Cell Structure and Function

Published on 02/03/2015 by admin

Filed under Basic Science

Last modified 22/04/2025

Print this page

rate 1 star rate 2 star rate 3 star rate 4 star rate 5 star
Your rating: none, Average: 4.5 (4 votes)

This article have been viewed 10239 times

Cell Structure and Function

WHY YOU NEED TO KNOW

HISTORY

Aristotle (384–322 bce), Paracelsus (1493–1541), and other ancient scientists and philosophers realized the existence of a continuity in the macroscopic parts of plants and animals. They observed structures that kept repeating themselves from animal to animal, from plant to plant, and from generation to generation. Centuries later with the advent of magnifying lenses and compound microscopes built by Robert Hooke (1635–1703) and Jan Swammerdam (1637–1680), descriptions of objects magnified ×20–×30 were possible. Then van Leeuwenhoek (1632–1723), using his skills at lens grinding and the use of light, improved the resolving powers of the microscope to ×200. In his letters to the newly formed Royal Society of London, he described “very little living animalcules very prettily a-moving” in samples taken from his own and other mouths. These are among the first descriptions of living bacteria. Other firsts for van Leeuwenhoek with his microscopic observations were the discovery of blood cells; the circulation in the capillaries of eels; and the first observations of living animal sperm cells, nematodes, and rotifers to mention a few. Later, Claude Bernard (1813–1878) described the cell as the primary representative of life. The cell is the basic unit of life, whether in single-unit organisms, as in protozoa, or in multicellular organisms, where cells group and differentiate into tissues and organs.

IMPACT

The arrival of the concept of functions for these microscopic structures marks the beginning of the study of the cell as an independent unit—cytology with its subsequent disciplines such as cytochemistry and cytogenetics. Moreover, the integrity of a cell as a functional unit becomes essential because if disease or physical disruption occurs to severely compromise or destroy it, the cell dies even though some of its functions such as enzyme activity remain. This persistence of cellular function allows the study of cellular contents from cell-free preparations in which cellular integrity is intentionally compromised. Moreover, the cell changes in response to its environment and the demands for its products. It is a chemical factory, the output or products of which are essential not only for the tissue and organism, but for the maintenance and repair of the cell itself. The cell doctrine—life exists in cellular form whether singular or multicellular; each cell can exist on its own; and all cells come from preexisting cells—guides scientific exploration of the role of cells in living organisms. Science is dependent on advances in technology to afford new ways of looking at new and at old persistent problems.

FUTURE

One of the recent developments in science focuses on progenitor pluripotent cells that differentiate to form the tissues of the human body. Embryonic stem cells form along the inside of blastocysts and create an inner cell mass during growth of the embryo. After being harvested, they grow on a layer of feeder cells in a culture medium. Potentially, these cells can develop into muscle, pancreatic, hepatic, nerve, and practically any other cell line. Uses of these new discoveries include replacement of tissues damaged, diseased, infected, maldeveloped, or otherwise undesirable and development of antimicrobial and other drugs. Because it is an entirely new and untested area of exploration, there is the potential for unknown consequences, which makes their use controversial. The understanding of cell structure and function is essential for human survival. Microorganisms, small though they are, can easily destroy human life. If science can proceed rationally with adequate suitable controls and limits, then as always, the future is bright. Without these qualifications the future, as always, may hold many problems.

General Structure of Prokaryotic and Eukaryotic Cells

Cells are the basic structural and functional unit of all living organisms. The name cell comes from the Latin word “cella,” a small room. Robert Hooke (see Chapter 1, Scope of Microbiology) selected this term in 1655, when he discovered cells in a piece of cork with his microscope, and compared the cork cells with small rooms. In 1838, Schleiden and Schwann proposed a formal cell theory, or cell doctrine, that marked the start of modern cell biology. The current cell theory states:

The major events leading to the knowledge of cell biology today are shown in Table 3.1.

TABLE 3.1

Major Events in Cell Biology

Year Name Event
1655 Robert Hooke Observes cells of a cork
1674 Antony van Leeuwenhoek Discovers protozoans
1683 Antony van Leeuwenhoek Discovers bacteria
1838 Schleiden and Schwann Propose the cell theory
1855 Carl Naegeli and Carl Cramer Describe the cell membrane as a barrier essential to explain osmosis
1857 Rudolph Albert von Kölliker Discovers mitochondria in muscle cells and describes them as conspicuous granules aligned between the striated myofibrils of muscle
1890 Richard Altman Develops mitochondrial stain and postulates genetic autonomy
1898 Carl Benda Develops crystal violet as mitochondria-specific stain and coins the name
1882 Robert Koch Identifies tuberculosis- and cholera-causing bacteria with aniline dyes
1898 Camillo Golgi Discovers the Golgi apparatus with a silver nitrate stain (Golgi stain)
1931 Ernst Ruska Builds the first transmission electron microscope
1965 Cambridge Instrument Company First commercial scanning electron microscope
1997 Roslin Institute, Scotland Sheep cloned

Two different types of cells exist, prokaryotes and eukaryotes, which are structurally and functionally different from one another (Table 3.2). However, both cell types share certain properties, and these are as follows:

TABLE 3.2

Comparison of Prokaryotic and Eukaryotic Cells

Characteristic Prokaryotic Cell Human Eukaryotic Cell
Size 0.2–60 µm 5–100 µm
Chromosome One Multiple
Cell membrane Yes Yes
Cell wall Yes Yes (prokaryotes)
No (eukaryotes)
Nucleus No Yes (except red blood cells)
Nucleoid area Yes No
Mitochondria No Yes
Endoplasmic reticulum No Yes
Golgi apparatus No Yes
Cytoskeleton No Yes
Ribosome 70S 80S
Mode of reproduction Asexual Asexual and sexual

Plasma Membrane and Cell Wall

In order to survive, all living cells need to be protected from the continuous changes of their surrounding environment. This is achieved by the plasma membrane and in some types of cells by an additional structure, a cell wall.

Plasma Membrane

All living cells, both prokaryotes and eukaryotes, are surrounded by a plasma membrane, about 8 nm (nanometer) thick, creating an intracellular and extracellular environment/compartment. The intracellular and extracellular fluid compartments (ICF and ECF, respectively) are aqueous and the plasma membrane provides the barrier between them. The membrane is composed primarily of phospholipids and proteins. Phospholipids are polar molecules with a polar (hydrophilic) head and a nonpolar (hydrophobic) tail (see Chapter 2, Chemistry of Life). Because the environment on each side of the plasma membrane is aqueous, the hydrophobic tails face each other and the hydrophilic heads are exposed to water on both surfaces of the membrane. This results in the formation of a phospholipid bilayer, illustrated in Figure 3.1.

Embedded in the phospholipid bilayer are peripheral and integral (transmembrane) proteins and cholesterol. The peripheral proteins are partially embedded on one side of the membrane, whereas integral proteins extend from one side through the membrane to the other side. Although the specific lipid and protein components vary in different membranes, the basic structure and composition of the plasma membrane are the same, as are those membranes that surround cell organelles. Because the plasma membrane is not solid, its components are freely movable, presenting a constantly changing fluid-mosaic membrane structure (see Figure 3.1, A).

Whereas the phospholipids are the main lipid component of the cell membrane, cholesterol is another major cellular membrane lipid and the amount varies with the type of plasma membrane. For example, the plasma membrane of animal cells contains almost one cholesterol molecule per phospholipid molecule, plant cells contain much less, and prokaryotic cells have no cholesterol but instead contain sterol-like molecules called hopanoids. Cholesterol and hopanoids (pentacyclic sterol–like compounds) immobilize the first few hydrocarbon groups of the phospholipid molecules. This makes the lipid bilayer less elastic and decreases permeability to small water-soluble molecules.

Glycolipids, also a component of membranes, project into the extracellular space. Functionally, these may protect, insulate, and serve as receptor-binding sites. In addition, the outer membrane of gram-negative cell walls (see Chapter 6, Bacteria and Archaea) contains lipopolysaccharide bacterial endotoxins that are released on lysis of these cells.

The functions of membrane proteins differ depending on their composition and location in the plasma membrane. They can:

Glycocalyx

Many cells have a matrix formed outside the plasma membrane. This extracellular fabric is composed of polymeric material called a glycocalyx, produced by some bacteria, epithelia, and other cells. In general the glycocalyx is a network of polysaccharides that project from the cellular surface and functions in/as:

In the case of bacteria the glycocalyx is a coating of macromolecules that protects the cell and sometimes helps the bacteria to adhere to its environment. The chemical composition, thickness, and organization of the glycocalyx differ among bacteria. One specialized function of the bacterial glycocalyx is the formation of capsules (Figure 3.2). Capsules protect pathogens, such as Streptococcus pneumoniae and Bacillus anthracis, from phagocytosis by white blood cells, thus adding to their pathogenicity.

Cell Wall

A cell wall is located immediately below the glycocalyx surrounding the plasma membranes. Bacteria, archaea, fungi, plants, and algae can all have cell walls of different chemical composition. The purpose of the cell wall is to maintain the shape of a cell and to protect the cell from any physical or chemical damage.

The bacterial cell wall is a unique rigid structure that maintains the shape of bacteria and protects them from hostile environments, including protection from the immune system of a host (see Chapter 20, The Immune System). The strength of the cell wall is due to the presence of peptidoglycan (mucopeptide or murein), a mixed polymer of hexose sugars (N-acetylglucosamine and N-acetylmuramic acid) cross-linked by short peptide fragments (Figure 3.3). The amount and composition of peptidoglycan vary among the major bacterial groups and provide the basis for Gram staining (see Chapter 4, Microbiological Laboratory Techniques), dividing bacteria into so-called gram-positive and gram-negative organisms.

However, some bacteria do not have a characteristic cell wall structure and some lack a cell wall altogether (i.e., Mycobacterium; see Chapter 6, Bacteria and Archaea). Differences in cell wall composition are an important consideration in the selection of antimicrobial drugs in treatment regimens of bacterial diseases. The cell wall structure is also important in choosing specific methods to control microbial growth in a variety of environments (see Chapter 19, Physical and Chemical Methods of Control).

Characteristics of cell walls in the different microorganisms are as follows:

Gram-positive bacteria have a thick peptidoglycan layer (20–80 nm) located external to the cell membrane (Figure 3.4, A). It also contains acidic polysaccharides such as teichoic acid and lipoteichoic acid, which aid in cell wall maintenance and contribute to an acidic cell surface. The small space between the plasma membrane and the cell wall is the periplasmic space, which can be minimal in some of the gram-positive bacteria.

Gram-negative bacteria have a thin (5- to 10-nm) peptidoglycan layer that is more complex because it has an outer membrane that provides a cover that is anchored to the lipoprotein molecules of the peptidoglycan layer (Figure 3.4, B). The outer membrane is similar in structure to the plasma membrane, but it contains lipopolysaccharides extending from its surface. These lipopolysaccharides can function as receptors or as antigens. In addition, the membrane contains porin proteins, which allow penetration only of small molecules. This serves as a defense mechanism against larger molecules such as antibiotics. The periplasmic space between the plasma membrane and the outer membrane may constitute up to 40% of the total cell volume. This space houses biochemical pathways for nutrient acquisition, peptidoglycan synthesis, electron transport, and detoxification of substances otherwise harmful to the cell.

Surface Appendages

Surface appendages are present in both prokaryotic and eukaryotic cells. Prokaryotic cells can have pili (fimbriae) and flagella, whereas cilia, flagella, and microvilli are common in eukaryotic cells. The functions of surface appendages include motility, attachment, absorption, and sensory capacity (Table 3.3).

TABLE 3.3

Surface Appendages of Prokaryotic and Eukaryotic Cells

Surface Appendage Cell Type Composition Function
Flagella Eukaryotic cell Nine pairs of peripheral microtubules made of protein tubulin + two central microtubules Motility through whiplike action
  Prokaryotic cell Single microtubule composed of flagellin subunits arranged in a helical formation around a hollow core Motility through propeller-like rotation
Pili Prokaryotic cell Pilin proteins helically arranged around a hollow core. Similar to flagella but shorter and more rigid

Cilia Eukaryotic cell Similar arrangement to flagella but much shorter

Microvilli Eukaryotic cell Extensions of the plasma membrane

image

Bacterial Flagella

Bacterial flagella are long (3 to 12 micrometers [µm]) helical filaments about 12 to 30 nm in diameter, which they use for movement in their environment. The protein components are flagellins and are assembled to form a cylindrical structure with a hollow core. A flagellum consists of three parts (Figure 3.5):

Flagellins are immunogenic and represent a group of antigens called the H antigens. These antigens are characteristic of a species, strain, or variant of an organism. The species specificity of the flagellins is the result of differences in the primary structures of the proteins (see Chapter 2, Chemistry of Life). The flagella may be present at the poles of the cells as a single polar structure at one end of the bacterium (monotrichous), or at each end (amphitrichous). Flagella also may be present in tufts at one or both ends of the organism, in which case such organisms are lophotrichous. If the flagella are distributed over the general cell surface, they are peritrichous (Figure 3.6).

Bacterial Mobility

The thrust by which the flagella propel the bacterial cell is produced by either clockwise or counterclockwise rotation of the basal body driven by energy from a proton-motive force rather than directly from adenosine-5′-triphosphate (ATP). Bacterial flagella do not rotate at a constant speed; the speed of rotation depends on the cell’s generation of energy. Bacteria can change speed and direction of their flagellar movements and as a result are capable of various patterns of motility. Movements are referred to as runs (swims) or tumbles. A bacterium moving in one direction for a period of time is said to run, and when the direction changes the bacterium tumbles (Figure 3.7). Any movement of bacteria toward or away from a particular stimulus is called taxis. As bacteria run, they may show properties of chemotaxis (response to chemical stimuli) or phototaxis (response to light). Both require sophisticated sensory receptors located in the cell surface and periplasm.

Not all prokaryotes have flagella, but some are still capable of movement by the process of gliding. In general, gliding prokaryotes are filamentous or rod shaped in contact with a solid surface (e.g., filamentous cyanobacteria).

Pili

Pili (fimbriae) are another form of bacterial surface projection (Figure 3.8) and are more rigid than flagella. They are composed of pilin proteins. In some organisms, such as Shigella species and Escherichia coli, as many as 200 pili are distributed over a single cell surface. Pili come in two types: short, abundant common pili, and a smaller number (one to six) of very long sex pili. Sex pili attach male to female bacteria during conjugation. Pili of many enteric bacteria have adhesive properties that can attach to various epithelial surfaces and to surfaces of yeast and fungal cells. These adhesive properties of piliated bacteria play an important role as factors in the colonization of epithelial surfaces. Adhesion to host cells is dependent on specific interactions between the adhesins and molecules in the host cell membranes. For example, the adhesins of E. coli chemically interact with molecules of mannose (a sugar monomer) on the surface of intestinal epithelial cells.

Cilia

Cilia and eukaryotic flagella are structurally identical hairlike appendages, about 0.25 µm in diameter, made from microtubules (Figure 3.9). They are motile and either move the cell or move substances over or around the cell. Protozoans use cilia to collect food particles and for locomotion (see Chapter 8, Eukaryotic Microorganisms). The primary action of cilia in most animal species is to move fluid, mucus, or cells over their surface. The major difference between cilia and eukaryotic flagella is their length, which is much greater for flagella. Cilia can either be motile, constantly beating in one direction, or nonmotile, nonbeating sensors. Motile cilia rarely occur singly and they are usually present in large numbers on cell surfaces, beating in coordinated waves. Nonmotile cilia, on the other hand, generally are single structures per cell.

Biofilms

A biofilm is a microbial community enclosed by an extracellular, mostly polysaccharide polymeric matrix (Figure 3.11). Noncellular materials such as mineral crystals, corrosive particles, blood, and other substances can also be found in the biofilm matrix. The kind of inclusions depends on the environment where a biofilm develops. Biofilms develop when free-floating microorganisms attach to a surface. The first colony initially adheres to a given surface by van der Waals forces (see Chapter 2, Chemistry of Life). If not immediately separated from that surface, the organisms can anchor themselves more permanently via adhesion molecules such as pili. Once adhered these microbes then provide various adhesion sites for other new incoming cells, resulting in the building of a matrix that holds the biofilm together. Microbial species that cannot attach to certain surfaces on their own can often attach to the biofilm matrix or to earlier colonized organisms. The biofilm then grows through a combination of cell division and recruitment. The cells on the surface of the biofilm are different from the ones within the matrix and the behavior of the embedded cells may change within the thickness of the biofilm.

Biofilms can form on a wide variety of surfaces including living tissues, indwelling medical devices and catheters, industrial or clean water system piping, natural aquatic systems, and rocks and pebbles, and on the surface of stagnant pools of water. Moreover, biofilms are important components of the food chains formed by the living matter of rivers and streams.

Microbial biofilms can cause equipment damage and therefore can be the source of product contamination and microbial infections. They do represent a concern in the food industry because they can arise from raw materials, surfaces, people, animals, and even the air. Once food or a surface in a food-processing plant is contaminated, microbial colonies and eventually biofilms can form. Cleaning materials commercially used to clean countertops will kill planktonic or single cells of bacteria, but might be unable to penetrate a biofilm. For example, chlorination of a biofilm is often unsuccessful because the agent only affects the outer layers, whereas the inner layers of the biofilm remain healthy and are able to reestablish the biofilm.

Biofilms also provide an ideal environment for the exchange of plasmids (extrachromosomal bacterial DNA; see Chapter 6). Certain plasmids encode resistance to antimicrobial agents and their inclusion within a biofilm provides an opportunity to promote and spread bacterial resistance to antimicrobial agents. Repeated use of antimicrobial agents actually increases the possibility of resistance to biocides (see Chapter 19, Physical and Chemical Methods of Control).

On the other hand, microbial processes at the surface of the biofilms can give opportunities for positive industrial and environmental effects. Biofilms can be used in:

Research investigations exploring new uses for biofilm processes now include scientists from a variety of disciplines such as engineers, microbiologists, biochemists, computer scientists, statisticians, and physicists, to mention a few.

MEDICAL HIGHLIGHTS

Biofilms and Infectious Diseases

According to the National Institutes of Health, over 80% of microbial infections in the body are related to biofilms. These include urinary tract infections, catheter infections, middle ear infections, formation of dental plaque, gingivitis, infections resulting from coating of contact lenses, infections of the peritoneal membrane and peritoneal devices, and even lethal conditions such as endocarditis and infections from permanent indwelling joint prostheses and heart valve devices.

Also, bacteria embedded in biofilms have a markedly increased resistance to antimicrobial drugs. This resistance can be 1000-fold higher than the same bacteria exhibit under planktonic conditions. These bacteria also resist the body’s nonspecific and specific pathogenic defense mechanisms (see Chapter 20, The Immune System). Because the immune response is generally directed only against the antigens present on the outer surface, antibodies often fail to penetrate to the deeper layers of the biofilms. Furthermore, phagocytic cells are unable to adequately engulf bacteria that grow within the complex biofilm matrix attached to a solid surface. Subsequently, the phagocytes release amounts of proinflammatory enzymes and cytokines resulting in inflammation and destruction of nearby tissue.

Technological advances in lasers, digital imaging, scanning electron microscopy, and new fluorescent probes will be used to assist scientists of different disciplines in developing new strategies for the prevention and treatment of microbial biofilm–derived diseases.

Cytoplasm, Cytoskeleton, Cell Organelles, and Inclusions

Other than a plasma membrane, cell wall, appendages, and glycocalyx, all of which play a specific role in the functioning and protection of cells, specific chemical processes (metabolism) occur within the cell. All metabolic processes in a cell are aided either by cell organelles or other structures and chemicals within the matrix of a cell.

Cytoskeleton

It was previously thought that the cytoskeleton was present only in eukaryotic cells, but more recent research shows that homologs of cytoskeletal proteins present in eukaryotic cells are also found in prokaryotes. Speculations indicate that constituents of today’s eukaryotic skeleton (tubulin and actin) evolved from prokaryotic precursor proteins closely related to today’s bacterial proteins. The studies indicate that the existence of a shape-preserving cytoskeleton is universally present in bacteria.

Structurally, the cytoskeleton (Figure 3.12) is not rigid but is dynamic and capable of rapid reorganization. This is essential for intracellular transport of vesicles and large molecules, as well as for the rearrangement of organelles. Furthermore, the cytoskeleton forms a spindle apparatus during cell division, which is discussed later in this chapter. The cytoskeleton is composed of actin filaments, intermediate filaments, and microtubules:

• Actin filaments (microfilaments) are made from the globular protein actin, which polymerizes in a helical fashion to form microfilaments. They are about 7 nm in diameter and provide mechanical support for the cell, determine the cell shape, and enable cell movements.

• Intermediate filaments are 8 to 11 nm in diameter and are the more stable components of the cytoskeleton. They organize the internal tridimensional structure of the cell. Intermediate filament proteins include keratins, vimentins, neurofilaments, and lamins.

• Microtubules are hollow cylinders with a diameter of about 24 nm and varying lengths, made of tubulin proteins. They serve as structural components within cells and take part in many cellular processes such as mitosis, cytokinesis, and vesicular transport.

Centrioles and centrosomes are also part of the cytoskeleton and are essential during cell division in eukaryotic cells. A centriole is a barrel-shaped microtubular structure present in most animal cells and cells of fungi and algae, but not frequent in plants. When two centrioles are arranged perpendicularly and are surrounded by additional proteins, they form the centrosome. The centrosome is known as the microtubule-organizing center and plays an important role in the microtubule organization of a cell. During the process of cell division centrioles form the mitotic spindle on which the chromosomes pull apart.

Nucleus

The nucleus (Figure 3.13) is the control center of the eukaryotic cell, containing the DNA. A double membrane, the nuclear envelope, surrounds the nucleus, interrupted by nuclear pores. The nuclear envelope regulates molecular transport between the nucleus and the cytoplasm. The outer membrane of the nuclear envelope is continuous with the cell’s endoplasmic reticulum (ER). Within the nucleus is a prominent structure, the nucleolus, that synthesizes ribosomal RNA (rRNA) and combines the rRNA with proteins to form the subunits of ribosomes.

Also visible in the nucleus is chromatin composed of DNA and its protein complex. Two different forms of chromatin exist in a eukaryotic cell: euchromatin and heterochromatin. Euchromatin is the least compact form of DNA and contains the genes expressed in a given cell (i.e., insulin gene in pancreatic cells). Heterochromatin is more tightly packed and it contains genes not expressed in that particular cell (i.e., thyroid hormone in pancreatic cells).

Endoplasmic Reticulum

The endoplasmic reticulum is part of the endomembrane system of eukaryotic cells. It consists of an elaborate network of membranous tubules and cisternae, continuous with the outer membrane of the nuclear envelope and held together by the cytoskeleton (Figure 3.14). The two forms of ER are rough ER (rER) and smooth ER (sER).

Golgi Apparatus

The Golgi apparatus (also called the Golgi complex) consists of flattened, disc-shaped sacs called cisternae with somewhat dilated peripheries that give rise to various secretory vesicles and lysosomes (Figure 3.15). The Golgi apparatus is always closely associated with the ER, the production site of organic molecules. In preparation for export, the Golgi complex modifies and packages the proteins, carbohydrates, and lipids that it receives from the ER. In addition, the Golgi apparatus produces lysosomes for the use within the cell.

Lysosomes

Lysosomes are membrane-bound organelles that pinch off from the Golgi apparatus. Lysosomes contain hydrolytic (digestive) enzymes for intracellular digestion of cell debris and food particles, as well as for the breakdown of invading microorganisms. Another function of lysosomes is to process products of receptor-mediated endocytosis (see Endocytosis, later in this chapter). Lysosomes can be classified into primary and secondary lysosomes, and residual bodies. The size and shape of lysosomes vary with the degree of degradative activity.

Whenever rupture of lysosomal membranes occurs, resulting in the subsequent release of their enzymes, cell death occurs.

Mitochondria

Mitochondria are the cell’s power source for energy, without which none of the cellular activities could proceed. Their double-membrane structure is distinct among cell organelles. The outer membrane encloses the entire organelle and has a phospholipid-to-protein ratio similar to that of the plasma membrane and an inner membrane that has folds reaching inward to form numerous cristae (Figure 3.17). This organelle contains proteins for different functions, including those that carry out the reactions of the respiratory chain for ATP production. The mitochondrial matrix is the space enclosed by the inner membrane and contains a mixture of hundreds of enzymes, mitochondrial ribosomes, transfer RNA (tRNA), and several copies of the mitochondrial DNA genome. The mitochondrial genome encodes primarily proteins for oxidative phosphorylation, as well as rRNAs, tRNAs, and proteins involved in protein synthesis. Mitochondrial ribosomes are of the 70S type (also found in bacteria), distinct from the 80S ribosomes found elsewhere in the eukaryotic cell.

Mitochondria replicate their DNA and divide primarily in response to the energy demands of the cell. Therefore, growth and division of mitochondria depend on the energy needs of cells and not on the cell cycle. They divide by binary fission similar to bacterial cell division. Mitochondrial DNA is different from the DNA of nuclear genes. Mitochondrial DNA comes from the egg because the single mitochondrion of the sperm is usually destroyed as the sperm enters the egg at fertilization.

Chloroplasts

Chloroplasts are organelles found in algae and plant cells and are able to convert sun energy into chemical energy through the process of photosynthesis (see Photosynthesis later in this chapter). Chloroplasts are spherical or oval, measuring about 3 to 8 µm in diameter, surrounded by inner membrane system and outer membranes (Figure 3.18). The primary components of the inner membrane system are the thylakoids which consist of flattened disks stacked in parallel arrangements. They contain pigments such as chlorophyll and the enzymes necessary for photosynthesis. The stroma is the fluid within the chloroplast and contains small circular DNA, containing the chloroplast genome, and ribosomes. Chloroplast genes code for photosynthesis and autotrophy. Other genes in the chloroplast genome encode rRNA used in chloroplast ribosomes, tRNA used in translation, some proteins used in transcription and translation, and some other proteins. However, some proteins necessary for chloroplast function are encoded in the nuclear genes.

Ribosomes

Ribosomes are cell organelles common to both prokaryotic and eukaryotic cells. They consist of two subunits of ribosomal RNA and ribosomal proteins (Figure 3.19). The subunits fit together and operate in unison with tRNA to translate mRNA into a polypeptide chain during protein synthesis. Ribosomes in eukaryotic cells either attach to the ER, identifying it as rough ER, or float freely in the cytoplasm. Free ribosomes produce proteins for maintenance and repair within the cell or in the organelle in which they occur (i.e., mitochondria). The ribosomes in eukaryotic cells are of the 80S type, whereas prokaryotic cells have 70S ribosomes such as are found in mitochondria and chloroplasts.

Vacuoles

Vacuoles are membrane-bound structures in eukaryotic cells that serve a variety of functions. In general, vacuoles in animal cells are smaller than those of plants. Vacuoles store a variety of fluids, soluble materials, and ingested particles. Many freshwater protozoans contain contractive vacuoles that pump excess water out of the cell. Mature plant cells have a single central vacuole surrounded by a membrane called the tonoplast (Figure 3.20). Here the central vacuole often takes up more than 80% of the cell interior, leaving the rest of the cytoplasm confined in the space between the tonoplast and the plasma membrane. The vacuole plays a major role in the growth of plant cells, in that the cells elongate as the vacuole absorbs water to increase cell size with minimal increase in cytoplasm. Also, some vacuoles contain pigments that give color to the cells.

Vesicles

A vesicle is a relatively small (typically less than 1 µm), membrane-bound compartment used by the cell in the organization of metabolism, in transport, for enzyme storage, and as a chamber for chemical reactions. They can be formed by the Golgi apparatus, endoplasmic reticulum, or the plasma membrane. Examples are as follows:

HEALTHCARE APPLICATION
Some Toxin-producing Bacteria That Cause Damage to Eukaryotic Cells

Organism/Toxin Target Damage Disease
Aeromonas hydrophila/aerolysin Glycophorin Plasma membrane Diarrhea
Clostridium perfringens/perfringolysin O Cholesterol Plasma membrane Gas gangrene
Escherichia coli/hemolysins Plasma membrane Plasma membrane Urinary tract infections
Staphylococcus aureus/α-toxin Plasma membrane Plasma membrane Abscesses
Streptococcus pneumoniae/pneumolysin Cholesterol Plasma membrane Pneumonia
Streptococcus pyogenes/streptolysin O Cholesterol Plasma membrane Strep throat
Corynebacterium diphtheriae/diphtheria toxin Elongation factor-2 Protein synthesis Diphtheria
E. coli, Shigella dysenteriae/Shiga toxins 28S rRNA Protein synthesis Hemorrhagic colitis; hemolytic uremic syndrome

image

Fluid Compartments and Membrane Transport Mechanisms

As previously stated, all living cells are separated from their external environment by plasma membranes and some organisms additionally by their cell walls. Prokaryotic cells and unicellular (single-cell) organisms are in direct contact with the external environment, whereas multicellular organisms form tissues, organs, and organ systems that are not in immediate contact with this environment. For a cell to survive, it must maintain a responsive, controllable, relatively stable, internal environment called homeostasis. In the human body cells are in contact with the extracellular fluid (ECF), which collectively is called the extracellular fluid compartment, or ECF compartment. The fluid inside the cell is the intracellular fluid (ICF), which collectively forms the intracellular fluid compartment or ICF compartment. In a 70-kg human under ideal conditions, the ideal ICF compartment volume is about 28 L and the ECF compartment is about 12 L. These are the volumes of fluid in which invading pathogens and antimicrobial drug treatments are distributed and in which effective concentration levels are reached to cause a disease or therapeutic levels for treatment are obtained. In order to understand the development of and the impact of infectious diseases on the body and drug treatments it is essential to have an understanding of the composition of the body’s fluid compartments. Furthermore, it is essential to know the cellular transport mechanisms required to maintain a dynamic balanced distribution of fluids and electrolytes (homeostasis).

Intracellular Fluid Compartment

ICF, the largest compartment, encompasses two thirds of the human body’s water. Normal values for body water, expressed as a percentage of total body weight, vary between 45% and 75%. These differences are due to age, fat content of the body, and gender. Adipose (fat) tissue contains the least amount of water among body tissues and therefore obese people have less body water per kilogram of weight than do slender ones. In contrast, muscle tissue contains the largest amount of water (65%). Old age is accompanied by loss of muscle mass and increases in body fat resulting in significant decreases in body water content, sometimes up to 45% of total body weight. The distribution of water in the body’s compartments is shown in Figure 3.21. Intracellular fluid has high concentrations of potassium, phosphate, and magnesium ions, and lesser amounts of sodium, chloride, and bicarbonate ions.

Extracellular Fluid Compartment

The ECF consists of plasma, lymph, the interstitial fluid (fluid surrounding the cells), and the transcellular fluid such as cerebrospinal fluid (CSF), joint fluids, humors of the eye, digestive juices, and mucus, to mention a few. The extracellular fluids have high concentrations of sodium, chloride, and bicarbonate ions, with lesser amounts of potassium, calcium, magnesium, phosphate, and sulfate ions.

Osmotic pressure markedly influences the movement of water and electrolytes throughout the body’s fluid compartments. Whereas the composition of the ECF and ICF differ, the total solute concentration and water amounts are normally equivalent. Thus, net gains of water or of solutes osmotically cause shifts affecting balances in the intracellular and extracellular fluids. These shifts can have a marked influence on uptake, redistribution, or removal of substances, including antibiotics or other drugs and chemicals (see Chapter 21, Pharmacology).

Passive Transport

In passive transport, molecules move from an area of higher concentration to an area of lower concentration without the use of cellular energy (ATP). The driving force of passive transport is atomic and molecular (Brownian) movement—the natural tendency of atoms and molecules to be in constant random movement. This motion can be demonstrated by the Brownian movement of small particles suspended in a liquid. For example, a few crystals of potassium permanganate put into a beaker of water and left undisturbed will slowly color the water in the entire beaker by the process of diffusion.

Facilitated Diffusion

Facilitated diffusion is a process necessary for molecules that cannot readily diffuse through cellular membrane barriers and require membrane transporters for movement across the membrane. As with diffusion, this type of transport does not require energy from ATP. The necessary transport molecules can either be specific membrane carriers or transmembrane channels. Carrier-mediated transport (Figure 3.23):

Protein channel–mediated membrane transport provides selective pathways for specific ions or other small water-soluble molecules. For example, sodium ions can only pass through sodium channels and potassium can only pass through potassium channels. Furthermore, these channels are gated, that is, they are either open or closed. Gated channels can be triggered to open or close by voltage, light, mechanical, or chemical stimuli, and they are named according to the stimuli activating them (i.e., sodium channels, voltage-gated channels, pressure-gated channels, etc.)

Osmosis

Osmosis is the diffusion of water across a selectively permeable membrane. The plasma membrane of cells is selectively permeable to water molecules, allowing them to diffuse but excluding substances dissolved in the water. When the concentration of water on one side of the plasma membrane is greater than that on the other side, water will follow its concentration gradient and diffuse to the area of lower water concentration (Figure 3.24). In living cells, however, isotonicity between the ICF and ECF avoids excessive movement of water either into or out of the cells (see Figure 2.11 in Chapter 2, Chemistry of Life). Although water is the solvent molecule that moves across the membrane during osmosis, the solute concentrations of the ECF and ICF dictate how much water is available for diffusion. For example, a 30% solution contains 30 parts of solute and 70 parts of water and a 3% solution contains 3 parts of solute and 97 parts of water. When these two solutions are placed on the opposite sides of a selectively permeable membrane, water will diffuse from an area of higher concentration of water molecules (the 3% solution) to an area of lower concentration of water molecules (the 30% solution).

Moreover, whereas isotonic conditions do not create stress on cells, hypotonic or hypertonic environments do. Some specific microbes adapt to different osmotic environments as illustrated by amoeba living in a freshwater pond, which is a hypotonic environment. In this case water continuously enters the cell through the plasma membrane. In order to prevent the eventual rupture of the cell, water is actively pumped out of the cell with the help of the contractile vacuole by the process of active transport.

Active Transport

The driving force for passive transport mechanisms is the concentration gradient. In active transport cellular energy (ATP) is required to move molecules uphill, against a concentration gradient. Active transport is also required when molecules need to be transported at a faster rate than is possible by diffusion, even if the molecules move with their concentration gradient. For example, some freshwater algae have an active transport system that is so efficient that nutrients are in much higher intracellular concentration than in the surrounding environment. Active transport mechanisms include transport by pumps or by vesicles such as occurs with endocytosis and exocytosis.

Pump Transport

Active transport of ions or molecules is due primarily to a pump transport mechanism working against a gradient. This work requires energy obtained by hydrolysis of ATP. Active transport carriers are composed of transmembrane proteins. The pump transport process requires:

Active pumping systems are important for cells because they allow the cells to move needed ions and other molecules to specific areas. The calcium pump, for example, moves calcium ions to specific areas within the muscle cell during muscle contraction and relaxation. The sodium–potassium pump removes three sodium ions (Na+) out of a cell while it transports two potassium ions (K+) into the cell. Both ions are transported against their chemical concentration gradient, which maintains an electrical charge on the cell membrane, referred to as membrane potential. Most cells contain many Na+ /K+ pumps that are constantly active to maintain the membrane potential. The steep gradient of sodium and potassium ion concentration across cell membranes is essential for conduction and transmission of electrochemical impulses in nerve and muscle cells. If the active extrusion of Na+ is blocked or markedly impeded, the increased Na+ concentration in the cell promotes osmosis, a challenge to isotonic conditions that could result in cell damage, including rupture of the membrane or crenation (see Figure 2.11 in Chapter 2, Chemistry of Life).

Endocytosis

Extracellular material may be brought into the cell by endocytosis, a bulk transport mechanism that uses vesicles. Some eukaryotic cells transport large molecules, particles, liquids, or other cells across the cell membrane via this process. Typically, a cell surrounds and encloses a particle with its membrane, forming an endocytotic vesicle, and then engulfs it. Whole cells or large particles of solid matter are brought into the cell by the process of phagocytosis (“cell eating”; Figure 3.25, A) (see Chapter 20, The Immune System). Liquids, or molecules dissolved in liquids, are transported into the cell by pinocytosis (“cell drinking”; Figure 3.25, B). If it is necessary for a molecule to bind to a membrane receptor before endocytosis can occur, it is called receptor-mediated endocytosis (Figure 3.25, C).

Exocytosis

Large molecules, such as polypeptides, proteins, and others, may be excreted from the cell via the process of exocytosis (Figure 3.26). Unicellular organisms (i.e., protozoans) use exocytosis for the elimination of waste products. In multicellular organisms exocytosis has a signaling or regulatory function. Vesicles for exocytosis are produced by the Golgi apparatus, from which they are transported to the cell membrane. Here the vesicles fuse with the cell membrane and their contents are emptied into the extracellular space or into a capillary. In addition to its transport function, exocytosis also adds material to the plasma membrane.

Cellular Metabolism

Cellular metabolism includes all chemical reactions within a cell, which are organized in sequences called metabolic pathways. Metabolic pathways that break down large molecules into smaller ones and release energy in the process are collectively part of catabolism. The pathways that produce larger molecules from smaller ones are part of anabolism and use energy released during catabolic reactions (also see Chapter 2, Chemistry of Life). Nutrients or sunlight supply the energy required to generate ATP molecules. Organisms that utilize energy from the breakdown of nutrient molecules are chemotrophs; those that use sunlight for photosynthesis are phototrophs and they release energy during the process. According to the first law of thermodynamics, energy is neither created nor destroyed; it is only transferred from one form into another. This includes energy for cellular metabolism and energy in general.

Enzymes

Enzymes are biological catalysts that initiate the chemical reactions necessary within the metabolic pathways of cells. All chemical reactions within a cell require enzymes. The mechanism of enzyme action is to lower the energy of activation needed to start a chemical reaction (Figure 3.27). Chemically, enzymes are proteins with a specific structure, usually in a tertiary or quaternary configuration (see Chapter 2, Chemistry of Life). The substances that enzymes act on are referred to as substrates, and the end result of the reaction is a product. Expressed in chemical notation:

< ?xml:namespace prefix = "mml" />E+S=P

image

(Enzyme+Substrate=Product)

image

The site to which substrates bind to an enzyme is called the active site and is specific for each substrate. After binding an enzyme–substrate complex is formed and the chemical reaction occurs resulting in a product. At the end of the reaction the enzyme itself is unchanged (Figure 3.28).

Classification and Naming of Enzymes

Enzymes are classified according to the type of reaction they catalyze and all metabolic reactions are catalyzed by a specific class of enzyme. Thus, individual enzymes are typically named according to the type of reaction they catalyze, the name of the substrate they act on, and using the suffix -ase.

Cofactors and Coenzymes

In many cases the enzymes require nonprotein coenzymes or cofactors, that is, other ions or molecules, in order to catalyze a reaction. The protein portion of an enzyme is the apoenzyme and the nonprotein portion is a coenzyme if it is an organic molecule (usually a derivative from water-soluble vitamins); if it is a metal ion, it is a cofactor. Holoenzymes are enzymes that may require one or more cofactors or coenzymes when combined (Figure 3.29).

Microorganisms require specific metal ions as trace elements and certain organic growth factors for survival (see Chapter 6, Bacteria and Archaea). The need for these substances occurs because of their roles as cofactors. Enzymes that require cofactors do not have an appropriate shape for the active site until they combine with the cofactor. Metals, therefore, activate enzymes by bringing the active site and the substrate closer together and then participate directly in the chemical reaction as a part of the enzyme–substrate complex. Coenzymes, on the other hand, remove a functional group from one substrate molecule and add it to another substrate. Coenzymes often function as intermediate carriers of hydrogen atoms, electrons, carbon dioxide, and amino groups. An example would be the role of nicotinamide adenine dinucleotide (NAD) in the transfer of electrons in coupled oxidation–reduction reactions (see Cellular Respiration, later in this chapter). Because coenzymes are vitamin derivatives, it might explain the importance of vitamins in our diet; a vitamin deficiency prevents the completion of the holoenzyme and the chemical reaction to be catalyzed fails to occur.

Regulation of Enzyme Activity

Rates at which enzyme reactions proceed within the metabolic pathways are influenced by temperature; pH; concentrations of substrate, enzyme, and product; and the presence or absence of cofactors and coenzymes.

• Temperature: Rates of most chemical reactions generally speed up at higher temperatures, and slow down at lower temperatures. However, high temperatures risk denaturing (changing the shape of) the protein portion of enzymes, causing them to lose specificity and functionality (see Chapter 2, Chemistry of Life).

• pH: Enzymes work best within an optimal pH range to promote the most rapid chemical reaction. For the majority of microbial enzymes, this is near pH 7 (neutral).

• Substrate concentration: If the amount of enzyme is fixed, then high concentrations of substrate speed the rate of reaction until the active sites on the enzymes saturate with substrate.

• Enzyme concentration: If the amount of substrate is constant, then the reaction rate initially rises but levels off as the substrate available for the reaction binds to the enzyme.

• Product concentration: Because many enzymes catalyze reactions in both forward and reverse directions, high concentrations may allow the enzymes to catalyze the reverse reaction. However, if the substrates are used in other reactions, then the direction of the first enzyme reaction remains forward.

• Cofactors and coenzymes: Depending on the presence (in sufficient numbers) or absence of cofactors and coenzymes, the rates of enzyme reactions will either increase or slow.

In addition, a variety of other substances such as drugs can either inhibit or reduce the rate of a reaction. The effectiveness of such an inhibitory substance depends on its reversible or irreversible binding properties with the enzyme. For example, in irreversible inhibition, the inhibitor competes with the substrate for the active site of the enzyme (Figure 3.30, A). The degree of inhibition is dependent on the relative concentrations of the inhibitor and the substrate; increasing the amount of substrate may result in reversible inhibition. Competitive enzyme inhibition is the basis for treatment of some infectious diseases by selectively inhibiting a specific metabolic pathway required for the pathogen but not necessary for the human. Antimetabolites function this way, and in so doing they provide an effective weapon against infectious diseases.

MEDICAL HIGHLIGHTS

Sulfa Drugs (Sulfonamides)

Sulfonamides are synthetic antimicrobial agents with a wide spectrum of activity (see Chapter 21 [Pharmacology] and Chapter 22 [Antimicrobial Drugs]) against both gram-positive and gram-negative bacteria. These drugs are an example of metabolic antagonism. All cells require folic acid for growth. Folic acid (vitamins in food) can be transported across human cells but cannot cross bacterial cell walls, and therefore must be synthesized by the bacteria. Sulfonamides are similar to and compete with para-aminobenzoic acid (PABA), a necessary intermediate in bacterial synthesis of folate—thereby killing bacteria.

Another mechanism that influences the rate of enzyme reactions is noncompetitive inhibition, in which the substance binds to an allosteric site of an enzyme, causing a conformational change of the enzyme’s active site (Figure 3.30, B). The process is irreversible and the degree of inhibition is dependent on the concentration of the inhibitor. Also, enzyme activity is often regulated at specific points of the metabolic pathways by the process of end-product inhibition—when the amount of product formed is sufficient the reaction shuts down. This is negative feedback inhibition and is an example of allosteric inhibition. In other words, in end-product inhibition, the inhibiting substance is the product formed by the initial enzyme activity. The amount of product formed then feeds back to slow down the reaction and prevent excessive accumulation of more final product (Figure 3.31).

Cellular Respiration and Photosynthesis

Prokaryotic and eukaryotic cells metabolize nutrients to generate ATP by respiration (cellular metabolic pathways), fermentation, and photosynthesis. Because ATP contains high-energy phosphate bonds for use in cellular respiration, it is a vital source of energy, and microorganisms utilize more than one process to produce it.

Cellular Respiration

Cellular respiration is the process by which the chemical energy of nutrient molecules is released and captured in the form of ATP. This energy transfer involves a series of oxidation–reduction reactions. In the breakdown of glucose and other nutrient molecules, some of the electrons originally present in these molecules are transferred to intermediate carriers and then to a final electron acceptor. When the breakdown is complete, with the end product being carbon dioxide and water, and oxygen the final electron acceptor, the pathway used is aerobic cellular respiration. If oxygen is not available cellular respiration continues with other inorganic molecules acting as the final electron acceptors, and this process is known as anaerobic cellular respiration or fermentation.

Aerobic cellular respiration requires three consecutive pathways: glycolysis, the Krebs cycle (citric acid cycle), and the electron transport chain (oxidative phosphorylation). The overall reaction for aerobic cellular respiration of glucose is as follows:

C6H12O6+6O6CO2+6H2O+energy

image

Glycolysis, the first pathway of the series, occurs in the cytoplasm and uses 10 enzyme reactions to break down glucose to produce a net yield of 2 molecules of pyruvic acid, 2 molecules of ATP, and 2 molecules of reduced NAD (Figure 3.32). The sequence of the principal steps of glycolysis that form ATP is as follows:

1. Phosphorylation (addition of a phosphate group to a molecule) of glucose to yield glucose 6-phosphate. Phosphorylated organic molecules do not cross the plasma membrane and glucose now remains in the cell.

2. Conversion of glucose 6-phosphate into fructose 6-phosphate.

3. Use of another ATP molecule to phosphorylate fructose 6-phosphate into fructose 1,6-diphosphate.

4. Splitting of fructose 1,6-diphosphate into two three-carbon molecules of 3-phosphoglyceraldehyde.

5. Conversion of each molecule of 3-phosphoglyceraldehyde to 1,3-biphosphoglyceric acid (coenzyme NAD picks up hydrogen from each 3-phosphoglyceraldehyde molecule and forms NADH to which inorganic phosphate [Pi] is added).

6. Biphosphoglyceric acid donates a high-energy phosphate to ADP via substrate-level phosphorylation, forming an ATP molecule. The product of this reaction is 3-phosphoglyceric acid.

7. Isomerization (rearrangement of the chemicals in a given compound) of 3-phosphoglyceric acid produces 2-phosphoglyceric acid.

8. The removal of a water molecule from 2-phosphoglyceric acid produces phosphoenolpyruvic acid.

9. In the final step of glycolysis a phosphate group is removed from each phosphoenolpyruvic acid molecule and donated to ADP, forming another ATP molecule. The product of this reaction is pyruvic acid.

The overall equation for glycolysis is as follows:

Glucose+2NAD+2ATP+2Pi2 pyruvic acid+2NADH+2ATP

image

This is followed by:

Pyruvic acidacetyl-CoA+CO2

image

Glycolysis is connected to the Krebs cycle by converting pyruvic acid into acetyl-coenzyme A (acetyl-CoA) and carbon dioxide (CO2) while transferring two hydrogen ions to NADH. The NADH formed during this reaction will be sent to the electron transport chain. Each acetyl-CoA molecule will enter the Krebs cycle.

The Krebs cycle is the second pathway in aerobic cellular respiration. It occurs in the plasma membrane of prokaryotes and within the matrix of mitochondria in eukaryotes. After acetyl-CoA enters the Krebs cycle it combines with oxaloacetic acid to form citric acid. Through a series of reactions involving the elimination of two carbons and four oxygens (two CO2 molecules) and the removal of hydrogens, citric acid is eventually converted to oxaloacetic acid to complete this metabolic cycle (Figure 3.33).

Steps in the Krebs cycle include the following:

For each glucose molecule entering aerobic respiration, the Krebs cycle runs twice because glycolysis produced two molecules of pyruvic acid. Molecules of NAD and FAD in glycolysis and the Krebs cycle are essential to pick up hydrogens and electrons. The reduced forms of these molecules then enter the electron transport chain, where they are oxidized in a series of reduction–oxidation reactions to regenerate coenzymes while producing ATP, oxygen, and water.

The electron transport chain, the last step in aerobic cellular respiration, takes place in the cristae of the inner mitochondrial membrane. In order to regenerate NAD+ and FAD+, electrons and hydrogen ions are transferred from reduced NAD and FAD molecules to carrier molecules in the electron transport chain. These molecules include flavoproteins, coenzyme Q, iron–sulfur enzymes, and cytochromes. The last carrier transfers the electrons to oxygen, the final electron acceptor of aerobic respiration. As the hydrogen ions and electrons are transferred in a stepwise fashion from NADH to the chain of carriers, energy is released. This energy is captured by ADP to generate many ATP molecules from each glucose molecule (Figure 3.34).

The breakdown of one glucose molecule in aerobic respiration is summarized in Table 3.4:

TABLE 3.4

Summary of Aerobic Cellular Respiration per Glucose Molecule

Produced Glycolysis Krebs Cycle Respiratory Chain Net Total
ATP 4 (2 used) 2 34 38
NADH 2 8 0 10
FADH 0 2 0 2
CO2 6 0 0 6
H2O 2 2 used 6 6

image

The total of 38 ATP molecules produced is theoretical, because the actual ATP yield is less in eukaryotic cells because of the energy expended in transporting NADH across the mitochondrial membrane. However, there are some aerobic bacteria that come close to achieving this theoretical total because they lack mitochondria and do not have to use ATP to transport NADH across the mitochondrial membrane.

The phosphogluconate pathway (also called the pentose phosphate pathway or hexose monophosphate shunt) is an alternative catabolic pathway followed by some bacteria (Figure 3.35). This pathway generates NADPH and synthesizes pentose sugars in two distinct phases. The first is the oxidative phase, which generates NADPH, and the second phase synthesizes pentose sugars. This is a common pathway for heterolactic fermentative bacteria. It yields various end products including lactic acid, ethanol, and carbon dioxide. In addition, this pathway is a significant source of pentose sugars during nucleic acid synthesis.

Anaerobic cellular respiration is a metabolic pathway by which glucose is converted to lactic acid if oxygen is not available. Some bacteria utilize anaerobic respiration (anaerobes) for energy production and the final electron acceptor is an inorganic molecule and not oxygen. Actually, oxygen is toxic to anaerobic microbes and exposure to it kills most of such bacteria. On the other hand, facultatively anaerobic bacteria can use anaerobic respiration when oxygen is absent or present in limited amounts only. The final electron acceptor of anaerobic bacteria can be CO2, the ion SO42–, or NO3 .

If the final electron acceptor is an organic molecule, then fermentation takes place. The process starts with glycolysis but the end product varies because of the variety of organic molecules that can act as final electron acceptor. Fermentative bacteria and yeast use specific enzymes to metabolize the end product of glycolysis, namely pyruvic acid, into the final product. In this process an electron transport system is not used and reduced NAD from glycolysis is oxidized by an organic molecule accepting electrons and hydrogens. With this pathway, the cell can catabolize sugar without oxygen, but this process is far less efficient than aerobic cellular respiration for the generation of ATP.

Fermentation end products produced by microorganisms can be beneficial to human life and various industries (i.e., dairy and brewing). Examples of bacteria and their fermentation end products are as follows:

Photosynthesis

Photosynthesis is a fundamental biochemical process by which plants, most algae, cyanobacteria, and some phototrophic bacteria convert light energy into chemical energy via ATP. Chlorophyll, a pigment found in chloroplasts (Figure 3.36, A) of plants and algae, is the site of photosynthesis. Photosynthetic bacteria do not have chloroplasts and photosynthesis takes place directly within the cell. Cyanobacteria contain thylakoid membranes (Figure 3.36, B) similar to those in chloroplasts and these are the only prokaryotes that perform oxygen-generating photosynthesis. Other photosynthetic bacteria contain a variety of different pigments, called bacteriochlorophylls, but they do not produce oxygen. Photosynthesis requires water for oxidation but some bacteria oxidize hydrogen sulfide instead and produce sulfur as a waste product. The rate of photosynthesis is affected by carbon dioxide, light intensity, and temperature.

Processes in photosynthesis are divided into two categories: reactions that require light, and reactions that occur in the dark and are light-independent. A simplified general equation for photosynthesis is as follows:

6CO2+12H2O+lightC6H12O6+6O2+6H2O

image

(Carbon dioxide+water+light energyglucose+oxygen+water)

image

The first stage of photosynthesis is a light-dependent reaction that requires solar energy that is converted to chemical energy. Chlorophyll absorbs light and drives a transfer of electrons and hydrogen from water to an acceptor called NADP+ (nicotinamide adenine dinucleotide phosphate), which stores the energized electrons for a short period of time. During this process water is split and oxygen gas is given off. The process of producing ATP from sunlight is called photophosphorylation. There are two forms of photophosphorylation: noncyclic photophosphorylation (Figure 3.37, A) and cyclic photophosphorylation (Figure 3.37, B).

Noncyclic photophosphorylation, the predominant route, includes two sets of pigments called photosystem I (PSI) and photosystem II (PSII). These pigments are sensitive to or are excited by slightly different wavelengths of light. Wavelengths of about 700 nm excite the chlorophyll in PSI (P700) whereas the chlorophyll of PSII (P680) is excited by wavelengths under 680 nm. The stages of noncyclic photophosphorylation are as follows:

1. Oxidation of chlorophyll in PSII by light causes an electron to rise to a higher energy level. This electron is detained by the primary electron acceptor. This oxidized P680 chlorophyll now has a “hole” that needs filling.

2. Electrons are enzymatically removed from water and transferred to P680 to replace the electrons lost when light was absorbed by chlorophyll. A water molecule is then split into two hydrogen ions and an oxygen atom. The oxygen atom combines with another one to form O2.

3. Each excited electron from the primary electron acceptor of PSII moves to PSI through the electron transport chain.

4. The energy released by electrons during the cascade down the electron transport chain is captured by the thylakoid membrane. This energy is used to produce ATP for synthesis of sugar during the second stage of photosynthesis—the Calvin cycle.

5. The electron “hole” in P700 of PSI, created by light energy driving an electron to the primary acceptor of PSI, is filled.

6. Electrons are passed by the primary acceptor to the iron-containing protein ferredoxin (Fd) in a second electron transport chain. The enzyme NADP+ reductase transfers the electrons from Fd to NADP+. This reduction reaction stores the electrons in NADPH, which is the molecule that will also provide energy for the synthesis of sugar in the Calvin cycle.

Cyclic photophosphorylation occurs when the photo-excited electrons utilize photosystem I as an alternative path. The electrons cycle back from Fd to the cytochrome complex and then to the P700 chlorophyll. This cyclic system does generate ATP but not NADPH and no oxygen is released.

The second stage of photosynthesis, the light-independent reactions or dark reactions, takes place in the stroma within the chloroplast. The enzyme RuBisCO (ribulose-1,5-biphosphate carboxylase/oxygenase) captures CO2 from the air and releases three-carbon sugars in a complex process called the Calvin-Benson cycle. The Calvin-Benson cycle is similar to the Krebs cycle because the starting material regenerates after molecules enter and leave the cycle. Carbon enters the cycle in the form of CO2 and leaves in the form of sugar. The energy source for the cycle is ATP, and NADPH is the reducing power for adding high-energy electrons to make three-carbon sugars (glyceraldehyde 3-phosphate) that later combine to form glucose. The Calvin-Benson cycle is organized into three phases: carbon fixation, reduction, and regeneration of the CO2 receptor (Figure 3.38).

Protein Synthesis

Cellular structure and function are dependent on proteins (see Chapter 2, Chemistry of Life). Amino acids are the monomers of proteins and 20 amino acids are the building blocks in the synthesis of the different polypeptide chains. The codes for the amino acid sequence in polypeptides and proteins lie in the genes of DNA and must be copied to direct their synthesis. In order for a gene to be expressed a copy must be made from the DNA template. This copying process is called transcription and the subsequent production of a particular amino acid sequence in a polypeptide chain is called translation.

Transcription

Prokaryotic transcription occurs in the cytoplasm and eukaryotic transcription takes place in the nucleus. The description or central dogma of the flow of information from genetic material was first introduced by Francis Crick in 1958 (Figure 3.39). It states that DNA transfers information to RNA and RNA then controls protein synthesis. Or, DNA makes RNA and RNA makes protein. DNA also controls its own replication (see DNA Replication, later in this chapter). The structures of DNA and RNA are discussed in Chapter 2.

The initiation of transcription is controlled by the enzyme RNA polymerase, which recognizes the beginning of a gene so that it can start mRNA synthesis at a particular DNA sequence that appears at the beginning of genes. This sequence is called a promoter. The promoter is a unidirectional sequence on one strand of the DNA and tells the RNA polymerase where to start and in which direction to continue synthesis. RNA polymerase stretches open the double helix at that point and then begins the synthesis of mRNA. This synthesis of mRNA follows the principle of complementary base pairing (see Chapter 2). Termination of transcription occurs when the polymerase recognizes a DNA sequence known as a terminator sequence. In prokaryotic cells ribosomes can begin protein synthesis with the mRNA immediately, whereas in eukaryotic cells the mRNA must leave the nucleus before it combines with the ribosomes in the cytoplasm. As stated earlier in this chapter, eukaryotic ribosomes may be free in the cytoplasm, or attached to rER.

Each mRNA molecule contains several hundred or more nucleotides complementary to the DNA template. Every three bases form a base triplet or codon and each codon is the code for a specific amino acid. DNA triplets, RNA codons, and their amino acid translation are shown in Table 3.5. However, some amino acids can be translated by different codon sequences as indicated in Table 3.6. As mRNA moves through the ribosome, the sequence of codons translates into a sequence of amino acids.

TABLE 3.5

DNA Base Triplets, mRNA Codons, and Their Amino Acids

DNA Triplet mRNA Codon Amino Acid
TAC AUG “Start” (methionine)
ATC UAG “Stop”
CGT GCA Alanine
GCT CGA Arginine
ACA UGU Cysteine
CCG GGC Glycine
CTC GAG Glutamic acid
GTG CAC Histidine
GAA CUU Leucine
AAA UUU Phenylalanine
GGG CCC Proline
AGG UCC Serine
ACC UGG Tryptophan

TABLE 3.6

Amino Acids Encoded by Multiple mRNA Codons

Amino Acid mRNA Codons
Alanine GCA GCC GCG GCU
Arginine AGA AGG CGA CGC CGG CGU
Asparagine AAC AAU
Aspartic acid GAC GAU
Cysteine UGC UGU
Glutamic acid GAA GAG
Glutamine CAA CAG
Glycine GGA GGC GGG GGU
Histidine CAC CAU
Isoleucine AUA AUC AUU
Leucine UUA UUG CUA CUC CUG CUU
Lysine AAA AAG
Methionine AUG
Phenylalanine UUC UUU
Proline CCA CCC CCG CCU
Serine AGC AGU UCA UCC UCG UCU
Threonine ACA ACC ACG ACU
Tryptophan UGG
Tyrosine UAC UAU
Valine GUA GUC GUG GUU

Translation

Specific enzymes and tRNA translate codons. Like other RNA molecules, tRNA is a single-stranded molecule; but it bends back in on itself to form a cloverleaf structure with an anticodon on one end (Figure 3.40). An anticodon consists of three nucleotides complementary to a specific codon on the mRNA molecule. Specific enzymes (aminoacyl-tRNA synthetase enzymes) in the cytoplasm bind specific amino acids to the ends of tRNA; and a tRNA molecule with a given anticodon can bind to a specific amino acid. Each tRNA molecule is therefore bound to one specific amino acid.

The binding of the tRNA anticodons to the codons of the mRNA, as it moves through the ribosome, forms a polypeptide chain. Translation occurs in three stages: initiation, elongation, and termination.

• Initiation: The ribosome assembles on the identified start codon (AUG—methionine) in mRNA and then each sequential base pair forms the next codon. The position of the start codon determines the open reading frame or order of codons that will be read to form a protein.

• Elongation: The first and second tRNAs bring the first and second amino acids close together. The first amino acid detaches from its tRNA and forms a peptide bond with the amino acid of the neighboring second tRNA, forming a dipeptide. When the third tRNA binds to the third codon, its amino acid binds to the second amino acid, which then detaches from its tRNA. The polypeptide chain grows as new amino acids are added to this tripeptide by the same process. This growing polypeptide chain remains attached by only one tRNA to the strand of mRNA (Figure 3.41).

• Termination: Elongation of the designated protein continues until a “stop” codon (UAA, UGA, or UAG) signals the end of the process. The building of the protein stops because there is no tRNA molecule that is complementary to the stop codon. A releasing enzyme frees the newly formed polypeptide chain from the last tRNA and the mRNA is released from the ribosome.

DNA Replication and Cell Division

DNA Replication

When a cell prepares to divide, the DNA duplicates, and therefore each cell after cell division contains identical DNA. This process is called replication or DNA synthesis. In eukaryotic cells it occurs during the S phase of the cell cycle and thus precedes mitosis and meiosis (see later in this chapter). The two resulting double strands are identical; each of them consists of one original strand and one newly synthesized strand of the genetic material. This process is called semiconservative replication.

The unwinding of DNA occurs through the action of the enzyme helicase, which separates the two strands of DNA. The site of separation is the replication fork (Figure 3.42). DNA synthesis begins when a short complementary strand of RNA (primer RNA) binds to each parent strand of DNA. The primer RNA serves as the point of attachment for the enzyme DNA polymerase III, which initiates the synthesis of complementary DNA. As the DNA chain lengthens, DNA polymerase I replaces the primer RNA. When the complementary nucleotides are in position opposite each of the nucleotides on the parent strand, covalent bonds are formed between the sugar and phosphate groups of adjacent deoxyribonucleotides on the new chain. Hydrogen bonds form, joining the complementary base pairs between the parental DNA strand and new DNA strand. DNA polymerase can synthesize DNA only in the 5′ → 3′ direction. It is a bidirectional process during which one strand is copied continuously and the other discontinuously. The continuous strand of DNA is the leading strand and the discontinuous strand is the lagging strand. The lagging strand of DNA consists of short stretches of RNA primer plus newly synthesized DNA approximately 100 to 1000 bases long. These DNA fragments are Okazaki fragments that attach to the lagging strand by the action of the enzyme DNA ligase. Replication continues until the entire chromosome is duplicated, forming two identical DNA molecules.

Cell Division

The process by which a parental cell divides into two daughter cells is called cell division. It is the biological basis of life. For unicellular organisms such as protozoans, one cell division reproduces an entire organism. The different classes of cells use different processes for cell division. Prokaryotic cells, bacteria, divide by binary fission and eukaryotic cells divide by mitosis and meiosis.

Binary Fission

The form of asexual reproduction by which all bacteria and most protists reproduce is binary fission. A single cell separates into two identical daughter cells, each containing an identical copy of the parental DNA. This process occurs in a stepwise fashion starting with elongation of the bacterium. Next, the bacterial chromosomal DNA, which is a single circular molecule, replicates. After this, a central transverse septum forms that divides the cell into two daughter cells (Figure 3.43). This type of reproduction normally results in two identical cells. However, bacterial DNA has a relatively high mutation rate, which is a rapid rate of genetic change that is the underlying reason for the development of antibiotic-resistant bacterial strains.

Cell Cycle and Mitosis

Eukaryotic cells undergo a cell cycle or a sequential series of events between cell divisions. The frequency of cell divisions varies between different cell types, and some cells never divide after they differentiate (i.e., nerve cells and muscle cells). The cell cycle consists of four distinct phases: the G1 phase, S phase, G2 phase, and M phase. The G1 phase, S phase, and G2 phase are collectively known as the interphase or the nondividing stage of the cell cycle (Figure 3.44). The M phase is composed of two tightly coupled processes: mitosis and cytokinesis.

• During the G1 phase, cells carry out the metabolic activities characteristic of the tissue to which they belong. In preparation for cell division, a cell might increase the amount of cytoplasm and the number of cell organelles. Chromosomes are in their extended form, and their genes actively direct RNA synthesis. At the end of G1, the cell commits to dividing and has reached the “point of no return.” In yeast it is called START and in multicellular eukaryotes, it is the restriction point.

• During the S phase DNA duplicates.

• During the G2 phase the cell continues to grow and its metabolic activities prepare for mitosis. The quantity of DNA increases and the chromatin condenses to form short, thick structures that indicate the end of the phase. Now, each chromosome consists of two strands called chromatids joined together by a centromere (Figure 3.45). The two chromatids contain identical DNA base sequences because each is produced by semiconservative replication of DNA. Each chromatid becomes a separate chromosome after mitotic cell division.

• During the M phase, mitosis and cytokinesis occur. Mitosis is divided into four stages: prophase, metaphase, anaphase, and telophase (Figure 3.46). The visible characteristics of these phases are as follows:

• Prophase: The chromosomes become clearly visible by light microscopy. The centrioles migrate toward the opposite poles of the cell and each centrosome has spindle fibers extending from it. The nuclear membrane starts to disappear and the nucleolus is no longer visible.

• Metaphase: The chromosomes line up at the equator of the cell. Spindle fibers from each centriole attach to the centromeres of the chromosomes. The nuclear membrane is now absent.

• Anaphase: The centrosomes split and the sister chromatids separate as each is pulled to an opposite pole.

• Telophase: The centrosomes become longer, thinner, and less distinct. New nuclear membranes form and the nucleolus reappears. The cell membrane develops a distinct furrowing at the cell midline. Two separate nuclei are now apparent within the same cell.

• Cytokinesis: The cytoplasm divides to complete cell division.

Meiosis

Eukaryotic cells maintain a specific number of chromosomes characteristic for that species. The chromosome set exists either in a haploid state (single, unpaired) or in a diploid state (matched pair). Cells of most fungi, many algae, and some protozoans exist in the haploid state during most of their life cycle. Gametes of organisms that sexually reproduce are haploid. Cells of animals, plants, some protozoans, fungi, and algae are diploid throughout most of their life cycle. This state begins at fertilization when a female haploid gamete (egg) fuses with a male haploid gamete (sperm) to create a diploid offspring (haploid + haploid = diploid). Mitosis, however, maintains the normal chromosome number in all eukaryotic cells. Moreover, sexual reproduction in diploid organisms requires haploid chromosome numbers to produce diploid offspring. The reduction of the diploid chromosome number to a haploid state in gametes occurs during reduction division or meiosis.

In meiosis, a single duplication of chromosomes occurs and two cell divisions follow. This type of cell division is diploid → haploid in nature, and occurs only during the production of gametes, that is, sex cells. The diploid cells in the sex organs undergo two cell divisions to produce haploid sex cells. The stages of meiosis subdivide depending on whether they occur in the first or in the second meiotic cell division. These stages are prophase I, metaphase I, anaphase I, and telophase I; and prophase II, metaphase II, anaphase II, and telophase II (Figure 3.47).

It is important during meiosis that the genetic basis of the evolutionary process occurs. In other words, the significance of meiosis goes beyond the reduction of the chromosome number for sexual reproduction. First, during the lining up of the homologous pairs of chromosomes in metaphase I, each member of the pair comes from a different parent and randomly shuffles to end up in one of the daughter cells. Second, crossing-over or additional exchanges of parts of homologous chromosomes can occur in prophase I. In addition, this, together with the random lining up in metaphase I, results in genetic recombination to ensure that the gametes produced in meiosis are genetically unique. All of this provides the genetic diversity for sexually reproducing organisms that is needed to provide the characteristics that promote the survival of species over time—evolution.

Summary

• The cell is the basic structural and functional unit of all living organisms. Two main categories of cells exist: prokaryotes, the cells without a nucleus, and eukaryotes, cells with a membrane-bound nucleus.

• All cells are surrounded by a plasma membrane composed of a phospholipid bilayer and embedded proteins. Functionally, the plasma membrane is selectively permeable and controls the movement of particles in and out of the cell. Many cells have a matrix surrounding the plasma membrane, the multifunctional glycocalyx.

• A cell wall is present around the plasma membrane of many types of cells. The bacterial cell wall is a unique structure, which maintains the rigid shape of bacteria and protects them from hostile environments. On the basis of the composition of their cell wall, most bacteria when stained with the Gram stain, are either gram-positive or gram-negative. This classification provides a major tool for the identification of bacteria.

• Surface appendages are present in both prokaryotic and eukaryotic cells. Prokaryotic cells can have pili or flagella, whereas cilia, flagella, and microvilli are common in eukaryotic cells.

• Biofilms are collections of surface-associated microbes enclosed by an extracellular matrix. Biofilms can form on a wide variety of surfaces and can be the source of product contamination and microbial infections.

• Enclosed by the plasma membrane is the cytoplasm, a gelatinous matrix and the site for biochemical activities of a cell. Within the cytoplasm are the nucleus, other cell organelles, inclusions, and the supportive cytoskeleton.

• Prokaryotic cells and unicellular organisms are in direct contact with the external environment, whereas the tissues of multicellular organisms are in contact with extracellular fluid (ECF) collectively referred to as the extracellular fluid compartment. The fluid inside the cells is the intracellular fluid or ICF, which collectively is the intracellular fluid compartment.

• Cells exhibit passive and active transport mechanisms by which they can move materials across the plasma membrane. These mechanisms include diffusion, facilitated diffusion, osmosis, filtration, pump transport, endocytosis, and exocytosis.

• Cellular metabolism includes all chemical reactions in a cell and is organized into metabolic pathways. Metabolic pathways that break down molecules and release energy are part of catabolism, and pathways that produce larger molecules from smaller ones, using energy, are anabolic pathways.

• All metabolic pathways require biological catalysts, called enzymes, to initiate biochemical reactions. Enzymes are proteins that often require nonprotein compounds such as coenzymes or cofactors for action.

• Prokaryotic and eukaryotic cells metabolize nutrients to generate ATP by respiration, fermentation, and photosynthesis. Cellular respiration in the presence of oxygen is aerobic cellular respiration with glycolysis, the Krebs cycle, and the respiratory chain. In the absence of oxygen, cells utilize anaerobic cellular respiration including fermentation. Some cells and organisms can convert light energy into chemical energy via ATP, in a process called photosynthesis.

• Cellular structure and function depend on many different types of proteins. Proteins are composed of different sequences of amino acids and the amino acid sequence of a given protein is encoded in the cell’s DNA. The process of protein synthesis requires RNA and involves two processes: transcription and translation.

• Before a cell divides, it duplicates its DNA to ensure that each cell after division contains identical DNA. This process is called replication or DNA synthesis.

• The different classes of cells use different processes for cell division. Prokaryotic cells, that is, bacteria, divide by binary fission and eukaryotic cells divide by mitosis and meiosis.

Review Questions

1. Which of the following is not found in all bacterial cells?

2. Bacterial capsules are important in:

3. The cell organelles responsible for the packaging of proteins are the:

4. The cell organelle found only in algae and plant cells is the:

5. The intracellular fluid has a high concentration of:

6. Which of the following is an active transport mechanism?

7. Which of the following is an enzyme?

8. The compound that enters the Krebs cycle and combines with oxaloacetic acid is:

9. The anticodons are located in:

10. In which phase of mitosis do the chromatids separate?

11. Cells without a nucleus are __________ cells.

12. The sterol-like molecules in bacterial plasma membranes are __________.

13. A cell organelle that contains digestive enzymes is a(n) __________.

14. The allosteric site is present in __________.

15. The organelle necessary for photosynthesis is a(n) __________.

16. Compare and contrast prokaryotic and eukaryotic cells.

17. Describe the cell wall of gram-positive and gram-negative cells.

18. Discuss the regulation of enzyme activity.

19. Name and describe the pathways of aerobic cellular respiration.

20. Describe the cell cycle and the phases of mitosis.