catalyzes the formation of peptide bonds between the amino acids. Some ribosomes are located in the cytosol and translate cytosolic proteins. Some ribosomes are attached to membranes; these ribosomes translate proteins destined to be inserted into the membrane or secreted outside the cell.
To summarize, three types of RNA carry out distinct roles in translation: mRNA encodes the protein, tRNA brings amino acids to mRNA, and rRNA is a component of the ribosome required for catalysis.
In the exercise below, you will simulate the process of transcription and translation and build a small protein. Complete the activity and then answer the questions below.
None of the processes discussed so far are free of errors. Errors can be made during transcription if incorrect nucleotides are incorporated into the growing RNA strands. Errors can happen during translation if incorrect amino acids are inserted into a growing protein, or if translation is terminated too soon. Errors during gene expression have generally short-lived effects and can be overcome by additional RNA or protein synthesis.
Errors during DNA replication can have a much longer lasting effect. If incorrect nucleotides are incorporated into a newly synthesized DNA strand, the error may be passed on to future generations. Such heritable changes are referred to as mutations. Some mutations happen due to large scale rearrangements of DNA molecules that involve thousands or millions of nucleotides. We will limit our discussion here to small-scale changes that involve one or a few nucleotides in the coding sequence of a gene. These mutations include substitution of a single nucleotide base and insertion or deletion of one or more bases.
Let us consider how different types of mutations affect the proteins encoded by DNA. The sentence below represents a protein-coding gene. Each three letter word carries some meaning – equivalent to each amino acid being encoded by a three-letter codon in DNA.
THE CAT ATE THE HEN |
THe CAT ATE THE HEN |
THE BAT ATE THE HEN |
THE stop |
Deletion of one nucleotide: introduces frame shift
TH|C ATA TET HEH EN (deletion of one base, the "E" in "THE") |
Deletion of three nucleotides removes information, but no frame shift.
THH ECA TAT ETH EHE N (insertion of one base, an extra "H" in "THE") |
There are many possible variations on the mutation types above. For example, three bases may be deleted in a group corresponding to one existing codon. In this rare instance, there would be no frame shift but a single amino acid would be missing from the encoded protein:
THE | ATE THE HEN |
Next, we apply the classification above to an actual mutation observed in human DNA.
The activity below deals with a mutation in the gene that encodes the oxygen transport protein hemoglobin. The mutation causes hemoglobin to form long rods inside red blood cells. As a result, it is difficult for the cells to pass through blood vessels. This leads to a disease called sickle cell anemia in many of those who inherit the mutation.
In this module you learned about the role of DNA as the molecule that stores genetic information. Any damage to the DNA molecule can have serious consequences for the cell. There are many different cellular mechanisms to repair DNA damage or to stop a cell with errors in its DNA from dividing. The BRCA genes code for proteins involved in detecting and repairing double strand breaks in the DNA molecule.
If there is a defect in either of the genes BRCA1 or BRCA2, that will result in increased chances of DNA errors, which ultimately can increase the likelihood of cancer. BRCA genes are therefore considered tumor suppressor genes. Mutations in these genes produce a hereditary breast-ovarian cancer syndrome in affected families.
Women with harmful mutations in either BRCA1 or BRCA2 have an increased risk of breast cancer that is about five times the normal risk, and a risk of ovarian cancer that is about 10 to 30 times normal. While mutations in BRCA1 and BRCA2 account for only five to ten percent of all breast cancer cases in women, it is suggested that women with familial breast or ovarian cancer (when several female close relatives have or had cancer) be tested for BRCA mutations. Testing positive for BRCA mutations does not mean the person will definitely get breast cancer, because there are many other genetic and environmental factors involved. However, having this information may help those individuals to take preventive measures, such as more frequent mammograms.
You have probably seen Siamese cats--those graceful creatures with darker fur on their faces, tails, and paws. This is an example of a phenotype dependent on the interaction between genes and environment. Modern Siamese cats have a defective tyrosinase enzyme; tyrosinase is involved in the synthesis of melanin, the dark pigment of the skin. This defective enzyme is inactive at normal body temperatures, but it becomes active in cooler areas of the skin. This results in dark coloration in the coolest parts of the body, including the extremities and the face, which is cooled by the passage of air through the sinuses.
In this module, we will explore the connection between DNA and phenotype on the molecular level. We will start by clarifying the distinctions between genes, genomes, alleles, and other related terms. We will compare prokaryotic and eukaryotic genes and genomes, and we will look at how DNA is packaged into chromosomes with the help of specialized proteins. We will consider how gene expression can be regulated in the cell by interactions with other genes and environment.
Throughout the genetics unit, we have been using the terms gene, allele, and DNA, along with several other related terms. Before we go on, we need to clarify the relationships among these terms. First, we need to distinguish between terms that relate to actual molecules and terms that relate to the information content of those molecules. The macromolecule that carries genetic information is DNA, which in the eukaryotic cell binds specialized proteins to create compact structures called chromosomes.
All the hereditary information contained within an organism is called a genome. The genome includes genes and noncoding DNA sequences. The human genome, for example, contains the DNA sequence (information) of 22 autosomal chromosomes, as well as the two sex chromosomes X and Y. The field of study that focuses on the properties of genomes is called genomics, which is related to but distinct from genetics, which focuses on individual genes or a group of genes.
Genes carry the information required to produce a functional product (protein or RNA). The information is contained in the order of bases in the DNA. This information is “read” by the cell to produce, first RNA, and then a protein, during a process called gene expression. When a protein affects an observable characteristic of an organism (remember, observable characteristics represent an organism’s phenotype), the gene encoding that particular protein can be linked to that characteristic or character (e.g., hair color, eye color, or height). Traits are specific variations in characteristics (e.g., black hair or blue eyes). For example, the gene encoding growth hormone (a protein) affects the height of an individual; in this case, height is the characteristic, and tall or short stature are the traits. Some genes can be linked to the development of diseases. Nearly all observable characteristics depend upon multiple genes. Therefore, when we hear someone refer to the “obesity gene,” “Alzheimer’s gene,” “breast cancer gene,” and so on, it is important to realize that these labels imply (or should imply) one of the multiple genes that contribute to each of these conditions.
What is the connection between genes and traits? We must now consider alleles, which are observed variations in the sequence of bases for a particular gene. Two alleles of a gene may differ by a single base. On one end of the spectrum, an allele may refer to a total deletion of a gene. A diploid organism has two alleles of each gene, one from the organism’s mother and one from its father. The combination of the two alleles of a particular gene is the genetic determinant of the phenotypic traits of the organism. Note that in addition to genetic factors, environment also contributes to traits. We will consider environmental effects in a later section.
What is the difference between allele and mutation? Any change in the DNA of an organism is called a mutation. Mutations can be advantageous, harmful, or neutral (in which case they have no effect on the phenotype). Mutations include large deletions and rearrangements in the DNA, but they can also be changes in a single base. Mutation is a more general term than allele because mutation does not necessarily involve a specific gene, but can include deletion or rearrangement of multiple genes. Multiple alleles of a gene arise by mutation, and they will be passed on to the offspring and will persist in the population.
In the previous module, we mentioned different kinds of genes. Some genes produce functional RNA (rRNA and tRNAs), but most genes code for proteins. Remember that a gene is information residing in the order of bases in a DNA molecule. The general organization of information of any gene can be represented as follows:
In this module, we will focus on the structure of the protein-coding genes. A protein-coding gene contains information for making the mRNA that is translated into a protein. First, let’s consider what kind of information in the mRNA is required to produce a protein during translation, and then we will step back and look at the information in the gene that is required for transcription.
To encode a protein, mRNA needs a protein-coding region and regulatory regions. The coding region of the gene (and the coding region of mRNA) has to start with a start codon: ATG in DNA and AUG in mRNA code for the amino acid methionine, Met. The coding region of a gene always ends with a stop codon TAA, TAG, or TGA (in mRNA, these will have a U instead of a T). The sequence between start and stop codons should contain nucleotides in multiples of three, encoding the sequence of amino acids in the protein. The untranslated regulatory regions (denoted by UTR) include a site for the ribosome to bind before the start codon (5’ UTR) and a region after the stop codon (3’ UTR). Both 5’ and 3’ UTRs can bind specific proteins involved in regulation of translation (how fast new protein product is built, how fast mRNA is degraded).
Note that mRNA in the figure above starts with 5’ UTR, not with the start codon. The start codon signals the start of translation, not transcription. Next, we will consider DNA regions required for transcription.
Genes include regulatory regions at the beginning and the end of the transcribed DNA:
Putting the regulatory signals associated with RNA production together with those signals on the mRNA, we can represent the central dogma of molecular biology with the following diagram:
After considering individual genes, let’s zoom out and consider a more global view: how genes are organized into genomes.
The prokaryotic genome resides on a circular piece of double-stranded DNA; there is no membrane to separate DNA from the rest of the cell. Most bacterial genomes are several million nucleotides long. The E. coli genome, for example, is approximately 5 million nucleotides long. Although this is a relatively small genome, the physical length of the DNA is longer than the length of the bacterial cell. In order to condense the DNA into a smaller size, the DNA is supercoiled. In supercoiled DNA, the double helix of the DNA contains additional turns beyond the normal one turn per 10 base pairs. These additional twists introduce strain into the DNA that is relieved by large-scale twists in the entire DNA molecule, or supercoils. A good model for supercoiling is an old-fashioned spiral phone cord. If you form the cord into a circle, and then rotate or twist one of the free ends relative to the other, you will introduce supercoils into the cord.
Many bacteria contain smaller, circular DNA molecules, called plasmids, in addition to the large DNA molecule that is their chromosome. Plasmids can replicate independently of the large DNA and can be transferred to other cells. Some naturally occurring plasmids carry genes encoding toxins and proteins that make bacteria resistant to antibiotics.
Plasmids provide a very useful tool for biotechnology. Scientists use them to insert “foreign” genes into bacteria so they can use the bacteria as “factories” to produce desired proteins. For example, to study a human disease caused by a malfunction in a specific protein, scientists need a large amount of the protein for experiments. Instead of purifying protein from human cells, scientists can combine coding DNA from the human genome with bacterial regulatory DNA regions on a plasmid, and introduce the plasmid carrying human gene into the bacteria. Growing a large amount of bacterial cells is much easier than obtaining large amounts of human cells. Combining DNA from different organisms is called recombinant DNA technology.
A eukaryotic genome resides on multiple linear DNA molecules in the nucleus. Each of these DNA molecules is called a chromosome. Eukaryotic cells also contain DNA in mitochondria and chloroplasts; mitochondrial and chloroplast genomes are circular like those of bacteria, and are typically considered separately from eukaryotic genomes.
In general, eukaryotic genomes are much larger than prokaryotic genomes, and they contain a lot more noncoding DNA. This presents a problem of packaging for eukaryotic -cells: how are they to fit very large DNA into a small compartment? Although eukaryotic DNA is supercoiled, this isn’t enough to solve the packaging problem. Eukaryotic cells use special proteins (histones) to wind up DNA and fold it into nucleosomes, which are compact structures. DNA can be transcribed in this state. Further packaging of DNA into highly compact fibers makes genes inaccessible for transcription, leading to gene silencing. Gene silencing is an important way of regulating gene expression.
The shape of the chromosomes changes throughout the cell cycle. For example, when a cell needs to divide, DNA is further compacted by the addition of scaffolding proteins that help divide copies of the chromosomes equally. The most compact shape of the chromosomes can be observed during cell division.
In the previous sections of this module, we discussed the structure and organization of genes and genomes. Throughout the molecular genetics module, we have been pointing out the connections between genotype and phenotype. Here, we will discuss the other important factor in determining the phenotype of an organism: environment. The effects of environment on phenotype can be fairly subtle or pretty dramatic. In certain groups of cold-blooded animals, such as Nile crocodiles, sex is determined by an environmental factor: average temperature during the middle third of the incubation period. Males will only hatch at temperatures between 89.1° and 94.1°F. Nests with higher or lower temperatures will produce predominantly females.
How does environment exert its influence on phenotype? Recall that at the molecular and cellular levels, phenotype depends on the molecules produced by the cell. Production of all molecules is catalyzed by protein enzymes that are encoded by the genome of the organism. Ultimately, if phenotype depends on protein function, then the environment must have a way of modifying the amount and the activity of proteins. Using the central dogma of molecular biology, we can deduce that environment may affect protein function at the level of protein, mRNA, or DNA.
Cells and organisms have to respond to changes in their environment. An organism that experiences a sudden increase in temperature will respond with increased expression of genes coding for heat-shock proteins; these proteins protect cells from damaged and unfolded proteins. A bacterial cell that depletes glucose in its environment will not die of starvation if an alternative energy source is available. In the presence of lactose, bacteria can "turn on" the lac operon containing genes required for metabolism of lactose. Regulation of this operon will be discussed in detail on the next page. The signaling occurs primarily through binding interactions between molecules and modifications of the proteins that often lead to changes in gene expression. These changes are short-lived and are not heritable.
Long-term heritable changes can be produced by the environment at the level of DNA. Any environmental influence that can damage DNA has the potential to create mutations that can result in changing amino acid sequence of a protein or regulatory information. For example, high-energy radiation causes double-stranded breaks in DNA that can lead to deletions. UV light causes single nucleotide changes in DNA. Environmental toxins and viruses also cause mutations. Cumulative DNA damage in somatic cells throughout an organism’s lifetime contributes to the aging process. Only mutations in germ cells result in heritable changes passed on to the next generation of offspring.
Heritable changes in gene expression can occur via changes in DNA sequence or in DNA packaging. Mutations in regulatory regions of the gene can lead to increased or decreased protein production without affecting the amino acid sequence of the protein. One example of such mutation is found in the regulatory region of the human LCT gene on chromosome 2. The LCT gene codes for the enzyme lactase, required for breaking down lactose into glucose and galactose. In many adults, expression of the LCT gene is turned off or turned down, resulting in little or no lactase production. Lack of lactase in the intestines can lead to gastrointestinal distress (bloating, cramps, diarrhea, etc.) after consumption of moderate to large amounts of dairy products, giving rise to a condition called lactose intolerance. Other people are able to produce lactase well into adulthood. This ability has been linked to mutations in the regulatory region of the lactase gene that allow gene expression to continue after childhood.
Heritable changes in DNA packaging do not involve any changes in DNA sequence. This type of inheritance is epigenetic (from the Greek epi-, meaning over, above, or outside). You will recall from the previous page that compact packaging of DNA leads to gene silencing (transcriptionally inactive genes). Compact packaging of DNA requires modifications to DNA itself and to histones, which are DNA-binding proteins involved in packaging. DNA modification involves the addition of a methyl group (-CH3) to specific bases on DNA (cytosine methylation), but does not change the sequence of DNA bases. The methylation pattern can be preserved during DNA replication, so the daughter cells will have the same methylation pattern as the parent cell. In germ cells, the pattern of DNA methylation can be passed through the gametes to the next generation of offspring.
One example of epigenetic silencing in somatic cells is common to all female mammals. One of the two X chromosomes in females is inactivated by epigenetic mechanisms (DNA methylation and histone modifications). The results of X-inactivation can be observed in calico or tortoiseshell cats. These cats are almost always female. The gene encoding the coat color resides on the X chromosome. A female cat (with two X chromosomes) could potentially carry one allele for a black coat and one allele for an orange coat. As a result, her phenotype will end up with patches of two different colors. This is because cells that inactivate the X chromosome with the black allele will express the orange allele, and after multiple cell divisions, a patch of orange fur will result. Cells that inactivate the X chromosome carrying the orange allele will give rise to black patches. The specific pattern of black and orange patches is not heritable because X-inactivation is random.
Epigenetic regulation and inheritance has been linked to cancer and obesity, aging and longevity, and other important processes. It is the key mechanism for interactions between genes and environment. Unlike the genome, which remains fairly unchanged throughout an organism’s life cycle, the epigenome (consisting of the pattern of DNA methylation and histone modifications) is dynamic, responds to the environment, and can be heritable.
In summary, environment can influence phenotype by interacting with the genome in the following ways:
In this spotlight, we will explore in more detail changes in gene expression (transcription) due to changes in the environment. Eukaryotes have complex regulatory mechanisms; consequently, we will focus on gene regulation in prokaryotes. However, common principles of regulation are used by both prokaryotic and eukaryotic organisms, and an understanding of regulation in prokaryotes will help you understand eukaryotic regulation in future studies.
Many bacterial genes are expressed at the same time as part of an operon. The expression of the genes in the operon is controlled by a repressor protein. The repressor protein will bind to the regulatory region at a DNA sequence. This sequence is the operator. When the repressor is bound to the DNA, RNA polymerase cannot transcribe the structural genes because it is blocked by the repressor protein bound on the DNA; under these conditions, none of the enzymes required for the pathway is produced. Whether or not the repressor binds to the operator is controlled by the binding of a small molecule to the repressor. The small molecule is either the compound that enters the pathway (for catabolic, or degradative pathways) or the compound that is produced by the pathway (for anabolic, or synthetic pathways). The compound binds to the repressor protein and causes the repressor either to bind to the DNA (this is called positive control), turning off transcription, or to be released from the operator site on the DNA (this is called negative control), turning on transcription. You will explore both of these control mechanisms in the spotlight below.
Bacteria can use a wide variety of carbon sources as fuel to extract energy for growth. Two common carbon sources are the sugars glucose and lactose (milk sugar). Glucose is a monosaccharide containing one six-membered ring, while lactose is a disaccharide made of glucose linked to galactose.
It is less efficient for bacteria to use lactose as an energy source because they must take additional steps to convert lactose into a form that can be used to extract energy. Each of these steps requires the synthesis of an enzyme, which is costly to the bacteria. Consequently, bacteria prefer to use glucose rather than lactose as a fuel source when both sugars are present. However, when lactose is present and glucose is not, the bacteria “turn on,” or synthesize, the enzymes required for lactose metabolism, allowing the bacteria to use lactose for energy.
The genes for lactose metabolism are contained in an operon known as the lac operon. The lac operon is controlled by the lac repressor, which is produced from the lacI gene. The lacI gene has its own promoter; consequently, low levels of the lac repressor protein are present all of the time. Three structural genes in the operon, lacZ, lacY, and lacA, encode enzymes required for lactose metabolism.
Thus far, we have described what happens when either glucose is present or lactose is present. If both are present, glucose is used first by the bacteria, and only very small amounts of the enzymes for lactose metabolism are made. Complete the following Learn by Doing to understand how glucose turns off the lac operon, even if lactose is present.
The enzymes for the synthesis of the amino acid tryptophan are contained in the Trp operon. The structure of this operon is shown below. The trp repressor binds the amino acid tryptophan, the end product of the biosynthesis of tryptophan (trp). As you can see, it has the same structure as the lac operon, but it contains more genes for the biosynthesis of trp.
The ability to manipulate DNA has led to dramatic changes in medicine, agriculture, and many other aspects of society. For example, genes from other species are transferred into crop plants and livestock to provide them with desired traits. And many drugs are produced by bacteria, yeast, or other organisms that have been engineered to express human genes.
In this module, you will learn about how biotechnology is used to produce human growth hormone (huGH) for medical applications. HuGH is a protein made of 191 amino acids. It is necessary for the normal growth and metabolism of an individual. Worldwide, one in 4,000 to 10,000 individuals is born without the ability to make sufficient human growth hormone, causing a host of medical problems. These people can be helped with growth hormone injections, but growth hormone from other species (pigs, cattle) cannot be used. Only human growth hormone will produce the desired effects. Historically, the only source of human growth hormone was a small amount that could be extracted from recently deceased individuals (cadavers). this severely limited the supply that was available for medical use.
You will also learn about how biotechnology has solved this problem and decreased the cost of medical treatments that significantly improve the quality of life for many people.
The basic steps in producing human growth hormone (HuGH) in bacteria are:
The synthetic gene provides the information that specifies the order of the amino acids in HuGH. The expression plasmid is a large closed circular piece of DNA. Various DNA sequences on the expression plasmid are required for maintenance of the plasmid in the cells. These sequences are present on all plasmids, regardless of the type of protein that is being made. The plasmid also contains DNA sequences (promoter for RNA polymerase, lac operator) that are required for the production of the mRNA that encodes HuGH. The bacterial cell provides the protein synthesizing machinery so that the mRNA can be translated to a protein. In order for the protein to be correctly made the mRNA must contain a ribosome binding site, a start codon, and a stop codon.
More recently, proteins used for medical treatment are produced in yeast or cultured mammalian cells that are grown in synthetic growth media. Protein production in yeast or mammalian cells avoids the possibility of having toxic bacterial impurities in the final product. The overall steps in producing the protein in yeast or mammalian cells are very similar to the steps described here, however the overall cost of producing the protein may be somewhat higher.
The first step in producing HuGH is to generate a synthetic gene whose nucleotide sequence codes for the amino acid sequence of the growth hormone. This sequence would contain the start codon, codons representing each amino acid in the growth hormone, and a stop codon. To do this, the known amino acid sequence for human growth hormone must be “back-translated” to a nucleic acid sequence. Since most amino acids can be coded for by more than one mRNA codon, this process is ambiguous. Because the genetic code is universal it would be possible to use the human codons to encode the sequence of amino acids. However, the host organism (which is bacteria in this example) often prefer certain codons over others, so the preferred codons must be used to optimize protein expression.
We will synthesize our HuGH gene using a chemical method called solid-phase synthesis. In this method, the first base of our DNA sequence is attached to a small glass bead. The next base is then added to the first, and so on, until the sequence is completed at which point the DNA is released from the glass bead. Any desired DNA sequence can be generated by this method. Current technology limits this approach to about 200 bases. If our gene is longer, we have to make it in segments and then join the individual segments together.
In addition to the codons, it is necessary to add additional sequences to the end of the gene to facilitate the insertion of our synthetic gene into the expression plasmid. These sequences are called restriction sites. Consequently the complete gene is:
After you have generated your synthetic gene, you must insert the gene into an expression plasmid so it can be placed inside a living bacterial cell and translated into the HuGH protein. We will look at that step in detail on the next page. For now, let’s assume that has been accomplished. Here we want to focus on the properties of the expression plasmid.
An expression plasmid is a closed circular molecule of DNA that has several unique properties associated with its DNA sequences. There are three general types of sequences that are found on expression vectors:
The step of inserting our synthetic DNA into the plasmid use two very important enzymes, restriction enzymes and DNA ligase. Restriction endonucleases are enzymes that recognize specific DNA sequences, such as GAATTC, and cut both strands of the DNA within these sequences. Typically four or six bases are recognized. Restriction enzymes are isolated from a number of bacterial species and the name of the enzyme reflects the original species. For example, the restriction enzyme EcoR1 was isolated from Escherichia coli. The normal biological function of restriction endonucleases is to protect the bacterial species from viruses by digesting the viral DNA as it enters the cell. Of course, the bacteria shouldn’t digest their own DNA. Consequently, they have companion enzymes called DNA methylases, which add methyl groups to the bacterial DNA, protecting it from digestion.
There are hundreds of different restriction enzymes, each of which recognizes and cleaves a different DNA sequence or “restriction site”. Examples of two such enzymes are shown below. A shorthand notation for the recognition site is given below each name. The “^” symbol indicates the site of cleavage. The DNA sequence that they recognize is shown on the left and the products of the digestion are shown on the right. The fragments that are produced after cleavage can contain single stranded overhangs (e.g. EcoR1) or have blunt ends, such as with PvuII.
The DNA sequences that are recognized by restriction endonucleases have a unique property in that the sequence on the top strand is identical to the sequence on the bottom strand, i.e. in the case of EcoR1 the sequence of the top strand is 5’-GAATTC. Taking the complement of that sequence to generate the bottom strand gives: 5’ GAATTC. The other notable feature is that the cutting position is in the same location on both strands. Using EcoR1 as an example, the enzyme cuts between the G and the A. The symmetry in the recognition sequence and the cutting location is due to the fact that the active form of these enzymes contain two identical polypeptide chains. You can imagine that one chain recognizes and cleaves the top strand. The second chain, because it is identical to the first, recognizes the same sequence on the other strand and cleaves in the same location.
If the restriction enzyme does not cut at the center of its recognition sequence the cleavage products will contain short single-stranded segments of DNA. These are often called “sticky-ends” because they can stick to other DNA molecules that have complementary sticky ends. Enzymes that cut in the middle of their recognition sequences produce blunt end fragments.
Below, the formation of sticky ends after EcoR1 cleavage is illustrated.
Because cleavage does not occur in the middle of the recognition sequence, the products have short single strand segments (highlighted). These sticky ends can “stick” to complementary single-stranded DNA sequences using normal DNA basepairing (A-T, G-C).
The presence of sticky ends on DNA fragments makes it easy to join DNA fragments. If two fragments are both cut with the same restriction enzyme, they will have complementary sticky ends. These will align and bind loosely to each other: hydrogen bonds will form between complementary bases.
At this point, the two fragments are not firmly held together by covalent bonds; the sugar-phosphate backbone is still broken on both strands. DNA ligases can be used to rejoin DNA after it has been cut with the restriction enzyme. This reaction requires energy (usually ATP) and requires that the two pieces of DNA that are being joined are close to each other. Since the sticky ends hold the two ends close together it is easy for DNA ligase to join the fragments, reforming the phosphodiester bonds on both strands.
Although most of the time DNA fragments are joined using sticky ends, it is possible to join fragments that are blunt ended. This is more difficult however because there are no hydrogen bonds to hold the DNA fragments together.
Insertion of the Human Growth Hormone Gene Into the Expression plasmid
Restriction endonucleases and DNA ligase are used to insert the synthetic gene into the correct location in the expression plasmid, just after the ribosome binding site.
The expression plasmid would have two different restriction sites after the ribosome binding site to facilitate the insertion of the synthetic gene. To utilize these sites it would be necessary to include these restriction sites at the end of our synthetic gene. The expression plasmid and the synthetic fragment would be cut with the restriction enzymes, mixed, and then treated with DNA ligase to re-seal the phosphodiester backbone:
Once the synthetic gene has been successfully spliced into the expression plasmid, the complete expression plasmid must be taken into the bacterial cell. This is a delicate process that must take place under special conditions. In these conditions, the expression plasmid is mixed with bacterial cells, allowing some of the cells to take up one circular DNA molecule. The origin of replication on the expression plasmid will cause the bacteria to replicate the plasmid along with its own chromosomal DNA. It is not possible to determine which cells have taken up the expression plasmid. Consequently, the cells are grown in the presence of an antibiotic and only those cells who have obtained the expression plasmid will live because of the gene for antibiotic resistance that is also present on the expression plasmid.
The cells containing the plasmid would be grown in culture media to a high density of cells to produce a high yield of huGH. Lactose would be added to the culture to cause the lac repressor to come off the DNA. This allows RNA polymerase to begin transcription of the growth hormone gene on the plasmid. The mRNA would be translated on the bacterial ribosomes, producing human growth hormone.
Thus far you have learned a great deal about how cells function. In the genetics units you’ve learned how biological information is inherited and used. You are beginning to appreciate that each cell and organism is a complex and highly integrated unit.
Evolution builds on what you already know. It is a scientific theory that explores life on a different scale and provides another layer of explanation (Table 1). First, it goes beyond the brief lifetimes of individual organisms. Instead, evolution considers how life changes from one generation to the next and how these changes accumulate over very long time scales. Second, evolution is focused on collections of individuals: populations, species, and larger taxonomic groups (for example, birds or flowering plants). Evolution also adds another layer of explanation to biological thinking. So far we have focused mainly on “how” questions. How does a cell maintain homeostasis? How is information passed from parents to offspring, and how is that information used to construct living cells? Evolution helps us answer “why” questions: Why do cells of very different organisms have so much in common? Why are cave fish blind? Rather than taking the features of life as a starting point, evolution asks why those features exist today and how they originated in the past.
Attribute | Characteristics |
---|---|
Time scale | Generations to much longer time scales |
Level of hierarchy | Populations to larger collections like taxonomic groups |
Level of explanation | Explores “why” questions. Why do features of life exist as they do? What is their history? |
As a scientific theory, evolution is more than a mere opinion or guess. In science, a “theory” is a broad explanation that is supported by a great deal of evidence and continues to inspire productive research. Scientists accept and use evolutionary theory because it is so strongly supported by evidence from many different sources. It fits our observations of the diversity and variability of physical traits in animals in different settings. The theory is now also supported by consistent and reproducible data gathered from fossils, a detailed understanding of the skeletal and organ structure of living organisms, and our twenty-first century abilities to examine and determine the structure and function of proteins and DNA in living things. Since the theory of evolution was first put forth in the 1850s and since some of the supporting evidence is based in ancient fossils, you might get the impression that evolution occurred in the past and is complete. That is not correct. In fact, evolution continues all around us today and evolutionary ideas are being applied to solve a huge range of problems of both academic and practical interest! Below are a few examples. You will learn about some of them in this unit. Evolutionary theory and models are used to:
Evolution is that and much more. In the first module, we will introduce and define the biological theory of evolution. In the second module we will focus on relatively fast and subtle changes that occur within populations (microevolution). In the third module we will consider the much more dramatic changes that occur over long time scales and generate the diversity of life (macroevolution). As you learn, try to maintain an open mind and evaluate the evidence on its own merits. You will gain a better understanding of a scientific theory that is very influential in our society.
People have been grappling with questions about the origins of life for nearly as long as we have been in existence. Every society has addressed these questions based on the information that was available at that time and place. Before the Scientific Revolution, people used many different sources of information to answer questions about the origins of life. Some of their ideas were based on direct observations, but many of them involved supernatural beings or spiritual forces. Stories were told to explain how the world and humans were made, but also to address questions of meaning: why the world exists and what human life is “for.” Ideas were passed down as oral tradition (myth and folklore) or in the written documents of organized religions. Scientific study is much more narrowly focused on close examination of processes and objects in the natural world. In order for an explanation to be considered scientific, it must be based on verifiable observations. It must be possible to test it through observations or experiments. Ideas or explanations that do not meet these criteria are not necessarily “bad” or “wrong,” but they are unscientific.
Scientific and non-scientific views are separate but potentially complementary ways of knowing. Scientific methods can be used to answer direct, factual questions about “what is” and “what probably was” through a clear procedure of observation and testing. Through an ongoing process scientists select ideas that best fit the available data. No religious tradition revealed the existence and function of DNA to humans: that was the fruit of scientific study and reasoning. However, answers derived from science are limited. They do not necessarily help with ultimate questions of meaning (why the Universe exists in the first place, what humans should be or do). Non-scientific ways of looking at the world (religion, philosophy) are more directly focused on these questions. Many scientists embrace religious ideas about the meaning and ultimate origins of life. However, they do not publish these ideas in scientific journals or used them to explain their detailed observations of the natural world.
In this course, we will focus on how the scientific community understands evolution and the origins of life. Like the community of science itself, we will restrict our scope to fairly narrow questions of how and when things probably happened, based on what we can observe today. Other explanations are better suited to philosophy or religion courses; they help us explore broader questions about why the Universe exists and what our place is within it.
Before defining the modern theory of evolution, we will consider its historic roots. You probably know that the scholar most widely recognized as the author of the theory of evolution was Charles Darwin. The first statement of modern evolutionary theory was his book On the Origin of Species, published in 1859. But Darwin didn’t make up the theory from nothing, and he wasn’t the first to ask questions about the ultimate origins of life. How did science deal with life’s big “why” questions before evolutionary theory developed? What clues led scientists to propose and embrace evolution? To gain some insights into these questions, explore the Learn By Doing activity below.
After Darwin, science has continued to reinforce, refine, and modify the theory of evolution. Two major developments were particularly important. In the early 1900s, scientists rediscovered Mendel’s genetic principles and applied them to evolution. Theorists developed models of how the gene pool of a population could change over time. Field biologists observed such changes in fruit flies and other short-lived organisms. Lab experiments also demonstrated the effects of natural selection. With a focus on genetics, these advances reinforced the idea of natural selection. In 1953 James Watson and Francis Crick discovered the structure of DNA. This laid the groundwork for an explosion of information that continues today. Inherited information now can be analyzed directly and in detail. Comparisons of DNA in different species support their common ancestry. DNA analysis also pinpoints the genetic mutations that produce new adaptive traits in organisms. To Learn More: A comprehensive database of Darwin-related publications, images, commentary and more are available free online at http://darwin-online.org.uk/. A more detailed history of evolutionary thought is available from the University of California Museum of Paleontology at http://evolution.berkeley.edu/evolibrary/article/0_0_0/history_01.
Before we begin our exploration of microevolution, let’s review a few principles that we have already covered in this course. The principles of evolution are based in genes; therefore, it is important that we remember that genes are segments of DNA that have "meaning." Genes are sequences of DNA that encode particular RNA and protein molecules, which work together to give a cell and an organism its characteristics and functions, whether they be physical, chemical, or behavioral. Most genes have multiple forms, called alleles,, discussed previously in the Heredity module of the Classical Genetics unit. For example, pea plants have a gene that determines seed color, and the different forms (alleles) of that gene are "green seed" and "yellow seed."
When you studied inheritance, you often were thinking about crosses: two individuals mating to produce offspring. As we study microevolution, we will be backing up to think about how inheritance works within larger groups. You will need to understand and use the following terms: species, population, and gene pool. The illustration below shows an example of how a species, population, and gene pool are related.
In this course we will define a species as a group of organisms whose members can and will breed with each other to produce fertile offspring. A population is all the individuals of the same species that occupy the same area and are likely to breed with one another. Some examples of a population might be all the people that live in your state, all of the frogs that live near a pond, all of the grizzly bears that live in Denali National Park, or all the dandelions that live in your hometown.
There are two defining aspects to a population. A population is strictly composed of members of the same species. For example, a population would always consist of people or of grizzly bears, but never of both people and grizzly bears. Second, the members of a population occupy the same area. Scientists define the area based upon the problem they are studying. In evolutionary studies, biologists try to define populations based on actual patterns of breeding, so that individuals most often breed within their population groups; migration and mixing between populations is less frequent.
You might wonder how a species and a population are different from one another. Populations are subsets of a species. The dandelions in Boston, Massachusetts and those in Annapolis, Maryland belong to separate populations. They are not likely to breed with one another because their pollen doesn't travel that far. However, the Boston dandelions and the Annapolis dandelions do belong to the same species. They could breed with one another if conditions allowed. For example, breeding could occur if a human being carried dandelion pollen from Boston to Annapolis.
The gene pool is all of the genes and alleles present in a population at some point in time. Just as the genotype is the genetic composition of an individual, the gene pool is the genetic composition of a population.
The pea plant example only looks at one trait of the plants, seed color, but there are many other genes present in this population. A gene pool includes all of the genes and alleles in a population, not just those for a particular trait. So the gene pool for this pea plant population actually contains genes and alleles that influence thousands of different traits. Some of these traits are visible, such as plant height and flower color. Other genes may determine how resistant the plants are to diseases and pests, how they respond to drought or freezing, and many more features that are important to peas - and to gardeners.
What is microevolution?
The word evolution means change, and things that change are said to "evolve." These terms are commonly used in our everyday lives (a work project evolves from the original idea to a finished product, people dealing with a tragedy evolve as they learn to cope), but in these settings they have a very different meaning than in biology. In biology, the words evolution and evolve are used in a precise way. These words apply to changes that occur in the gene pool of a population. Evolution occurs in populations, not in individual organisms. In biological terms, for example, individual people don't evolve, but the human race does. When evolution occurs in a population, different alleles (and genotypes) become more or less frequent within the gene pool. As a result, we observe a gradual change in the inherited characteristics of the population as one generation succeeds another.
Microevolution occurs when the type or frequency of the alleles and genotypes in a population change over one to many generations of time. But what does that actually mean? Let's look at a hypothetical example. We’ll observe two different populations of foxes, one in Maryland and one in Wisconsin. We’ll narrow our focus to a single gene with two alleles that determine coat thickness. And we will visit the populations at two different times.
As you can see from the Fox population example, evolution is not determined by the increase or decrease in the number of individuals within a population. Both populations saw an increase in the total number of foxes from 1962 to 2202. Microevolution occurs when there are changes in the genetic makeup of the population: when the frequency (percentage) of each genotype changes over time.
This example of microevolution covers a few hundred years, but in organisms that have shorter generation times, like bacteria, microevolution can be observed on an even smaller time scale. Antibiotic resistance has become a major problem for the medical community worldwide.
Now that we have a better understanding of what microevolution is, we must consider in more detail how microevolution happens. What are the mechanisms that allow for changes in the gene pool over time? In this page we will learn about selection, a central process in evolution. Selection occurs whenever some genetic types in a population reproduce more than others in a given environment. We will discuss three types of selection: natural, sexual, and artificial.
Working independently of one another in the 1850s, Charles Darwin and Alfred Wallace were the first scientists to posit the theory of evolution by natural selection. Ideas about evolution had been proposed before, but Darwin and Wallace added a crucial element to the theory: they provided a mechanism for how evolution could work.
Natural selection is a process by which nature (i.e. the environment and all of its components) impacts the evolution of a population. Individuals vary in their inherited traits. Those with inherited phenotypes that are better suited to the environment will have a greater likelihood of passing on their genes to subsequent generations. To put it another way, the driving force behind evolution is the interaction of genes and the environment.
Sometimes people use the phrase “survival of the fittest” to sum up natural selection. This catchphrase may be helpful, but we encourage you to think carefully. “Survival” is important, but only because it enables reproduction: once they reach reproductive age, some animals and plants literally kill themselves to reproduce. Examples include many salmon and annual plants, which put all of their energy into a single reproductive effort.
Second, what do we mean by “fittest”? We don’t mean physical fitness. Evolutionary fitness is based on genes and inheritance and is measured by lifetime reproductive success. The genes of an organism determine many of its features, and these features help determine how likely an organism is to reproduce in its particular environment. There are other factors at work too, including luck: nature does not line up all the individuals in each population and kill them off systematically, leaving only those with the fittest genes to reproduce. Instead, we say that on average the individuals with some phenotypes will reproduce better than others in a given environment. These fitter individuals will be more likely to pass on their alleles (through reproduction) to the next generation.
Selection occurs whenever some genetic types in a population reproduce more than others in a given environment; in natural selection, the favored types do best because:
The genetic types that reproduce more than others are more likely to pass on their alleles to the next generation. Traits that give an organism a reproductive advantage in a particular environment are called adaptive traits. For a white-footed mouse in a forest, a long tail could be adaptive because it allows a mouse to escape from predators by climbing trees. It could also be adaptive because it allows a mouse to obtain more food. In either case, the result is that the long-tailed mice reproduce more, on average, than other types in the population.
Each individual birth or death changes a population’s gene pool by a tiny fraction: some individuals pass on their alleles through reproduction, others do not. This filtering process affects the composition of the gene pool of the population, and ultimately of the species as a whole.
Let's return now to our example of the Wisconsin foxes to see how genes and their environment might interact to drive the evolution of a population. Remember that this population evolved, showing a dramatic increase of ff foxes with thin coats from 1962 to 2202:
In our fox example it is important to remember that the environment did not change the alleles of individual foxes. The foxes did not all individually lose their coats as the weather got warmer. Instead, the environmental conditions changed the likelihood that a particular set of alleles would be passed from individuals in one generation to their descendants. The environment increased the odds that the f allele would be passed on to the next generation, and decreased the odds that the F allele would be passed on. As a result, the frequency of the FF genotypes in the population's gene pool decreased while the frequency of the ff types increased. We see this change in frequency reflected in an increase in the percentage of foxes in the population that are born with genes for a thin coat (i.e., we observe a change in the frequency of the thin coat phenotype). Evolution takes longer than a lifetime. Second, the warmer climate did not cause a mutation to occur so that thinner coats could exist. Instead, selection is limited to filtering existing variation. Diversity already exists within the alleles of a population and the environment places a selective pressure on that population. The evolution of a population is a reactive and automatic process, not a planned or deliberate one.
Sexual selection is a subtype of natural selection. In natural selection, some genetic types in a population reproduce more than others because they are more likely to survive or because they obtain more resources. In sexual selection, some types reproduce more than others because they have traits that allow them to:
Can you think of some traits in animals or plants that would increase their likelihood of mating? Here are some examples.
Female peafowl are more likely to mate with males that have fuller, more colorful, and more grandiose tail plumage. During mating season, males display their plumage when females are in the area in the hopes of attracting the females for mating. The males with the better tail feathers are more likely to successfully mate and pass on their genes.
These are just two examples of many thousands of traits that are driven by natural selection in nature. Whenever the males and females of a species are noticeably different in appearance, sexual selection is probably at work. Some additional examples of sexually selected traits include:
The typical pattern in sexual selection is that males compete with each other for mates; sometimes they do so through displays that are meant to attract females, while in other cases they may battle each other directly. Females tend to choose mates carefully.
The perplexing thing about sexual selection is that the showy traits that are used to attract a mate can also interfere with survival. For instance, the elaborate plumage of the peacock can make the males more likely targets for predators. Elk with larger antlers may get them stuck in fences or brambles, which can lead to the untimely death of the stuck elk. And fighting is never good for survival.
So why do these dangerous traits exist? First, survival is only a means to an end: individuals that fail to mate have zero fitness, even if they are very good survivors. The gene pool will contain traits of those who have survived, yes, but also of those who have mated successfully. Second, many sexually selected traits (e.g. elaborate plumage, bright colors, large antlers) are known to be indicative of a healthier mate. Males who can “carry off” a good display and a successful battle are also fit in other ways. Females who are choosy are doubly successful in their own reproduction: they produce “sexy sons” who will themselves be successful in mating, and they also produce offspring that are generally healthy.
Sexual selection is a powerful force in evolution. To understand the impact of sexual selection consider what happens to the genes of the individual who does not mate.
In the figure above, Mark’s genes have been passed down to a second generation because his children have mated and have children of their own. Mark is successful evolutionarily speaking and his alleles have become more common through his reproduction and that of his offspring. The evolutionary success of Matt’s genes is zero because he has no offspring. Sexual selection is a powerful driving force within evolution.
Selection occurs whenever some genetic types in a population reproduce more than others in a given environment. In natural selection, the types that succeed are those that survive and/or grow the best. In sexual selection, mating success is key. In artificial selection, humans decide which types reproduce best: we deliberately breed certain individuals with desired traits.
Artificial selection is sometimes referred to as selective breeding. It has been in use for thousands of years as humans have domesticated plants and animals to suit their needs. As we discussed at the beginning of this module, corn as we know it today (with large, full ears) has been selectively bred by humans from its wild grass ancestor (with small, sparse ears). In the same way, dairy cows have been selectively bred to produce more milk; beef cattle have been bred to have more muscle; horses have been bred to run faster; tomatoes have been bred to produce larger fruit; and wild mustard has been bred to produce cabbage, Brussels sprouts, kale, broccoli, cauliflower, and kohlrabi.
There is perhaps no better example of artificial selection than the domestic dog. Dogs are descended from gray wolves, but after thousands of years of humans directing the breeding of dogs, we have well over one hundred different breeds. There are dogs with good noses that can track game and search for missing persons. Some dogs have been bred to herd or protect domestic animals. Some dog breeds are known for happy personalities and soft fur that make them great companions while others are known for their strength and intelligence which makes them good guard or police dogs. Consider the image below to get a sense of the power of artificial selection.
Selective breeding is not as simple as it may appear. Let’s say you take a dairy cow who is a high milk producer. You mate her with a male who has the red spotted coat you like. Unfortunately, you are not guaranteed to get a female calf who is a high milk producer and has red spots. You are not just dealing with the two traits you desire in a dairy cow (high milk yield and red spots), but every trait (i.e. every allele) that each parent cow has. When you breed these two animals, the offspring will have some combination of all of the parents’ traits. Because of this, it often takes many generations of artificial selection to consistently produce offspring with the desired traits.
In the process of selecting for desired traits, undesirable traits often “tag along”, and thus we see an increased likelihood of genetic disorders in many domestic animals. This is a result of inbreeding (mating of individuals that are closely related). For example, inbreeding among dalmatians has resulted in a tendency toward aggressive behavior. Additionally, sometimes selective breeding can magnify a desired trait to such an extreme that it interferes with survival or reproduction. American domestic turkeys have been bred to have larger breast muscles to satisfy consumers’ preference for white meat. As a result of generations of selective breeding, today’s male turkeys are literally too large-chested to mount and copulate with female turkeys. To produce more turkeys, farmers must artificially inseminate females. In this and other examples, selective breeding can produce plant and animal varieties that are totally dependent on our care and intervention, exhibiting traits that would never be adaptive in the wild.
Artificial selection is a powerful tool that drives the evolution of domesticated plants and animals, but it is one that must be utilized with great thought and consideration so as to avoid the problems that come with inbreeding. Does it surprise you that farmers and animal breeders need to understand a good deal about biology to be successful in their professions?
We have just learned how natural selection, sexual selection, and artificial selection can change the gene pool of a population and result in evolution. In this section we will consider other processes that can change the gene pool of a population: mutation, gene flow, and genetic drift.
Remember from the gene expression module that a mutation is a permanent, irreversible, and heritable change in DNA. We often think of mutations as a “bad” thing, in part because of how they are portrayed in TV shows and movies. Mutant creatures are often bizarre (three eyed fish in Lake Springfield of The Simpsons) or dangerous (mutant dinosaurs of Jurassic Park). Even when the mutations create superheros like the Teenage Mutant Ninja Turtles or Spider Man, the mutation is never really portrayed as a good thing. While a certain amount of that bad reputation is deserved, as some mutations do indeed have negative consequences for living organisms, not all mutations are "bad." All of the following are true statements about mutations:
Mutations in genes are random and have the potential to change the gene pool of a population. When parents produce sperm or eggs, errors (mutations) can occur so that the gametes carry new versions (alleles) of existing genes. Fertilization, growth, and development can result in an adult individual that carries a new allele and is able to pass it on to offspring. Technically, the birth of a single individual carrying a new mutation qualifies as microevolution: as soon as the new allele appears, the gene pool has been altered.
Let's consider for a moment the gene that encodes the alpha-amylase enzyme. This enzyme is found in saliva and begins the process of digesting starches in your mouth. Assume that you were born with a new mutant version of the alpha-amylase gene. You produce a form of the enzyme that very efficiently digests starch. No one else in the world carries this form of the alpha-amylase gene in their cells. So as a result of this mutation, a new allele has been introduced in the gene pool of your population and of the entire human race. The type and frequency of the alleles and genotypes in your population has changed (ever so slightly)… so evolution has occurred. Will this change have a meaningful effect on the entire population or species? That depends on what happens next.
The alleles and genotypes in a gene pool can also change via two processes that are sometimes confused with one another: gene flow and genetic drift. Gene flow refers to the movement of alleles from one population to another, as a result of migration followed by breeding. Remember that when living organisms move from one place to another, they take their alleles with them! Consider the average height difference between people in East Asia (e.g. China) and those in Western nations (Europe and the Americas). As westerners migrate to East Asia, the gene pool there changes to include more alleles that confer tall height. Conversely, alleles characteristic of Asian populations move along with immigrants to the United States. Gene flow involves breeding between two or more populations and it tends to make them more similar to each other.
Genetic drift refers to random events that change the frequency of alleles and genotypes in a population. It is similar to selection in some ways. Recall that in selection, certain genetic types become more frequent in the population because they survive well, grow well, mate successfully, or even just because humans like them and choose to breed them. In genetic drift, certain genetic types become more frequent through dumb luck.
Genetic drift often reduces the diversity of a gene pool and has the biggest effect when there is a low number of organisms in the population. In two different situations, genetic drift can profoundly affect a gene pool. The first is known as the bottleneck effect. When events like fires, earthquakes, hurricanes, and other disasters kill a large percent of the population, the surviving population is unlikely to have the same gene pool as the original population. It is very likely that some alleles will be underrepresented or overrepresented in the surviving population as compared to the original. An analogy may help. Half the cards in a deck are red (hearts and diamonds) and the other half are black (spades and clubs). But if you draw six cards from the deck at random, will you always get 3 red and 3 black cards? No. You might well draw six black or six red cards. Similarly, consider a grove of 1000 oak trees. If a flash flood washes away and kills 900 of the trees, the alleles of these trees have been lost from the population gene pool. Compared to the original group, the 100 lucky survivors will have higher percentages of some genetic types, a lower percentage of others, and some alleles might be entirely missing. The flood altered the population’s gene pool in a random way.
Genetic drift can also occur through the founder effect. In the founder effect, a small number of individuals from a population settle in a new area. This small group is not likely to contain all of the alleles of the original population, just as is seen with the survivors in the bottleneck effect. Imagine a field of dandelions. Dandelions are self-fertile, so they do not require cross pollination from other dandelions to produce offspring. If just one of the white fluffy dandelion seeds is carried by the breeze to a new location that does not have dandelions, the plant from this single seed could start a new population. As a result of the founder effect, however, the gene pool of the new population would be very limited. It would have only those alleles that were found in the original seed, drawn at random from a much larger source population.
As you can see from the examples presented here, mutations, gene flow and genetic drift are mechanisms that can alter the gene pool of a population leading to microevolution. While these mechanisms of evolution have been presented separately from the mechanisms of natural, artificial, and sexual selection, in reality it is likely that several or all of these mechanisms are at work at any given time to drive evolution. Evolution is a complicated process. It involves some relatively predictable elements (selection and gene flow) but also has many unpredictable aspects (mutation, genetic drift).
Many species have very short generation times ranging from minutes (some bacteria) to a few weeks or months (many insects, some weeds). We can track evolutionary changes in populations of these organisms. They may change in ways that are important to us. For example, disease organisms may shift to human hosts, evolve to evade vaccines, or evolve to resist drugs. Pests and weeds may also evolve in ways that make them more harmful or harder to control. Scientists and doctors can use what we know about microevolution and the mechanisms that drive evolution to help tackle these problems in the “real world” not just in the realm of academia.
In our first example of microevolution and infectious disease, we will explore the Human Immunodeficiency Virus (HIV), which causes Acquired Immune Deficiency Syndrome (AIDS). In AIDS, HIV disables the human immune system by depleting CD4 T-cells. The HIV virus prefers to infect these cells. HIV is an interesting virus in that it uses RNA to encode its genetic material instead of DNA like most viruses.
HIV has special enzymes to help it make copies of itself within the host cell. Some of these are reverse transcriptase, integrase, and protease. Each one is essential to the virus:
AIDS was first recognized and tied to the HIV virus in the early 1980s. At that time a diagnosis of HIV infection was considered a death sentence as there were no available treatments. The first breakthrough in HIV treatment came about in 1986 when zidovudine (also known as AZT) was introduced in the United States. AZT works by blocking the activity of the HIV reverse transcriptase enzyme. AZT was hailed as a HUGE medical breakthrough, and it worked well at keeping the viral infection in check. AZT would remain the only treatment available for HIV until the early 1990s when other reverse transcriptase inhibitor drugs as well as drugs that inhibited the HIV protease were introduced. With AZT and the new drugs available the scientific and medical communities thought that the struggles of managing this disease were over! Having HIV was no longer a death sentence! Infected patients could simply be given any one of these drugs, and live fairly normal lives as long as they continued treatment. However, by the mid 1990s, it became apparent that there was a problem. Patients whose infections had been well controlled (even for many years) started once again to exhibit symptoms of AIDS and their health deteriorated rapidly. While this turn of events may seem perplexing, the explanation has everything to do with microevolution.
By the late 1990s, doctors and scientists came to the conclusion that in order to continue to treat HIV effectively, they needed to approach the problem from multiple fronts - as you just discovered in the above Learn By Doing activity. New drugs were developed that targeted different parts of the virus, current drugs were administered together, and when it became clear that a drug was no longer working for a patient, the patient was given a different one. The end result of all of this work was the development of what is known today as Highly Active Antiretroviral Therapy or HAART. HAART generally consists of one protease inhibitor given in concert with one or two other anti-HIV drugs that target different aspects of the virus. To date, there are over 30 drugs approved for use in the treatment of HIV.
While doctors and scientists have made great strides in combating HIV, the “war” on this virus is far from over. In 2010, there were approximately 34 million people infected with HIV worldwide and 1.8 million deaths associated with HIV. Rates of HIV infection and the death toll from AIDS are disproportionately high in Sub-Saharan Africa and in other areas without access to HAART. Even in wealthier nations, the problem of drug resistance continues. As of 2009 researchers were tracking 93 different common drug-resistance mutations worldwide. Today the World Health Organization (WHO) and other groups are working to raise HIV awareness, reduce infection rates, and increase access to drugs among those already infected. Researchers are also working to develop new drugs and vaccines.
Now that we have learned how our understanding of evolution can be used to help fight HIV, let’s see if we can apply that knowledge to an even more complicated infectious disease: malaria. Malaria is an ancient disease that has been around for over 50,000 years! Malaria is caused by four parasites of the Plasmodium genus: Plasmodium falciparum, Plasmodium vivax, Plasmodium ovale, and Plasmodium malariae. These protists (eukaryotes) have complicated life-cycles that require both a mosquito host and a human host, and the parasite undergoes many developmental stages within these two hosts. Humans become infected with the malarial parasites when they are bitten by an infected mosquito. Similarly, uninfected mosquitoes can pick up the parasites when they ingest blood from an infected human.
Because of its complex life cycle and the eukaryotic nature of this organism, malaria has been a difficult disease to combat. However, this complexity also means that there are more potential ways to intervene, prevent, and treat this disease. The parasite could be targeted with anti-malarial drugs in the human host. Other methods could be used to prevent the parasite from surviving in the mosquito host. Or measures can be taken to prevent the human host from being bitten by mosquitoes.
We have now explored two corners of the “Interactions in Malaria” triangle - the mosquito vector and the malaria parasite. As you can see from the question above, there are several ways in which malarial parasites can evolve resistance to anti-malarial medications. This is a real world problem in the fight against malaria. Resistance to almost every drug available to prevent or treat malaria has been reported, and the rates of resistance are only increasing. Insecticide resistant mosquitoes are also contributing to the problem because many areas where malaria is rampant rely heavily on pesticide control of the mosquito populations to curtail the spread of malaria. If the mosquitoes are no longer killed by the pesticide, then there will be more mosquitoes around to transmit malaria.
The last corner of our triangle is the human host. The alleles of many human genes have been determined to alter a person’s susceptibility to malarial infections. The most famous and arguably the most important is the sickle cell allele. People homozygous for the sickle-cell allele suffer from sickle cell anemia. Their red blood cells are misshapen, resembling a sickle.
Sickle cell anemia is caused by a recessive allele (we will call it “a”). Three genotypes are possible:
AA | Aa | aa |
homozygous dominant | heterozygous carriers | homozygous recessive |
Normal red blood cells, with no sickled blood cells. No negative health effects related to sickling of blood cells. | Most blood cells normal, 20-40% are sickled; few negative health effects. | Suffer from sickle-cell anemia, a condition in which sickled red blood cells are abundant and cause pain, swelling, poor circulation, and many other negative health effects. Without modern care, typically die without producing offspring. |
Good work. You’ve cracked the mystery! There is a situation called heterozygote advantage at work here. Why does a gene that alters red blood cells have anything to do with malaria? The malaria parasite spends a large part of its life cycle in red blood cells. The red blood cells of those with the Aa genotype are less hospitable to Plasmodium than those of an average (AA) individual. For example, in an Aa person the red blood cells may sickle in response to being invaded by Plasmodium. This will lead to the death of the newly invaded red blood cell, and the parasites will die with it.
AA | Aa | aa | |
Without malaria | Highest fitness; no sickle cell anemia symptoms. | Slightly reduced fitness; mild symptoms related to sickled cells. | Lowest fitness; full effect of sickle cell anemia disorder. |
With malaria | Moderate fitness; more likely to die of malaria or suffer reduced fertility due to its effects. | Highest overall fitness; more likely to survive malaria and reproduce well. | Lowest fitness; full effect of sickle cell anemia disorder. |
Malaria is a serious disease with an estimated 216 million cases and over 600,000 deaths in 2010. In Sub-Saharan Africa, where malaria rates are highest, a child under the age of five dies every minute from this disease. Doctors and scientists are working to help prevent and treat this disease by designing new treatments, developing vaccines, working to prevent mosquitoes from transmitting the malaria parasites, and by studying the interrelatedness between the parasite, the insect host and the human host. At the center of the connection between parasite, insect, and human lies evolution.
HIV and malaria are just two examples where an understanding of evolutionary principles has proven to be a critical component in the effort to fight infectious disease. Other prominent examples include tuberculosis, methicillin-resistant Staphylococcus aureus (MRSA), avian influenza, and pertussis. Evolution is at work all around us, and every population of every organism is subjected to it.
Thus far, we have been exploring how subtle changes can occur within a population. We’ve seen that these changes can lead to adaptation. The genetic makeup of populations also can change through migration and interbreeding (gene flow) and even through the vagaries of chance (genetic drift). But how does this relate to macroevolution? For example, can microevolution possibly explain the origin of new and different species?
Macroevolution is a long-term process and includes the origin and extinction of species and larger taxonomic groups. However, each and every event of macroevolution occurs through the processes of microevolution you just studied. The two processes are continuous and are part of a unified whole.
As you should recall, microevolution occurs at the population level. But a population is only a part of a larger collection that we refer to here as a biological lineage. A lineage is a group of populations that evolves independently of other groups. Members of the same lineage can move between its populations. Mating may occur between members of different populations in the same lineage. Lineages change as their member populations evolve and exchange genes and individuals with each other. To get a clearer sense for the nature of a biological lineage, do the Learn By Doing activity below.
To recap, a lineage is a group of populations that can exchange genes and individuals and that evolves as a unit as shown in the Figure below. The lineage evolves through changes that occur in its component populations.
Evolution within a biological lineage.
Within a biological lineage, evolution occurs as the component populations evolve. The diagram shows evolution of a biological lineage starting at an arbitrary reference point (“baseline generation”). Each chevron represents a snapshot of the lineage at one point in time. Circles (A-F) represent populations of various sizes linked by gene flow (white lines). Population (A) goes extinct. A mutation produces a new adaptive trait that spreads first through population E, then moves through gene flow to the remaining populations.
In practice, biologists often use the term “species” to describe a biological lineage. The definition of “species” you learned earlier in this Unit fits the idea of a lineage. Consider the Roseate Spoonbill (Platalea ajaja). These birds use their sensitive, flat bills to capture small animal prey in shallow water. There are populations of Roseate Spoonbill in coastal and wetland areas within South America, Central America, and the U.S. Gulf Coast. Individual Spoonbills can move from one population to another, taking their genes and adaptive traits with them. Changes in the genetic makeup of one population can affect the genetic makeup of other populations over time, so the species as a whole evolves as a unit. Yet the evolution of the Roseate Spoonbill does not affect the genetic makeup of other bird species. Five additional species of Spoonbill live in different parts of Africa, Australia, and Eurasia. Each is a distinct biological lineage.
Roseate Spoonbill, Platalea ajaja.
Speciation is the formation of new biological lineages. It occurs when an existing species splits or branches into two or more new species that evolve independently of each other. Over very long periods of time, repeated branching of a lineage can create groups of related species called clades. A clade is a complete group of all species that descend from some common ancestor. The six species of Spoonbill form a clade; all of them share unique adaptive traits not found in other birds, most notably the distinctive bill. They all likely descended from a single ancestral species that possessed these unique features.
![]() |
![]() |
Two additional members of the Spoonbill clade, genus Platalea. Royal Spoonbill (left), Eurasian Spoonbill (right). All spoonbill species have a similar bill shape and mode of feeding, but species vary in distribution, behavior, genetics, coloration, structure, and plumage. Left image by Fir0002(Wikipedia) (Royal Spoonbill mouth open) GFDL 1.2. Right image by Andreas Trepte (Eurasian Spoonbill) CC BY-SA 2.5.
In the early 1700s, Carolus Linnaeus established the modern system for naming and classifying organisms. In it, species are placed in nested groups based on their shared characteristics. Species correspond to biological lineages. Higher levels of classification are the genus, family, order, class, phylum, and kingdom. Each level is nested within the next higher level; all members of a genus, for example, are members of the same family, order, and so on. Below, the Linnaean hierarchy is shown for the Roseate Spoonbill.
Table 2.Linnean classification is a system of nested groups.
Species Platalea ajaja | Roseate Spoonbill | One lineage, many populations |
Genus Platalea | Spoonbills | One clade containing 6 related species |
Family Threskiornithidae | Ibises and Spoonbills | One clade containing 13 related genera |
Order Pelecaniformes | Herons, Ibises, and Pelicans | One clade containing 5 related families |
Class Aves | Birds | One clade containing 33 related orders |
Phylum Chordata | chordates | One clade containing 14 related classes |
Kingdom Animalia | animals | One clade containing 36 related phyla, millions of species |
How does this classification system relate to macroevolution? Today, biologists strive to place closely related species (members of a clade) within the same taxonomic groups. We hypothesize that the members of a genus had a recent common ancestor; members of genera within a family can be traced to an earlier common ancestor; and so forth. Thus the taxonomic system is not merely a convenient way to organize species. It partly reflects our understanding of how biological diversity has evolved over time.
The idea of common ancestry is very strongly supported by many lines of evidence. Like plagiarized term papers, our genes contain many sequences that closely match those in other species. Body structures and patterns of embryological development tell the same story. And fossils provide a historic record of the traits of our ancestors. You will see some such evidence in this Module.
To Learn More / Credits:
All taxonomic data based on current status in the Integrated Taxonomic Information System, http://www.itis.gov/
Speciation is the formation of new species. We will assume in this course that a species is a biological lineage, although there can be considerable debate among biologists about how species are best defined in practice.
How can new species form? Recall that changes that occur in the genetic makeup of one lineage do not directly influence the genetics of another lineage as we move forward through time. Therefore speciation is the establishment of a new group that does not exchange individuals or genes with its source lineage.
What could break up a biological lineage? Species can form if they are isolated from the source lineage by geography. This is called allopatric speciation. Allopatric translates to “different country.” Species may also form through genetic or other changes that create a subgroup within a lineage that does not breed or exchange individuals with the source population. A new species “pops up” right within, and surrounded by, its parent species. This is called sympatric speciation (translates to “same-country”).
Allopatric speciation begins when populations are separated in space by some kind of geographic barrier. It can occur if geological changes break a species’ range into separate “islands” of habitat. The diagram below illustrates this process:
Allopatric speciation can occur in another way as well: through long-distance dispersal and colonization of a new habitat. We can see the result of this process in islands far out at sea: a few hardy colonists “make it” to an island to start a population; after many generations they are distinct from their cousins on the mainland. This mechanism is illustrated in the diagram below:
The map of the world is a jigsaw puzzle of broken-up habitats (continents, oceans, islands, mountain ranges, lakes, caves, river systems, forests, deserts...) occupied by different regional or local species. Allopatric speciation helps us make sense of these patterns: as Earth’s landscapes were reshaped and broken up by geologic processes, living lineages also became separated and diverged into locally adapted species.
In some cases many closely-related species all live in the same geographic area and there is no known barrier that divides them today or in the recent geologic past. These species likely formed through sympatric speciation, which occurs without help from a physical barrier to dispersal. The diagram below shows a general model for sympatric speciation:
How could sympatric speciation happen? The drive to reproduce generally maintains the integrity of species by assuring that all their populations will exchange genes with each other. Sympatric speciation goes against this tendency; it requires the formation of a group of finicky, “cliquish” breeders. Most of the well-known examples involve genetic errors that produce offspring with extra chromosomes, a condition known as polyploidy. Sometimes polyploid types form new species that can breed with each other but are incompatible with the original species. Second, changes in ecology and breeding habits can also isolate a group within a species.
To Learn More:
You have probably noticed certain features that “run in families:” dark curly hair, an oddly shaped ear, bushy eyebrows, freckles... Siblings are often similar in appearance because they have inherited their genes from a common source.
In evolutionary biology, the same principle is extended to related lineages. Species with a recent common ancestor are close relatives, like brother and sister. Their lineages split apart not too long ago. Other species are much more distantly related, like second cousins. They can be traced to a common ancestor, but this ancestor lived very long ago. We can tell their degree of relationship by looking at homologous features: shared features that were inherited from the same source and reflect common ancestry. Two species that are closely related share many homologous features. Species that are distantly related share only a few.
Not every similarity between two species is homologous. We are looking for features that were inherited from a common source. Therefore the features should be based in genes (not acquired through learning, for example). Homologous features will also be similar in their details; there should be more than a passing resemblance.
An analogy might help you to recognize what we’re getting at here. Two students are brought into a professor’s office. They are shown their term papers, side by side. Bill’s paper begins: “Phylogenetic biology is the exhaustive comparison of taxonomic units in an attempt to elucidate the ancestral relationships among all life forms on the planet Earth.” Sarah’s paper starts with the exact same sentence. At several points, the papers have identical or extremely similar wording.
The students are charged with plagiarism. Both claim to have written their own papers. After all, the professor asked for a paper on phylogenetic biology, so of course the papers will be similar! The professor counters this with a quick calculation: the odds that two students’ papers would have such similar language by chance are one in several trillion. The academic integrity board agrees: either one student copied from the other, or they both copied language from the same outside source.
In biology, the same sorts of judgments are made about similarities among species. When inherited features are very similar in their details, biologists assume they were likely inherited from a common source. One classic example involves the skeletal structure of vertebrates.
It seems unlikely that these four animals (not to mention thousands of others) would develop such similar limb structures by chance. It also does not appear to be necessary. Why should a whale flipper have embedded fingers, wrist bones, and three major limb bones? If adaptive evolution was starting “from scratch,” it might “design” a flipper with just a few bony plates. Instead, evidence supports the idea that all four-limbed vertebrates descend from a common ancestor; they all inherit their curious and specific arrangement of limb bones from that source. In separate lineages, adaptation has reshaped the limbs and suited them to very different functions. At first glance, a whale flipper and a hummingbird wing have little in common, but the details of the structures point to shared ancestry. Let’s take a moment to assess your understanding of homologous features.
Homologous Molecules and Cellular Structures
It is not just four-limbed animals that are related. According to current evolutionary thinking, all life on Earth is related by ancestry. You, a bacterium, a fly, a fern, and a Portabello mushroom all belong to a single clade. To some, this is a difficult or even preposterous concept. What could you and a bacterium or mushroom possibly have in common? And if you do have something in common, why do biologists think it was inherited from the same source?
We may look very different, but unifying features are evident within our cells. For example, all cells contain DNA. And to make use of their genes, all cells have ribosomes: clusters of protein and RNA that carry out the synthesis of proteins.
Every cell has ribosomes, but each organism’s ribosomes are built according to a slightly different set of specifications. Here, then, is a characteristic that we can compare and contrast among every single living species. That is just what Carl Woese, a microbiologist at the University of Illinois, set out to do in the 1970s. He obtained gene sequences for ribosomes of a wide range of species. He focused on part of the small subunit, shown above for one species of heat-loving bacterium. Amazingly, he could line up matching gene sequences (long series of the nucleotide bases A,G,C,T) among various microbes, plants, animal, and fungi. This was painstaking work, done without the aid of modern computers and software.
In 1990 Woese published a paper proposing a change in the classification system. He reported that there were consistent differences in ribosomal gene sequences among three major groups. Above the level of Kingdom, he proposed three Domains of life. Two of them are made up of prokaryotes (Bacteria and Archaea). The third is called Eukarya and includes all organisms whose cells have nuclei, including fungi, animals, plants, and several groups of protists. Woese’s data have since been greatly expanded, and the new data are consistent with the idea that the three Domains are the first three branches on a unified tree of life.
Why do scientists think that the chemical similarities of cells reflect shared ancestry? As with vertebrate limb structure, the genetic similarities among life forms are more than a passing resemblance. DNA sequences are detailed, complex, and contain a great deal of information. In a court of law, DNA evidence can be used to tie a suspect to a crime scene sample: it is extremely unlikely that two long DNA sequences would show an exact match unless they came from the same source. In evolutionary biology, the same sort of reasoning applies: long chunks of DNA sequences match closely among species and are evidence for shared ancestry. Thus we can now analyze inherited characteristics that have been present for not just millions but billions of years. In these ancient gene sequences we see the deepest roots of the tree of life.
Homologous features are the key sources of information for the construction of phylogenetic trees. A phylogenetic tree, or phylogeny, is a diagram that shows how a biological lineage may have branched and formed clades over time. A tree diagram is a hypothesis about how different species are related and is subject to change as more data are gathered and as analysis techniques improve. To get a feel for how phylogenetic trees work and what they mean, do the following activity.
As you have learned, shared characteristics can be used to organize species into groups, with clades defined by unique inherited traits. Unfortunately, this procedure is often quite difficult: there are millions of species to organize and an infinite number of possible characteristics to compare! Many traits may be lost or reversed over time. Consider the examples of tetrapods with one or both pairs of limbs reduced or missing (snakes, some lizards, whales and dolphins, manatees, seals, etc.). Some changes also occur quite predictably as adaptations to special environments. In the Arctic, foxes, birds, mink, hares, bears, and more all have snow-white winter coats or feathers. They are not closely related and did not all inherit their winter coloration from a common ancestor. Instead, “snow white winter color” developed independently in each of these lineages at different times and starting with different mutations.
Unlike many physical or behavioral traits, DNA sequences are an excellent source of data on relationships. By looking at DNA, biologists have a way to directly and objectively compare inherited information. For example, virtually all eukaryote species have a gene in their mitochondria that codes for cytochrome C oxidase, an enzyme that is vital to the process of aerobic respiration. This widespread gene is quite similar in most species, but mutations do occur and tend to be preserved within a lineage. Subunit 2 is a portion of the cytochrome C oxidase protein; its gene is composed of about 700 nucleotide base pairs. What does this gene say about the relationships among Carp, Salamander, Sea Turtle, Human, and Mouse? The image below gives a glimpse of the data.
As you can see, there is a great deal of similarity in this gene among the five species. This shouldn’t be too surprising: they are all vertebrate animals, close relatives in the grand scheme of life’s diversity. There are clearly sections of the gene where all the bases tend to agree; other regions show more differences. Do the differences between the genes match our expectations about how Carp, Salamander, Sea Turtle, Human, and Mouse are related? It is not easy for a human to sort through the sequences and count differences one by one, but a computer can complete this task in a flash. Below is a summary of how the full gene sequence for each species compares to that of the Common Carp:
Tiger Salamander | 162 differences | 76% identical to Carp |
Green Sea Turtle | 176 differences | 74% identical to Carp |
House Mouse | 200 differences | 71% identical to Carp |
Human | 229 differences | 67% identical to Carp |
At least from the Carp’s perspective, the genetic data agree with our tree: the Tiger Salamander is Carp’s closest match; Sea Turtle is in between; and Mouse and Human are more distant!
Using modern computers and a variety of analytical techniques, biologists can now compare vast amounts of genetic data among hundreds or even thousands of species. They can sometimes even include DNA sequence data extracted from fossils! The result has been a huge improvement in our ability to identify the clades within the tree of life.
To Learn More
Tree of Life web project: http://tolweb.org/tree/
Travels in the Great Tree of Life, Peabody Museum: http://archive.peabody.yale.edu/exhibits/treeoflife/index.html
Wellcome Trust Tree of Life website: http://www.wellcometreeoflife.org/
Why does macroevolution matter? Is it just ancient history? Far from it! “Tree thinking” is basic to progress in modern biology. It helps us deal with the vast diversity of life in an organized fashion. We don’t have to start from “square one” with each species of interest. Instead, we can use natural groupings (clades) to our advantage.
A phylogenetic perspective can help us deal with threats. When there is an outbreak of food poisoning, genetic analysis can show that the harmful bacteria came from a particular farm. If we are attacked by a new virus, we can compare its genes to those of known pathogens. Once we place the virus within a clade, we can quickly get a handle on what we’re facing, how it is likely to affect us, and how we can best combat it.
Phylogenetics can also help us get the most out of helpful species. You may be surprised to learn that the familiar daffodil is a plant with medicinal potential. Although daffodils are poisonous, drugs might be developed from chemicals that can be extracted from the plant. One wild daffodil species in Spain (Narcissus confusus) has been found to contain a chemical that slows the progression of Alzheimer’s disease and another that inhibits replication of HIV, the AIDS virus.
Narcissus confusus (inset) in its habitat in Spain.
Daffodils are poisonous but drugs might be developed from chemicals extracted or
modified from those found in the plant. Photos by A. Barra,
CC BY 3.0.
Based on these discoveries, scientists are interested in learning more about the chemistry and therapeutic potential of daffodils. But where to look? There are up to 150 different species of Narcissus and countless varieties. It would be inefficient and much too costly to do a full analysis on every species. Instead, researchers are using “tree thinking” to narrow down the search. Future testing will focus on close relatives of species that are already known to be chemically rich.
A second principle of macroevolution, common ancestry, is important for another reason. Medical research focuses on the biology of humans. But let’s face it: humans make terrible study subjects. We live too long, take up too much space, and have too many inalienable rights to be of much use in the lab! Therefore biomedical researchers also study other species as far-flung as worms, fruit flies, mice, fish, or rhesus monkeys. They can tell us a great deal about our bodies, our genes, and even our diseases. Why? Because these species have many inherited features that are very similar to our own. When we learn something about a worm or a wallaby, we learn more about ourselves in the process. Below we summarize just a few examples of the power of this approach.
Discovery in Model Organism | Application to Humans | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
![]() |
![]() |
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
![]() |
![]() |
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
![]() |
![]() |
References / To Learn More:
Willcox, B.J. and others. 2008. FOXO3A genotype is strongly associated with human longevity. Proceedings of the National Academy of Sciences 105: 13987-13992.
Spencer, G. 2002. Background on mouse as a model organism. National Human Genome Research Institute website.
Partridge, L. and others. 2011. The new science of ageing. Philosophical Transactions of the Royal Society of London Biological Sciences 366: 6-8.
National Science Foundation / American Museum of Natural History. 2000. Assembling the Tree of Life: Harnessing Life’s History to Benefit Science and Society.
In exploring biology thus far, we have focused on life from the “skin in”—you have learned about the structure and function of cells and have explored the chemical processes that enable life. But biology does not stop with this inward-looking perspective. Biology also looks from the skin out at the individual’s life in context. Biologists explore how you and other organisms get food, how you change the world around you and in turn are influenced by factors like temperature, environmental toxins, and other organisms.
Ecology is the branch of biology that studies life at population and higher levels of organization, emphasizing how groups of living organisms interact with each other and with nonliving factors in their environment. The main levels of analysis in ecology are population, community, ecosystem, and biosphere. Let's examine the meaning of these terms to see how they differ from one another, and yet build upon one another. The diagram below summarizes how the levels of analysis relate and gives a specific example (the forest) to guide your understanding.
Within an area defined by the ecologist, individuals of a single species that live together and may breed with each other make up a population. All of the different populations of living organisms living within an area make up an ecological community. In the example above, the squirrels and sugar maples of a region, together with humans plus hundreds of additional populations of plants, animals, fungi, bacteria, and other organisms would make up a forest community. The forest ecosystem also includes non-living factors such as chemicals in the soil, rainfall and other climate factors, and countless other features of the environment that interact with living things in the region. Additional ecosystems (lakes, grasslands, oceans, etc.) share our Earth with the forest ecosystem. Together, all of the zones of Earth that support life make up a giant interconnected system called the biosphere.
Many students have some difficulty distinguishing between the definitions of ecological community and ecosystem. The community and ecosystem levels can cover exactly the same area: communities are not literally “smaller than” or “contained within” ecosystems. Instead, the ecosystem is a different way of looking at an area that includes a broader set of factors and processes than the community level of analysis. The community ecologist looks at a forest and sees particular types of birds, trees, fungi, and insects eating each other and otherwise interacting. The ecosystem ecologist looks at the same forest and may see the organisms in less detail, but also considers how the trees build the soil or how floods shape the landscape.
Population ecology is focused on how populations — of plants, animals, bacteria, and other types of organisms — change over time. A population is a group that includes all individuals of a particular species within a given area at one point in time.
Population size, often represented by the variable N, is the total number of individuals in a population at a particular point in time. Population size can be an abstract concept, particularly for small organisms whose numbers can be astronomical. A single species of mosquito in a swamp might easily number in the millions or even billions of individuals. It is often better and more practical to describe such populations in terms of their density.
Population density is the total number of individuals of a species per unit area or volume within a specified habitat or region. In many cases this number can be pictured and measured more easily than population size itself. It also can be used to compare populations that occupy areas of different size, allowing you to judge where a population is more or less crowded.
population size = N = total number in population |
population density = N / area or volume occupied |
As illustrated by the case of Valley Forge’s deer population, population size and density are dynamic, or changeable, numbers. As humans we care about changes in the density and sizes of other populations. We notice when the density of mosquitoes skyrockets after a summer rain. We are concerned about the population sizes and densities of animals we hunt, fish we harvest for food, and endangered species we seek to conserve. Much of population ecology is directed toward predicting, understanding, and controlling changes that occur in the populations of other species over time.
Populations change through four primary processes. New members can arise by birth through reproduction of existing population members. In the photo on the previous page, a white-tailed deer has given birth to two offspring (family sizes vary widely among species). New population members may also immigrate (move into the population) from other areas outside the boundaries set by the researcher. A population declines through the death of its existing members. The population may also decrease if members emigrate (move away to areas outside the boundaries enclosing the population). To simplify calculations ecologists often assume that a population is closed - that emigration and immigration do not occur. For this course we will adopt this idea and assume that, unless otherwise stated, the populations we analyze are closed.
Researchers studying populations usually describe birth and death in terms of rates so that they can compare different populations to each other on a similar basis. According toFederal Bureau of Investigation Uniform Crime Report statistics, in 2010 there were roughly 530 cases of murder in New York City and about 180 murder cases in New Orleans, Louisiana. Think New York City is more dangerous? Raw numbers don’t tell the full story. In 2010, New York City had a population of over 8 million people compared to New Orleans’ population of about 350,000. Therefore the rate of murder was almost 8 times greater in New Orleans than in NYC.
By calculating rates we can make fair comparisons between populations regardless of their size. The simplest way to do this is by using per capita (literally “per head”) rates where population size is included in the denominator of a rate calculation. Per capita rates can be calculated for events like birth and death or even occurrences like murders or car accidents. In each case, the calculation is equivalent.
If N = population size at the beginning of a time period, the per capita rate of some event is calculated as shown below for birth and death rates:
per capita rate of birth =
per capita rate of death =
Now you have some mathematical tools for describing a population’s status in terms of its size, density, and birth and death rates. As we will see in the next page, the rates of birth and death can be very helpful in projecting how a population will change over time.
(1) Source: "The City of New York, Summary of Vital Statistics 2009"
Now that we know a bit about how to describe a population, how can we figure out how its numbers will change over time? Let’s consider the relatively simple case of a closed population, where there is no immigration or emigration. We can predict how a population’s size will change over time simply by taking the difference between the rates of birth and death:
A population will grow if its rate of birth is greater than its rate of death. The population will remain constant if the two rates are equal—that is, if births are balanced by deaths. It will decrease if the rate of death is greater than the rate of birth.
Think of a population as a parade of generations, with each new crop of offspring replacing the previous generation. For sexual species, the average female must produce two offspring that live to reproduce: one to replace herself, and another to replace her mate. Yet we know that most species attempt to produce more—often far more—than one or two offspring. Furthermore, they often attempt to live for several reproductive seasons, each time generating another set of offspring. As long as conditions permit good survival and successful reproduction, any population’s size will inevitably increase.
For any population, the actual quantity of increase (number added to the population) will depend on how many you already have in the group. It can be calculated as follows:
That is, the per captita rate of increase multiplied by the starting population size (N).
Now you see that a rate of increase predicts how much a population will grow, and that all species tend to increase if given good conditions for growth. Below are a table and a graph that continue the process of growth for the Valley Forge NHP deer population, using our estimated annual per capita growth rate of 0.39 deer added per deer per year.
The population is projected to grow at a constant annual rate of increase (0.39 deer added per deer per year, or 39% annual growth). Each year, the starting N is updated to reflect the growth of the previous year. For example, the starting population for 1986 (236) = the base population from 1985 (170) plus the growth from 1985 (66). Although the per capita rate of growth remains constant, the number added each year increases with the base population.
Year | N at Start of Year | Number Added |
---|---|---|
1985 | 170 | 66 |
1986 | 236 | 92 |
1987 | 328 | 128 |
1988 | 457 | 178 |
1989 | 635 | 247 |
1990 | 882 | 344 |
1991 | 1226 | 478 |
1992 | 1704 | 665 |
1993 | 2369 | 924 |
1994 | 3293 | 1284 |
1995 | 4577 | 1785 |
The table and graph above depict a pattern of change called exponential growth. In exponential growth, a quantity grows by a constant percentage each time step. As the quantity gets larger, so does the actual amount of growth that occurs. Some familiar examples of exponential growth involve finances. A bank account may earn 1% of its value in interest each year.When the bank account’s value is $100, 1% amounts to a dollar. But if the account already contains $1 million, a gain of 1% brings in $10,000!
You can recognize exponential growth as a pattern in which the numbers increase faster and faster with each time step. A series of numbers with this property would be 2, 4, 8, 16, 32, 64, and so on. Visually, you can imagine exponential growth as a snowball rolling down a hill, picking up snow as it goes: the bigger it gets, the faster it grows. A wildfire spreading through dry grass might also grow exponentially. In a graph, exponential growth shows up as a J-shaped curve.If you imagine the graph as a staircase, each individual step is taller than the one that came before it.
A contrasting model of growth is linear growth. Here a quantity grows by a constant number (not percentage) in each time interval. A series of numbers with this growth pattern would be 2, 4, 6, 8, 10, 12, and so on. Each time step the quantity increases by 2, no matter how big it might already be. In a graph, linear growth appears as a straight line. A staircase is linear: the risers are evenly spaced and you gain a set amount of height with each subsequent step.
Compared to linear growth, the exponential growth pattern is a better starting point for predicting how populations will change over time. It captures a little bit of reality, in that a bigger group will have more individuals capable of producing offspring. However, it is important to emphasize that the table and graph above are based on a “what-if” exercise. They show numbers generated by a model of how populations increase. This model assumes that the growth rate will remain constant and predicts how the population will change if this condition is met.
All populations have the potential and tendency to increase. For sexual species, each adult female must produce two adult offspring to replace herself and her mate and thereby maintain a steady population size. But if conditions are right, individuals of any type of organism will produce offspring in excess of this replacement number.
The exponential model of growth predicts what will happen if this potential is realized each generation for an extended period of time. The population will grow faster and faster the larger it gets, producing a characteristic J-shaped curve on a graph of population size over time.
What actually happened to the deer population at Valley Forge NHP from 1985 onward? Data are not available for the late 1980s or early 1990s. However, the record that is available looks like this:
Estimated changes in deer population at Valley Forge NHP. Data are not available for the time period from 1985 to 1996.Source:
A few things are evident in the graph. First, the population did not increase to over 4500 by 1995 as predicted by the simple exponential growth model. Instead, it increased to about 1400 until the early 2000s and then began to fluctuate.
The exponential growth model assumes that populations will maintain constant rates of birth, death, and increase as they grow larger. In reality, we often see that populations stop growing—and sometimes even crash—due to the effects of limiting factors. Limiting factors are biotic or abiotic factors that limit populations by reducing birth rates, increasing death rates, or both. Limiting factors may also increase rates of emigration or decrease immigration.
Sometimes limiting factors act in an unpredictable way to reduce population sizes. Hard freezes, fires, droughts, and floods are all examples of factors that could act to quickly reduce a population. These factors most strongly affect small-bodied organisms that are easy to kill such as insects, microbes, and weedy plants. Their populations often show an erratic boom-and-bust pattern: population sizes shoot up when times are good and crash when conditions are bad. Ecologists agree that it is difficult to predict or model the changes in such populations, except by taking into account the weather or other outside factors known to control them.
Other types of limiting factors act more gradually and have their greatest effects when populations are crowded. Scarce resources, limited space, and infectious diseases are important examples. These factors often play the greatest role in limiting large-bodied organisms with extended life spans such as large animals and woody plants. For example, deer survive frost well but are more likely to starve in a hard winter if there are many mouths to feed. Over time, populations of such species limit themselves: they remain close to a stable size called the carrying capacity. When a population is at its carrying capacity, it does not tend to increase or decrease but tends to stay essentially the same over many generations.
At least in principle, we can predict what will happen to a self-limiting population over time: it should grow until it reaches its carrying capacity and then remain stable. This pattern of change is calledlogistic growth. It is simple to describe logistic growth with a math equation that allows a population to grow when it is small but causes it to level off when it reaches a predetermined value (the carrying capacity). A logistic model for the Valley Forge NHP is shown below, superimposed on the actual population data from the park.
A graph of logistic growth shows up as a slightly S-shaped curve. Initial growth is relatively slow because the population is small and has few individuals to give birth. At moderate population sizes the population grows fastest: there are many parents but the effects of crowding are not yet severe. As the population approaches its carrying capacity, the effects of crowding start to reduce its growth. The rate of growth slows and then stops.
Principles of population ecology are critically important to many enterprises including game and fisheries management, conservation of rare and endangered species, and pest control. How can a population be “managed”? Managers seek to control conditions for survival and reproduction.
Management of threatened or endangered species may simply involve increasing birth and reducing death rates as much as possible, though this is often easier said than done. To help such species, conservationists may breed them in captivity and release individuals back into the wild. Captive breeding programs have been used to aid in the recovery of many species, including the peregrine falcon and the black-footed ferret. Most endangered species management focuses on improving survival rates (reducing death rates) by banning harvest of these species or trade in their products, by eliminating introduced predators, or by improving habitat conditions.
The situation is more complicated when the goal is to manage a resource so that it can be harvested. Consider a population of fish that supports a commercial fishery. Here, the goal is to maximize the harvest rate while maintaining a healthy population size. How can principles of population ecology help in choosing the best rate of harvest?
The general idea is that a population is like a bank account. Every so often, we can remove the interest that has accumulated in a bank account. This can continue indefinitely so long as we do not remove any of the principal that is earning the interest. Similarly, every so often we can remove a population’s growth without causing the population to decline: it will be reset to its beginning level and the harvested individuals will be replaced by further reproduction and growth. Ideally, we should be able to find the population size that provide the maximum level of growth (and harvest) each year. Population modeling can help with this effort.
As you have just seen, the logistic model indicates that a harvestable population should be maintained at a moderate size of about half its carrying capacity. At this point there is a large increment of growth each year and these individuals can be harvested. This is the general framework used by fisheries managers as they seek to maintain and improve the productivity of fish populations. They often set regulations and allowable harvest limits to try to keep populations close to the “sweet spot” at about 50% of carrying capacity.
Population models help officials manage fisheries for maximum benefits to commercial fishermen and recreational anglers. However, this approach can be risky. Models are not the same as reality. Outside factors such as weather events, disease epidemics, and so on can sometimes unpredictablydecimate fish populations. With fewer adults in the population, the population cannot rebound after a harvest that would normally be sustainable. Even just one bad year can set up a series of events in which the population dwindles over time. Because of this unpredictability, many fisheries managers try to maintain populations at a level that is a bit “too high” (say, 75% of carrying capacity) and limit harvest rates accordingly. This reduces the amount of fish that can be harvested but it also reduces the risk that harvests will drive the population downward.
Endangered species and fisheries management are just a few of the areas in which principles of population ecology are applied to solve real-world problems. Pest management is often a very challenging exercise because pest species tend to reproduce very rapidly and show a boom-and-bust pattern of growth. Even if we can induce a bust—say, by spraying an insecticide—the population is likely to rebound with another boom unless we can reduce the birth rate. Effective pest management often requires that several strategies be used together with careful monitoring of the pest population.
Population management can also raise ethical issues. It involves life-and-death decisions that can be very emotionally charged. At Valley Forge NHP, park managers recognized the need to act as deer populations climbed past 1,000 and serious ecological damage began to occur. The park’s plant diversity was severely threatened by the high deer numbers; few tree seedlings or wildflowers were able to survive except within small fenced areas (deer exclosures).
Most agreed that something should be done, but how should the deer population be controlled? It would be unsafe to allow the public to hunt in the park. Some citizens argued that birth control injections should be used to manage the population, but evidence indicated that this would be expensive and ineffective. After extensive deliberation and community input, a deer control program was begun. In the winters of 2010 and 2011, sharpshooters were contracted to kill deer at the park; meat was provided to an area food bank. Sharpshooters removed 600 deer in 2010 and 377 deer in 2011. As a result, the deer population fell from 1,277 in 2009 to 374 in the spring of 2012. Initial data indicate that many tree, shrub, and wildflower species are already beginning to recover as a result.
To learn more / sources: Valley Forge NHP White-Tailed Deer Management Page
Until now, we have focused on specific populations responding to limiting factors such as weather, crowding, or fishing by humans. Community ecology broadens the focus to explore how populations of different species interact with each other. An ecological community is the collection of all interacting species populations within some defined area. Because animals move in and out of any given area, it can be tricky to determine what populations belong within a community. For example, birds such as herons may interact with pond communities by removing fish. They are therefore included as part of the pond community because they can have an important effect on pond life.
Scientists learn about species interactions first by direct observation. For example, it is simple enough to discover a predator-prey relationship by watching a lion take down an antelope, or by finding prey in a predator’s stomach. However, community ecologists often want to go a step further and determine the consequences of an interaction. How will a community change if a new species is added, or if an existing population dies out? To find out, ecologists may do comparative observations or experiments, as we will see in this module.
In several examples we will also see that community ecology is important to people. Managers of nature preserves and wildlife parks seek to increase or maintain the diversity of native species in the areas under their care. Community interactions may also control the abundance of pests, disease agents, pollinators, fish, game animals, and other species that have a direct effect on human well-being.
An ecological community can be described in many different ways. One statistic that is often of interest is the species richness of a community: the number of different species of organisms present in a given area. Some communities, such as tropical rainforest or coral reef communities, have high species richness with thousands of species of plants, animals, and microorganisms coexisting in a small area. Other communities with harsher conditions (deserts, arctic tundra, salt lakes) may hold far fewer species. Within any region, a given type of habitat will have a fairly predictable level of species richness. Human activities that alter living conditions, such as pollution or habitat destruction, can change the membership of communities and usually will cause species richness to drop. Thus a naturally high level of species richness is one indicator that a community is healthy and relatively unaffected by human activities.
An interspecific interaction is an effect of one population on another in a community. Although we are often focused on how one population affects the size of another, ecologists also may look at how species interact in terms of their activity, growth rates, body size, or other aspects of their well-being. Interactions are classified according to their consequences for each species involved. The most important types of interactions are summarized in the table below.
Interspecific interactions classified by consequences for interacting species. | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
|
In a community, any pair of species may potentially interact. However, some interactions are much clearer and more easily predictable than others.
Direct interactions occur when two species make contact and affect each other’s well-being without the participation of any third species. Some pairs of species are intimately and physically connected, with a smaller species (symbiont) living in or on the body of a larger species (host). These direct interactions are described using the term symbiosis,which translates to “together living.” In other cases the contact between direct interactors is more fleeting, as when a bee visits a flower or when a hawk swoops down on its prey.
Mutualistic, or mutually beneficial, relationships are widespread and important features of communities. Sometimes mutualism also involves symbiosis. Many animals and plants are host to microorganisms that provide them with vital functions. In other cases, equally important mutualistic transactions are carried out at a distance: not all mutualisms are symbiotic. Plants, for example, reward animals for transportation services that allow them to mate (pollination) and travel as seeds (seed dispersal). Animals also may help each other in subtle ways.
Predation can be broadly defined as any win-lose interaction, though it is most commonly used to describe situations where one species kills and eats another. You are no doubt familiar with many spectacular examples, as when a pack of wolves takes down a moose. More subtle examples of predation also exist. Sometimes the predator will eat only part of the prey, allowing it to live another day. Though we often use the term “herbivory” for the consumption of plants, the moose might itself be considered a predator if it harms the plants on which it feeds. Many forms of predation involve symbiosis. Such interactions are called parasitism and may involve animal or a plant hosts that provide both habitat and food to their harmful symbionts.
Competition is a lose-lose, or mutually harmful, relationship between two species. Two species are competitors if each would be better off alone. Clear examples of competition are often seen among plants that require very similar resources, such as light, water, and certain mineral nutrients. Animals may compete for space, sometimes fighting directly for territory or simply excluding each other from living sites. Animals that eat similar foods may also sometimes make life harder for each other, though here the degree of competition is often limited because each species may eat slightly different foods.
Whenever ecologists study species interactions, they try to get clear data answering “what-if” questions. What happens to prey populations if a predator is removed from a community, or if a new predator shows up? Do two species really compete; that is, do they each do better if they are alone than if they are living together in a habitat? Can a species and its mutualist live apart, or are the two species entirely dependent on each other? Sometimes we can learn about these issues simply by observing what happens when introduced species—species brought into an area by human activity—establish themselves in a community. We can also see what happens to the larger community if a particular population is eradicated by human activity. On a smaller scale, ecologists sometimes carry out true experiments. In these studies ecologists deliberately manipulate the presence or abundance of species in enclosures or in natural environments and then track what happens to other species as a result.
Explore some examples of species interactions below. In each case you may be asked to identify the type of interaction—mutualism, predation, or competition—and to clarify exactly how the species interact.
Ecological interactions range from harmful to helpful and can take on many forms. Some interactions involve exchanges of materials; others are more subtle and involve services or behaviors. Mutualistic interactions often develop when two partners have complementary abilities. For example, aphids are excellent plant sap drillers but ants are much tougher and more mobile, so the trade between aphid and ant is quite logical. Predation is a nearly universal interaction: almost every organism is food for something else, and most animals eat living prey of some kind. Competition often develops where individuals or species have similar requirements, like the plants that will crowd together in a dense Alabama thicket if conditions permit it.
Ecological communities can be dizzyingly complex. A community may be home to thousands of species, and every single one of them may be linked to several others through a variety of interactions. Among all this complexity, ecologists have worked to discover the most meaningful and important interactions: those that determine the stability and diversity of the whole community.
One way to simplify the problem is to focus on a few particularly important species. Species that are physically dominant (largest and/or most abundant) in communities are sometimes called foundation species. They include the most abundant trees within terrestrial forests, common grasses within grasslands, and reef-building corals. Foundation species directly provide habitat and often food for most of the other species in the community. Anything that threatens their well-being is likely to harm many other species as well.
Along the coast of western North America, rocky areas are often home to a spectacular community called the kelp forest. It is dominated by kelp, big brown algae (seaweeds) whose fronds can stretch to dozens of meters in length. Kelp are clearly the foundation species of the kelp forest: they provide food and shelter to a huge diversity of fish and invertebrates including such economically important species as lobster and rockfish.
Most communities are clearly dependent on one or a few foundation species. But not every important species has effects that are so obvious. In ecological communities, many interactions are indirect.Indirect interactions, sometimes called “ripple effects,” occur through chains of cause and effect. The effect of one species on another is said to be indirect if one or more additional species are involved in the interaction as intermediaries. Complexity of this sort is a source of fascination and challenge for ecologists attempting to unravel how communities work.
Indirect interactions can produce surprises in communities. Some species, called keystone species, may be very influential in communities even though they are not particularly abundant or large. In architecture, a keystone is a single wedge-shaped stone at the top of a stone arch that maintains the integrity of the structure; its removal will cause both sides of the arch to collapse. A keystone species, then is a species that has an unexpectedly strong effect on community stability or diversity. Keystone species play unique roles and are not easily replaced by other species within the community. We often discover their importance only after they are eliminated by human actions.
Within kelp forests, human hunting revealed the importance of a keystone species. The story begins with urchins, which are abundant bottom-dwelling grazers and scavengers. Kelp are tough and fast-growing, but urchins (animals related tostarfish) can eat them with their hard rasp-like mouth parts. Urchins are protected by their prickly spines and a hard outer shell. In the absence of factors controlling their abundance, urchins can increase to very high densities, producing what are sometimes called “urchin barrens” devoid of kelp. Unlike kelp forests, urchin barrens do not support much life.
What factors normally keep urchins in check? In the cold waters off the west coast of Alaska and Canada, evidence points to one factor as particularly critical: the presence of sea otters. Historically, humans hunted sea otters for their fur, nearly eliminating them by the late 1800s. More recently, sea otters have been reintroduced to many areas where they had been absent. Since the 1980s, researchers, including James Estes of the University of California Santa Cruz, have been tracking sea otter numbers and documenting their effects on marine communities in Alaska and off the coast of British Columbia. They have found that the sea otter is a keystone species in the kelp forest.
Sea otters are uniquely suited to eating the urchins. With their dexterous paws and long incisors, they can crack open the urchins, eating the soft flesh inside. Sometimes they even use rocks as tools to crush tough prey. After eating an urchin, a sea otter discards its empty shell, or test. A study published in 2011 by Estes and Jane Watson reported that when broken tests sink to the bottom, remaining urchins will flee from the immediate area. Thus, sea otters can reduce urchin numbers and change urchin behavior. In this way, otters safeguard kelp and help to maintain healthy kelp forests.
Kelp forests provide a glimpse into the many facets of community ecology. They are complex, with more than twenty species of kelp plus hundreds of animal species. They also are structured in an understandable way, with some species interactions more important than others. As ecologists continue to unravel community interactions, we are improving in our ability to predict how communities will respond to factors like hunting. Kelp forests are a treasured resource for recreational diving and fishing. Kelp also are harvested for use in aquaculture and industry. The science of ecology can help us learn more about kelp forests and supports our efforts to safeguard these dazzling communities.
Source and To Learn More:Whipple, WJ and RL Beschta. 2012. Trophic cascades in Yellowstone: The first 15 years after wolf reintroduction. Biological Conservation 145: 205-213.
People are concerned about ecological community interactions for many reasons. For example, animal pollinators help us produce crops including almonds, soybeans, apples, tomatoes, coffee, cotton, and many more. In fact, over a third of global crop production depends on animal pollinators. Many of these crops are pollinated by domesticated honeybees. Originally native to Eurasia, honeybees were brought to North America by the earliest European settlers. Native Americans are said to have called them the “white man’s fly.” Honeybees and other pollinators transfer pollen from one flower to another, helping to increase seed and fruit yield.
Ongoing declines in bee populations are a major focus for ecological research. Since about 2006, North American beekeepers have been plagued by a complex problem called colony collapse disorder, in which worker bees abandon their hives. Research has implicated parasitic mites, viruses, pesticides, and other factors in causing bee colony declines. In 2012 scientists documented another threat: honeybees are being parasitized by a fly that was not previously known to interact with them. The fly lays its eggs on worker bees and its larvae live within bees, feeding on their body tissues. When the fly larvae are mature, the worker bee hostswander away from their hives at night and die away from the colony. The fly larvae then abandon the host, beheading it as they exit.
Even the gory interaction with decapitating flies probably cannot explain colony collapse disorder all by itself. Instead, experts agree that many factors have likely worked together to threaten bees. As we learn more about the habitat needs of bees and threats to their existence, we may be able to find ways to protect them. In this way, biologists hope to reinforce the ancient mutualism between bees, crop plants, and food producers.
To learn more review the following sources:
Up to this point, we have focused on the living components of the environment. Consider a suburban backyard that is home to diverse species, including robins, squirrels, ants, oak trees, grass, rabbits, daisies, spiders, toads, mushrooms, and raccoons. These living, or biotic, components together make up an ecological community. Its populations affect each other directly and indirectly as we just explored. But each population also interacts with a host of nonliving factors. A grass plant, for instance, requires light, carbon dioxide gas, soil nutrients, and water. It is affected by temperature and by factors like snow cover or fires. In turn, it influences nonliving factors as well. Its roots, for example, hold soil in place.
To fully appreciate an organism’s role in the environment, we must go beyond the web of life. We must also consider nonliving, or abiotic, components of the environment: rocks, water, gases in the air, chemicals in soil and water, light, temperature, wind, and waves. The combination of biotic and abiotic components forms an ecosystem. A community is restricted to the populations of living organisms in an area; an ecosystem also includes all of the nonliving physical and chemical factors that are important to these populations.
In a healthy ecosystem, there is a continuous flow of energy that provides resources to sustain a diverse community of life. As we will explore in this module, energy starts in a nonliving form, usually as sunlight, and is captured and stored as chemical energy by producers. It then moves through chains of consumers, dissipating as heat along the way. A continuing supply of solar energy keeps this system in motion.
As they take up and release energy, growing organisms also absorb chemical nutrients and release chemical wastes. There is a continuous exchange of material between the living and nonliving portions of the ecosystem. Each chemical element retains its integrity but takes on many different forms as chemical bonds are formed and broken. To see how this material recycling takes place, think about a leaf on a tree. Its living tissue contains organic molecules rich in carbon (C), hydrogen (H), nitrogen (N), oxygen (O), phosphorus (P), sulfur (S), and other elements. In autumn, the leaf dies and falls to the ground. Over time, the leaf decays. Bacteria and fungi feed on the leaf material. In the process, its large organic molecules are broken down to simpler inorganic molecules. Focusing on just one element, the nitrogen atoms that were in the leaf are now released to the soil as ammonium (NH4+). To complete the cycle back to life, plants take up ammonium through their roots and use its nitrogen to make proteins and nucleic acids. Thus, nitrogen has cycled between the biotic and abiotic components of the ecosystem.
Energy and nutrients take many complicated paths through an ecosystem. Feeding relationships are a good place to start; they describe how atoms and calories move through the living community. Take as an example the simple act of eating a tomato. A tomato plant grows and produces fruit only if it has plenty of light and nutrients. Chemical energy is stored in the bonds that join together the tomato’s organic molecules. After you eat and digest a tomato, some of its molecules are broken down into their subunits. Your cells can use some of that energy to make new molecules or run processes in your cells. Much of it is released as heat.
The simple example above illustrates two transfers of energy: sunlight → tomato and tomato → human. Every organism in an ecosystem must obtain its energy through one or more such transfers. Some organisms are close to the source of energy (the tomato); others are several steps down the line (a hawk that eats a bird that ate a caterpillar that ate the tomato plant). We can break a community up into a series of food-based groups called trophic levels. Trophic level 1 is composed of producers that form the base of all ecosystems. Organisms in level 1 are eaten by organisms in level 2; organisms in level 2 are eaten by organisms in level 3; organisms in level 3 are by organisms in level 4, and so on.
The energy that fuels most ecosystems comes from the sun. Photosynthetic producers such as plants are found in trophic level 1. Organisms in level 1 do not consume other living organisms to obtain nutrients. They can make simple organic food molecules from inorganic raw materials using light energy from the sun. Plants, algae, and cyanobacteria are the photosynthetic producers that make up trophic level 1 in most ecosystems. Organisms in level 2 and beyond are called consumers because they obtain food by eating plants or by eating organisms that have eaten plants. Animals, fungi, and many microbes are consumers. Consumers are identified as primary, secondary and tertiary depending on their trophic level. Cows eat grass and are primary consumers. When you consume a salad, you too are acting as a primary consumer. Humans who eat meat from cattle act as secondary consumers.
In addition to the chains of consumers that eat living prey, many consumers feed on dead organic material. Dead plants, dead animals, shed parts of animals such as fur or skin, shed parts of plants such as leaves, and feces all contain nutrients and calories. Animals that feed on dead organic matter are called detritivores; some of them specialize on eating dead animals (e.g., vultures) and others feed on dead plant material (e.g., earthworms, which consume decaying leaves). Bacteria and fungi that colonize dead material and absorb food molecules from it are called decomposers. Detritivores and decomposers extract remaining calories and nutrients found in dead organisms and dung. In the process, they release simple inorganic wastes to the environment, making chemical elements available to producers.
A food chain is a group of organisms that are joined in a linear series of feeding relationships. Each species in a food chain gains energy from a single source and is, in turn, a source of energy to at most one consumer as shown in the picture below:
In the food chain above, the plant is the producer, the grasshopper is the primary consumer, the mouse the secondary consumer and the snake is the top predator or tertiary consumer. Food chains help us trace the flow of energy and materials through ecosystems. Based on the transfers shown above, plants are in trophic level 1, grasshoppers are in trophic level 2, mice are in level 3, and snakes are in level 4. Each level is dependent on the previous level as a source of food and may provide nutrients and energy to a subsequent level.
Food chains are not very realistic. In real ecosystems, food chains overlap and are interconnected as food webs. Consumers rarely specialize on only one type of food. Many of them are omnivores, feeding on more than one trophic level. Mice, for example, may eat:
In addition, most organisms are fed upon by more than one species. A mice could be eaten by a snake, but it could also be eaten by an owl, a hawk, or a house cat. A complete food web shows all of these possible feeding relationships. In diagrams that depict food chains or food webs, the arrow points from the organism being eaten to the organism doing the eating. In other words, the arrows follow the flow of energy and nutrients through the ecosystem.
We have seen how food chains, trophic levels, and food webs can be used to describe how energy and materials move through an ecosystem. Although they are not as realistic as food webs, food chains and trophic levels do provide us with some important insights. They help us understand, for example, why top predators are rare and why meat costs more than grain.
All organisms require energy for growth, reproduction, and metabolism. But as energy flows through a food web, much of it is lost at each transfer along the way. Let’s trace how energy moves through a simple ecosystem consisting of tomato plants, caterpillars, and birds.
Transfer 1: sunlight → tomato plant. Most of the sunlight energy that reaches the plant is reflected or absorbed and re-emitted as heat. Some is captured by photosynthesis. Most of this is used by the plant for aerobic respiration and ATP generation; ultimately this energy is released as heat. Only about 1% of the incoming solar energy is stored as calories in the molecules within the tissues of the growing plant.
Transfer 2: tomato plant → caterpillar. Much of the energy in the leaf material eaten by a caterpillar is not digested and is excreted in feces. Of the energy that is absorbed, most is used in aerobic respiration and is released as heat. Only about 10% is stored as calories in the molecules (fats, proteins, etc.) that make up the tissues of the caterpillar.
Transfer 3: caterpillar → bird. Some of the energy in the caterpillar is excreted in the bird’s feces. Most of what is absorbed is used in aerobic respiration and is released as heat. Only about 10% is stored as calories in the tissues of the bird.
If you are a bird-eating consumer such as a snake or hawk, very little energy is available to you. Based on the estimates above, you would have access to only 0.01% of the solar energy that fuels the ecosystem as a whole.
As you can see, only about 1% of incoming solar energy is stored as calories in the tissues of growing photosynthetic producers. This energy is then passed up to higher trophic levels through a series of transfers. On average, only about 10% of the available energy in one trophic level is incorporated and stored as calories in the bodies of the next level up. The rest of the energy is released undigested in feces or is dissipated as heat. The result is that all ecosystems show a pattern known as the energy pyramid in which energy availability decreases with increasing trophic level. The bottom of the pyramid contains producers (usually plants or algae). Each trophic level of consumers (level 2 and above) becomes progressively smaller. The amount of energy available to consumers at the top of the pyramid is much smaller that what is available to organisms at lower levels.
The diagram below shows each trophic level as a block of color. In keeping with real patterns as seen in many ecosystems, each block’s area is 10% or one-tenth that of the preceding block. That leaves mighty slim pickings for level 5!
In real ecosystems, energy availability can be measured in many different ways. The mass of organisms at each trophic level can be measured, or their total annual calorie requirements can be estimated. In either case, the overall energy pyramid pattern remains quite consistent. This helps to explain why huge hunting grounds support only small numbers of top-level predators such as killer whales, great white sharks, or tigers.
The energy pyramid also has implications for humans. Our diet is very flexible. Strict vegetarians live on trophic level 2. Those who eat a lot of animal products live closer to trophic level 3. When grain is fed to animals much of its calorie content is used by the animals themselves, and only a small fraction is available to us in the food that is produced. Therefore people who eat a diet rich in animal products generally have a greater total environmental impact—in terms of crops consumed, fertilizers used, land area farmed, etc.—than those who eat plant foods. And as seen in the largely vegetarian traditional diets of rural China or India, a given land area can support more people at lower cost if they live on trophic level 2.
Energy enters ecosystems as sunlight and dissipates from all living things as heat. Producers (trophic level 1) have access to the greatest amount of energy (sunlight). As they grow, they store up food calories that supply all consumers with energy. Consumers feed on producers directly (trophic level 2) or feed on each other through a series of transfers that can be described as a food chain. The amount of energy available to consumers declines by about 90% with each transfer, so top predators at the end of long food chains tend to be rare.
All living things are made of carbohydrates, lipids, proteins, and nucleic acids, and they all contain roughly the same blend of chemical elements. The most important are carbon (C), hydrogen (H), oxygen (O), nitrogen (N), phosphorus (P), and sulfur (S). Earth has a finite stock of each of these chemical elements, but the atoms are immortal. In any transfer or reaction the total number of atoms of each element will stay the same. Matter does not disappear or appear from nothing; it is conserved.
Earth’s immortal atoms move through living things as we feed, breathe, and excrete solid and liquid wastes. The same atoms are also found in abiotic reservoirs, stored up within rocks, soil, water, and air. They bond with other elements to take shape as many different materials. Carbon, for example, makes up the backbone of all organic macromolecules. Outside of life, it exists as carbon dioxide gas that mixes into the air and dissolves into water. This gas gives carbonated beverages their fizz. Limestone rock, coal, and diamonds are some solid materials rich in carbon.
Organisms must obtain chemical nutrients in very specific forms. You would starve without suitable organic macromolecules (food), even if you were surrounded by carbon-rich limestone, air, and wood. A plant has very different needs. It would starve without access to carbon dioxide gas, even if you planted its roots in a juicy porterhouse steak.
Organisms’ needs are met through continuous recycling of chemical elements. Direct biotic cycling occurs when producers use wastes excreted by consumers; urine contains nitrogen and phosphorus compounds that can act as plant fertilizer. Cycling also occurs through decomposition. When an organism dies, its macromolecules are broken down by bacteria or fungi and the elements are returned to soil, air or water as simpler compounds. Physical processes (weather, formation and erosion of rocks, etc.) also transform and transport elements. Biogeochemical cycles are the pathways that chemical elements follow through ecosystems; they involve life (bio) and nonliving components (geo) .
We will discuss four cycles that are key for living things: the water, carbon, nitrogen, phosphorus cycles. Key features of each cycle are summarized below in the table:
Water H2O | Carbon C | Nitrogen N | Phosphorus P | |
main gaseous form | water vapor (H2O) | carbon dioxide (CO2) | Nitrogen gas (N2) | none |
importance to life | solvent that makes up most of the mass of all living cells; stabilizes temperature | forms the backbone of carbohydrates, lipids, proteins, and nucleic acids | found in proteins and nucleic acids | found in phospholipids, ATP, and nucleic acids |
abiotic reservoirs | air, groundwater, ice, salt and fresh water bodies | air, dissolved in water, soil, rock | air, dissolved in water, soil | dissolved in water, soil, rock |
available to producers as | liquid water | carbon dioxide (CO2) in air or dissolved in water | ammonium (NH4+) or nitrate (NO3−) dissolved in water | phosphate (PO43−) dissolved in water |
major driving processes | physical processes (evaporation, precipitation) plus plant transpiration | aerobic respiration and photosynthesis; formation and erosion of rocks | bacterial action; for example, bacteria convert N2 gas from air to usable form for producers | formation and erosion of rocks |
Water is an input to photosynthesis and a waste product of aerobic respiration, so it is biotically cycled. These flows are relatively small. Solar energy powers the major processes driving Earth’s water cycle. Water evaporates, condenses in the air, returns to Earth as precipitation, and is transported through wind, river flow, and ocean currents. Land plants play an important role in the water cycle. They take up water through their roots and release it as vapor through tiny holes in their leaves called stomata. This process, called transpiration, greatly accelerates the transfer of water from soil to the atmosphere. A forest can increase humidity and even boost rainfall in a region through this process.
Humans also influence the water cycle by damming rivers and by pumping groundwater. We slow the return of water to the ocean and redistribute it to meet our needs; by far the largest human use of water is crop irrigation.
Carbon is cycled through a variety of biological and physical processes. Photosynthesis removes carbon dioxide from the atmosphere; aerobic respiration by living plants and animals returns carbon to the air. Decomposition, a combination of aerobic respiration and fermentation carried out by microbes, takes carbon from dead material and returns it to the atmosphere. In the much longer term, geologic processes store carbon in fossil fuels and in rocks such as limestone. This stored carbon is slowly returned to active circulation by erosion of rock and by volcanic eruptions.
Nitrogen is cycled mainly through biological processes. The atmosphere is a huge reservoir of nitrogen; it is almost 80% N2 gas. However, most life forms cannot use nitrogen gas. Plants and algae obtain nitrogen as nitrate (NO3−) or ammonium (NH4+). These compounds dissolve in water and are absorbed directly (algae) or taken up through the roots (plants). Bacteria are the key intermediaries that drive the nitrogen cycle. Some bacteria, called nitrogen fixers, transform N2 gas to ammonium that plants can use. Bacterial and fungal decomposers break down dead organic matter and also generate ammonium. In traditional and organic agriculture, plant and animal wastes (particularly manure) are collected and composted to favor the activity of decomposers and nitrifying bacteria. The resulting nitrate-rich compost is spread on fields, providing usable nitrogen to crop plants.
Phosphorus does not have a gaseous form, so it is not cycled globally through the atmosphere. Instead it circulates among living organisms, soil, water, and rocks. You can think of the phosphorus cycle as a combination of two cycles: one fast and local, the other very slow and global. The fast cycle is driven by producers, consumers, and decomposers. Plants and algae take up phosphate dissolved in water and use it to make organic compounds. Consumers feed on them and excrete phosphorus in their feces and urine. Bacteria and fungi decompose dead organic matter and return phosphates to the soil and water, completing the loop. Producers are extremely efficient in taking up phosphate from water and phosphorus is often a key resource that limits the growth of plants. Phosphorus does not have a gaseous form, so it is not cycled globally through the atmosphere. Instead it circulates among living organisms, soil, water, and rocks. You can think of the phosphorus cycle as a combination of two cycles: one fast and local, the other very slow and global. The fast cycle is driven by producers, consumers, and decomposers. Plants and algae take up phosphate dissolved in water and use it to make organic compounds. Consumers feed on them and excrete phosphorus in their feces and urine. Bacteria and fungi decompose dead organic matter and return phosphates to the soil and water, completing the loop.
The slow phosphorus cycle involves the buildup, movement, and erosion of rocks. Over time, phosphorus tends to follow the flow of water downhill. Phosphorus-rich sediments build up on ocean and lake bottoms, forming phosphate rocks over time. Seabird and bat colonies can also produce concentrated deposits of guano (excrement) that is very rich in phosphates. Through slow geologic processes, rocks containing phosphorus can be lifted up to form new land masses and their erosion can make phosphates available to producers once again.
Humans now play a huge role in the global nitrogen and phosphorus cycles. Today, artificial fertilizer production fixes more nitrogen than all natural processes combined. Phosphate mining removes phosphorus from storage in rocks much more quickly than natural uplift and erosion. Fertilizers run off farm fields and enter water bodies directly; human and animal wastes also deliver excess nitrogen and phosphorus to water bodies. The result is eutrophication, the overfertilization of aquatic ecosystems by humans. Excess ammonium can kill fish. Extra phosphates and nitrates fuel blooms of algae and cyanobacteria. These turn water pea-soup green and can lead to fish kills and other negative effects.
The stuff of life is continually recycled by biological and physical processes. Each element follows its own unique pathway through an ecosystem and is concentrated in specific abiotic reservoirs. Human activities can strongly modify how natural chemical cycles work, sometimes leading to the accumulation of wastes and environmental harm.
In the past, the science of ecology focused on nonhuman species. Rather than studying cities or farms, ecologists have worked in “wild” places like forests, mountain streams, coral reefs, and grasslands. Ecologists sought to unravel how plants and animals in these places interacted with each other. The activities of humans were viewed as a separate issue, outside the normal set of processes at work in “natural” communities and ecosystems.
Today, however, ecologists recognize that humans have an influence on all communities and ecosystems. In some ways, we are simply participants in and members of ecosystems. Like other animals, we feed on the bodies of plants and other animals; excrete gaseous, liquid, and solid wastes; and disturb small areas of soil when we travel across the landscape on foot. In most of our activities, however, humans have impacts that go far beyond those of other animal species. What makes us different?
Each of the above factors promotes greater environmental impact if all other factors are held equal. In real societies, however, all three factors change over time and influence each other. This leads to some surprising patterns of impact. For example, some environmental conditions are better in high-income nations than in lower-income countries. In general, the air and water are cleaner, industries are better-regulated, and forests are less heavily exploited in higher-income nations. We can account for this in two ways. First, goods can be consumed in one place but generate their impacts elsewhere. With global trade, many of the impacts of consumption occur in developing nations where goods are produced, dumped, or recycled. Second, more advanced technology is not always more disruptive; technology can also be developed to reduce impact. Examples of advanced low-impact technologies include renewable energy technologies (solar panels, windmill generators), biodegradable plastics, smokestack “scrubbers” that remove pollutants from power plant emissions, and energy-efficient appliances and buildings.
Although there are many competing priorities, environmental concerns have become more prominent in our society since the 1970s. Corporations have adopted environmental goals as part of their mission statements. Nonprofit groups advocate for the environment. Government programs promote specific environmental goals. We use buzzwords like “sustainability” and “stewardship.” Products and ideas are marketed as being “ecosafe,” “earth-friendly,” or “green.”
As a consumer, voter, employee, or business leader, you have some control over environmental impact in your daily decisions. You should have a clear sense for what sustainability really means. Sustainable technologies meet human needs—economic and social—and preserve the productivity or biodiversity of ecosystems over a long period of time. In any given case, different options may maximize individual economic benefit, public welfare, or environmental integrity. The most sustainable choice often involves finding a compromise among these competing priorities.
One step toward sustainability is the careful use of renewable resources: those that are replenished quickly enough to replace what we consume. Resources are renewed through growth or other ongoing processes. Most products of plants and animals are potentially renewable, so long as we do not consume them too quickly. Many physical processes (solar energy input, river flow, tidal flow, wind, flow of heat from Earth’s interior) also occur at a high and sustained rate.Nonrenewable resources, by contrast, are replenished very slowly. Such a resource can be thought of as a fixed stock. Once depleted, a nonrenewable resource cannot be replaced in useful quantities. Most mineral resources and fossil fuels are nonrenewable. Over time, if humans continue to use these resources they will become increasingly scarce. A second factor in sustainability involves the disposal or ultimate fate of materials and wastes we produce. Biodegradable materials are more sustainable over time because natural decomposition processes will break them down and recycle any nutrients they contain. Nonbiodegradable materials tend to accumulate over time and may cause negative environmental impacts. Some examples include plastics and persistent toxic chemicals.
Renewable resources can be used in a sustainable way so long as the rate of use does not exceed the rate of renewal. Nonrenewable resources cannot be used indefinitely and technologies that rely heavily on them are unsustainable. In the table below we summarize several examples of renewable and nonrenewable resources.
Renewable | Nonrenewable |
---|---|
Foods produced from plants, animals, and other living organisms. Grains, nuts, vegetable oils, fish, poultry, meat, dairy products, vegetables, fruits, etc. | Petrochemicals produced from oil and/or natural gas. Examples include plastics, nylon, polyester, most pesticides, many food additives, adhesives, solvents, and much more. |
Fertilizers made from animal manure, animal byproducts, and composted plant wastes. | Artificial fertilizers made using natural gas, mined phosphate rock, and mined potash (a source of potassium). |
Construction and industrial materials from living plants, including wood, paper, rayon cloth, plant-based plastics, natural rubber, bark mulch, and medicinal compounds. | Metal, cement, and glass, all of which are made using mined materials and are usually produced with a high input of fossil fuel energy. |
Fuels produced from plants or plant-derived by-products such as sawdust or crop wastes. Power supplied by human labor or livestock. Electrical energy produced from sunlight, tides, river flows, wind, and geothermal heat. | Fuels derived from coal, oil, or natural gas. Coal and natural gas are used directly; oil is refined to create products like gasoline and jet fuel along with many petrochemicals. |
Over time, human activities may have permanent negative effects on the environment’s capacity to support life or provide resources to future generations. Ultimately the degree and nature of our impact will depend on how population size, affluence levels, and technologies change in the future. We will explore these issues in more depth in this module. In the next page, we will examine the history of the global human population, tracing some of the factors that have enabled our dramatic expansion in numbers and in impact.
To grasp our collective impact as human beings, we must begin with some analysis of our population’s size and technological reach. How many of us are there? How did we get to our current abundance and affluence? What were some of the major milestones along the way?
Based on fossil and genetic evidence, the first anatomically modern Homo sapiens arose in Africa about 200,000 years ago. The history of humanity can be divided into three broad periods that represent widely differing ways of life. We will call them the Stone Age, the Agricultural Age, and the Industrial Age.
Stone Age. By 50,000 years ago, humans had developed many features of Stone Age technology and culture. By 20,000 years ago, modern humans had dispersed through much of the world. Stone Age humans obtained food by fishing, hunting wild game, and gathering wild plant products. Food sources were unreliable and most humans were nomadic, traveling widely to find food. Shelter, clothing, and tools were made from stone, animal hides, and wood. Fire was used as a source of heat and to modify the landscape for better hunting or travel. Human labor was the main energy source.
Agricultural Age. This age started around 10,000 years ago. It involved many (though not all) human cultures worldwide and extended through the late 1700s. People began to domesticate plants and animals. Food supplies became more reliable and abundant and diets expanded to include grains, poultry, meat and dairy products from livestock, improved fruits and nuts, and more. Agricultural Age humans stayed close to productive farms and pastures and built the first permanent towns and cities. Humans began to extract and forge metal to make durable tools, weapons, and ornaments. Energy sources expanded to include labor from farm animals such as oxen and horses. Windmills and water mills were also used as a source of power. Some cultures developed extensive water control systems for irrigation of crops. Governments, organized religions, and written languages flourished.
Industrial Age. The Industrial Age began in Europe in the 1700s and continues today. Agriculture continued to improve and cities continued to grow. Global trade moved people and ideas around the world. Science helped to spark new technologies. Machines harnessed external sources of energy to accomplish all kinds of work. Fossil fuels became the main source of energy for transportation, farm work, heating, cooling, manufacturing, mining, and other activities. Scientific principles were applied to improve public health, beginning with vast improvements to sanitation in cities. Sewage treatment, safe municipal water supplies, public garbage disposal, and soap are all Industrial Age innovations. Modern medical practices, including aseptic surgery, antibiotics, and vaccinations, were developed. At the same time, the reliability and quality of food supplies continued to improve.
Thus the world’s humans have gone through dramatic changes in their technologies and living conditions. In the activity below, explore and learn more about how rates of birth, death, and population growth changed as a result.
Thus far, we have seen how human technology and population size have changed over the past 200,000 years. How have humans impacted the environment over this time span? Humans have had an impact on Earth’s ecosystems at every stage of our history.
During the Stone Age, for example, human use of fire may have had a profound effect on many ecosystems.By setting forest and grassland fires, humans favor fire-adapted plant and animal species. Hunting may have depleted some food resources and probably contributed to past animal extinctions as well.
During the Agricultural Age, humans cleared forests and created dams and irrigation systems. These actions harmed many species and benefited others. In some cases, resources like forests or fertile soils were depleted on a local or regional scale, contributing to the collapse of human societies. Throughout this history, however, human impacts on the environment were limited in their scope and severity.
Starting with the Industrial Revolution, human impacts on the environment have intensified. Many of them have become particularly extreme since about 1950. Human activities generate impacts through three major mechanisms summarized below.
The scope of human actions and impacts is truly mind-boggling. The map below illustrates our global impact using a “human footprint index” that summarizes several different indicators.
We humans are incredibly ingenious as individuals. But as part of an organized culture that has accumulated knowledge since the Stone Age, our power is far greater. Consider a gadget like a laptop computer. No one human really knows how to make a laptop, and certainly no individual could manufacture one from scratch working alone. Instead, thousands of individuals must contribute specialized knowledge to make the parts and assemble them in a working unit. Making a laptop also depends on extraction of metals and oil from deposits around the world, and requires heavy inputs of fossil energy. Similar feats of cooperation and innovation allow us to make airplanes, smart phones, heart monitors, reliable automobiles, and thousands of other useful objects. Unfortunately, we also have the power to do great damage to the environment and its ability to sustain life. We simplify and reshape ecosystems. We overuse resources so that they become scarce and costly. We make atomic weapons and synthesize chemicals that are both persistent and toxic. We overload natural geochemical cycles, leading to the accumulation of wastes. These and other impacts will be described further in this module.
Next we will focus on one particular aspect of human impact that is very relevant to the field of biology: how humans have changed biodiversity on a global scale.
From earliest childhood, humans are fascinated by the diversity of life. Tourists flock to African savannas, coral reefs, and tropical rain forests, cameras at the ready. Kids’ picture books are populated with bright, beautiful, and bizarre animal drawings. Television nature programs, zoos, botanical gardens, and natural history museums provide an up-close look at the real thing. At some level each of us feels interested in, inspired by, or connected to nature in some form. And we are saddened to learn that much of this diversity is threatened by human activity. What is at stake? What has been lost? What can we do to protect what’s left?
Biodiversity is simply the diversity of life in an area. As seen in the photo below, it can be measured at a number of different levels.
Earth is home to a diverse and vibrant riot of life. Over the past 400 years, biologists have discovered, described, and named over 1.3 million species of animals, plants, fungi, protists, and prokaryotes. Millions more exist but have not yet been described.
Biodiversity is of great value to humans. Functioning natural ecosystems supply us with countless benefits, many of which we take for granted. When was the last time you paid—or even expressed thanks—for oxygen? Some of the most important benefits provided by ecosystems and by specific wild species within them include:
As a whole, biodiversity may seem limitless. But when we look at specific communities and species, we can see evidence to the contrary. Biodiversity is reduced when populations are eliminated from local communities (local extinction). Local extinctions can severely disrupt ecological communities and may even lead to additional extinctions. This is particularly likely if keystone species are lost, as you learned in the community ecology module. Over time, local extinctions can lead to the loss of an entire species when the last reproducing group dies out (global extinction). Global extinction is irreversible: When a species goes extinct, we lose all future opportunities to learn from it, use it, and appreciate it.
Human activities have affected biodiversity for thousands of years. In fact, human hunting is one probable cause for the extinction of many large animal species that ended about 10,000 years ago. We may never know for sure if humans caused the demise of the woolly mammoth or the saber-toothed tiger, but we can be quite sure that our activities severely threaten many species today. Tracking and clarifying this problem is a major challenge for ecologists. Below is a diagram showing how many species are threatened (or already extinct) among some well-studied groups of plants and animals.
Extinction is a naturally occurring process; thousands of species known only from fossils attest to that fact. However, current rates of endangerment and extinction are much faster than those estimated from a study of the fossil record. Today, the vast majority of extinctions can be clearly attributed to the environmental impacts generated by human activities. The major causes of extinction can be categorized as follows:
In summary, species and communities worldwide face many threats. Very often, a single species will be affected by multiple factors. In most cases, humans do not set out to deliberately eradicate species. Instead, we cause extinction inadvertently when our actions generate unexpected consequences.
In the past, humans were often unaware of the effects of their actions or did not try to change course until it was too late. Today, with advances in technology and ecological science, we can do much better. We have begun to effectively regulate or prohibit harvesting, reduce some forms of pollution, protect remaining habitats, and control introduced species. As a result, the status of many populations in high-income nations or in priority areas is improving. In the United States, several species, including the American beaver, the American bison, and the gray wolf, have been protected from overharvesting and have recovered and returned to much of their former range. The bald eagle and peregrine falcon were nearly driven extinct by the pesticide DDT and related compounds that entered food chains, concentrated in the birds, and impaired their introduction. Both species have recovered strongly since DDT was banned for use in the U.S. in 1972. As seen in the graph below, these and similar cases have led to an overall increase in many vertebrate animal populations within high-income nations in North America, Europe, Australia, and Japan since 1970.
Much of the world’s remaining biodiversity is in the tropics within low- and moderate-income nations. Here, the biodiversity situation remains much more troubling. Resources for conservation efforts are limited and high rates of human population growth continue to drive habitat destruction and overharvesting in many of these nations. Poaching (illegal hunting), forest destruction, poorly regulated mining, overfishing, and destructive farming and livestock grazing practices are all contributing to biodiversity decline that threatens thousands of unique and irreplaceable species. On the positive side, conservationists are working very hard to protect wild habitats within “biodiversity hotspots” that harbor a large diversity of species. Nongovernmental organizations from high-income nations are spearheading these efforts. Through investments and education, they are working with communities to develop sustainable businesses that provide income while protecting the integrity of local ecosystems.
To Learn More / References:
In this page we will focus on human effects on the atmosphere and climate as a chief example of our ecosystem-level impacts. You have probably heard and learned about global warming and global climate change. Through this course we aim to help you improve your knowledge about the science behind this issue and clear up any misconceptions you may have about it.
A continuing increase in Earth’s ocean and land surface temperatures since the early 1900s is known as global warming. The phrase “global warming” may leave the false impression that the Earth simply gets a bit cozier over time as this process continues. Earth’s climate is a complex system. As the climate warms, temperatures are not evenly adjusted upward by a fixed number of degrees across seasons and locations. Instead, we are also seeing many dramatic weather anomalies in recent decades, including strong hurricanes and tornadoes, protracted droughts, unusually severe heat waves, and even record snowfalls. Instead of global warming, some like to say we are experiencing “global weirding.” Anthropogenic global climate change is a more appropriate description of how climate is responding to human activities on a global scale. It is the full range of climatic disruptions that have occurred as a probable result of human alteration of the atmosphere. In other words, it is recent climate warming—and weirdness—that is probably caused by humans. The case for global climate change can be summarized as follows:
Let’s look at some of this theory and evidence, beginning by thinking about how the Earth’s overall temperature is determined.
Energy enters the Earth system—its land, water, and air—as sunlight. Some of the light is reflected back into space, but much of it is absorbed by land and water. When materials absorb light, they warm up. Over time, they release energy as heat, also known as infrared radiation. Eventually, all of the energy that entered as sunlight escapes back into space as infrared radiation. Energy that enters the Earth system as light is balanced by energy that exits as heat.
If the Earth had no atmosphere, heat would escape very rapidly. The Earth’s night-time surface temperatures would be like those on the moon: −153° Celsius, −243° Fahrenheit. Below is a diagram showing what Earth’s energy balance might look like if there were no clouds to reflect light and no gases that could interfere with heat’s escape.
Lucky for us, Earth has an atmosphere comprising gases that make our atmosphere act something like an insulating blanket. The atmosphere is a mixture of many different gases that are held close to Earth’s surface by gravity. The two most abundant gases in the atmosphere are N2,or nitrogen gas (78% of air by volume), and O2, or oxygen gas (21%). These gases are transparent to light and to heat radiation; that is, they do not interfere with sunlight energy coming into Earth or with the heat radiating back into space. Several much less abundant gases, however, do have a major effect on Earth’s energy budget. These greenhouse gases absorb and reemit heat waves (infrared radiation). Three important greenhouse gases are water vapor (H2O), carbon dioxide (CO2), and methane (CH4). Much of the heat (infrared) radiation that leaves Earth’s surface is absorbed by greenhouse gases and reemitted back toward Earth. As a result, these gases retain heat in the atmosphere. Instead of escaping immediately to space, heat is delayed near the Earth’s surface and continues to warm our atmosphere.
As you just learned, greenhouse gases in the Earth’s atmosphere retain heat and keep surface temperatures warm and relatively steady over time. This phenomenon is called the greenhouse effect. It is a natural part of the Earth’s climate system and helps make our planet livable. Currently, however, the greenhouse effect is strengthened by a high and increasing level of greenhouse gases in the atmosphere. Evidence indicates that human activity is responsible for this increase.
In terms of its total effect on climate, the most significant greenhouse gas is carbon dioxide. As of May 2012, the concentration of carbon dioxide in the atmosphere is about 397 ppm (parts per million). How does this compare to past levels?
The situation is similar for other greenhouse gases including methane and nitrous oxide, each of which has increased dramatically since about 1900.
Clearly, greenhouse gases are increasing. And you already know that human population and impact have increased greatly through the Industrial Age, so it is natural to suspect that human activities are the cause of recent changes in atmospheric gases. Many lines of evidence confirm this. Specific human activities and practices are known to emit greenhouse gases and are also known to have increased in step with rising greenhouse gas concentrations.
Many ongoing natural processes also influence levels of greenhouse gases, but these processes have not changed in a way that can account for the extra CO2 in the air. Typically natural processes cancel each other out; they are part of cycle that is more or less in balance. Human activities, by contrast, have a history of rapid increase that tracks very closely with the observed increase in the air’s greenhouse gas levels over the past century.
As previously discussed, greenhouse gases absorb and reemit heat. There is a clear theoretical basis for the idea that increases in greenhouse gas levels should warm the climate. There is also an abundance of direct evidence that climate has warmed in association with higher concentrations of greenhouse gases over time. Coinciding with the observed increase in greenhouse gases, scientists have documented substantial increases in the average temperatures of Earth’s air and oceans, particularly since about 1950.
Many other ongoing natural processes have an influence on the climate. Why do scientists think greenhouse gas increases are the factor driving the recent warming? Again, natural processes have not changed in a way that can explain the observed warming of the climate. For example, the sun’s energy output changes over time, and this can influence our climate. But these changes cannot account for recent global warming. In fact, the sun’s energy output was considerably lower than normal from 2005 to 2010, but the climate continued to warm over that period.
Taken together, thousands of observations and trends provide evidence for global climate change and indicate some of its consequences for life on planet Earth. The following are a few of the many strong indicators of change:
Warm weather and ocean temperatures. In a record of weather and ocean temperature data stretching back to 1900, the top five warmest years were 2005, 2010, 1998, 2003, and 2002. Studies based on weather stations and ocean temperature measurements around the world have repeatedly found a trend of increasing temperatures since about 1900 with steep increases since around 1970. The data for land weather stations are summarized in the graph below. As average temperatures have increased, so has the frequency of extreme heat waves. These are particularly dangerous in low-income nations like India, where people lack air conditioning.
Sea level rise. The average sea level has increased by an estimated 0.2 m since 1900, both because the ocean water is expanding as it warms and because glaciers on land are melting. Sea level rise is expected to continue and accelerate with continued warming of the climate. Particularly when coupled with more intense storms, increasing sea levels threaten human settlements and natural ecosystems in low-lying coastal areas worldwide.
Ice and permafrost melting. The extent and thickness of floating sea ice in the Arctic has declined in recent decades. Based on satellite data, the area covered by thick multiyear ice decreased by about half between 1980 and 2012. A majority of mountain glaciers are also shrinking in thickness and area over time. Perpetually frozen soils (permafrost) have been melting to increasing depths each summer in the Arctic. Such changes have many implications. Sea ice is important to the ecology of polar bears and other arctic animals; it acts as a platform for travel and foraging. Mountain glaciers supply summer melt water to major river systems in Asia and elsewhere; glacier retreat may reduce the reliability of water supplies for irrigation and other uses. And melting of permafrost leads to landslides and releases additional greenhouse gases to the atmosphere.
Increased intensity or duration of drought. Climate change is likely to reduce rainfall in some regions. Even if rainfall does not decline, warmer temperatures lead directly to drier soils and lower water levels by speeding up evaporation and boosting the rate at which plants use water. Record-breaking droughts and wildfire seasons are some indicators that this effect is already under way in some regions. Droughts threaten human well-being by limiting crop production and by reducing the availability of water for municipal use, navigation, power generation, and recreation.
Changes in species distributions. Many species’ geographic ranges are shifting, following climate conditions toward historically cooler areas as their existing habitats get warmer. Species are shifting toward the poles: further north in the Northern Hemisphere and further south in the Southern Hemisphere. In mountainous regions, species’ distributions are shifting upward to higher elevations. Range shifts may lead to endangerment of species, loss of species that provide economic resources, and northward spread of tropical pathogens and their vectors (particularly mosquitoes).
Changes in seasonal patterns of life. As the climate warms, springtime events—leaf emergence and flowering, nesting and migration of birds, etc.—are occurring at earlier dates than in the past. A 2003 study reviewed data from studies on 172 different species of plants, birds, butterflies, and amphibians and found that springtime events had shifted toward earlier dates by an average of 2.3 days per ten years; most of the studies extended over a period of about 50 years. Such shifts can disrupt community interactions, particularly since some species respond more quickly than others to changes in climate. Studies have found that some birds arrive at summer breeding grounds only to find that their prey (insects) have already completed their life cycles. Similarly, some butterfly species are in trouble because their food plants are developing tough leaves and beginning to wither earlier in the season, before the caterpillars have hatched. To learn more about how butterflies are responding to climate changes view the video in the “Changing Planet” series by NBC Learn on “ The Adaption of Butterflies .”
We began this module by considering how three factors have changed over time: human population size, affluence or material consumption, and human use of high-impact technologies. Next we examined how these factors have led, through various mechanisms, to effects on biodiversity and climate. But what does all this say about the future? Here we will conclude by looking at how population, affluence, and technology may change, and how these changes are likely to affect humanity’s impact on the global environment.
Based on our graph showing how population has exploded during the Industrial Age, you may assume that it will continue to increase dramatically with no end in sight. Recall that world population began to grow rapidly in the 1800s. At this time death rates began to fall rapidly with advances in public health worldwide. Some of the most important factors were improved sanitation, basic medical care, and nutrition; later the trend was reinforced by immunization and the use of antibiotics. This health-driven reduction in death rates started in western Europe, North America, and other industrialized nations. Today, all nations have benefited to some degree from these trends; total longevity is increasing worldwide and rates of infant and child mortality continue to fall. Death rates are low and headed lower. Does this mean populations are growing equally fast everywhere?
Fertility (which corresponds to the birth rate) is the main factor that determines modern and future rates of population growth. Since about 1900, birth rates have declined in industrialized, high-income nations. Why? Education and economic opportunity for women, access to contraception, and changing economic pressures have all contributed to reduced fertility in these countries. With low rates of death and birth, populations are stable or even declining in many of the world’s most affluent nations. In less developed low- and middle-income nations, fertility rates are still high enough to generate rapid population growth. In addition, these countries have populations with many young people just now entering reproductive age. Therefore the populations of these nations will continue to grow rapidly for years into the future.
Overall, the global population will continue to increase for at least a few decades, and this increase will be driven mainly by growth in the low- and middle-income nations of Africa and Asia. Experts predict that fertility will continue to fall in most nations as they develop economically, as contraception becomes more available, and as educational and economic opportunities for women improve. If so, the world population is projected to peak at around 9 to 11 billion in 2050 and begin to decline thereafter.
As their populations continue to expand, trends and projections indicate that low- and middle-income nations will also continue to increase in their affluence. Automobiles, electronic gadgets, and household conveniences like air conditioning and indoor plumbing are likely to reach more and more people in populous and fast-growing economies like those of Brazil, India, and China. What does this mean for the global environment? By 2050, projections suggest that the world’s humans may use a total of 140 billion tons of minerals, ores, fossil fuels, and biomass per year—three times the current consumption rate (United Nations Environmental Programme, International Resource Panel, 2011). Will Earth be able to sustain these high levels of consumption spread across a population of 9 to 11 billion people? Much will depend on the next factor, technology.
After high standards of living have been achieved, societies have tended to shift toward use of lower-impact and greener technologies. Take, for example, the use of fossil fuels. Within the United States and Europe several trends such as greater efficiency and a shift toward renewable and nuclear power have led to reduced emissions of carbon dioxide per dollar of goods and services generated.
Carbon dioxide emissions per dollar of national economic activity in China, the United States, and France in 2009. For each unit of economic activity, China generated four times as much carbon dioxide than did France.
China: 0.81 kg CO2 / $ GDP | U.S.A.: 0.41 kg CO2 / $ GDP | France: 0.18 kg CO2 / $ GDP |
---|---|---|
Heavy reliance on coal, older and less efficient industrial and consumer technology | Continued primary reliance on fossil fuels, newer and more efficient technology | Strong shift toward nuclear power, more energy-efficient technology |
Several important low-impact technologies are currently in various stages of development. Some examples include renewable electric power (mainly solar and wind power), use of plant and algal biomass as a source of renewable fuel, and improvements in the sustainability of appliances and buildings. In many industries, improvements are occurring gradually through better management practices that reduce waste output, reduce energy and resource use, and reduce harm to the environment. These efforts are often supported by certification systems and ecolabels that help consumers choose lower-impact products. Examples include the Forest Stewardship Council for sustainable forest products, the Marine Stewardship Council for sustainable seafood, LEED certification for sustainable buildings, and the U.S. EPA Energy Star and Water Sense labels for energy- and water-efficient appliances.
If these trends continue and spread globally, humanity may be able to limit some of its future environmental impacts through technological progress. However, new and more sustainable green technologies are not guaranteed to succeed. They are usually developed first in industrialized nations and their use may be limited by cost. Ultimately, green technologies will reduce our collective environmental impact only if they are widely used. To have a large effect, they must be adopted widely in industrialized nations and then quickly transferred to the fast-growing and populous nations of Asia and Africa. Major public and private investments may be needed to meet this goal.
Climate change is the most troubling environmental impact of humanity because of the global and potentially severe nature of its effects. How can humans tackle such a big and complex problem? Efforts to deal with global climate change can be placed in two broad categories: mitigation and adaptation.
Climate change mitigation is the effort to reduce emissions of greenhouse gases or increase the rate at which they are removed from the air, thereby slowing the pace and reducing the severity of climate change. Some of the actions that could be taken to accomplish this include energy conservation, widespread adoption of renewable energy technologies, increased use of nuclear power, switching from coal to natural gas, and capture and storage of CO2 from coal-burning power plants. We can also work to increase the extent and growth of carbon-absorbing forests. Changes to agriculture and waste disposal practices could help reduce emissions of methane and nitrous oxides. At a fundamental level, efforts to slow population growth may mitigate climate change simply by reducing population growth, which is a key underlying factor driving all human impacts on the environment.
Changes to mitigate climate change can be costly. Deciding whether to make such changes requires detailed scientific analysis. How much will a given action help to reduce climate change? What other costs and benefits will be generated?
Unfortunately, there is a limit to what we can reasonably expect to achieve with mitigation. Some degree of climate warming has already occurred. Experts agree that even with our best efforts to mitigate growth in greenhouse gas emissions, these gases will continue to build up in the atmosphere and will continue to modify climate for many decades.
Climate change adaptation is the effort to make humans and ecosystems less vulnerable to the effects of ongoing and future climate change. To reduce the consequences of climate change, we can develop crops that are better adapted to a warmer and drier world; improve irrigation systems to compensate for expected droughts; protect coasts from rising sea levels with sea walls or with buffer zones of natural vegetation; improve emergency response systems and preparedness; and deliberately move people, plants, and other species to areas where the climate is likely to be more tolerable in the future. Scientific assessments help to predict future changes and test the effectiveness of adaptation strategies.
Use this table as a quick reference for the major elements.
Atomic Properties of the major biological atoms | ||||||||||||||||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
|
The following is a list of the key terms discussed in this Introduction to Biology course and their meaning.
glucose + fructose = Sucrose (table sugar) |
glucose + galactose = Lactose (milk sugar) |
grass → grasshopper → rat → owl |
acid + base → salt + water |
Biologists have different ways of representing chemical structures. Each type conveys different information.
Portions of this course are built on materials developed and generously provided by University of Maryland University College, made available with permission under a CC-BY-NC license. Direct use of specific activities and media elements are noted throughout the course.
The development and ongoing improvement of an OLI course is always a collaborative effort, built on the time, talent and commitment of many individuals. This OLI Biology Course is no exception.
Course Authors: | Affiliation |
---|---|
Ana Marie Barral | National University |
Beth Carpenter | University of Maryland University College |
Anya Goodman | California Polytechnic State University |
Jon Hoekstra | Heartland Community College |
Brian Kram | Prince George Community College |
Debra McLaughlin | University of Maryland University College |
Wendy Riggs | Redwood Community College |
Gordon Rule | Carnegie Mellon University |
Suzanne Wakim | Butte Community College |
Kathy Warner | University of Maryland University College |
Other Contributors: | Affiliation |
Diana Bajzek | Carnegie Mellon University |
Norman Bier | Carnegie Mellon University |
Michaele Brown | Carnegie Mellon University |
Renee Fisher | Carnegie Mellon University |
Jim Greeno | Carnegie Mellon University |
Michelle Meyers | Editor |
Candace Orsetti | University of Maryland University College |
Sandy Raysor | Carnegie Mellon University |
Robin Searles-Adenegan | University of Maryland University College |
Keris Simmonds | University of Maryland University College |
Cheryl Templeton | Carnegie Mellon University |
Daniel Williams | Winston Salem State University |