Bacteria are small single-celled organisms. The human body is full of bacteria, and in fact is estimated to contain more bacterial cells than human cells. Most bacteria in the body are harmless, and some are even helpful. A relatively small number of species cause disease.
Bacterial cells of Staphylococcus aureus
Image by Eric Erbe, Christopher Pooley
What Is a Bacteria?
Bacteria
Image by National Human Genome Research Institute (NHGRI)
Bacteria
Bacteria are small single-celled organisms.
Image by National Human Genome Research Institute (NHGRI)
What Is a Bacteria?
Bacteria are small single-celled organisms. Bacteria are found almost everywhere on Earth and are vital to the planet's ecosystems. Some species can live under extreme conditions of temperature and pressure. The human body is full of bacteria, and in fact is estimated to contain more bacterial cells than human cells. Most bacteria in the body are harmless, and some are even helpful. A relatively small number of species cause disease.
Bacteria are microorganisms that come in various shapes. They can be spheres, they can be rods, or they can be spirals. There are bacteria that are bad, that we call pathogenic, and they will cause diseases, but there's also good bacteria. As an example, in our digestive system, in the gut, we have bacteria that are very necessary to help our bodies function in a normal way. What's interesting about bacteria is that in our bodies we have 10 times more bacterial cells than we have human cells. Bacteria are also important in biotechnology. They are also important in that they, again, will help the body maintain itself in a healthy manner.
Bettie J. Graham, Ph.D.
Source: National Human Genome Research Institute (NHGRI)
Common Structures
Bacteria
Image by Mariana Ruiz Villarreal, LadyofHats
Bacteria
Average prokaryote cell
Image by Mariana Ruiz Villarreal, LadyofHats
The Bacterial and Archaeal Cell: Common Structures
Introduction to the basic cell stucture
Bacteria and archaea (Figure) are unicellular organisms, which lack internal membrane-bound structures that are disconnected from the plasma membrane. Therefore, they do not have a nucleus, but instead their genetic material is located in an area of the cell called the nucleoid. The bacterial and archaeal chromosome is often a single covalently closed circular double stranded DNA molecule. However, some bacteria have linear chromosomes and some bacteria and archaea have more than one chromosome. Besides the nucleoid, the next common feature is the cytoplasm (or cytosol), which is the "aqueous" jelly-like region encompassing the internal portion of the cell. The cytoplasm is where the soluble (non-membrane) reactions occur, and which contains the ribosomes - the protein-RNA complex where proteins are made. Both bacteria and archaea have a cell membrane, the phospholipid bi-layer that defines the boundary of the cell. In prokaryotes, the cytoplasmic membrane also contains all membrane-bound reactions, including the electron transport chain, ATP synthase, and photosynthesis. Finally, bacteria and archaea have cell walls, the rigid structural feature that gives definition to the cell shape.
The features of a typical prokaryotic cell are shown.
Constraints on the bacterial and archaeal cell
One common, almost universal feature of bacteria and archaea is that they are small, microscopic to be be exact. Even the two examples given as exceptions, Epulopiscium fishelsoni and Thiomargarita namibiensis, still face the basic constraints all bacteria and archaea face; they simply found unique strategies around the problem. So what is the largest constraint when it comes to dealing with the size of bacteria and archaea? Think about what the cell must do to survive.
Basic common cellular functionsSo what do cells have to do to survive? They need to transform energy into a usable form. This involves making ATP, maintaining an energized membrane, and maintaining NAD+/NADH2 ratios. Cells also need to be able to synthesize the appropriate macromolecules (proteins, lipids, polysaccharides, etc.) and other cellular structural components. To do this, they need to be able to either make the monomers, or get them from the environment. Therefore, cells need a way to bring in nutrients from the environment, move compounds around inside the cell, and get rid of waste products that accumulate internally. So how do they do this? What is the mechanism by which these processes occur in bacteria and archaea? The answer is diffusion. And using diffusion to move compounds around leads to major constraints. Why?
Diffusion and its importance to bacteria and archaeaMovement by diffusion is passive and proceeds down the concentration gradient. For compounds to move from the outside to the inside of the cell, the compound must be able to cross the phospholipid bilayer. Then, as long as the concentration is lower inside, the compound will want to move into the cell. Nutrients that enter the cell tend to be broken down into useable pieces. As a result, the internal concentration of these nutrients is low. This is one of the mechanisms bacteria and archaea use to aid in the transport of metabolites; by keeping the internal pool of metabolites low, the desired compounds will continue to enter the cell by passive transport.
Diffusion is also used to get rid of waste materials, such as CO2. As waste products accumulate, their concentration rises compared to the outside environment and the waste product can leave the cell. Movement within the cell works the same way, compounds will move down their concentration gradient, away from where they are synthesized to places where their concentration is low and therefore may be needed. Remember that the process is random, blind to destination. The ability of two different compounds to interact, therefore, becomes a meeting of chance. In small confined spaces, random interactions or collisions can occur more frequently than in large spaces.
The ability of a compound to diffuse is dependent upon the viscosity of the solvent. For example, it is a lot easier for you to move around in air than in water (think about moving around underwater in a pool). If you put a drop of food coloring into a glass of water, it quickly diffuses until the entire glass has changed color. Now what do you think would happen if you put that same drop of food coloring into a glass of corn syrup (very viscous and sticky)? It will take a lot longer for the glass of corn syrup to change color. As you will soon read, the cytoplasm is very viscous. It contains many proteins, metabolites, small molecules, etc. and has the viscosity more like corn syrup than water. So diffusion in cells is slower and more limited than you might have expected. So, if cells rely on diffusion to move compounds around, what do you think happens as the internal volume of the cell gets bigger? Do you see the problem?
So how do cells get bigger?So for cells that rely on diffusion, like bacteria and archaea, size does matter. Then how do you suppose Epulopiscium fishelsoni and Thiomargarita namibiensis got so big? Take a look at these links and see what these bacteria look like morphologically and structurally. Epulopiscium fishelsoni and Epulopiscium fishelsoni and Thiomargarita namibiensis.
Based on what we have just discussed, in order for cells to get bigger, that is, for their volume to increase, intracellular transport must somehow become independent of diffusion. One of the great evolutionary leaps was the ability of cells (eukaryotic cells) to transport compounds and material intracellularly independent of diffusion. These mechanisms will be discussed in Modules 10.2 and 10.3. Compartmentalization also provided a way to localize processes to smaller organelles, which overcame another problem caused by the large size. Compartmentalization and the complex intracellular transport systems have allowed eukaryotic cells to become very large in comparison to the diffusion-limited bacterial and archael cells.
Source: CNX OpenStax
Additional Materials (1)
Microbiology - Bacteria (Structure)
Video by Armando Hasudungan/YouTube
9:07
Microbiology - Bacteria (Structure)
Armando Hasudungan/YouTube
Bacterial Pathogens
Mycobacterium tuberculosis
Image by National Institute of Allergy and Infectious Diseases (NIAID)
Mycobacterium tuberculosis
Produced by the National Institute of Allergy and Infectious Diseases (NIAID), this digitally colorized scanning electron microscopic (SEM) image, depicts a grouping of red-colored, rod shaped, Mycobacterium tuberculosis bacteria, which cause tuberculosis (TB) in human beings.
Image by National Institute of Allergy and Infectious Diseases (NIAID)
Bacterial Pathogens
The following tables list the species, and some higher groups, of pathogenic Eubacteria mentioned in the text. The classification of Bacteria, one of the three domains of life, is in constant flux as relationships become clearer through sampling of genetic sequences. Many groups at all taxonomic levels still have an undetermined relationship with other members of the phylogenetic tree of Bacteria. Bergey’s Manual of Systematics of Archaea and Bacteria maintains a published list and descriptions of prokaryotic species. The tables here follow the taxonomic organization in the Bergey’s Manual Taxonomic Outline.
We have divided the species into tables corresponding to different bacterial phyla. The taxonomic rank of kingdom is not used in prokaryote taxonomy, so the phyla are the subgrouping below domain. Note that many bacterial phyla not represented by these tables. The species and genera are listed only under the class within each phylum. The names given to bacteria are regulated by the International Code of Nomenclature of Bacteria as maintained by the International Committee on Systematics or Prokaryotes.
This image depicts a close view of the surface of an unidentified Petri dish culture dish growth medium, which had been inoculated with the organism Mycobacterium tuberculosis, highlighting the bacterium’s colonial morphology. Note the colorless, rough surface, which are typical morphologic characteristics seen in M. tuberculosis colonies. Macroscopic examination of colonial growth patterns is still one of the ways microorganisms are often identified.
Image by CDC/ Dr. George Kubica
Mycobacterium tuberculosis
This thin section transmission electron microscopic (TEM) image depicted the ultrastructural details displayed by a number of Gram-positive, Mycobacterium tuberculosis bacilli, the causative agent for tuberculosis.
Image by CDC/ Elizabeth "Libby" White
Mycobacterium tuberculosis
This illustration depicts a photomicrograph of a sputum specimen, processed using the Ziehl-Neelsen staining method, and revealed the presence of numerous, Mycobacterium tuberculosis bacteria. This bacterium can attack any part of the body, though usually the lungs, causing tuberculosis, and is spread through inhalation of infected sputum from a coughing, or sneezing individual.
Image by CDC
Mycobacterium fortuitum bacteria
Under a magnification of 400X, this is a photomicrograph revealed the presence of Gram-positive, Mycobacterium fortuitum bacteria, which had been harvested from an isolate, grown on a growth medium of tap water agar. The morphology of M. fortuitum resembles that of a Nocardia spp., except its branches are noticeably shorter, and there are no aerial mycelia are present. This species has been found in natural and processed water sources, as well as in sewage and dirt.
Image by CDC/ Dr. David Berd
Mycobacterium tuberculosis MEB (1)
Mycobacterium tuberculosis
Image by NIAID on Flickr./Wikimedia
Colorized scanning electron micrograph of Mycobacterium tuberculosis
Colorized scanning electron micrograph of Mycobacterium tuberculosis
Image by Clifton E. Barry III, Ph.D., NIAID, NIH
Mycobacterium tuberculosis
CDC/ Dr. George Kubica
Mycobacterium tuberculosis
CDC/ Elizabeth "Libby" White
Mycobacterium tuberculosis
CDC
Mycobacterium fortuitum bacteria
CDC/ Dr. David Berd
Mycobacterium tuberculosis MEB (1)
NIAID on Flickr./Wikimedia
Colorized scanning electron micrograph of Mycobacterium tuberculosis
Clifton E. Barry III, Ph.D., NIAID, NIH
Bleach vs. Bacteria
Chlorine gas in high concentration
Image by Larenmclane/Wikimedia
Chlorine gas in high concentration
I mixed bleach and muriatic acid and received a high concentration of chlorine gas.
Image by Larenmclane/Wikimedia
Bleach Vs. Bacteria
Spring cleaning often involves chlorine bleach, which has been used as a disinfectant for hundreds of years. But our bodies have been using bleach's active component, hypochlorous acid, to help clean house for millennia. As part of our natural response to infection, certain types of immune cells produce hypochlorous acid to help kill invading microbes, including bacteria.
Researchers funded by the National Institutes of Health have made strides in understanding exactly how bleach kills bacteria—and how bacteria's own defenses can protect against the cellular stress caused by bleach. The insights gained may lead to the development of new drugs to breach these microbial defenses, helping our bodies fight disease.
"When we started looking into how bleach actually kills bacteria, there was very little known about it," says Ursula Jakob of the University of Michigan. In a series of experiments, her team showed that hypochlorous acid causes bacterial proteins to unfold and stick to one another, making them nonfunctional and leading to cell death.
By investigating how bacteria respond to stressful conditions, the Jakob lab has uncovered several ways that bacteria in our bodies—and on our kitchen counters—can survive attack by hypochlorous acid. One such survival mechanism uses a protein called Hsp33, which is a molecular chaperone that helps other proteins fold into and maintain their normal forms. Protection by Hsp33 lets bacteria refold their proteins once a stressful situation has passed, thereby allowing the cells to survive. The Jakob lab also has discovered several bacterial proteins that sense hypochlorous acid and, in response, activate genes that help the bacteria eliminate toxins produced by exposure to the noxious chemical.
Recently, the team discovered that a simple inorganic molecule called polyphosphate also serves as a molecular chaperone within bacterial cells. Polyphosphate, which likely existed before life arose on Earth and is produced by almost all organisms, from bacteria to humans, may be one of the oldest molecular chaperones in existence. Bacteria lacking polyphosphate are very sensitive to the cellular stress caused by bleach and are less likely to cause infection.
Together, these results provide insights into how modern-day bacteria defend against immune attack and how early organisms survived environmental challenges. The studies also point to potential targets for antimicrobial drug development. "Many of these protective mechanisms that bacteria use in response to bleach are specific to bacteria," said Jakob, potentially making it possible to target these defenses without harming human cells. She and her team hope to find drugs to exploit this specificity and disarm bacterial defenses against bleach, allowing our immune systems to finish cleaning house.
By Sharon Reynolds
Source: National Institute of General Medical Sciences (NIGMS)
Additional Materials (4)
Water
The pH scale measures the concentration of hydrogen ions (H+) in a solution. (credit: modification of work by Edward Stevens)
Image by CNX Openstax
The pH Scale
Image by Tara Gross, USGS
Atoms, Isotopes, Ions, and Molecules: The Building Blocks
In the formation of an ionic compound, metals lose electrons and nonmetals gain electrons to achieve an octet.
Image by CNX Openstax
The Secret Life of Bleach (Short Version)
Video by AmericanChemistry/YouTube
Water
CNX Openstax
The pH Scale
Tara Gross, USGS
Atoms, Isotopes, Ions, and Molecules: The Building Blocks
CNX Openstax
2:16
The Secret Life of Bleach (Short Version)
AmericanChemistry/YouTube
Stop the Spread of Superbugs
Scanning Electron Micrograph of Vancomycin Resistant Enterococci
Image by CDC/ Janice Haney Carr
Scanning Electron Micrograph of Vancomycin Resistant Enterococci
This scanning electron microscopic (SEM) image depicts a cluster of vancomycin-resistant, Gram-positive, Enterococci sp. bacteria.
Image by CDC/ Janice Haney Carr
Stop the Spread of Superbugs: Help Fight Drug-Resistant Bacteria
For nearly a century, bacteria-fighting drugs known as antibiotics have helped to control and destroy many of the harmful bacteria that can make us sick. But in recent decades, antibiotics have been losing their punch against some types of bacteria. In fact, certain bacteria are now unbeatable with today’s medicines. Sadly, the way we’ve been using antibiotics is helping to create new drug-resistant “superbugs.”
Superbugs are strains of bacteria that are resistant to several types of antibiotics. Each year these drug-resistant bacteria infect more than 2 million people nationwide and kill at least 23,000, according to the U.S. Centers for Disease Control and Prevention (CDC). Drug-resistant forms of tuberculosis, gonorrhea, and staph infections are just a few of the dangers we now face.
Antibiotics are among the most commonly prescribed drugs for people. They’re also given to livestock to prevent disease and promote growth. Antibiotics are effective against bacterial infections, such as strep throat and some types of pneumonia, diarrheal diseases, and ear infections. But these drugs don’t work at all against viruses, such as those that cause colds or flu.
Unfortunately, many antibiotics prescribed to people and to animals are unnecessary. And the overuse and misuse of antibiotics helps to create drug-resistant bacteria.
Here’s how that might happen. When used properly, antibiotics can help destroy disease-causing bacteria. But if you take an antibiotic when you have a viral infection like the flu, the drug won’t affect the viruses making you sick. Instead, it’ll destroy a wide variety of bacteria in your body, including some of the “good” bacteria that help you digest food, fight infection, and stay healthy. Bacteria that are tough enough to survive the drug will have a chance to grow and quickly multiply. These drug-resistant strains may even spread to other people.
Over time, if more and more people take antibiotics when not necessary, drug-resistant bacteria can continue to thrive and spread. They may even share their drug-resistant traits with other bacteria. Drugs may become less effective or not work at all against certain disease-causing bacteria.
“Bacterial infections that were treatable for decades are no longer responding to antibiotics, even the newer ones,” says Dr. Dennis Dixon, an NIH expert in bacterial and fungal diseases. Scientists have been trying to keep ahead of newly emerging drug-resistant bacteria by developing new drugs, but it’s a tough task.
“We need to make the best use of the drugs we have, as there aren’t many in the antibiotic development pipeline,” says Dr. Jane Knisely, who oversees studies of drug-resistant bacteria at NIH. “It’s important to understand the best way to use these drugs to increase their effectiveness and decrease the chances of resistance to emerge.”
You can help slow the spread of drug-resistant bacteria by taking antibiotics properly and only when needed. Don’t insist on an antibiotic if your health care provider advises otherwise. For example, many parents expect doctors to prescribe antibiotics for a child’s ear infection. But experts recommend delaying for a time in certain situations, as many ear infections get better without antibiotics.
NIH researchers have been looking at whether antibiotics are effective for treating certain conditions in the first place. One recent study showed that antibiotics may be less effective than previously thought for treating a common type of sinus infection. This kind of research can help prevent the misuse and overuse of antibiotics.
“Treating infections with antibiotics is something we want to preserve for generations to come, so we shouldn’t misuse them,” says Dr. Julie Segre, a senior investigator at NIH.
In the past, some of the most dangerous superbugs have been confined to health care settings. That’s because people who are sick or in a weakened state are more susceptible to picking up infections. But superbug infections aren’t limited to hospitals. Some strains are out in the community and anyone, even healthy people, can become infected.
One common superbug increasingly seen outside hospitals is methicillin-resistant Staphylococcus aureus (MRSA). These bacteria don’t respond to methicillin and related antibiotics. MRSA can cause skin infections and, in more serious cases, pneumonia or bloodstream infections.
A MRSA skin infection can appear as one or more pimples or boils that are swollen, painful, or hot to the touch. The infection can spread through even a tiny cut or scrape that comes into contact with these bacteria. Many people recover from MRSA infections, but some cases can be life-threatening. The CDC estimates that more than 80,000 aggressive MRSA infections and 11,000 related deaths occur each year in the United States.
When antibiotics are needed, doctors usually prescribe a mild one before trying something more aggressive like vancomycin. Such newer antibiotics can be more toxic and more expensive than older ones. Eventually, bacteria will develop resistance to even the new drugs. In recent years, some superbugs, such as vancomycin-resistant Enterococci bacteria, remain unaffected by even this antibiotic of last resort.
“We rely on antibiotics to deliver modern health care,” Segre says. But with the rise of drug-resistant bacteria, “we’re running out of new antibiotics to treat bacterial infections,” and some of the more potent ones aren’t working as well.
Ideally, doctors would be able to quickly identify the right antibiotic to treat a particular infection. But labs need days or even weeks to test and identify the bacteria strain. Until the lab results come in, antibiotic treatment is often an educated guess.
“We need to know how to treat for a favorable outcome, but knowledge about the infection can be several days away,” explains Dr. Vance Fowler, an infectious disease expert at Duke University School of Medicine.
Fowler says faster diagnostic testing offers one of the best hopes for treating infectious diseases. Technology is catching up, he says, and new research in this area looks promising.
Genetic studies by NIH-supported researchers such as Segre and Fowler are also helping us understand the unique characteristics of antibiotic-resistant bacteria. Their findings could point the way to innovative new treatments.
While scientists search for ways to beat back these stubborn bacteria, you can help by preventing the spread of germs so we depend less on antibiotics in the first place.
The best way to prevent bacterial infections is by washing your hands frequently with soap and water. It’s also a good idea not to share personal items such as towels or razors. And use antibiotics only as directed. We can all do our part to fight drug-resistant bacteria.
Blocking Harmful Bacteria
Wash your hands often with soap and water, or use an alcohol-based hand sanitizer.
If you’re sick, make sure your doctor has a clear understanding of your symptoms. Discuss whether an antibiotic or a different type of treatment is appropriate for your illness.
If antibiotics are needed, take the full course exactly as directed. Don’t save the medicine for a future illness, and don’t share with others.
Maintain a healthy lifestyle—including proper diet, exercise, and good hygiene—to help prevent illness, thereby helping to prevent the overuse or misuse of medications.
Source: NIH News in Health
Additional Materials (10)
How to prevent antibiotic resistance
Video by Healthy Canadians/YouTube
Slowing the Spread of Superbugs
Video by Lee Health/YouTube
CDC Vital Signs: Stop the Spread of Antibiotic Resistance (Extended)
Video by Centers for Disease Control and Prevention (CDC)/YouTube
How can we solve the antibiotic resistance crisis? - Gerry Wright
Video by TED-Ed/YouTube
Antimicrobial Resistance
Figure in review article from Plos One Pathogens on agricultural spread of antimicrobial resistance. Caption: Antibiotic-resistant Staphylococcus aureus is a growing public health concern, but tracing the origins of the bacterium is complicated."
Image by Tara C. Smith
Antimicrobial Resistance
Infographic from the CDC Threat Report 2013, Antibiotic resistance threats in the United States
Image by Centers for Disease Control and Prevention
Non-resistant bacteria vs drug resistant bacteria
Diagram showing the difference between non-resistant bacteria and drug resistant bacteria. Non-resistant bacteria multiply, and upon drug treatment, the bacteria die. Drug resistant bacteria multiply as well, but upon drug treatment, the bacteria continue to spread.
Image by NIAID
Host infection stimulates antibiotic resistance
This illustration shows pathogenic bacteria behave like a Trojan horse: switching from antibiotic susceptibility to resistance during infection. Salmonella are vulnerable to antibiotics while circulating in the blood (depicted by fire on red blood cell) but are highly resistant when residing within host macrophages. This leads to treatment failure with the emergence of drug-resistant bacteria.
This image was chosen as a winner of the 2016 NIH-funded research image call, and the research was funded in part by NIGMS.
Image by Mike Mahan and Peter Allen, UC Santa Barbara
Diagram depicting antibiotic resistance through alteration of the antibiotic's target site, modeled after MRSA's resistance to penicillin. Beta-lactam antibiotics permanently inactivate PBP enzymes, which are essential for bacterial life, by permanently binding to their active sites. MRSA, however, expresses a PBP that does not allow the antibiotic into its active site.
Diagram depicting one recognized way in which bacteria may be resistant to antibiotics: if the antibiotic functions by blocking the active site of the enzyme, the bacteria may evolve to produce an enzyme that will not allow the antibiotic to bind to its active site. (This diagram is modeled specifically on the way some strains of Staphylococcus aureus have evolved to resist the beta-lactam antibiotic methicillin by expressing the mecA gene.)
1- Both enzymes are structurally similar, but differ in the kind of substances they will allow in their active sites. (The differences between the two may actually be much more subtle than what is implied by the image.)
2- Both enzymes carry out normal functions in the bacterial cell. (In this case, the enzymes are depicted making cross-links in the bacterial cell wall, a function which is crucial to bacterial cell survival and replication.)
3- The beta-lactam antibiotic fits in the active site of the antibiotic-sensitive enzyme, but not in that of the resistant enzyme.
4- Once in the active site, the beta-lactam ring of the antibiotic springs open, permanently inactivating the sensitive enzyme. The resistant enzyme, however, is totally unaffected; it is free to continue carrying out its normal function in the bacteria.
Image by Mcstrother
Antibiotic targets and mechanisms of resistance.
Antibiotic targets and mechanisms of resistance.
Image by Gerard D Wright
2:30
How to prevent antibiotic resistance
Healthy Canadians/YouTube
1:41
Slowing the Spread of Superbugs
Lee Health/YouTube
3:24
CDC Vital Signs: Stop the Spread of Antibiotic Resistance (Extended)
Centers for Disease Control and Prevention (CDC)/YouTube
6:23
How can we solve the antibiotic resistance crisis? - Gerry Wright
TED-Ed/YouTube
Antimicrobial Resistance
Tara C. Smith
Antimicrobial Resistance
Centers for Disease Control and Prevention
Non-resistant bacteria vs drug resistant bacteria
NIAID
Host infection stimulates antibiotic resistance
Mike Mahan and Peter Allen, UC Santa Barbara
Diagram depicting antibiotic resistance through alteration of the antibiotic's target site, modeled after MRSA's resistance to penicillin. Beta-lactam antibiotics permanently inactivate PBP enzymes, which are essential for bacterial life, by permanently binding to their active sites. MRSA, however, expresses a PBP that does not allow the antibiotic into its active site.
Mcstrother
Antibiotic targets and mechanisms of resistance.
Gerard D Wright
Types of Microorganisms
Gut Flora
Image by TheVisualMD
Gut Flora
The Benefits of Prebiotics & Probiotics : Probiotics are enjoying a spot on the nutritional A List these days, thanks to preliminary evidence of their health benefits. They may support a healthy immune system and protect against disease-causing microorganisms living in the intestinal tract. Recognized as a "good" bacteria, probiotics are similar to microorganisms already found in the human intestinal tract, especially in those of breastfed infants. From external sources they can be derived from supplements or more naturally from their lesser-known precursors, prebiotics - which are most readily available in high-fiber carbohydrates such as bananas, artichokes, berries, flax, garlic, legumes, and whole grains.
Image by TheVisualMD
Types of Microorganisms
Most microbes are unicellular and small enough that they require artificial magnification to be seen. However, there are some unicellular microbes that are visible to the naked eye, and some multicellular organisms that are microscopic. An object must measure about 100 micrometers (µm) to be visible without a microscope, but most microorganisms are many times smaller than that. For some perspective, consider that a typical animal cell measures roughly 10 µm across but is still microscopic. Bacterial cells are typically about 1 µm, and viruses can be 10 times smaller than bacteria for units of length used in microbiology.
Figure 1.12 The relative sizes of various microscopic and nonmicroscopic objects. Note that a typical virus measures about 100 nm, 10 times smaller than a typical bacterium (~1 µm), which is at least 10 times smaller than a typical plant or animal cell (~10–100 µm). An object must measure about 100 µm to be visible without a microscope.
Units of Length Commonly Used in Microbiology
Metric Unit
Meaning of Prefix
Metric Equivalent
meter (m)
—
1 m = 100 m
decimeter (dm)
1/10
1 dm = 0.1 m = 10−1 m
centimeter (cm)
1/100
1 cm = 0.01 m = 10−2 m
millimeter (mm)
1/1000
1 mm = 0.001 m = 10−3 m
micrometer (μm)
1/1,000,000
1 μm = 0.000001 m = 10−6 m
nanometer (nm)
1/1,000,000,000
1 nm = 0.000000001 m = 10−9 m
Table1.1
Microorganisms differ from each other not only in size, but also in structure, habitat, metabolism, and many other characteristics. While we typically think of microorganisms as being unicellular, there are also many multicellular organisms that are too small to be seen without a microscope. Some microbes, such as viruses, are even acellular (not composed of cells).
Microorganisms are found in each of the three domains of life: Archaea, Bacteria, and Eukarya. Microbes within the domains Bacteria and Archaea are all prokaryotes (their cells lack a nucleus), whereas microbes in the domain Eukarya are eukaryotes (their cells have a nucleus). Some microorganisms, such as viruses, do not fall within any of the three domains of life. In this section, we will briefly introduce each of the broad groups of microbes. Later chapters will go into greater depth about the diverse species within each group.
Source: CNX OpenStax
Additional Materials (14)
Gut Flora
Prebiotics & Probiotics : Probiotics are enjoying a spot on the nutritional A List these days, thanks to preliminary evidence of their health benefits. They may support a healthy immune system and protect against disease-causing microorganisms living in the intestinal tract. Recognized as a "good" bacteria, probiotics are similar to microorganisms already found in the human intestinal tract, especially in those of breastfed infants. From external sources they can be derived from supplements or more naturally from their lesser-known precursors, prebiotics - which are most readily available in high-fiber carbohydrates such as bananas, artichokes, berries, flax, garlic, legumes, and whole grains.
Image by TheVisualMD
Biofilm formed by the opportunistic pathogen Pseudomonas aeruginosa
A biofilm is a highly organized community of microorganisms that develops naturally on certain surfaces. These communities are common in natural environments and generally do not pose any danger to humans. Many microbes in biofilms have a positive impact on the planet and our societies. Biofilms can be helpful in treatment of wastewater, for example. This dime-sized biofilm, however, was formed by the opportunistic pathogen Pseudomonas aeruginosa. Under some conditions, this bacterium can infect wounds that are caused by severe burns. The bacterial cells release a variety of materials to form an extracellular matrix, which is stained red in this photograph. The matrix holds the biofilm together and protects the bacteria from antibiotics and the immune system. A biofilm is a highly organized community of microorganisms that develops naturally on certain surfaces. These communities are common in natural environments and generally do not pose any danger to humans. Many microbes in biofilms have a positive impact on the planet and our societies. Biofilms can be helpful in treatment of wastewater, for example. This dime-sized biofilm, however, was formed by the opportunistic pathogen Pseudomonas aeruginosa. Under some conditions, this bacterium can infect wounds that are caused by severe burns. The bacterial cells release a variety of materials to form an extracellular matrix, which is stained red in this photograph. The matrix holds the biofilm together and protects the bacteria from antibiotics and the immune system.
Image by NIGMS/Scott Chimileski, Ph.D., and Roberto Kolter, Ph.D., Harvard Medical School
Gingivitis - Actinomyces israelii - one of the types of microorganisms that cause gingivitis
Actinomyces Israelii _ Scanning electron micrograph of Actinomyces israelii. Actinomyces israelii - one of the types of microorganisms that cause gingivitis
Image by GrahamColm
Microbes
A colorized scanning electron micrograph of bacteria. Scanning electron microscopes allow scientists to see the three-dimensional surface of their samples.
Image by NIH National Institute of General Medical Sciences publication, Inside the Cell. Credit Tina Carvalho, University of Hawaii at Manoa.
Mouse colon with gut bacteria
A section of mouse colon with gut bacteria (center, in green) residing within a protective pocket. Understanding how microorganisms colonize the gut could help devise ways to correct for abnormal changes in bacterial communities that are associated with disorders like inflammatory bowel disease. More information about the research behind this image can be found in a Biomedical Beat Blog posting from September 2013.
Image by S. Melanie Lee, Caltech; Zbigniew Mikulski and Klaus Ley, La Jolla Institute for Allergy and Immunology
Herpesviridae Infections
Under a high magnification of 150,000X, these negatively-stained transmission electron microscopic (TEM) images revealed some of the ultrastructural morphology exhibited by a number of different microorganisms. Panel A represents a composite micrograph, created for the purpose of comparing the size difference between a poxvirus at the top, a bacillus in the middle, and a herpesvirus at the bottom. Panels B, C, and D, are TEMs depicting the sequential degeneration of variola virus particles.
Image by CDC/ Dr. Kenneth
Animal Diseases
Under a magnification of 475X, this photomicrograph of a lactophenol cotton blue-stained specimen, revealed some of the ultrastructural morphology exhibited by a Fusarium sp. fungal organism. Of importance here, were the septate, filamentous hyphae, and the floral arrangement of the fusiform shaped macroconidia. See PHIL 17972, for a higher magnification of this image.
Image by CDC/ Dr. Hardin
Agar Plates
A bacteria mix is spread on an agar plate. From that plate, a recombinant clone containing a gene of interest is lifted. Then large amounts of the bacteria are grown and the plasma is harvested. The DNA is then extracted and used for studying genes, including oncogenes. Also in the same setting is a male scientist holding an agar plate.
Image by Dr. Stuart Aaronson. Laboratory Of Cellular And Molecular Biology / Bill Branson (Photographer)
Lacticaseibacillus paracasei
Image by Dr. Horst Neve, Max Rubner-Institute
Meningococcal Infections
The picture shows the growth of Neisseria meningitidis on NYC media. It is used as a selective media for gonococci.
Image by Xishan01
Streptococcus pneumoniae
This scanning electron microscopic (SEM) image depicts two, round-shaped, Gram-positive, Streptococcus pneumoniae bacteria.
Image by CDC/ Dr. Richard Facklam; Photo credit: Janice Haney Carr
Enterococcus sp. bacteria
This photomicrograph reveals numerous, Gram-positive, Enterococcus sp. bacteria, which were found in a tissue sample harvested from a pneumonia patient.
Image by CDC/ Dr. Mike Miller
Cross-Section of Colon Revealing Digested Food Particles and Bacteria
This image features a cross-section of the colon, revealing digested food particles and bacteria (purple and pink) inside. Prebiotics, sometimes known as fermentable fiber, are a type of carbohydrate that ferment in the large intestine. The fermentation process stimulates the growth of various strains of colon bacteria, including probiotics - live microorganisms that support healthy intestinal function and may play a role in preventing disease.
Image by TheVisualMD
Microorganisms | Genetics | Biology | FuseSchool
Video by FuseSchool - Global Education/YouTube
Gut Flora
TheVisualMD
Biofilm formed by the opportunistic pathogen Pseudomonas aeruginosa
NIGMS/Scott Chimileski, Ph.D., and Roberto Kolter, Ph.D., Harvard Medical School
Gingivitis - Actinomyces israelii - one of the types of microorganisms that cause gingivitis
GrahamColm
Microbes
NIH National Institute of General Medical Sciences publication, Inside the Cell. Credit Tina Carvalho, University of Hawaii at Manoa.
Mouse colon with gut bacteria
S. Melanie Lee, Caltech; Zbigniew Mikulski and Klaus Ley, La Jolla Institute for Allergy and Immunology
Herpesviridae Infections
CDC/ Dr. Kenneth
Animal Diseases
CDC/ Dr. Hardin
Agar Plates
Dr. Stuart Aaronson. Laboratory Of Cellular And Molecular Biology / Bill Branson (Photographer)
Lacticaseibacillus paracasei
Dr. Horst Neve, Max Rubner-Institute
Meningococcal Infections
Xishan01
Streptococcus pneumoniae
CDC/ Dr. Richard Facklam; Photo credit: Janice Haney Carr
Enterococcus sp. bacteria
CDC/ Dr. Mike Miller
Cross-Section of Colon Revealing Digested Food Particles and Bacteria
TheVisualMD
5:05
Microorganisms | Genetics | Biology | FuseSchool
FuseSchool - Global Education/YouTube
Prokaryotic Microorganisms
E. Coli
Image by Mark Ellisman and Thomas Deerinck
National Center for Microscopy and Imaging Research
University of California, San Diego
E. Coli
Image by Mark Ellisman and Thomas Deerinck
National Center for Microscopy and Imaging Research
University of California, San Diego
Prokaryotic Microorganisms
Bacteria are found in nearly every habitat on earth, including within and on humans. Most bacteria are harmless or helpful, but some are pathogens, causing disease in humans and other animals. Bacteria are prokaryotic because their genetic material (DNA) is not housed within a true nucleus. Most bacteria have cell walls that contain peptidoglycan.
Bacteria are often described in terms of their general shape. Common shapes include spherical (coccus), rod-shaped (bacillus), or curved (spirillum, spirochete, or vibrio). Figure 1.13 shows examples of these shapes.
Figure 1.13 Common bacterial shapes. Note how coccobacillus is a combination of spherical (coccus) and rod-shaped (bacillus). (credit “Coccus”: modification of work by Janice Haney Carr, Centers for Disease Control and Prevention; credit “Coccobacillus”: modification of work by Janice Carr, Centers for Disease Control and Prevention; credit “Spirochete”: Centers for Disease Control and Prevention)
They have a wide range of metabolic capabilities and can grow in a variety of environments, using different combinations of nutrients. Some bacteria are photosynthetic, such as oxygenic cyanobacteria and anoxygenic green sulfur and green nonsulfur bacteria; these bacteria use energy derived from sunlight, and fix carbon dioxide for growth. Other types of bacteria are nonphotosynthetic, obtaining their energy from organic or inorganic compounds in their environment.
Archaea are also unicellular prokaryotic organisms. Archaea and bacteria have different evolutionary histories, as well as significant differences in genetics, metabolic pathways, and the composition of their cell walls and membranes. Unlike most bacteria, archaeal cell walls do not contain peptidoglycan, but their cell walls are often composed of a similar substance called pseudopeptidoglycan. Like bacteria, archaea are found in nearly every habitat on earth, even extreme environments that are very cold, very hot, very basic, or very acidic (Figure 1.14). Some archaea live in the human body, but none have been shown to be human pathogens.
Figure 1.14 Some archaea live in extreme environments, such as the Morning Glory pool, a hot spring in Yellowstone National Park. The color differences in the pool result from the different communities of microbes that are able to thrive at various water temperatures.
Source: CNX OpenStax
Additional Materials (5)
Structure of Prokaryotes
Archaea are separated into four phyla: the Korarchaeota, Euryarchaeota, Crenarchaeota, and Nanoarchaeota. (credit “Halobacterium”: modification of work by NASA; credit “Nanoarchaeotum equitans”: modification of work by Karl O. Stetter; credit “korarchaeota”: modification of work by Office of Science of the U.S. Dept. of Energy; scale-bar data from Matt Russell)
Image by CNX Openstax (credit “Halobacterium”: modification of work by NASA; credit “Nanoarchaeotum equitans”: modification of work by Karl O. Stetter; credit “korarchaeota”: modification of work by Office of Science of the U.S. Dept. of Energy; scale-bar data from Matt Russell)
Prokaryotic Metabolism
Prokaryotes play a key role in the nitrogen cycle. (credit: Environmental Protection Agency)
Image by CNX Openstax (credit: Environmental Protection Agency)
Prokaryotic Diversity
Five stages of biofilm development are shown. During stage 1, initial attachment, bacteria adhere to a solid surface via weak van der Waals interactions. During stage 2, irreversible attachment, hairlike appendages called pili permanently anchor the bacteria to the surface. During stage 3, maturation I, the biofilm grows through cell division and recruitment of other bacteria. An extracellular matrix composed primarily of polysaccharides holds the biofilm together. During stage 4, maturation II, the biofilm continues to grow and takes on a more complex shape. During stage 5, dispersal, the biofilm matrix is partly broken down, allowing some bacteria to escape and colonize another surface. Micrographs of a Pseudomonas aeruginosa biofilm in each of the stages of development are shown. (credit: D. Davis, Don Monroe, PLoS)
Image by CNX Openstax (credit: D. Davis, Don Monroe, PLoS)
Prokaryotic Metabolism
Prokaryotes play a significant role in continuously moving carbon through the biosphere. (credit: modification of work by John M. Evans and Howard Perlman, USGS)
Image by CNX Openstax (credit: modification of work by John M. Evans and Howard Perlman, USGS)
Prokaryotic microorganisms
Video by Natalie Hewitt/YouTube
Structure of Prokaryotes
CNX Openstax (credit “Halobacterium”: modification of work by NASA; credit “Nanoarchaeotum equitans”: modification of work by Karl O. Stetter; credit “korarchaeota”: modification of work by Office of Science of the U.S. Dept. of Energy; scale-bar data from Matt Russell)
CNX Openstax (credit: modification of work by John M. Evans and Howard Perlman, USGS)
3:16
Prokaryotic microorganisms
Natalie Hewitt/YouTube
Unique Characteristics of Prokaryotic Cells
Prokaryotic Cells
Image by CNX Openstax
Prokaryotic Cells
This figure shows relative sizes of microbes on a logarithmic scale (recall that each unit of increase in a logarithmic scale represents a 10-fold increase in the quantity being measured).
Image by CNX Openstax
Unique Characteristics of Prokaryotic Cells
Cell theory states that the cell is the fundamental unit of life. However, cells vary significantly in size, shape, structure, and function. At the simplest level of construction, all cells possess a few fundamental components. These include cytoplasm (a gel-like substance composed of water and dissolved chemicals needed for growth), which is contained within a plasma membrane (also called a cell membrane or cytoplasmic membrane); one or more chromosomes, which contain the genetic blueprints of the cell; and ribosomes, organelles used for the production of proteins.
Beyond these basic components, cells can vary greatly between organisms, and even within the same multicellular organism. The two largest categories of cells—prokaryotic cells and eukaryotic cells—are defined by major differences in several cell structures. Prokaryotic cells lack a nucleus surrounded by a complex nuclear membrane and generally have a single, circular chromosome located in a nucleoid. Eukaryotic cells have a nucleus surrounded by a complex nuclear membrane that contains multiple, rod-shaped chromosomes.
All plant cells and animal cells are eukaryotic. Some microorganisms are composed of prokaryotic cells, whereas others are composed of eukaryotic cells. Prokaryotic microorganisms are classified within the domains Archaea and Bacteria, whereas eukaryotic organisms are classified within the domain Eukarya.
The structures inside a cell are analogous to the organs inside a human body, with unique structures suited to specific functions. Some of the structures found in prokaryotic cells are similar to those found in some eukaryotic cells; others are unique to prokaryotes. Although there are some exceptions, eukaryotic cells tend to be larger than prokaryotic cells. The comparatively larger size of eukaryotic cells dictates the need to compartmentalize various chemical processes within different areas of the cell, using complex membrane-bound organelles. In contrast, prokaryotic cells generally lack membrane-bound organelles; however, they often contain inclusions that compartmentalize their cytoplasm. Figure 3.12 illustrates structures typically associated with prokaryotic cells. These structures are described in more detail in the next section.
Figure 3.12 A typical prokaryotic cell contains a cell membrane, chromosomal DNA that is concentrated in a nucleoid, ribosomes, and a cell wall. Some prokaryotic cells may also possess flagella, pili, fimbriae, and capsules.
Common Cell Morphologies and Arrangements
Individual cells of a particular prokaryotic organism are typically similar in shape, or cell morphology. Although thousands of prokaryotic organisms have been identified, only a handful of cell morphologies are commonly seen microscopically. Figure 3.13 names and illustrates cell morphologies commonly found in prokaryotic cells. In addition to cellular shape, prokaryotic cells of the same species may group together in certain distinctive arrangements depending on the plane of cell division. Some common arrangements are shown in Figure 3.14.
Figure 3.13 (credit “Coccus” micrograph: modification of work by Janice Haney Carr, Centers for Disease Control and Prevention; credit “Coccobacillus” micrograph: modification of work by Janice Carr, Centers for Disease Control and Prevention; credit “Spirochete” micrograph: modification of work by Centers for Disease Control and Prevention)
Figure 3.14
In most prokaryotic cells, morphology is maintained by the cell wall in combination with cytoskeletal elements. The cell wall is a structure found in most prokaryotes and some eukaryotes; it envelopes the cell membrane, protecting the cell from changes in osmotic pressure (Figure 3.15). Osmotic pressure occurs because of differences in the concentration of solutes on opposing sides of a semipermeable membrane. Water is able to pass through a semipermeable membrane, but solutes (dissolved molecules like salts, sugars, and other compounds) cannot. When the concentration of solutes is greater on one side of the membrane, water diffuses across the membrane from the side with the lower concentration (more water) to the side with the higher concentration (less water) until the concentrations on both sides become equal. This diffusion of water is called osmosis, and it can cause extreme osmotic pressure on a cell when its external environment changes.
The external environment of a cell can be described as an isotonic, hypertonic, or hypotonic medium. In an isotonic medium, the solute concentrations inside and outside the cell are approximately equal, so there is no net movement of water across the cell membrane. In a hypertonic medium, the solute concentration outside the cell exceeds that inside the cell, so water diffuses out of the cell and into the external medium. In a hypotonic medium, the solute concentration inside the cell exceeds that outside of the cell, so water will move by osmosis into the cell. This causes the cell to swell and potentially lyse, or burst.
The degree to which a particular cell is able to withstand changes in osmotic pressure is called tonicity. Cells that have a cell wall are better able to withstand subtle changes in osmotic pressure and maintain their shape. In hypertonic environments, cells that lack a cell wall can become dehydrated, causing crenation, or shriveling of the cell; the plasma membrane contracts and appears scalloped or notched (Figure 3.15). By contrast, cells that possess a cell wall undergo plasmolysis rather than crenation. In plasmolysis, the plasma membrane contracts and detaches from the cell wall, and there is a decrease in interior volume, but the cell wall remains intact, thus allowing the cell to maintain some shape and integrity for a period of time (Figure 3.16). Likewise, cells that lack a cell wall are more prone to lysis in hypotonic environments. The presence of a cell wall allows the cell to maintain its shape and integrity for a longer time before lysing (Figure 3.16).
Figure 3.15 In cells that lack a cell wall, changes in osmotic pressure can lead to crenation in hypertonic environments or cell lysis in hypotonic environments.
Figure 3.16 In prokaryotic cells, the cell wall provides some protection against changes in osmotic pressure, allowing it to maintain its shape longer. The cell membrane is typically attached to the cell wall in an isotonic medium (left). In a hypertonic medium, the cell membrane detaches from the cell wall and contracts (plasmolysis) as water leaves the cell. In a hypotonic medium (right), the cell wall prevents the cell membrane from expanding to the point of bursting, although lysis will eventually occur if too much water is absorbed.
CHECK YOUR UNDERSTANDING
Explain the difference between cell morphology and arrangement.
What advantages do cell walls provide prokaryotic cells?
The Nucleoid
All cellular life has a DNA genome organized into one or more chromosomes. Prokaryotic chromosomes are typically circular, haploid (unpaired), and not bound by a complex nuclear membrane. Prokaryotic DNA and DNA-associated proteins are concentrated within the nucleoid region of the cell (Figure 3.17). In general, prokaryotic DNA interacts with nucleoid-associated proteins (NAPs) that assist in the organization and packaging of the chromosome. In bacteria, NAPs function similar to histones, which are the DNA-organizing proteins found in eukaryotic cells. In archaea, the nucleoid is organized by either NAPs or histone-like DNA organizing proteins.
Figure 3.17 The nucleoid region (the area enclosed by the green dashed line) is a condensed area of DNA found within prokaryotic cells. Because of the density of the area, it does not readily stain and appears lighter in color when viewed with a transmission electron microscope.
Plasmids
Prokaryotic cells may also contain extrachromosomal DNA, or DNA that is not part of the chromosome. This extrachromosomal DNA is found in plasmids, which are small, circular, double-stranded DNA molecules. Cells that have plasmids often have hundreds of them within a single cell. Plasmids are more commonly found in bacteria; however, plasmids have been found in archaea and eukaryotic organisms. Plasmids often carry genes that confer advantageous traits such as antibiotic resistance; thus, they are important to the survival of the organism.
Ribosomes
All cellular life synthesizes proteins, and organisms in all three domains of life possess ribosomes, structures responsible for protein synthesis. However, ribosomes in each of the three domains are structurally different. Ribosomes, themselves, are constructed from proteins, along with ribosomal RNA (rRNA). Prokaryotic ribosomes are found in the cytoplasm. They are called 70S ribosomes because they have a size of 70S (Figure 3.18), whereas eukaryotic cytoplasmic ribosomes have a size of 80S. (The S stands for Svedberg unit, a measure of sedimentation in an ultracentrifuge, which is based on size, shape, and surface qualities of the structure being analyzed). Although they are the same size, bacterial and archaeal ribosomes have different proteins and rRNA molecules, and the archaeal versions are more similar to their eukaryotic counterparts than to those found in bacteria.
Figure 3.18 Prokaryotic ribosomes (70S) are composed of two subunits: the 30S (small subunit) and the 50S (large subunit), each of which are composed of protein and rRNA components.
Inclusions
As single-celled organisms living in unstable environments, some prokaryotic cells have the ability to store excess nutrients within cytoplasmic structures called inclusions. Storing nutrients in a polymerized form is advantageous because it reduces the buildup of osmotic pressure that occurs as a cell accumulates solutes. Various types of inclusions store glycogen and starches, which contain carbon that cells can access for energy. Volutin granules, also called metachromatic granules because of their staining characteristics, are inclusions that store polymerized inorganic phosphate that can be used in metabolism and assist in the formation of biofilms. Microbes known to contain volutin granules include the archaea Methanosarcina, the bacterium Corynebacterium diphtheriae, and the unicellular eukaryotic alga Chlamydomonas. Sulfur granules, another type of inclusion, are found in sulfur bacteria of the genus Thiobacillus; these granules store elemental sulfur, which the bacteria use for metabolism.
Occasionally, certain types of inclusions are surrounded by a phospholipid monolayer embedded with protein. Polyhydroxybutyrate (PHB), which can be produced by species of Bacillus and Pseudomonas, is an example of an inclusion that displays this type of monolayer structure. Industrially, PHB has also been used as a source of biodegradable polymers for bioplastics. Several different types of inclusions are shown in Figure 3.19.
Figure 3.19 Prokaryotic cells may have various types of inclusions. (a) A transmission electron micrograph of polyhydroxybutryrate lipid droplets. (b) A light micrograph of volutin granules. (c) A phase-contrast micrograph of sulfur granules. (d) A transmission electron micrograph of gas vacuoles. (e) A transmission electron micrograph of magnetosomes. (credit b, c, d: modification of work by American Society for Microbiology)
Some prokaryotic cells have other types of inclusions that serve purposes other than nutrient storage. For example, some prokaryotic cells produce gas vacuoles, accumulations of small, protein-lined vesicles of gas. These gas vacuoles allow the prokaryotic cells that synthesize them to alter their buoyancy so that they can adjust their location in the water column. Magnetotactic bacteria, such as Magnetospirillum magnetotacticum, contain magnetosomes, which are inclusions of magnetic iron oxide or iron sulfide surrounded by a lipid layer. These allow cells to align along a magnetic field, aiding their movement (Figure 3.19). Cyanobacteria such as Anabaena cylindrica and bacteria such as Halothiobacillus neapolitanus produce carboxysome inclusions. Carboxysomes are composed of outer shells of thousands of protein subunits. Their interior is filled with ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) and carbonic anhydrase. Both of these compounds are used for carbon metabolism. Some prokaryotic cells also possess carboxysomes that sequester functionally related enzymes in one location. These structures are considered proto-organelles because they compartmentalize important compounds or chemical reactions, much like many eukaryotic organelles.
Endospores
Bacterial cells are generally observed as vegetative cells, but some genera of bacteria have the ability to form endospores, structures that essentially protect the bacterial genome in a dormant state when environmental conditions are unfavorable. Endospores (not to be confused with the reproductive spores formed by fungi) allow some bacterial cells to survive long periods without food or water, as well as exposure to chemicals, extreme temperatures, and even radiation. Table 3.1 compares the characteristics of vegetative cells and endospores.
Characteristics of Vegetative Cells versus Endospores
Vegetative Cells
Endospores
Sensitive to extreme temperatures and radiation
Resistant to extreme temperatures and radiation
Gram-positive
Do not absorb Gram stain, only special endospore stains
Normal water content and enzymatic activity
Dehydrated; no metabolic activity
Capable of active growth and metabolism
Dormant; no growth or metabolic activity
Table3.1
The process by which vegetative cells transform into endospores is called sporulation, and it generally begins when nutrients become depleted or environmental conditions become otherwise unfavorable (Figure 3.20). The process begins with the formation of a septum in the vegetative bacterial cell. The septum divides the cell asymmetrically, separating a DNA forespore from the mother cell. The forespore, which will form the core of the endospore, is essentially a copy of the cell’s chromosomes, and is separated from the mother cell by a second membrane. A cortex gradually forms around the forespore by laying down layers of calcium and dipicolinic acid between membranes. A protein spore coat then forms around the cortex while the DNA of the mother cell disintegrates. Further maturation of the endospore occurs with the formation of an outermost exosporium. The endospore is released upon disintegration of the mother cell, completing sporulation.
Figure 3.20 (a) Sporulation begins following asymmetric cell division. The forespore becomes surrounded by a double layer of membrane, a cortex, and a protein spore coat, before being released as a mature endospore upon disintegration of the mother cell. (b) An electron micrograph of a Carboxydothermus hydrogenoformans endospore. (c) These Bacillus spp. cells are undergoing sporulation. The endospores have been visualized using Malachite Green spore stain. (credit b: modification of work by Jonathan Eisen)
Endospores of certain species have been shown to persist in a dormant state for extended periods of time, up to thousands of years. However, when living conditions improve, endospores undergo germination, reentering a vegetative state. After germination, the cell becomes metabolically active again and is able to carry out all of its normal functions, including growth and cell division.
Not all bacteria have the ability to form endospores; however, there are a number of clinically significant endospore-forming gram-positive bacteria of the genera Bacillus and Clostridium. These include B. anthracis, the causative agent of anthrax, which produces endospores capable of survive for many decades; C. tetani (causes tetanus); C. difficile (causes pseudomembranous colitis); C. perfringens (causes gas gangrene); and C. botulinum (causes botulism). Pathogens such as these are particularly difficult to combat because their endospores are so hard to kill. Special sterilization methods for endospore-forming bacteria are discussed in Control of Microbial Growth.
CHECK YOUR UNDERSTANDING
What is an inclusion?
What is the function of an endospore?
Plasma Membrane
Structures that enclose the cytoplasm and internal structures of the cell are known collectively as the cell envelope. In prokaryotic cells, the structures of the cell envelope vary depending on the type of cell and organism. Most (but not all) prokaryotic cells have a cell wall, but the makeup of this cell wall varies. All cells (prokaryotic and eukaryotic) have a plasma membrane (also called cytoplasmic membrane or cell membrane) that exhibits selective permeability, allowing some molecules to enter or leave the cell while restricting the passage of others.
The structure of the plasma membrane is often described in terms of the fluid mosaic model, which refers to the ability of membrane components to move fluidly within the plane of the membrane, as well as the mosaic-like composition of the components, which include a diverse array of lipid and protein components (Figure 3.21). The plasma membrane structure of most bacterial and eukaryotic cell types is a bilayer composed mainly of phospholipids formed with ester linkages and proteins. These phospholipids and proteins have the ability to move laterally within the plane of the membranes as well as between the two phospholipid layers.
Figure 3.21 The bacterial plasma membrane is a phospholipid bilayer with a variety of embedded proteins that perform various functions for the cell. Note the presence of glycoproteins and glycolipids, whose carbohydrate components extend out from the surface of the cell. The abundance and arrangement of these proteins and lipids can vary greatly between species.
Archaeal membranes are fundamentally different from bacterial and eukaryotic membranes in a few significant ways. First, archaeal membrane phospholipids are formed with ether linkages, in contrast to the ester linkages found in bacterial or eukaryotic cell membranes. Second, archaeal phospholipids have branched chains, whereas those of bacterial and eukaryotic cells are straight chained. Finally, although some archaeal membranes can be formed of bilayers like those found in bacteria and eukaryotes, other archaeal plasma membranes are lipid monolayers.
Proteins on the cell’s surface are important for a variety of functions, including cell-to-cell communication, and sensing environmental conditions and pathogenic virulence factors. Membrane proteins and phospholipids may have carbohydrates (sugars) associated with them and are called glycoproteins or glycolipids, respectively. These glycoprotein and glycolipid complexes extend out from the surface of the cell, allowing the cell to interact with the external environment (Figure 3.21). Glycoproteins and glycolipids in the plasma membrane can vary considerably in chemical composition among archaea, bacteria, and eukaryotes, allowing scientists to use them to characterize unique species.
Plasma membranes from different cells types also contain unique phospholipids, which contain fatty acids. As described in Using Biochemistry to Identify Microorganisms, phospholipid-derived fatty acid analysis (PLFA) profiles can be used to identify unique types of cells based on differences in fatty acids. Archaea, bacteria, and eukaryotes each have a unique PFLA profile.
Membrane Transport Mechanisms
One of the most important functions of the plasma membrane is to control the transport of molecules into and out of the cell. Internal conditions must be maintained within a certain range despite any changes in the external environment. The transport of substances across the plasma membrane allows cells to do so.
Cells use various modes of transport across the plasma membrane. For example, molecules moving from a higher concentration to a lower concentration with the concentration gradient are transported by simple diffusion, also known as passive transport (Figure 3.22). Some small molecules, like carbon dioxide, may cross the membrane bilayer directly by simple diffusion. However, charged molecules, as well as large molecules, need the help of carriers or channels in the membrane. These structures ferry molecules across the membrane, a process known as facilitated diffusion (Figure 3.23).
Active transport occurs when cells move molecules across their membrane against concentration gradients (Figure 3.24). A major difference between passive and active transport is that active transport requires adenosine triphosphate (ATP) or other forms of energy to move molecules “uphill.” Therefore, active transport structures are often called “pumps.”
Figure 3.22 Simple diffusion down a concentration gradient directly across the phospholipid bilayer. (credit: modification of work by Mariana Ruiz Villareal)
Figure 3.23 Facilitated diffusion down a concentration gradient through a membrane protein. (credit: modification of work by Mariana Ruiz Villareal)
Figure 3.24 Active transport against a concentration gradient via a membrane pump that requires energy. (credit: modification of work by Mariana Ruiz Villareal)
Group translocation also transports substances into bacterial cells. In this case, as a molecule moves into a cell against its concentration gradient, it is chemically modified so that it does not require transport against an unfavorable concentration gradient. A common example of this is the bacterial phosphotransferase system, a series of carriers that phosphorylates (i.e., adds phosphate ions to) glucose or other sugars upon entry into cells. Since the phosphorylation of sugars is required during the early stages of sugar metabolism, the phosphotransferase system is considered to be an energy neutral system.
Photosynthetic Membrane Structures
Some prokaryotic cells, namely cyanobacteria and photosynthetic bacteria, have membrane structures that enable them to perform photosynthesis. These structures consist of an infolding of the plasma membrane that encloses photosynthetic pigments such as green chlorophylls and bacteriochlorophylls. In cyanobacteria, these membrane structures are called thylakoids; in photosynthetic bacteria, they are called chromatophores, lamellae, or chlorosomes.
Cell Wall
The primary function of the cell wall is to protect the cell from harsh conditions in the outside environment. When present, there are notable similarities and differences among the cell walls of archaea, bacteria, and eukaryotes.
The major component of bacterial cell walls is called peptidoglycan (or murein); it is only found in bacteria. Structurally, peptidoglycan resembles a layer of meshwork or fabric (Figure 3.25). Each layer is composed of long chains of alternating molecules of N-acetylglucosamine (NAG) and N-acetylmuramic acid (NAM). The structure of the long chains has significant two-dimensional tensile strength due to the formation of peptide bridges that connect NAG and NAM within each peptidoglycan layer. In gram-negative bacteria, tetrapeptide chains extending from each NAM unit are directly cross-linked, whereas in gram-positive bacteria, these tetrapeptide chains are linked by pentaglycine cross-bridges. Peptidoglycan subunits are made inside of the bacterial cell and then exported and assembled in layers, giving the cell its shape.
Since peptidoglycan is unique to bacteria, many antibiotic drugs are designed to interfere with peptidoglycan synthesis, weakening the cell wall and making bacterial cells more susceptible to the effects of osmotic pressure. In addition, certain cells of the human immune system are able “recognize” bacterial pathogens by detecting peptidoglycan on the surface of a bacterial cell; these cells then engulf and destroy the bacterial cell, using enzymes such as lysozyme, which breaks down and digests the peptidoglycan in their cell walls.
Figure 3.25 Peptidoglycan is composed of polymers of alternating NAM and NAG subunits, which are cross-linked by peptide bridges linking NAM subunits from various glycan chains. This provides the cell wall with tensile strength in two dimensions.
The Gram staining protocol is used to differentiate two common types of cell wall structures (Figure 3.26). Gram-positive cells have a cell wall consisting of many layers of peptidoglycan totaling 30–100 nm in thickness. These peptidoglycan layers are commonly embedded with teichoic acids (TAs), carbohydrate chains that extend through and beyond the peptidoglycan layer. TA is thought to stabilize peptidoglycan by increasing its rigidity. TA also plays a role in the ability of pathogenic gram-positive bacteria such as Streptococcus to bind to certain proteins on the surface of host cells, enhancing their ability to cause infection. In addition to peptidoglycan and TAs, bacteria of the family Mycobacteriaceae have an external layer of waxy mycolic acids in their cell wall; as described in Staining Microscopic Specimens, these bacteria are referred to as acid-fast, since acid-fast stains must be used to penetrate the mycolic acid layer for purposes of microscopy (Figure 3.27).
Figure 3.26 Bacteria contain two common cell wall structural types. Gram-positive cell walls are structurally simple, containing a thick layer of peptidoglycan with embedded teichoic acid external to the plasma membrane.20 Gram-negative cell walls are structurally more complex, containing three layers: the inner membrane, a thin layer of peptidoglycan, and an outer membrane containing lipopolysaccharide. (credit: modification of work by “Franciscosp2”/Wikimedia Commons)
Figure 3.27 (a) Some gram-positive bacteria, including members of the Mycobacteriaceae, produce waxy mycolic acids found exterior to their structurally-distinct peptidoglycan. (b) The acid-fast staining protocol detects the presence of cell walls that are rich in mycolic acid. Acid-fast cells are stained red by carbolfuschin. (credit a: modification of work by “Franciscosp2”/Wikimedia Commons; credit b: modification of work by Centers for Disease Control and Prevention)
Gram-negative cells have a much thinner layer of peptidoglycan (no more than about 4 nm thick) than gram-positive cells, and the overall structure of their cell envelope is more complex. In gram-negative cells, a gel-like matrix occupies the periplasmic space between the cell wall and the plasma membrane, and there is a second lipid bilayer called the outer membrane, which is external to the peptidoglycan layer (Figure 3.26). This outer membrane is attached to the peptidoglycan by murein lipoprotein. The outer leaflet of the outer membrane contains the molecule lipopolysaccharide (LPS), which functions as an endotoxin in infections involving gram-negative bacteria, contributing to symptoms such as fever, hemorrhaging, and septic shock. Each LPS molecule is composed of Lipid A, a core polysaccharide, and an O side chain that is composed of sugar-like molecules that comprise the external face of the LPS (Figure 3.28). The composition of the O side chain varies between different species and strains of bacteria. Parts of the O side chain called antigens can be detected using serological or immunological tests to identify specific pathogenic strains like Escherichia coli O157:H7, a deadly strain of bacteria that causes bloody diarrhea and kidney failure.
Figure 3.28 The outer membrane of a gram-negative bacterial cell contains lipopolysaccharide (LPS), a toxin composed of Lipid A embedded in the outer membrane, a core polysaccharide, and the O side chain.
Archaeal cell wall structure differs from that of bacteria in several significant ways. First, archaeal cell walls do not contain peptidoglycan; instead, they contain a similar polymer called pseudopeptidoglycan (pseudomurein) in which NAM is replaced with a different subunit. Other archaea may have a layer of glycoproteins or polysaccharides that serves as the cell wall instead of pseudopeptidoglycan. Last, as is the case with some bacterial species, there are a few archaea that appear to lack cell walls entirely.
Glycocalyces and S-Layers
Although most prokaryotic cells have cell walls, some may have additional cell envelope structures exterior to the cell wall, such as glycocalyces and S-layers. A glycocalyx is a sugar coat, of which there are two important types: capsules and slime layers. A capsule is an organized layer located outside of the cell wall and usually composed of polysaccharides or proteins (Figure 3.29). A slime layer is a less tightly organized layer that is only loosely attached to the cell wall and can be more easily washed off. Slime layers may be composed of polysaccharides, glycoproteins, or glycolipids.
Glycocalyces allows cells to adhere to surfaces, aiding in the formation of biofilms (colonies of microbes that form in layers on surfaces). In nature, most microbes live in mixed communities within biofilms, partly because the biofilm affords them some level of protection. Biofilms generally hold water like a sponge, preventing desiccation. They also protect cells from predation and hinder the action of antibiotics and disinfectants. All of these properties are advantageous to the microbes living in a biofilm, but they present challenges in a clinical setting, where the goal is often to eliminate microbes.
Figure 3.29 (a) Capsules are a type of glycocalyx composed of an organized layer of polysaccharides. (b) A capsule stain of Pseudomonas aeruginosa, a bacterial pathogen capable of causing many different types of infections in humans. (credit b: modification of work by American Society for Microbiology)
The ability to produce a capsule can contribute to a microbe’s pathogenicity (ability to cause disease) because the capsule can make it more difficult for phagocytic cells (such as white blood cells) to engulf and kill the microorganism. Streptococcus pneumoniae, for example, produces a capsule that is well known to aid in this bacterium’s pathogenicity. As explained in Staining Microscopic specimens, capsules are difficult to stain for microscopy; negative staining techniques are typically used.
An S-layer is another type of cell envelope structure; it is composed of a mixture of structural proteins and glycoproteins. In bacteria, S-layers are found outside the cell wall, but in some archaea, the S-layer serves as the cell wall. The exact function of S-layers is not entirely understood, and they are difficult to study; but available evidence suggests that they may play a variety of functions in different prokaryotic cells, such as helping the cell withstand osmotic pressure and, for certain pathogens, interacting with the host immune system.
CLINICAL FOCUS
Part 3
After diagnosing Barbara with pneumonia, the PA writes her a prescription for amoxicillin, a commonly-prescribed type of penicillin derivative. More than a week later, despite taking the full course as directed, Barbara still feels weak and is not fully recovered, although she is still able to get through her daily activities. She returns to the health center for a follow-up visit.
Many types of bacteria, fungi, and viruses can cause pneumonia. Amoxicillin targets the peptidoglycan of bacterial cell walls. Since the amoxicillin has not resolved Barbara’s symptoms, the PA concludes that the causative agent probably lacks peptidoglycan, meaning that the pathogen could be a virus, a fungus, or a bacterium that lacks peptidoglycan. Another possibility is that the pathogen is a bacterium containing peptidoglycan but has developed resistance to amoxicillin.
How can the PA definitively identify the cause of Barbara’s pneumonia?
What form of treatment should the PA prescribe, given that the amoxicillin was ineffective?
Filamentous Appendages
Many bacterial cells have protein appendages embedded within their cell envelopes that extend outward, allowing interaction with the environment. These appendages can attach to other surfaces, transfer DNA, or provide movement. Filamentous appendages include fimbriae, pili, and flagella.
Fimbriae and Pili
Fimbriae and pili are structurally similar and, because differentiation between the two is problematic, these terms are often used interchangeably. The term fimbriae commonly refers to short bristle-like proteins projecting from the cell surface by the hundreds. Fimbriae enable a cell to attach to surfaces and to other cells. For pathogenic bacteria, adherence to host cells is important for colonization, infectivity, and virulence. Adherence to surfaces is also important in biofilm formation.
The term pili (singular: pilus) commonly refers to longer, less numerous protein appendages that aid in attachment to surfaces (Figure 3.30). A specific type of pilus, called the F pilus or sex pilus, is important in the transfer of DNA between bacterial cells, which occurs between members of the same generation when two cells physically transfer or exchange parts of their respective genomes.
Figure 3.30 Bacteria may produce two different types of protein appendages that aid in surface attachment. Fimbriae typically are more numerous and shorter, whereas pili (shown here) are longer and less numerous per cell. (credit: modification of work by American Society for Microbiology)
MICRO CONNECTIONS
Group A Strep
Before the structure and function of the various components of the bacterial cell envelope were well understood, scientists were already using cell envelope characteristics to classify bacteria. In 1933, Rebecca Lancefield proposed a method for serotyping various β-hemolytic strains of Streptococcus species using an agglutination assay, a technique using the clumping of bacteria to detect specific cell-surface antigens. In doing so, Lancefield discovered that one group of S. pyogenes, found in Group A, was associated with a variety of human diseases. She determined that various strains of Group A strep could be distinguished from each other based on variations in specific cell surface proteins that she named M proteins.
Today, more than 80 different strains of Group A strep have been identified based on M proteins. Various strains of Group A strep are associated with a wide variety of human infections, including streptococcal pharyngitis (strep throat), impetigo, toxic shock syndrome, scarlet fever, rheumatic fever, and necrotizing fasciitis. The M protein is an important virulence factor for Group A strep, helping these strains evade the immune system. Changes in M proteins appear to alter the infectivity of a particular strain of Group A strep.
Flagella
Flagella are structures used by cells to move in aqueous environments. Bacterial flagella act like propellers. They are stiff spiral filaments composed of flagellin protein subunits that extend outward from the cell and spin in solution. The basal body is the motor for the flagellum and is embedded in the plasma membrane (Figure 3.31). A hook region connects the basal body to the filament. Gram-positive and gram-negative bacteria have different basal body configurations due to differences in cell wall structure.
Different types of motile bacteria exhibit different arrangements of flagella (Figure 3.32). A bacterium with a singular flagellum, typically located at one end of the cell (polar), is said to have a monotrichous flagellum. An example of a monotrichously flagellated bacterial pathogen is Vibrio cholerae, the gram-negative bacterium that causes cholera. Cells with amphitrichous flagella have a flagellum or tufts of flagella at each end. An example is Spirillum minor, the cause of spirillary (Asian) rat-bite fever or sodoku. Cells with lophotrichous flagella have a tuft at one end of the cell. The gram-negative bacillus Pseudomonas aeruginosa, an opportunistic pathogen known for causing many infections, including “swimmer’s ear” and burn wound infections, has lophotrichous flagella. Flagella that cover the entire surface of a bacterial cell are called peritrichous flagella. The gram-negative bacterium E. coli shows a peritrichous arrangement of flagella.
Figure 3.31 The basic structure of a bacterial flagellum consists of a basal body, hook, and filament. The basal body composition and arrangement differ between gram-positive and gram-negative bacteria. (credit: modification of work by “LadyofHats”/Mariana Ruiz Villareal)
Figure 3.32 Flagellated bacteria may exhibit multiple arrangements of their flagella. Common arrangements include monotrichous, amphitrichous, lophotrichous, or peritrichous.
Directional movement depends on the configuration of the flagella. Bacteria can move in response to a variety of environmental signals, including light (phototaxis), magnetic fields (magnetotaxis) using magnetosomes, and, most commonly, chemical gradients (chemotaxis). Purposeful movement toward a chemical attractant, like a food source, or away from a repellent, like a poisonous chemical, is achieved by increasing the length of runs and decreasing the length of tumbles. When running, flagella rotate in a counterclockwise direction, allowing the bacterial cell to move forward. In a peritrichous bacterium, the flagella are all bundled together in a very streamlined way (Figure 3.33), allowing for efficient movement. When tumbling, flagella are splayed out while rotating in a clockwise direction, creating a looping motion and preventing meaningful forward movement but reorienting the cell toward the direction of the attractant. When an attractant exists, runs and tumbles still occur; however, the length of runs is longer, while the length of the tumbles is reduced, allowing overall movement toward the higher concentration of the attractant. When no chemical gradient exists, the lengths of runs and tumbles are more equal, and overall movement is more random (Figure 3.34).
Figure 3.33 Bacteria achieve directional movement by changing the rotation of their flagella. In a cell with peritrichous flagella, the flagella bundle when they rotate in a counterclockwise direction, resulting in a run. However, when the flagella rotate in a clockwise direction, the flagella are no longer bundled, resulting in tumbles.
Figure 3.34 Without a chemical gradient, flagellar rotation cycles between counterclockwise (run) and clockwise (tumble) with no overall directional movement. However, when a chemical gradient of an attractant exists, the length of runs is extended, while the length of tumbles is decreased. This leads to chemotaxis: an overall directional movement toward the higher concentration of the attractant.
Source: CNX OpenStax
Additional Materials (2)
Prokaryotic Cells
This figure shows the generalized structure of a prokaryotic cell. All prokaryotes have chromosomal DNA localized in a nucleoid, ribosomes, a cell membrane, and a cell wall. The other structures shown are present in some, but not all, bacteria.
Image by CNX Openstax
Bacteria
Video by Khan Academy/YouTube
Prokaryotic Cells
CNX Openstax
18:26
Bacteria
Khan Academy/YouTube
Proteobacteria
Dvulgaris micrograph
Image by Graham Bradley assumed/Wikimedia
Dvulgaris micrograph
transmission electron micrograph of bacterium;
Desulfovibrio vulgaris
bar = 0.5 microns
I am creator
Image by Graham Bradley assumed/Wikimedia
Proteobacteria
In 1987, the American microbiologist Carl Woese (1928–2012) suggested that a large and diverse group of bacteria that he called “purple bacteria and their relatives” should be defined as a separate phylum within the domain Bacteria based on the similarity of the nucleotide sequences in their genome. This phylum of gram-negative bacteria subsequently received the name Proteobacteria. It includes many bacteria that are part of the normal human microbiota as well as many pathogens. The Proteobacteria are further divided into five classes: Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria, Deltaproteobacteria, and Epsilonproteobacteria.
Alphaproteobacteria
The first class of Proteobacteria is the Alphaproteobacteria. The unifying characteristic of this class is that they are oligotrophs, organisms capable of living in low-nutrient environments such as deep oceanic sediments, glacial ice, or deep undersurface soil.
Among the Alphaproteobacteria are rickettsias, obligate intracellular pathogens, that require part of their life cycle to occur inside other cells called host cells. When not growing inside a host cell, Rickettsia are metabolically inactive outside the host cell. They cannot synthesize their own adenosine triphosphate (ATP), and, therefore, rely on cells for their energy needs.
Rickettsia spp. include a number of serious human pathogens. For example, R. rickettsii causes Rocky Mountain spotted fever, a life-threatening form of meningoencephalitis (inflammation of the membranes that wrap the brain). R. rickettsii infects ticks and can be transmitted to humans via a bite from an infected tick (Figure 4.4).
Figure 4.4 Rickettsias require special staining methods to see them under a microscope. Here, R. rickettsii, which causes Rocky Mountain spotted fever, is shown infecting the cells of a tick. (credit: modification of work by Centers for Disease Control and Prevention)
Another species of Rickettsia, R.prowazekii, is spread by lice. It causes epidemic typhus, a severe infectious disease common during warfare and mass migrations of people. R. prowazekii infects human endothelium cells, causing inflammation of the inner lining of blood vessels, high fever, abdominal pain, and sometimes delirium. A relative, R. typhi, causes a less severe disease known as murine or endemic typhus, which is still observed in the southwestern United States during warm seasons.
C. trachomatis is a human pathogen that causes trachoma, a disease of the eyes, often leading to blindness. C. trachomatis also causes the sexually transmitted disease lymphogranuloma venereum (LGV). This disease is often mildly symptomatic, manifesting as regional lymph node swelling, or it may be asymptomatic, but it is extremely contagious and is common on college campuses.
Table 4.2 summarizes the characteristics of important genera of Alphaproteobacteria.
Class Alphaproteobacteria
Genus
Microscopic Morphology
Unique Characteristics
Agrobacterium
Gram-negative bacillus
Plant pathogen; one species, A. tumefaciens, causes tumors in plants
Facultative intracellular bacteria, transmitted by lice and fleas, cause trench fever and cat scratch disease in humans
Brucella
Gram-negative, small, flagellated coccobacillus
Facultative intracellular bacteria, transmitted by contaminated milk from infected cows, cause brucellosis in cattle and humans
Caulobacter
Gram-negative bacillus
Used in studies on cellular adaptation and differentiation because of its peculiar life cycle (during cell division, forms “swarm” cells and “stalked” cells)
Coxiella
Small, gram-negative bacillus
Obligatory intracellular bacteria; cause Q fever; potential for use as biological weapon
Ehrlichia
Very small, gram-negative, coccoid or ovoid bacteria
Obligatory intracellular bacteria; can be transported from cell to cell; transmitted by ticks; cause ehrlichiosis (destruction of white blood cells and inflammation) in humans and dogs
Hyphomicrobium
Gram-negative bacilli; grows from a stalk
Similar to Caulobacter
Methylocystis
Gram-negative, coccoid or short bacilli
Nitrogen-fixing aerobic bacteria
Rhizobium
Gram-negative, rectangular bacilli with rounded ends forming clusters
Nitrogen-fixing bacteria that live in soil and form symbiotic relationship with roots of legumes (e.g., clover, alfalfa, and beans)
Rickettsia
Gram-negative, highly pleomorphic bacteria (may be cocci, rods, or threads)
Obligate intracellular bacteria; transmitted by ticks; may cause Rocky Mountain spotted fever and typhus
Table4.2
Betaproteobacteria
Unlike Alphaproteobacteria, which survive on a minimal amount of nutrients, the class Betaproteobacteria are eutrophs (or copiotrophs), meaning that they require a copious amount of organic nutrients. Betaproteobacteria often grow between aerobic and anaerobic areas (e.g., in mammalian intestines). Some genera include species that are human pathogens, able to cause severe, sometimes life-threatening disease. The genus Neisseria, for example, includes the bacteria N. gonorrhoeae, the causative agent of the STI gonorrhea, and N. meningitides, the causative agent of bacterial meningitis.
Neisseria are cocci that live on mucosal surfaces of the human body. They are fastidious, or difficult to culture, and they require high levels of moisture, nutrient supplements, and carbon dioxide. Also, Neisseria are microaerophilic, meaning that they require low levels of oxygen. For optimal growth and for the purposes of identification, Neisseria spp. are grown on chocolate agar (i.e., agar supplemented by partially hemolyzed red blood cells). Their characteristic pattern of growth in culture is diplococcal: pairs of cells resembling coffee beans (Figure 4.5).
Figure 4.5 Neisseria meningitidis growing in colonies on a chocolate agar plate. (credit: Centers for Disease Control and Prevention)
The pathogen responsible for pertussis (whooping cough) is also a member of Betaproteobacteria. The bacterium Bordetella pertussis, from the order Burkholderiales, produces several toxins that paralyze the movement of cilia in the human respiratory tract and directly damage cells of the respiratory tract, causing a severe cough.
Table 4.3 summarizes the characteristics of important genera of Betaproteobacteria.
Class Betaproteobacteria
Example Genus
Microscopic Morphology
Unique Characteristics
Bordetella
A small, gram-negative coccobacillus
Aerobic, very fastidious; B. pertussis causes pertussis (whooping cough)
Burkholderia
Gram-negative bacillus
Aerobic, aquatic, cause diseases in horses and humans (especially patients with cystic fibrosis); agents of nosocomial infections
Leptothrix
Gram-negative, sheathed, filamentous bacillus
Aquatic; oxidize iron and manganese; can live in wastewater treatment plants and clog pipes
Require moisture and high concentration of carbon dioxide; oxidase positive, grow on chocolate agar; pathogenic species cause gonorrhea and meningitis
Thiobacillus
Gram-negative bacillus
Thermophilic, acidophilic, strictly aerobic bacteria; oxidize iron and sulfur
Table4.3
CLINICAL FOCUS
Part 2
When Marsha finally went to the doctor’s office, the physician listened to her breathing through a stethoscope. He heard some crepitation (a crackling sound) in her lungs, so he ordered a chest radiograph and asked the nurse to collect a sputum sample for microbiological evaluation and cytology. The radiologic evaluation found cavities, opacities, and a particular pattern of distribution of abnormal material (Figure 4.6).
What are some possible diseases that could be responsible for Marsha’s radiograph results?
Figure 4.6 This anteroposterior radiograph shows the presence of bilateral pulmonary infiltrate (white triangles) and “caving formation” (black arrows) present in the right apical region. (credit: Centers for Disease Control and Prevention)
Gammaproteobacteria
The most diverse class of gram-negative bacteria is Gammaproteobacteria, and it includes a number of human pathogens. For example, a large and diverse family, Pseudomonaceae, includes the genus Pseudomonas. Within this genus is the species P. aeruginosa, a pathogen responsible for diverse infections in various regions of the body. P. aeruginosa is a strictly aerobic, nonfermenting, highly motile bacterium. It often infects wounds and burns, can be the cause of chronic urinary tract infections, and can be an important cause of respiratory infections in patients with cystic fibrosis or patients on mechanical ventilators. Infections by P. aeruginosa are often difficult to treat because the bacterium is resistant to many antibiotics and has a remarkable ability to form biofilms. Other representatives of Pseudomonas include the fluorescent (glowing) bacterium P. fluorescens and the soil bacteria P. putida, which is known for its ability to degrade xenobiotics (substances not naturally produced or found in living organisms).
The Pasteurellaceae also includes several clinically relevant genera and species. This family includes several bacteria that are human and/or animal pathogens. For example, Pasteurella haemolytica causes severe pneumonia in sheep and goats. P. multocida is a species that can be transmitted from animals to humans through bites, causing infections of the skin and deeper tissues. The genus Haemophilus contains two human pathogens, H. influenzae and H. ducreyi. Despite its name, H. influenzae does not cause influenza (which is a viral disease). H. influenzae can cause both upper and lower respiratory tract infections, including sinusitis, bronchitis, ear infections, and pneumonia. Before the development of effective vaccination, strains of H. influenzae were a leading cause of more invasive diseases, like meningitis in children. H. ducreyi causes the STI known as chancroid.
The order Vibrionales includes the human pathogen Vibrio cholerae. This comma-shaped aquatic bacterium thrives in highly alkaline environments like shallow lagoons and sea ports. A toxin produced by V. cholerae causes hypersecretion of electrolytes and water in the large intestine, leading to profuse watery diarrhea and dehydration. V. parahaemolyticus is also a cause of gastrointestinal disease in humans, whereas V. vulnificus causes serious and potentially life-threatening cellulitis (infection of the skin and deeper tissues) and blood-borne infections. Another representative of Vibrionales, Aliivibrio fischeri, engages in a symbiotic relationship with squid. The squid provides nutrients for the bacteria to grow and the bacteria produce bioluminescence that protects the squid from predators (Figure 4.7).
Figure 4.7 (a) Aliivibrio fischeri is a bioluminescent bacterium. (b) A. fischeri colonizes and lives in a mutualistic relationship with the Hawaiian bobtail squid (Euprymna scolopes). (credit a: modification of work by American Society for Microbiology; credit b: modification of work by Margaret McFall-Ngai)
The genus Legionella also belongs to the Gammaproteobacteria. L. pneumophila, the pathogen responsible for Legionnaires disease, is an aquatic bacterium that tends to inhabit pools of warm water, such as those found in the tanks of air conditioning units in large buildings (Figure 4.8). Because the bacteria can spread in aerosols, outbreaks of Legionnaires disease often affect residents of a building in which the water has become contaminated with Legionella. In fact, these bacteria derive their name from the first known outbreak of Legionnaires disease, which occurred in a hotel hosting an American Legion veterans’ association convention in Philadelphia in 1976.
Figure 4.8 (a) Legionella pneumophila, the causative agent of Legionnaires disease, thrives in warm water. (b) Outbreaks of Legionnaires disease often originate in the air conditioning units of large buildings when water in or near the system becomes contaminated with L. pneumophila. (credit a: modification of work by Centers for Disease Control and Prevention)
Enterobacteriaceae is a large family of enteric (intestinal) bacteria belonging to the Gammaproteobacteria. They are facultative anaerobes and are able to ferment carbohydrates. Within this family, microbiologists recognize two distinct categories. The first category is called the coliforms, after its prototypical bacterium species, Escherichia coli. Coliforms are able to ferment lactose completely (i.e., with the production of acid and gas). The second category, noncoliforms, either cannot ferment lactose or can only ferment it incompletely (producing either acid or gas, but not both). The noncoliforms include some notable human pathogens, such as Salmonella spp., Shigella spp., and Yersinia pestis.
E. coli has been perhaps the most studied bacterium since it was first described in 1886 by Theodor Escherich (1857–1911). Many strains of E. coli are in mutualistic relationships with humans. However, some strains produce a potentially deadly toxin called Shiga toxin. Shiga toxin is one of the most potent bacterial toxins identified. Upon entering target cells, Shiga toxin interacts with ribosomes, stopping protein synthesis. Lack of protein synthesis leads to cellular death and hemorrhagic colitis, characterized by inflammation of intestinal tract and bloody diarrhea. In the most severe cases, patients can develop a deadly hemolytic uremic syndrome. Other E. coli strains may cause traveler’s diarrhea, a less severe but very widespread disease.
The genus Salmonella, which belongs to the noncoliform group of Enterobacteriaceae, is interesting in that there is still no consensus about how many species it includes. Scientists have reclassified many of the groups they once thought to be species as serotypes (also called serovars), which are strains or variations of the same species of bacteria. Their classification is based on patterns of reactivity by animal antisera against molecules on the surface of the bacterial cells. A number of serotypes of Salmonella can cause salmonellosis, characterized by inflammation of the small and the large intestine, accompanied by fever, vomiting, and diarrhea. The species S. enterobacterica (serovar typhi) causes typhoid fever, with symptoms including fever, abdominal pain, and skin rashes (Figure 4.9).
Figure 4.9 Salmonella typhi is the causative agent of typhoid fever. (credit: Centers for Disease Control and Prevention)
Table 4.4 summarizes the characteristics of important genera of Gammaproteobacteria.
Class Gammaproteobacteria
Example Genus
Microscopic Morphology
Unique Characteristics
Beggiatoa
Gram-negative bacteria; disc-shaped or cylindrical
Aquatic, live in water with high content of hydrogen disulfide; can cause problems for sewage treatment
Enterobacter
Gram-negative bacillus
Facultative anaerobe; cause urinary and respiratory tract infections in hospitalized patients; implicated in the pathogenesis of obesity
Erwinia
Gram-negative bacillus
Plant pathogen causing leaf spots and discoloration; may digest cellulose; prefer relatively low temperatures (25–30 °C)
Escherichia
Gram-negative bacillus
Facultative anaerobe; inhabit the gastrointestinal tract of warm-blooded animals; some strains are mutualists, producing vitamin K; others, like serotype E. coli O157:H7, are pathogens; E. coli has been a model organism for many studies in genetics and molecular biology
Hemophilus
Gram-negative bacillus
Pleomorphic, may appear as coccobacillus, aerobe, or facultative anaerobe; grow on blood agar; pathogenic species can cause respiratory infections, chancroid, and other diseases
Klebsiella
Gram-negative bacillus; appears rounder and thicker than other members of Enterobacteriaceae
Facultative anaerobe, encapsulated, nonmotile; pathogenic species may cause pneumonia, especially in people with alcoholism
Legionella
Gram-negative bacillus
Fastidious, grow on charcoal-buffered yeast extract; L. pneumophila causes Legionnaires disease
Methylomonas
Gram-negative bacillus
Use methane as source of carbon and energy
Proteus
Gram-negative bacillus (pleomorphic)
Common inhabitants of the human gastrointestinal tract; motile; produce urease; opportunistic pathogens; may cause urinary tract infections and sepsis
Pseudomonas
Gram-negative bacillus
Aerobic; versatile; produce yellow and blue pigments, making them appear green in culture; opportunistic, antibiotic-resistant pathogens may cause wound infections, hospital-acquired infections, and secondary infections in patients with cystic fibrosis
Serratia
Gram-negative bacillus
Motile; may produce red pigment; opportunistic pathogens responsible for a large number of hospital-acquired infections
Shigella
Gram-negative bacillus
Nonmotile; dangerously pathogenic; produce Shiga toxin, which can destroy cells of the gastrointestinal tract; can cause dysentery
Vibrio
Gram-negative, comma- or curved rod-shaped bacteria
Inhabit seawater; flagellated, motile; may produce toxin that causes hypersecretion of water and electrolytes in the gastrointestinal tract; some species may cause serious wound infections
Yersinia
Gram-negative bacillus
Carried by rodents; human pathogens; Y. pestis causes bubonic plague and pneumonic plague; Y. enterocolitica can be a pathogen causing diarrhea in humans
Table4.4
CHECK YOUR UNDERSTANDING
List two families of Gammaproteobacteria.
Deltaproteobacteria
The Deltaproteobacteria is a small class of gram-negative Proteobacteria that includes sulfate-reducing bacteria (SRBs), so named because they use sulfate as the final electron acceptor in the electron transport chain. Few SRBs are pathogenic. However, the SRB Desulfovibrio orale is associated with periodontal disease (disease of the gums).
Deltaproteobacteria also includes the genus Bdellovibrio, species of which are parasites of other gram-negative bacteria. Bdellovibrio invades the cells of the host bacterium, positioning itself in the periplasm, the space between the plasma membrane and the cell wall, feeding on the host’s proteins and polysaccharides. The infection is lethal for the host cells.
Another type of Deltaproteobacteria, myxobacteria, lives in the soil, scavenging inorganic compounds. Motile and highly social, they interact with other bacteria within and outside their own group. They can form multicellular, macroscopic “fruiting bodies” (Figure 4.10), structures that are still being studied by biologists and bacterial ecologists.11 These bacteria can also form metabolically inactive myxospores.
Figure 4.10 Myxobacteria form fruiting bodies. (credit: modification of work by Michiel Vos)
Table 4.5 summarizes the characteristics of several important genera of Deltaproteobacteria.
Class Deltaproteobacteria
Genus
Microscopic Morphology
Unique characteristics
Bdellovibrio
Gram-negative, comma-shaped rod
Obligate aerobes; motile; parasitic (infecting other bacteria)
Desulfovibrio (formerly Desufuromonas)
Gram-negative, comma-shaped rod
Reduce sulfur; can be used for removal of toxic and radioactive waste
Live in soil; can move by gliding; used as a model organism for studies of intercellular communication (signaling)
Table4.5
CHECK YOUR UNDERSTANDING
What type of Deltaproteobacteria forms fruiting bodies?
Epsilonproteobacteria
The smallest class of Proteobacteria is Epsilonproteobacteria, which are gram-negative microaerophilic bacteria (meaning they only require small amounts of oxygen in their environment). Two clinically relevant genera of Epsilonproteobacteria are Campylobacter and Helicobacter, both of which include human pathogens. Campylobacter can cause food poisoning that manifests as severe enteritis (inflammation in the small intestine). This condition, caused by the species C. jejuni, is rather common in developed countries, usually because of eating contaminated poultry products. Chickens often harbor C. jejuni in their gastrointestinal tract and feces, and their meat can become contaminated during processing.
Within the genus Helicobacter, the helical, flagellated bacterium H. pylori has been identified as a beneficial member of the stomach microbiota, but it is also the most common cause of chronic gastritis and ulcers of the stomach and duodenum (Figure 4.11). Studies have also shown that H. pylori is linked to stomach cancer.H. pylori is somewhat unusual in its ability to survive in the highly acidic environment of the stomach. It produces urease and other enzymes that modify its environment to make it less acidic.
Figure 4.11 Helicobacter pylori can cause chronic gastritis, which can lead to ulcers and stomach cancer.
Table 4.6 summarizes the characteristics of the most clinically relevant genera of Epsilonproteobacteria.
Class Epsilonproteobacteria
Example Genus
Microscopic Morphology
Unique Characteristics
Campylobacter
Gram-negative, spiral-shaped rod
Aerobic (microaerophilic); often infects chickens; may infect humans via undercooked meat, causing severe enteritis
Helicobacter
Gram-negative, spiral-shaped rod
Aerobic (microaerophilic) bacterium; can damage the inner lining of the stomach, causing chronic gastritis, peptic ulcers, and stomach cancer
Table4.6
Source: CNX OpenStax
Nonproteobacteria Gram-Negative
Cinematic scientific visualization of chromatophore
Image by Kalinalinkalina/Wikimedia
Cinematic scientific visualization of chromatophore
Cinematic scientific visualization of chromatophore. Simulation by Theoretical Computational Biophysics Group, Beckman Institute, University of Illinois at Urbana-Champaign. Visualization by Advanced Visualization Lab, National Center for Supercomputing Applications, University of Illinois at Urbana-Champaign.
Image by Kalinalinkalina/Wikimedia
Nonproteobacteria Gram-Negative Bacteria and Phototrophic Bacteria
The majority of the gram-negative bacteria belong to the phylum Proteobacteria, discussed in the previous section. Those that do not are called the nonproteobacteria. In this section, we will describe four classes of gram-negative nonproteobacteria: Chlamydia, the spirochetes, the CFB group, and the Plantomycetes. A diverse group of phototrophic bacteria that includes Proteobacteria and nonproteobacteria will be discussed at the end of this section.
Chlamydia
Chlamydia is another taxon of the Alphaproteobacteria. Members of this genus are gram-negative, obligate intracellular pathogens that are extremely resistant to the cellular defenses, giving them the ability to spread from host to host rapidly via elementary bodies. The metabolically and reproductively inactive elementary bodies are the endospore-like form of intracellular bacteria that enter an epithelial cell, where they become active. Figure 4.12 illustrates the life cycle of Chlamydia.
Figure 4.12 Chlamydia begins infection of a host when the metabolically inactive elementary bodies enter an epithelial cell. Once inside the host cell, the elementary bodies turn into active reticulate bodies. The reticulate bodies multiply and release more elementary bodies when the cell dies after the Chlamydia uses all of the host cell’s ATP. (credit: modification of work by Centers for Disease Control and Prevention)
Spirochetes
Spirochetes are characterized by their long (up to 250 μm), spiral-shaped bodies. Most spirochetes are also very thin, which makes it difficult to examine gram-stained preparations under a conventional brightfield microscope. Darkfield fluorescent microscopy is typically used instead. Spirochetes are also difficult or even impossible to culture. They are highly motile, using their axial filament to propel themselves. The axial filament is similar to a flagellum, but it wraps around the cell and runs inside the cell body of a spirochete in the periplasmic space between the outer membrane and the plasma membrane (Figure 4.13).
Figure 4.13 Spirochetes are typically observed using darkfield microscopy (left). However, electron microscopy (top center, bottom center) provides a more detailed view of their cellular morphology. The flagella found between the inner and outer membranes of spirochetes wrap around the bacterium, causing a twisting motion used for locomotion. (credit “spirochetes” micrograph: modification of work by Centers for Disease Control and Prevention; credit “SEM/TEM”: modification of work by Guyard C, Raffel SJ, Schrumpf ME, Dahlstrom E, Sturdevant D, Ricklefs SM, Martens C, Hayes SF, Fischer ER, Hansen BT, Porcella SF, Schwan TG)
Several genera of spirochetes include human pathogens. For example, the genus Treponema includes a species T. pallidum, which is further classified into four subspecies: T. pallidum pallidum, T. pallidum pertenue, T. pallidum carateum, and T. pallidum endemicum. The subspecies T. pallidum pallidum causes the sexually transmitted infection known as syphilis, the third most prevalent sexually transmitted bacterial infection in the United States, after chlamydia and gonorrhea. The other subspecies of T. pallidum cause tropical infectious diseases of the skin, bones, and joints.
Another genus of spirochete, Borrelia, contains a number of pathogenic species. B. burgdorferi causes Lyme disease, which is transmitted by several genera of ticks (notably Ixodes and Amblyomma) and often produces a “bull’s eye” rash, fever, fatigue, and, sometimes, debilitating arthritis. B. recurrens causes a condition known as relapsing fever. Appendix D lists the genera, species, and related diseases for spirochetes.
Cytophaga, Fusobacterium, and Bacteroides
The gram-negative nonproteobacteria of the genera Cytophaga, Fusobacterium, and Bacteroides are classified together as a phylum and called the CFB group. Although they are phylogenetically diverse, bacteria of the CFB group share some similarities in the sequence of nucleotides in their DNA. They are rod-shaped bacteria adapted to anaerobic environments, such as the tissue of the gums, gut, and rumen of ruminating animals. CFB bacteria are avid fermenters, able to process cellulose in rumen, thus enabling ruminant animals to obtain carbon and energy from grazing.
Cytophaga are motile aquatic bacteria that glide. Fusobacteria inhabit the human mouth and may cause severe infectious diseases. The largest genus of the CFB group is Bacteroides, which includes dozens of species that are prevalent inhabitants of the human large intestine, making up about 30% of the entire gut microbiome (Figure 4.14). One gram of human feces contains up to 100 billion Bacteroides cells. Most Bacteroides are mutualistic. They benefit from nutrients they find in the gut, and humans benefit from their ability to prevent pathogens from colonizing the large intestine. Indeed, when populations of Bacteroides are reduced in the gut—as often occurs when a patient takes antibiotics—the gut becomes a more favorable environment for pathogenic bacteria and fungi, which can cause secondary infections.
Figure 4.14 Bacteroides comprise up to 30% of the normal microbiota in the human gut. (credit: NOAA)
Only a few species of Bacteroides are pathogenic. B.melaninogenicus, for example, can cause wound infections in patients with weakened immune systems.
Planctomycetes
The Planctomycetes are found in aquatic environments, inhabiting freshwater, saltwater, and brackish water. Planctomycetes are unusual in that they reproduce by budding, meaning that instead of one maternal cell splitting into two equal daughter cells in the process of binary fission, the mother cell forms a bud that detaches from the mother cell and lives as an independent cell. These so-called swarmer cells are motile and not attached to a surface. However, they will soon differentiate into sessile (immobile) cells with an appendage called a holdfast that allows them to attach to surfaces in the water (Figure 4.15). Only the sessile cells are able to reproduce.
Figure 4.15 (a) Sessile Planctomycetes have a holdfast that allows them to adhere to surfaces in aquatic environments. (b) Swarmers are motile and lack a holdfast. (credit: modification of work by American Society for Microbiology)
Table 4.7 summarizes the characteristics of some of the most clinically relevant genera of nonproteobacteria.
Nonproteobacteria
Example Genus
Microscopic Morphology
Unique Characteristics
Chlamydia
Gram-negative, coccoid or ovoid bacterium
Obligatory intracellular bacteria; some cause chlamydia, trachoma, and pneumonia
Bacteroides
Gram-negative bacillus
Obligate anaerobic bacteria; abundant in the human gastrointestinal tract; usually mutualistic, although some species are opportunistic pathogens
Cytophaga
Gram-negative bacillus
Motile by gliding; live in soil or water; decompose cellulose; may cause disease in fish
Fusobacterium
Gram-negative bacillus with pointed ends
Anaerobic; form; biofilms; some species cause disease in humans (periodontitis, ulcers)
Leptospira
Spiral-shaped bacterium (spirochetes); gram negative-like (better viewed by darkfield microscopy); very thin
Aerobic, abundant in shallow water reservoirs; infect rodents and domestic animals; can be transmitted to humans by infected animals’ urine; may cause severe disease
Borrelia
Gram-negative-like spirochete; very thin; better viewed by darkfield microscopy
B. burgdorferi causes Lyme disease and B. recurrens causes relapsing fever
Treponema
Gram-negative-like spirochete; very thin; better viewed by darkfield microscopy
Motile; do not grow in culture; T. pallidum (subspecies T. pallidum pallidum) causes syphilis
Table4.7
Phototrophic Bacteria
The phototrophic bacteria are a large and diverse category of bacteria that do not represent a taxon but, rather, a group of bacteria that use sunlight as their primary source of energy. This group contains both Proteobacteria and nonproteobacteria. They use solar energy to synthesize ATP through photosynthesis. When they produce oxygen, they perform oxygenic photosynthesis. When they do not produce oxygen, they perform anoxygenic photosynthesis. With the exception of some cyanobacteria, the majority of phototrophic bacteria perform anoxygenic photosynthesis.
One large group of phototrophic bacteria includes the purple or green bacteria that perform photosynthesis with the help of bacteriochlorophylls, which are green, purple, or blue pigments similar to chlorophyll in plants. Some of these bacteria have a varying amount of red or orange pigments called carotenoids. Their color varies from orange to red to purple to green (Figure 4.16), and they are able to absorb light of various wavelengths. Traditionally, these bacteria are classified into sulfur and nonsulfur bacteria; they are further differentiated by color.
Figure 4.16 Purple and green sulfur bacteria use bacteriochlorophylls to perform photosynthesis.
The sulfur bacteria perform anoxygenic photosynthesis, using sulfites as electron donors and releasing free elemental sulfur. Nonsulfur bacteria use organic substrates, such as succinate and malate, as donors of electrons.
The purple sulfur bacteria oxidize hydrogen sulfide into elemental sulfur and sulfuric acid and get their purple color from the pigments bacteriochlorophylls and carotenoids. Bacteria of the genus Chromatium are purple sulfur Gammaproteobacteria. These microorganisms are strict anaerobes and live in water. They use carbon dioxide as their only source of carbon, but their survival and growth are possible only in the presence of sulfites, which they use as electron donors. Chromatium has been used as a model for studies of bacterial photosynthesis since the 1950s.
The green sulfur bacteria use sulfide for oxidation and produce large amounts of green bacteriochlorophyll. The genus Chlorobium is a green sulfur bacterium that is implicated in climate change because it produces methane, a greenhouse gas. These bacteria use at least four types of chlorophyll for photosynthesis. The most prevalent of these, bacteriochlorophyll, is stored in special vesicle-like organelles called chlorosomes.
Purple nonsulfur bacteria are similar to purple sulfur bacteria, except that they use hydrogen rather than hydrogen sulfide for oxidation. Among the purple nonsulfur bacteria is the genus Rhodospirillum. These microorganisms are facultative anaerobes, which are actually pink rather than purple, and can metabolize (“fix”) nitrogen. They may be valuable in the field of biotechnology because of their potential ability to produce biological plastic and hydrogen fuel.
The green nonsulfur bacteria are similar to green sulfur bacteria but they use substrates other than sulfides for oxidation. Chloroflexus is an example of a green nonsulfur bacterium. It often has an orange color when it grows in the dark, but it becomes green when it grows in sunlight. It stores bacteriochlorophyll in chlorosomes, similar to Chlorobium, and performs anoxygenic photosynthesis, using organic sulfites (low concentrations) or molecular hydrogen as electron donors, so it can survive in the dark if oxygen is available. Chloroflexus does not have flagella but can glide, like Cytophaga. It grows at a wide range of temperatures, from 35 °C to 70 °C, thus can be thermophilic.
Another large, diverse group of phototrophic bacteria compose the phylum Cyanobacteria; they get their blue-green color from the chlorophyll contained in their cells (Figure 4.17). Species of this group perform oxygenic photosynthesis, producing megatons of gaseous oxygen. Scientists hypothesize that cyanobacteria played a critical role in the change of our planet’s anoxic atmosphere 1–2 billion years ago to the oxygen-rich environment we have today.
Figure 4.17 (a) Microcystis aeruginosa is a type of cyanobacteria commonly found in freshwater environments. (b) In warm temperatures, M. aeruginosa and other cyanobacteria can multiply rapidly and produce neurotoxins, resulting in blooms that are harmful to fish and other aquatic animals. (credit a: modification of work by Dr. Barry H. Rosen/U.S. Geological Survey; credit b: modification of work by NOAA)
Cyanobacteria have other remarkable properties. Amazingly adaptable, they thrive in many habitats, including marine and freshwater environments, soil, and even rocks. They can live at a wide range of temperatures, even in the extreme temperatures of the Antarctic. They can live as unicellular organisms or in colonies, and they can be filamentous, forming sheaths or biofilms. Many of them fix nitrogen, converting molecular nitrogen into nitrites and nitrates that other bacteria, plants, and animals can use. The reactions of nitrogen fixation occur in specialized cells called heterocysts.
Photosynthesis in Cyanobacteria is oxygenic, using the same type of chlorophyll a found in plants and algae as the primary photosynthetic pigment. Cyanobacteria also use phycocyanin and cyanophycin, two secondary photosynthetic pigments that give them their characteristic blue color. They are located in special organelles called phycobilisomes and in folds of the cellular membrane called thylakoids, which are remarkably similar to the photosynthetic apparatus of plants. Scientists hypothesize that plants originated from endosymbiosis of ancestral eukaryotic cells and ancestral photosynthetic bacteria. Cyanobacteria are also an interesting object of research in biochemistry,with studies investigating their potential as biosorbents and products of human nutrition.
Unfortunately, cyanobacteria can sometimes have a negative impact on human health. Genera such as Microcystis can form harmful cyanobacterial blooms, forming dense mats on bodies of water and producing large quantities of toxins that can harm wildlife and humans. These toxins have been implicated in tumors of the liver and diseases of the nervous system in animals and humans.
Table 4.8 summarizes the characteristics of important phototrophic bacteria.
Phototrophic Bacteria
Phylum
Class
Example Genus or Species
Common Name
Oxygenic or Anoxygenic
Sulfur Deposition
Cyanobacteria
Cyanophyceae
Microcystisaeruginosa
Blue-green bacteria
Oxygenic
None
Chlorobi
Chlorobia
Chlorobium
Green sulfur bacteria
Anoxygenic
Outside the cell
Chloroflexi (Division)
Chloroflexi
Chloroflexus
Green nonsulfur bacteria
Anoxygenic
None
Proteobacteria
Alphaproteobacteria
Rhodospirillum
Purple nonsulfur bacteria
Anoxygenic
None
Betaproteobacteria
Rhodocyclus
Purple nonsulfur bacteria
Anoxygenic
None
Gammaproteobacteria
Chromatium
Purple sulfur bacteria
Anoxygenic
Inside the cell
Table4.8
Source: CNX OpenStax
Additional Materials (2)
Gram Positive and Gram Negative Bacteria
Video by Biology Professor/YouTube
Bacteria shapes
A colorized scanning electron micrograph of bacteria. Scanning electron microscopes allow scientists to see the three-dimensional surface of their samples.
Appears in the NIGMS booklet Inside the Cell.
Image by Tina Carvalho, University of Hawaii at Manoa
5:50
Gram Positive and Gram Negative Bacteria
Biology Professor/YouTube
Bacteria shapes
Tina Carvalho, University of Hawaii at Manoa
Gram-Negative Bacteria Infections
Gram Stain
Image by Paityn Donaldson/Wikimedia
Gram Stain
The gram staining procedure and its effects on gram (-) and gram (+) cell wall types, respectively.
Image by Paityn Donaldson/Wikimedia
Gram-Negative Bacteria Infections in Healthcare Settings
Gram-negative Bacteria Infections in Healthcare Settings
General Information about gram-negative bacteria
Gram-negative bacteria cause infections including pneumonia, bloodstream infections, wound or surgical site infections, and meningitis in healthcare settings. Gram-negative bacteria are resistant to multiple drugs and are increasingly resistant to most available antibiotics. These bacteria have built-in abilities to find new ways to be resistant and can pass along genetic materials that allow other bacteria to become drug-resistant as well. CDC’s aggressive recommendations, if implemented, can prevent the spread of gram-negatives.
Gram-negative infections include those caused by Klebsiella, Acinetobacter, Pseudomonas aeruginosa, and E. coli., as well as many other less common bacteria.
Outbreak investigations
Outbreak investigations have led to a better understanding of how to control these bacteria in healthcare. In the past 3 years, the Division of Healthcare Quality Promotion has assisted in at least 10 investigations of outbreaks of gram negative infections.
CDC has collaborated with state health departments in Maryland and Arizona to successfully control outbreaks of Multidrug-resistant-Acinetobacter infections occurring among intensive care unit patients.
CDC has worked with the Puerto Rico health department to control an outbreak of highly resistant Klebsiella at a neonatal intensive-care unit in Puerto Rico.
CDC assisted the Ohio health department’s investigation of infections caused by Acinetobacter. These outbreaks have occurred in various healthcare facilities in the state of Ohio and have been controlled by aggressive infection control interventions.
CDC worked with the state health department of Texas on separate outbreaks of B. cepacia and Pseudomonas.
Additionally, CDC worked with the state health department in Georgia on an unrelated outbreak of B. cepacia.
CDC worked with the Department of Defense to investigate and control Acinetobacter infections occurring in soldiers injured in the Middle East. This collaboration led to important improvements in infection control in military medical facilities.
In addition to these outbreaks, CDC’s reference laboratory has confirmed carbapenemase resistance in bacteria for 32 other U.S. states.
Laboratory tests for detecting resistance
CDC is collaborating with laboratory standards-setting institutions to identify and recommend laboratory tests for the accurate detection of carbapenemase-mediated resistance.
CDC is working with states to identify isolates with unusual resistance and to determine new mechanisms of resistance among multidrug-resistant gram-negatives, including the recent identification of a new mechanism of resistance in patients returning from Asia.
Source: Centers for Disease Control and Prevention (CDC)
Additional Materials (3)
Micrococcus luteus
Under a high magnification of 10965x, this scanning electron micrograph (SEM) depicted some of the ultrastructural morphologic features displayed by this group of Gram-positive Micrococcus luteus bacteria. The specimen was obtained from a pure culture that was raised on a polycarbonate filter, for the purpose of identification of the organism.
Peter Raven, Kenneth Mason, Jonathan Losos, Susan Singer - McGraw-Hill Education, Ali Zifan
Gram-Positive Bacteria
(From L to R) Purple stained Gram-positive and pink stained Gram-negative
Image by Scientific Animations, Inc.
(From L to R) Purple stained Gram-positive and pink stained Gram-negative
Gram-positive bacteria(left) with thick peptidoglycan layer stains.
Image by Scientific Animations, Inc.
Gram-Positive Bacteria
Prokaryotes are identified as gram-positive if they have a multiple layer matrix of peptidoglycan forming the cell wall. Crystal violet, the primary stain of the Gram stain procedure, is readily retained and stabilized within this matrix, causing gram-positive prokaryotes to appear purple under a brightfield microscope after Gram staining. For many years, the retention of Gram stain was one of the main criteria used to classify prokaryotes, even though some prokaryotes did not readily stain with either the primary or secondary stains used in the Gram stain procedure.
Advances in nucleic acid biochemistry have revealed additional characteristics that can be used to classify gram-positive prokaryotes, namely the guanine to cytosine ratios (G+C) in DNA and the composition of 16S rRNA subunits. Microbiologists currently recognize two distinct groups of gram-positive, or weakly staining gram-positive, prokaryotes. The class Actinobacteria comprises the high G+C gram-positive bacteria, which have more than 50% guanine and cytosine nucleotides in their DNA. The class Bacilli comprises low G+C gram-positive bacteria, which have less than 50% of guanine and cytosine nucleotides in their DNA.
Actinobacteria: High G+C Gram-Positive Bacteria
The name Actinobacteria comes from the Greek words for rays and small rod, but Actinobacteria are very diverse. Their microscopic appearance can range from thin filamentous branching rods to coccobacilli. Some Actinobacteria are very large and complex, whereas others are among the smallest independently living organisms. Most Actinobacteria live in the soil, but some are aquatic. The vast majority are aerobic. One distinctive feature of this group is the presence of several different peptidoglycans in the cell wall.
The genus Actinomyces is a much studied representative of Actinobacteria. Actinomyces spp. play an important role in soil ecology, and some species are human pathogens. A number of Actinomyces spp. inhabit the human mouth and are opportunistic pathogens, causing infectious diseases like periodontitis (inflammation of the gums) and oral abscesses. The species A. israelii is an anaerobe notorious for causing endocarditis (inflammation of the inner lining of the heart) (Figure 4.18).
Figure 4.18 (a) Actinomyces israelii (false-color scanning electron micrograph [SEM]) has a branched structure. (b) Corynebacterium diphtheria causes the deadly disease diphtheria. Note the distinctive palisades. (c) The gram-variable bacterium Gardnerella vaginalis causes bacterial vaginosis in women. This micrograph shows a Pap smear from a woman with vaginosis. (credit a: modification of work by “GrahamColm”/Wikimedia Commons; credit b: modification of work by Centers for Disease Control and Prevention; credit c: modification of work by Mwakigonja AR, Torres LM, Mwakyoma HA, Kaaya EE)
The genus Mycobacterium is represented by bacilli covered with a mycolic acid coat. This waxy coat protects the bacteria from some antibiotics, prevents them from drying out, and blocks penetration by Gram stain reagents. Because of this, a special acid-fast staining procedure is used to visualize these bacteria. The genus Mycobacterium is an important cause of a diverse group of infectious diseases. M. tuberculosis is the causative agent of tuberculosis, a disease that primarily impacts the lungs but can infect other parts of the body as well. It has been estimated that one-third of the world’s population has been infected with M. tuberculosis and millions of new infections occur each year. Treatment of M. tuberculosis is challenging and requires patients to take a combination of drugs for an extended time. Complicating treatment even further is the development and spread of multidrug-resistant strains of this pathogen.
Another pathogenic species, M. leprae, is the cause of Hansen’s disease (leprosy), a chronic disease that impacts peripheral nerves and the integrity of the skin and mucosal surface of the respiratory tract. Loss of pain sensation and the presence of skin lesions increase susceptibility to secondary injuries and infections with other pathogens.
Bacteria in the genus Corynebacterium contain diaminopimelic acid in their cell walls, and microscopically often form palisades, or pairs of rod-shaped cells resembling the letter V. Cells may contain metachromatic granules, intracellular storage of inorganic phosphates that are useful for identification of Corynebacterium. The vast majority of Corynebacterium spp. are nonpathogenic; however, C. diphtheria is the causative agent of diphtheria, a disease that can be fatal, especially in children (Figure 4.18). C. diphtheria produces a toxin that forms a pseudomembrane in the patient’s throat, causing swelling, difficulty breathing, and other symptoms that can become serious if untreated.
The genus Bifidobacterium consists of filamentous anaerobes, many of which are commonly found in the gastrointestinal tract, vagina, and mouth. In fact, Bifidobacterium spp. constitute a substantial part of the human gut microbiota and are frequently used as probiotics and in yogurt production.
The genus Gardnerella, contains only one species, G. vaginalis. This species is defined as “gram-variable” because its small coccobacilli do not show consistent results when Gram stained (Figure 4.18). Based on its genome, it is placed into the high G+C gram-positive group. G. vaginalis can cause bacterial vaginosis in women; symptoms are typically mild or even undetectable, but can lead to complications during pregnancy.
Table 4.9 summarizes the characteristics of some important genera of Actinobacteria.
Actinobacteria: High G+C Gram-Positive
Example Genus
Microscopic Morphology
Unique Characteristics
Actinomyces
Gram-positive bacillus; in colonies, shows fungus-like threads (hyphae)
Facultative anaerobes; in soil, decompose organic matter; in the human mouth, may cause gum disease
Arthrobacter
Gram-positive bacillus (at the exponential stage of growth) or coccus (in stationary phase)
Obligate aerobes; divide by “snapping,” forming V-like pairs of daughter cells; degrade phenol, can be used in bioremediation
Bifidobacterium
Gram-positive, filamentous actinobacterium
Anaerobes commonly found in human gut microbiota
Corynebacterium
Gram-positive bacillus
Aerobes or facultative anaerobes; form palisades; grow slowly; require enriched media in culture; C. diphtheriae causes diphtheria
Frankia
Gram-positive, fungus-like (filamentous) bacillus
Nitrogen-fixing bacteria; live in symbiosis with legumes
Gardnerella
Gram-variable coccobacillus
Colonize the human vagina, may alter the microbial ecology, thus leading to vaginosis
Micrococcus
Gram-positive coccus, form microscopic clusters
Ubiquitous in the environment and on the human skin; oxidase-positive (as opposed to morphologically similar S. aureus); some are opportunistic pathogens
Mycobacterium
Gram-positive, acid-fast bacillus
Slow growing, aerobic, resistant to drying and phagocytosis; covered with a waxy coat made of mycolic acid; M. tuberculosis causes tuberculosis; M. leprae causes leprosy
Nocardia
Weakly gram-positive bacillus; forms acid-fast branches
May colonize the human gingiva; may cause severe pneumonia and inflammation of the skin
Propionibacterium
Gram-positive bacillus
Aerotolerant anaerobe; slow-growing; P. acnes reproduces in the human sebaceous glands and may cause or contribute to acne
Rhodococcus
Gram-positive bacillus
Strict aerobe; used in industry for biodegradation of pollutants; R. fascians is a plant pathogen, and R. equi causes pneumonia in foals
Streptomyces
Gram-positive, fungus-like (filamentous) bacillus
Very diverse genus (>500 species); aerobic, spore-forming bacteria; scavengers, decomposers found in soil (give the soil its “earthy” odor); used in pharmaceutical industry as antibiotic producers (more than two-thirds of clinically useful antibiotics)
Table4.9
Low G+C Gram-positive Bacteria
The low G+C gram-positive bacteria have less than 50% guanine and cytosine in their DNA, and this group of bacteria includes a number of genera of bacteria that are pathogenic.
Clostridia
One large and diverse class of low G+C gram-positive bacteria is Clostridia. The best studied genus of this class is Clostridium. These rod-shaped bacteria are generally obligate anaerobes that produce endospores and can be found in anaerobic habitats like soil and aquatic sediments rich in organic nutrients. The endospores may survive for many years.
Clostridium spp. produce more kinds of protein toxins than any other bacterial genus, and several species are human pathogens. C. perfringens is the third most common cause of food poisoning in the United States and is the causative agent of an even more serious disease called gas gangrene. Gas gangrene occurs when C. perfringens endospores enter a wound and germinate, becoming viable bacterial cells and producing a toxin that can cause the necrosis (death) of tissue. C. tetani, which causes tetanus, produces a neurotoxin that is able to enter neurons, travel to regions of the central nervous system where it blocks the inhibition of nerve impulses involved in muscle contractions, and cause a life-threatening spastic paralysis. C. botulinum produces botulinum neurotoxin, the most lethal biological toxin known. Botulinum toxin is responsible for rare but frequently fatal cases of botulism. The toxin blocks the release of acetylcholine in neuromuscular junctions, causing flaccid paralysis. In very small concentrations, botulinum toxin has been used to treat muscle pathologies in humans and in a cosmetic procedure to eliminate wrinkles. C. difficile is a common source of hospital-acquired infections (Figure 4.19) that can result in serious and even fatal cases of colitis (inflammation of the large intestine). Infections often occur in patients who are immunosuppressed or undergoing antibiotic therapy that alters the normal microbiota of the gastrointestinal tract.
Figure 4.19 Clostridium difficile , a gram-positive, rod-shaped bacterium, causes severe colitis and diarrhea, often after the normal gut microbiota is eradicated by antibiotics. (credit: modification of work by Centers for Disease Control and Prevention)
Lactobacillales
The order Lactobacillales comprises low G+C gram-positive bacteria that include both bacilli and cocci in the genera Lactobacillus, Leuconostoc, Enterococcus, and Streptococcus. Bacteria of the latter three genera typically are spherical or ovoid and often form chains.
Streptococcus, the name of which comes from the Greek word for twisted chain, is responsible for many types of infectious diseases in humans. Species from this genus, often referred to as streptococci, are usually classified by serotypes called Lancefield groups, and by their ability to lyse red blood cells when grown on blood agar.
S. pyogenes belongs to the Lancefield group A, β-hemolytic Streptococcus. This species is considered a pyogenic pathogen because of the associated pus production observed with infections it causes (Figure 4.20). S. pyogenes is the most common cause of bacterial pharyngitis (strep throat); it is also an important cause of various skin infections that can be relatively mild (e.g., impetigo) or life threatening (e.g., necrotizing fasciitis, also known as flesh eating disease), life threatening.
Figure 4.20 (a) A gram-stained specimen of Streptococcus pyogenes shows the chains of cocci characteristic of this organism’s morphology. (b) S. pyogenes on blood agar shows characteristic lysis of red blood cells, indicated by the halo of clearing around colonies. (credit a, b: modification of work by American Society for Microbiology)
The nonpyogenic (i.e., not associated with pus production) streptococci are a group of streptococcal species that are not a taxon but are grouped together because they inhabit the human mouth. The nonpyogenic streptococci do not belong to any of the Lancefield groups. Most are commensals, but a few, such as S. mutans, are implicated in the development of dental caries.
S. pneumoniae (commonly referred to as pneumococcus), is a Streptococcus species that also does not belong to any Lancefield group. S. pneumoniae cells appear microscopically as diplococci, pairs of cells, rather than the long chains typical of most streptococci. Scientists have known since the 19th century that S. pneumoniae causes pneumonia and other respiratory infections. However, this bacterium can also cause a wide range of other diseases, including meningitis, septicemia, osteomyelitis, and endocarditis, especially in newborns, the elderly, and patients with immunodeficiency.
Bacilli
The name of the class Bacilli suggests that it is made up of bacteria that are bacillus in shape, but it is a morphologically diverse class that includes bacillus-shaped and cocccus-shaped genera. Among the many genera in this class are two that are very important clinically: Bacillus and Staphylococcus.
Bacteria in the genus Bacillus are bacillus in shape and can produce endospores. They include aerobes or facultative anaerobes. A number of Bacillus spp. are used in various industries, including the production of antibiotics (e.g., barnase), enzymes (e.g., alpha-amylase, BamH1 restriction endonuclease), and detergents (e.g., subtilisin).
Two notable pathogens belong to the genus Bacillus. B. anthracis is the pathogen that causes anthrax, a severe disease that affects wild and domesticated animals and can spread from infected animals to humans. Anthrax manifests in humans as charcoal-black ulcers on the skin, severe enterocolitis, pneumonia, and brain damage due to swelling. If untreated, anthrax is lethal. B. cereus, a closely related species, is a pathogen that may cause food poisoning. It is a rod-shaped species that forms chains. Colonies appear milky white with irregular shapes when cultured on blood agar (Figure 4.21). One other important species is B. thuringiensis. This bacterium produces a number of substances used as insecticides because they are toxic for insects.
Figure 4.21 (a) In this gram-stained specimen, the violet rod-shaped cells forming chains are the gram-positive bacteria Bacillus cereus . The small, pink cells are the gram-negative bacteria Escherichia coli . (b) In this culture, white colonies of B. cereus have been grown on sheep blood agar. (credit a: modification of work by “Bibliomaniac 15”/Wikimedia Commons; credit b: modification of work by Centers for Disease Control and Prevention)
The genus Staphylococcus also belongs to the class Bacilli, even though its shape is coccus rather than a bacillus. The name Staphylococcus comes from a Greek word for bunches of grapes, which describes their microscopic appearance in culture (Figure 4.22). Staphylococcus spp. are facultative anaerobic, halophilic, and nonmotile. The two best-studied species of this genus are S. epidermidis and S. aureus.
Figure 4.22 This SEM of Staphylococcus aureus illustrates the typical “grape-like” clustering of cells. (credit: modification of work by Centers for Disease Control and Prevention)
S. epidermidis, whose main habitat is the human skin, is thought to be nonpathogenic for humans with healthy immune systems, but in patients with immunodeficiency, it may cause infections in skin wounds and prostheses (e.g., artificial joints, heart valves). S. epidermidis is also an important cause of infections associated with intravenous catheters. This makes it a dangerous pathogen in hospital settings, where many patients may be immunocompromised.
Strains of S. aureus cause a wide variety of infections in humans, including skin infections that produce boils, carbuncles, cellulitis, or impetigo. Certain strains of S. aureus produce a substance called enterotoxin, which can cause severe enteritis, often called staph food poisoning. Some strains of S. aureus produce the toxin responsible for toxic shock syndrome, which can result in cardiovascular collapse and death.
Many strains of S. aureus have developed resistance to antibiotics. Some antibiotic-resistant strains are designated as methicillin-resistant S. aureus (MRSA) and vancomycin-resistant S. aureus (VRSA). These strains are some of the most difficult to treat because they exhibit resistance to nearly all available antibiotics, not just methicillin and vancomycin. Because they are difficult to treat with antibiotics, infections can be lethal. MRSA and VRSA are also contagious, posing a serious threat in hospitals, nursing homes, dialysis facilities, and other places where there are large populations of elderly, bedridden, and/or immunocompromised patients.
Mycoplasmas
Although Mycoplasma spp. do not possess a cell wall and, therefore, are not stained by Gram-stain reagents, this genus is still included with the low G+C gram-positive bacteria. The genus Mycoplasma includes more than 100 species, which share several unique characteristics. They are very small cells, some with a diameter of about 0.2 μm, which is smaller than some large viruses. They have no cell walls and, therefore, are pleomorphic, meaning that they may take on a variety of shapes and can even resemble very small animal cells. Because they lack a characteristic shape, they can be difficult to identify. One species, M. pneumoniae, causes the mild form of pneumonia known as “walking pneumonia” or “atypical pneumonia.” This form of pneumonia is typically less severe than forms caused by other bacteria or viruses.
Table 4.10 summarizes the characteristics of notable genera low G+C Gram-positive bacteria.
Bacilli: Low G+C Gram-Positive Bacteria
Example Genus
Microscopic Morphology
Unique Characteristics
Bacillus
Large, gram-positive bacillus
Aerobes or facultative anaerobes; form endospores; B. anthracis causes anthrax in cattle and humans, B. cereus may cause food poisoning
Clostridium
Gram-positive bacillus
Strict anaerobes; form endospores; all known species are pathogenic, causing tetanus, gas gangrene, botulism, and colitis
Enterococcus
Gram-positive coccus; forms microscopic pairs in culture (resembling Streptococcus pneumoniae)
Anaerobic aerotolerant bacteria, abundant in the human gut, may cause urinary tract and other infections in the nosocomial environment
Lactobacillus
Gram-positive bacillus
Facultative anaerobes; ferment sugars into lactic acid; part of the vaginal microbiota; used as probiotics
Leuconostoc
Gram-positive coccus; may form microscopic chains in culture
Fermenter, used in food industry to produce sauerkraut and kefir
Mycoplasma
The smallest bacteria; appear pleomorphic under electron microscope
Have no cell wall; classified as low G+C Gram-positive bacteria because of their genome; M. pneumoniae causes “walking” pneumonia
Staphylococcus
Gram-positive coccus; forms microscopic clusters in culture that resemble bunches of grapes
Tolerate high salt concentration; facultative anaerobes; produce catalase; S. aureus can also produce coagulase and toxins responsible for local (skin) and generalized infections
Streptococcus
Gram-positive coccus; forms chains or pairs in culture
Diverse genus; classified into groups based on sharing certain antigens; some species cause hemolysis and may produce toxins responsible for human local (throat) and generalized disease
Ureaplasma
Similar to Mycoplasma
Part of the human vaginal and lower urinary tract microbiota; may cause inflammation, sometimes leading to internal scarring and infertility
Table4.10
CLINICAL FOCUS
Resolution
Marsha’s sputum sample was sent to the microbiology lab to confirm the identity of the microorganism causing her infection. The lab also performed antimicrobial susceptibility testing (AST) on the sample to confirm that the physician has prescribed the correct antimicrobial drugs.
Direct microscopic examination of the sputum revealed acid-fast bacteria (AFB) present in Marsha’s sputum. When placed in culture, there were no signs of growth for the first 8 days, suggesting that microorganism was either dead or growing very slowly. Slow growth is a distinctive characteristic of M.tuberculosis.
After four weeks, the lab microbiologist observed distinctive colorless granulated colonies (Figure 4.23). The colonies contained AFB showing the same microscopic characteristics as those revealed during the direct microscopic examination of Marsha’s sputum. To confirm the identification of the AFB, samples of the colonies were analyzed using nucleic acid hybridization, or direct nucleic acid amplification (NAA) testing. When a bacterium is acid-fast, it is classified in the family Mycobacteriaceae. DNA sequencing of variable genomic regions of the DNA extracted from these bacteria revealed that it was high G+C. This fact served to finalize Marsha’s diagnosis as infection with M. tuberculosis. After nine months of treatment with the drugs prescribed by her doctor, Marsha made a full recovery.
Figure 4.23 M. tuberculosis grows on Löwenstein-Jensen (LJ) agar in distinct colonies. (credit: Centers for Disease Control and Prevention)
EYE ON ETHICS
Biopiracy and Bioprospecting
In 1969, an employee of a Swiss pharmaceutical company was vacationing in Norway and decided to collect some soil samples. He took them back to his lab, and the Swiss company subsequently used the fungus Tolypocladiuminflatum in those samples to develop cyclosporine A, a drug widely used in patients who undergo tissue or organ transplantation. The Swiss company earns more than $1 billion a year for production of cyclosporine A, yet Norway receives nothing in return—no payment to the government or benefit for the Norwegian people. Despite the fact the cyclosporine A saves numerous lives, many consider the means by which the soil samples were obtained to be an act of “biopiracy,” essentially a form of theft. Do the ends justify the means in a case like this?
Nature is full of as-yet-undiscovered bacteria and other microorganisms that could one day be used to develop new life-saving drugs or treatments. Pharmaceutical and biotechnology companies stand to reap huge profits from such discoveries, but ethical questions remain. To whom do biological resources belong? Should companies who invest (and risk) millions of dollars in research and development be required to share revenue or royalties for the right to access biological resources?
Compensation is not the only issue when it comes to bioprospecting. Some communities and cultures are philosophically opposed to bioprospecting, fearing unforeseen consequences of collecting genetic or biological material. Native Hawaiians, for example, are very protective of their unique biological resources.
For many years, it was unclear what rights government agencies, private corporations, and citizens had when it came to collecting samples of microorganisms from public land. Then, in 1993, the Convention on Biological Diversity granted each nation the rights to any genetic and biological material found on their own land. Scientists can no longer collect samples without a prior arrangement with the land owner for compensation. This convention now ensures that companies act ethically in obtaining the samples they use to create their products.
Source: CNX OpenStax
Additional Materials (7)
Gram Positive vs. Gram Negative Bacteria
Video by Beverly Biology/YouTube
Structure of a bacterium
Image by JrPol/Wikimedia
Bacteria
Average prokaryote cell
Image by Mariana Ruiz Villarreal, LadyofHats
Bacteria
A diagram of a typical prokaryotic bacteria cell
Image by Peter Raven, Kenneth Mason, Jonathan Losos, Susan Singer - McGraw-Hill Education, Ali Zifan
Acinetobacter baumannii
Under a magnification of 27600X, this scanning electron microscopic (SEM) image depicted a cluster of Gram-negative, non-motile, Acinetobacter baumannii bacteria. Members of the genus Acinetobacter are nonmotile rods, 1.0-1.5µm in diameter, and 1.5-2.5µm in length, becoming spherical during their stationary growth phase. They are widely distributed in nature, and are normal flora on the skin. See PHIL 10095 for a colorized version of this image.
Image by CDC/ Matthew J. Arduino, DRPH; Jana Swenson; Photo credit: Janice Haney Carr
Bacillus cereus
Bacillus cereus, SEM image
Image by Mogana Das Murtey and Patchamuthu Ramasamy
Streptococcus Pyogenes (Group A Strep)
Colorized scanning electron micrograph of Group A Streptococcus (Streptococcus pyogenes) bacteria (blue) and a human neutrophil (purple).
Image by NIAID
9:19
Gram Positive vs. Gram Negative Bacteria
Beverly Biology/YouTube
Structure of a bacterium
JrPol/Wikimedia
Bacteria
Mariana Ruiz Villarreal, LadyofHats
Bacteria
Peter Raven, Kenneth Mason, Jonathan Losos, Susan Singer - McGraw-Hill Education, Ali Zifan
Acinetobacter baumannii
CDC/ Matthew J. Arduino, DRPH; Jana Swenson; Photo credit: Janice Haney Carr
Bacillus cereus
Mogana Das Murtey and Patchamuthu Ramasamy
Streptococcus Pyogenes (Group A Strep)
NIAID
Deeply Branching Bacteria
Bacteria
Image by qimono/Pixabay
Bacteria
Image by qimono/Pixabay
Deeply Branching Bacteria
On a phylogenetic tree, the trunk or root of the tree represents a common ancient evolutionary ancestor, often called the last universal common ancestor (LUCA), and the branches are its evolutionary descendants. Scientists consider the deeply branching bacteria, such as the genus Acetothermus, to be the first of these non-LUCA forms of life produced by evolution some 3.5 billion years ago. When placed on the phylogenetic tree, they stem from the common root of life, deep and close to the LUCA root—hence the name “deeply branching” (Figure 4.24).
Figure 4.24 The star on this phylogenetic tree of life shows the position of the deeply branching bacteria Acetothermus. (credit: modification of work by Eric Gaba)
The deeply branching bacteria may provide clues regarding the structure and function of ancient and now extinct forms of life. We can hypothesize that ancient bacteria, like the deeply branching bacteria that still exist, were thermophiles or hyperthermophiles, meaning that they thrived at very high temperatures. Acetothermus paucivorans, a gram-negative anaerobic bacterium discovered in 1988 in sewage sludge, is a thermophile growing at an optimal temperature of 58 °C.22 Scientists have determined it to be the deepest branching bacterium, or the closest evolutionary relative of the LUCA (Figure 4.24).
The class Aquificae includes deeply branching bacteria that are adapted to the harshest conditions on our planet, resembling the conditions thought to dominate the earth when life first appeared. Bacteria from the genus Aquifex are hyperthermophiles, living in hot springs at a temperature higher than 90 °C. The species A. pyrophilus thrives near underwater volcanoes and thermal ocean vents, where the temperature of water (under high pressure) can reach 138 °C. Aquifex bacteria use inorganic substances as nutrients. For example, A. pyrophilus can reduce oxygen, and it is able to reduce nitrogen in anaerobic conditions. They also show a remarkable resistance to ultraviolet light and ionizing radiation. Taken together, these observations support the hypothesis that the ancient ancestors of deeply branching bacteria began evolving more than 3 billion years ago, when the earth was hot and lacked an atmosphere, exposing the bacteria to nonionizing and ionizing radiation.
The class Thermotogae is represented mostly by hyperthermophilic, as well as some mesophilic (preferring moderate temperatures), anaerobic gram-negative bacteria whose cells are wrapped in a peculiar sheath-like outer membrane called a toga. The thin layer of peptidoglycan in their cell wall has an unusual structure; it contains diaminopimelic acid and D-lysine. These bacteria are able to use a variety of organic substrates and produce molecular hydrogen, which can be used in industry. The class contains several genera, of which the best known is the genus Thermotoga. One species of this genus, T. maritima, lives near the thermal ocean vents and thrives in temperatures of 90 °C; another species, T. subterranea, lives in underground oil reservoirs.
Finally, the deeply branching bacterium Deinococcus radiodurans belongs to a genus whose name is derived from a Greek word meaning terribleberry. Nicknamed “Conan the Bacterium,” D. radiodurans is considered a polyextremophile because of its ability to survive under the many different kinds of extreme conditions—extreme heat, drought, vacuum, acidity, and radiation. It owes its name to its ability to withstand doses of ionizing radiation that kill all other known bacteria; this special ability is attributed to some unique mechanisms of DNA repair.
Figure 4.25 Deinococcus radiodurans, or “Conan the Bacterium,” survives in the harshest conditions on earth.
Source: CNX OpenStax
Additional Materials (1)
Bacteria
Bacteria are small single-celled organisms.
Image by National Human Genome Research Institute (NHGRI)
Bacteria
National Human Genome Research Institute (NHGRI)
Methanobrevibacter
Archaebacteria
Image by K. Gottlieb, V. Wacher, J. Sliman and M. Pimentel
Archaebacteria
Diagrammatic view of Methanobrevibacter smithii, showing the cell membrane (ochre, with inset) and cell wall (purple).
Image by K. Gottlieb, V. Wacher, J. Sliman and M. Pimentel
Methanobrevibacter
A genus of gram-positive, anaerobic, cocci to short rod-shaped archaea, in the family methanobacteriaceae, order methanobacteriales. They are found in the gastrointestinal tract or other anoxic environments.
Source: National Center for Biotechnology Information (NCBI)
Additional Materials (2)
Protocyte of a Methanobrevibacter gottschalkii
Protocyte of a Methanobrevibacter gottschalkii
Image by Wolfgang Gelbricht/Wikimedia
Lora Hooper (UT Southwestern) 1: Mammalian gut microbiota: Mammals and their symbiotic gut microbes
Video by iBiology/YouTube
Protocyte of a Methanobrevibacter gottschalkii
Wolfgang Gelbricht/Wikimedia
34:55
Lora Hooper (UT Southwestern) 1: Mammalian gut microbiota: Mammals and their symbiotic gut microbes
iBiology/YouTube
Archaea
Prokaryotes: Bacteria and Archaea
Image by CNX Openstax (credit: modification of work by Jon Sullivan)
Prokaryotes: Bacteria and Archaea
Certain prokaryotes can live in extreme environments such as the Morning Glory pool, a hot spring in Yellowstone National Park. The spring’s vivid blue color is from the prokaryotes that thrive in its very hot waters. (credit: modification of work by Jon Sullivan)
Image by CNX Openstax (credit: modification of work by Jon Sullivan)
Archaea
Like organisms in the domain Bacteria, organisms of the domain Archaea are all unicellular organisms. However, archaea differ structurally from bacteria in several significant ways. To summarize:
The archaeal cell membrane is composed of ether linkages with branched isoprene chains (as opposed to the bacterial cell membrane, which has ester linkages with unbranched fatty acids).
Archaeal cell walls lack peptidoglycan, but some contain a structurally similar substance called pseudopeptidoglycan or pseudomurein.
The genomes of Archaea are larger and more complex than those of bacteria.
Domain Archaea is as diverse as domain Bacteria, and its representatives can be found in any habitat. Some archaea are mesophiles, and many are extremophiles, preferring extreme hot or cold, extreme salinity, or other conditions that are hostile to most other forms of life on earth. Their metabolism is adapted to the harsh environments, and they can perform methanogenesis, for example, which bacteria and eukaryotes cannot.
The size and complexity of the archaeal genome makes it difficult to classify. Most taxonomists agree that within the Archaea, there are currently five major phyla: Crenarchaeota, Euryarchaeota, Korarchaeota, Nanoarchaeota, and Thaumarchaeota. There are likely many other archaeal groups that have not yet been systematically studied and classified.
With few exceptions, archaea are not present in the human microbiota, and none are currently known to be associated with infectious diseases in humans, animals, plants, or microorganisms. However, many play important roles in the environment and may thus have an indirect impact on human health.
Crenarchaeota
Crenarchaeota is a class of Archaea that is extremely diverse, containing genera and species that differ vastly in their morphology and requirements for growth. All Crenarchaeota are aquatic organisms, and they are thought to be the most abundant microorganisms in the oceans. Most, but not all, Crenarchaeota are hyperthermophiles; some of them (notably, the genus Pyrolobus) are able to grow at temperatures up to 113 °C.
Archaea of the genus Sulfolobus (Figure 4.26) are thermophiles that prefer temperatures around 70–80°C and acidophiles that prefer a pH of 2–3. Sulfolobus can live in aerobic or anaerobic environments. In the presence of oxygen, Sulfolobus spp. use metabolic processes similar to those of heterotrophs. In anaerobic environments, they oxidize sulfur to produce sulfuric acid, which is stored in granules. Sulfolobus spp. are used in biotechnology for the production of thermostable and acid-resistant proteins called affitins. Affitins can bind and neutralize various antigens (molecules found in toxins or infectious agents that provoke an immune response from the body).
Figure 4.26 Sulfolobus, an archaeon of the class Crenarchaeota, oxidizes sulfur and stores sulfuric acid in its granules.
Another genus, Thermoproteus, is represented by strictly anaerobic organisms with an optimal growth temperature of 85 °C. They have flagella and, therefore, are motile. Thermoproteus has a cellular membrane in which lipids form a monolayer rather than a bilayer, which is typical for archaea. Its metabolism is autotrophic. To synthesize ATP, Thermoproteus spp. reduce sulfur or molecular hydrogen and use carbon dioxide or carbon monoxide as a source of carbon. Thermoproteus is thought to be the deepest-branching genus of Archaea, and thus is a living example of some of our planet’s earliest forms of life.
Euryarchaeota
The phylum Euryarchaeota includes several distinct classes. Species in the classes Methanobacteria, Methanococci, and Methanomicrobia represent Archaea that can be generally described as methanogens. Methanogens are unique in that they can reduce carbon dioxide in the presence of hydrogen, producing methane. They can live in the most extreme environments and can reproduce at temperatures varying from below freezing to boiling. Methanogens have been found in hot springs as well as deep under ice in Greenland. Some scientists have even hypothesized that methanogens may inhabit the planet Mars because the mixture of gases produced by methanogens resembles the makeup of the Martian atmosphere.
Methanogens are thought to contribute to the formation of anoxic sediments by producing hydrogen sulfide, making “marsh gas.” They also produce gases in ruminants and humans. Some genera of methanogens, notably Methanosarcina, can grow and produce methane in the presence of oxygen, although the vast majority are strict anaerobes.
The class Halobacteria (which was named before scientists recognized the distinction between Archaea and Bacteria) includes halophilic (“salt-loving”) archaea. Halobacteria require a very high concentrations of sodium chloride in their aquatic environment. The required concentration is close to saturation, at 36%; such environments include the Dead Sea as well as some salty lakes in Antarctica and south-central Asia. One remarkable feature of these organisms is that they perform photosynthesis using the protein bacteriorhodopsin, which gives them, and the bodies of water they inhabit, a beautiful purple color (Figure 4.27).
Figure 4.27 Halobacteria growing in these salt ponds gives them a distinct purple color. (credit: modification of work by Tony Hisgett)
Notable species of Halobacteria include Halobacterium salinarum, which may be the oldest living organism on earth; scientists have isolated its DNA from fossils that are 250 million years old. Another species, Haloferax volcanii, shows a very sophisticated system of ion exchange, which enables it to balance the concentration of salts at high temperatures.
MICRO CONNECTIONS
Finding a Link Between Archaea and Disease
Archaea are not known to cause any disease in humans, animals, plants, bacteria, or in other archaea. Although this makes sense for the extremophiles, not all archaea live in extreme environments. Many genera and species of Archaea are mesophiles, so they can live in human and animal microbiomes, although they rarely do. As we have learned, some methanogens exist in the human gastrointestinal tract. Yet we have no reliable evidence pointing to any archaean as the causative agent of any human disease.
Still, scientists have attempted to find links between human disease and archaea. For example, in 2004, Lepp et al. presented evidence that an archaean called Methanobrevibacter oralis inhabits the gums of patients with periodontal disease. The authors suggested that the activity of these methanogens causes the disease. However, it was subsequently shown that there was no causal relationship between M. oralis and periodontitis. It seems more likely that periodontal disease causes an enlargement of anaerobic regions in the mouth that are subsequently populated by M. oralis.
There remains no good answer as to why archaea do not seem to be pathogenic, but scientists continue to speculate and hope to find the answer.
Source: CNX OpenStax
Additional Materials (5)
Archaebacteria
Diagrammatic view of Methanobrevibacter smithii, showing the cell membrane (ochre, with inset) and cell wall (purple).
Image by K. Gottlieb, V. Wacher, J. Sliman and M. Pimentel
Structure of Prokaryotes
Archaeal phospholipids differ from those found in Bacteria and Eukarya in two ways. First, they have branched phytanyl sidechains instead of linear ones. Second, an ether bond instead of an ester bond connects the lipid to the glycerol.
Image by CNX Openstax
Organizing Life on Earth
Both of these phylogenetic trees shows the relationship of the three domains of life—Bacteria, Archaea, and Eukarya—but the (a) rooted tree attempts to identify when various species diverged from a common ancestor while the (b) unrooted tree does not. (credit a: modification of work by Eric Gaba)
Image by CNX Openstax
Themes and Concepts of Biology
This phylogenetic tree was constructed by microbiologist Carl Woese using data obtained from sequencing ribosomal RNA genes. The tree shows the separation of living organisms into three domains: Bacteria, Archaea, and Eukarya. Bacteria and Archaea are prokaryotes, single-celled organisms lacking intracellular organelles. (credit: Eric Gaba; NASA Astrobiology Institute)
Image by CNX Openstax (credit: Eric Gaba; NASA Astrobiology Institute)
Send this HealthJournal to your friends or across your social medias.
Bacteria
Bacteria are small single-celled organisms. The human body is full of bacteria, and in fact is estimated to contain more bacterial cells than human cells. Most bacteria in the body are harmless, and some are even helpful. A relatively small number of species cause disease.