signal&mito

24
6/4/2014 1/24 Histology & Cell Biology: Examination & Board Review, 5e > Chapter 2. The Plasma Membrane & Cytoplasm Objectives This chapter should help the student to: Perceive the inseparability of structure and function in living organisms. Know the names and functions of cytoplasmic components. Know the subunits of each cytoplasmic component and their roles in its function. List the cell's functions and explain the role of each cytoplasmic component in each function. Recognize a cell's cytoplasmic components in a micrograph and hence predict the cell's function(s). Predict which structures are present in a cell from its function. Predict the functional deficit(s) that accompany specific structural aberrations. Predict the cytoplasmic component(s) likely to be involved in a functional deficit. Explain and give examples of cell differentiation. MAX-Yield™ Study Questions 1. Compare prokaryotic and eukaryotic cells in terms of the presence of a nucleus, histones, and cytoplasmic organelles, as well as overall size (I.B). 2. List several functions of cell membranes (II.C). 3. List the major biochemical constituents of cell membranes and sketch their organization as described in the fluid mosaic model of membrane structure (II.A and B; Fig. 2–1). 4. Explain why a membrane's phospholipid bilayer has a trilaminar appearance in transmission electron micrographs (TEMs) (II.A.1). 5. Compare the locations of peripheral and integral membrane proteins in relation to the lipid bilayer and the methods required to isolate them from cell membranes (II.A.2.a and b; Fig. 2–1). 6. Name three membrane receptor types that transduce signals across the plasma membrane. Compare them in terms of the number of passes they make through the plasma membrane, the effects of ligand binding on their conformation, and the role of enzymes in signal transduction (II.C.2.a–c; Figs. 2–2, 2–3). 7. Explain how signal transduction by steroid hormone receptors differs from that by membrane receptors in terms of ligand type, receptor location, and binding between the receptor and DNA (II.C.2.a–d). 8. Compare phagocytosis and pinocytosis in terms of the way that vacuoles or vesicles are formed, the types of materials endocytosed, and the relative size of the vacuoles and vesicles (II.C.3.a and b). 9. List the steps in receptor-mediated endocytosis, beginning with ligand-receptor binding and ending with receptor return to the cell surface (II.C.3.c; Fig. 2–4). 10. Compare organelles and cytoplasmic inclusions in terms of the presence of limiting membranes, enzyme content, active or passive role in cell function, and their relative constancy in the cytoplasm (III).

Upload: richard-bradshaw

Post on 16-Jan-2016

4 views

Category:

Documents


0 download

DESCRIPTION

MITOCONDRIA

TRANSCRIPT

Page 1: Signal&Mito

6/4/2014

1/24

Histology & Cell Biology: Examination & Board Review, 5e >

Chapter 2. The Plasma Membrane & Cytoplasm

Objectives

This chapter should help the student to:

Perceive the inseparability of structure and function in living organisms.

Know the names and functions of cytoplasmic components.

Know the subunits of each cytoplasmic component and their roles in its function.

List the cell's functions and explain the role of each cytoplasmic component in each function.

Recognize a cell's cytoplasmic components in a micrograph and hence predict the cell's function(s).

Predict which structures are present in a cell from its function.

Predict the functional deficit(s) that accompany specific structural aberrations.

Predict the cytoplasmic component(s) likely to be involved in a functional deficit.

Explain and give examples of cell differentiation.

MAX-Yield™ Study Questions

1. Compare prokaryotic and eukaryotic cells in terms of the presence of a nucleus, histones, and cytoplasmic

organelles, as well as overall size (I.B).

2. List several functions of cell membranes (II.C).

3. List the major biochemical constituents of cell membranes and sketch their organization as described in the

fluid mosaic model of membrane structure (II.A and B; Fig. 2–1).

4. Explain why a membrane's phospholipid bilayer has a trilaminar appearance in transmission electron

micrographs (TEMs) (II.A.1).

5. Compare the locations of peripheral and integral membrane proteins in relation to the lipid bilayer and the

methods required to isolate them from cell membranes (II.A.2.a and b; Fig. 2–1).

6. Name three membrane receptor types that transduce signals across the plasma membrane. Compare them

in terms of the number of passes they make through the plasma membrane, the effects of ligand binding on

their conformation, and the role of enzymes in signal transduction (II.C.2.a–c; Figs. 2–2, 2–3).

7. Explain how signal transduction by steroid hormone receptors differs from that by membrane receptors in

terms of ligand type, receptor location, and binding between the receptor and DNA (II.C.2.a–d).

8. Compare phagocytosis and pinocytosis in terms of the way that vacuoles or vesicles are formed, the types of

materials endocytosed, and the relative size of the vacuoles and vesicles (II.C.3.a and b).

9. List the steps in receptor-mediated endocytosis, beginning with ligand-receptor binding and ending with

receptor return to the cell surface (II.C.3.c; Fig. 2–4).

10. Compare organelles and cytoplasmic inclusions in terms of the presence of limiting membranes, enzyme

content, active or passive role in cell function, and their relative constancy in the cytoplasm (III).

Page 2: Signal&Mito

6/4/2014

2/24

11. What is the major function of mitochondria (III.A.2)?

12. Sketch a mitochondrion (III.A; Fig. 2–5) and label or show the location of the following:

a. Outer mitochondrial membrane

b. Inner mitochondrial membrane

c. Cristae

d. Inner membrane (F1) subunits

e. Intermembrane space (III.A.1.c)

f. Intercristal space

g. Intracristal space

h. Matrix

i. Matrix granules

j. ATP synthase (III.A.1.b)

k. Citric acid cycle enzymes

l. Electron transport system (III.A.1.b)

13. Name the substances and structures in the mitochondrial matrix that resemble those in prokaryotic cells

and duplicate eukaryotic cell components found elsewhere in eukaryotic cells (III.A.1.d).

14. Compare the mitochondrial cristae of most cells with those of steroid-secreting cells and cells with a high

metabolic rate (III.A.1.b).

15. Briefly describe four major steps in oxidative phosphorylation (III.A.2).

16. What is the primary function of ribosomes (III.B)?

17. List the major biochemical and structural components of ribosomes, and name their sites of synthesis and

assembly (III.B.1).

18. Compare the ribosomes’ appearance in light and electron microscopy (III.B.1).

19. Explain how individual ribosomes assemble to form a polyribosome (III.B.2).

20. What is the difference between the functions of free polyribosomes and those of polyribosomes attached to

the rough endoplasmic reticulum (RER) (III.B.2)?

21. Compare the RER with smooth endoplasmic reticulum (SER) in terms of the presence of ribosomes

(III.C.1.a and 2.a), shape of cisternae (III.C.1.a and 2.a), functions (III.C.1.b and 2.b), and cell types in which

each is typically abundant; include examples (III.C.1.c and 2.c).

22. List the steps in RER-associated protein synthesis and posttranslational modification, beginning with

ribosome attachment to mRNAs for proteins destined for secretion and ending with the transfer vesicle budding

for transport to the Golgi complex (II.D; III.B.2, C.1.a and b).

23. List the functions of the Golgi complex (III.D.2).

24. Sketch a Golgi complex (III.D.1) and label the following:

a. Cisternae

b. Cis face

Page 3: Signal&Mito

6/4/2014

3/24

c. Trans face

d. Transfer vesicles

e. Condensing vacuoles

f. Secretory granules

g. Site of selective osmium deposition

25. Compare primary and secondary lysosomes in terms of size, appearance, and contents (III.F.1 and 2).

26. Compare primary lysosomes (III.F) and peroxisomes (III.G) in terms of size, content, appearance, and

function.

27. Compare autophagosomes with heterophagosomes (III.E).

28. Trace the steps in the ingestion and digestion of extracellular materials, beginning with endocytosis (II.C.3)

and ending with residual body formation (III.E and F.1–3).

29. Describe how an inherited deficiency or lack of a particular lysosomal enzyme affects the intracellular

concentrations of substrates of that enzyme (III.F).

30. Compare microfilaments and microtubules in terms of their:

a. Diameters (III.I.1.a and 2.a)

b. Major protein components (III.I.1.a and 2.a)

c. Functions (III.I.1.b and 2.b)

d. Subunit polymerization and depolymerization (III.I.1.a,b and 2.a,b)

e. Contractile capacity (III.I.1.b and 2.b)

f. Location in the cell (III.I.1.c and 2.c)

g. Associated motor proteins (III.I.1.b and 2.b)

31. Compare dynein and kinesin in terms of their typical direction of movement along microtubules and their

association with the RER and Golgi complex (III.I.1.b).

32. Explain the main function of intermediate filaments and give their diameter in nanometers (III.I.3.a and b).

33. List five intermediate filament protein types and the cell types in which each is found (III.I.3.a).

34. Compare centrioles, basal bodies, cilia, and flagella in terms of microtubule number and organization (III.J–

N).

35. Explain the roles of centrioles in cell function (III.N).

36. Compare cilia and flagella in terms of their structure, length, motion, and typical number per cell (III.J–L).

37. Explain the concept of membrane trafficking and the role of SNAREs and coat proteins in the process (II.D).

38. Explain the process of cellular differentiation (IV.A).

39. Explain the process of apoptosis (IV.D).

Synopsis

I. General Features of Cells

A. Subunits of Life

Page 4: Signal&Mito

6/4/2014

4/24

Cells are the structural and functional units of life (and of disease processes) in all tissues, organs, and organ

systems. Each cell's capabilities and limitations are implicit in its structure.

B. Prokaryotes and Eukaryotes

There are two basic cell types. Prokaryotes are small, single-celled organisms (e.g., bacteria) that lack a

nuclear envelope, histones, and membranous organelles. Eukaryotic cells exist primarily as components of

multicellular organisms. This chapter covers the basic structural and functional features of eukaryotic cells.

Specific human cell types are described in later chapters.

C. Cellular Components

Eukaryotic cells have three major components:

1. Cell membranes (II) separate a cell from its environment and form distinct functional compartments

(nucleus, organelles) in the cell. The outer cell membrane is called the plasma membrane, or

plasmalemma.

2. The cytoplasm (III) surrounds the nucleus and is enclosed by the plasma membrane. It contains the

structures and substances that decode the instructions of DNA and carry on the cell's activities.

3. The membrane-limited nucleus (Chapter 3) contains the DNA, which carries the genetic code for protein

synthesis and thus for all cell activities. It also has components that determine which parts of the genetic

code are used and that deliver coded information to the cytoplasm.

D. Cellular Functions

Three activities basic to living organisms are nourishment, growth, and reproduction. Functions directed toward

these activities are described in this chapter and in Chapter 3. More specialized cell functions receive detailed

treatment in subsequent chapters.

II. Cell Membranes

A. Biochemical Components (Fig. 2–1)

1. Lipids in cell membranes include phospholipids (Fig. 2–1, E), sphingolipids, and cholesterol (Fig. 2–

1, A). Phospholipids (e.g., lecithin) are the most abundant form. Each phospholipid molecule has a polar

(hydrophilic), phosphate-containing head group (Fig. 2–1, G) and a nonpolar (hydrophobic) pair of fatty

acid tails (Fig. 2–1, F). Membrane phospholipids are arranged in a bilayer, with their tails directed toward

one another at the center of the membrane. In electron micrographs (EMs) of osmium-stained tissue, a

single membrane, or unit membrane, has two dark outer lines with a lighter layer between them. This

trilaminar appearance reflects the deposition of reduced osmium on the hydrophilic head groups.

2. Proteins may contribute more than 50% of membrane weight. Most membrane proteins are globular and

belong to one of two groups:

a. Integral membrane proteins (Fig. 2–1, C and D) are tightly lodged in the lipid bilayer; detergents

are required to extract them. They are folded, with hydrophilic amino acids in contact with the

membrane phospholipids’ phosphate groups and hydrophobic amino acids in contact with the fatty

acid tails. Some protrude from only one membrane surface (Fig. 2–1, D). Others, called

transmembrane proteins (Fig. 2–1, C), penetrate the entire membrane and protrude from both

sides. Some transmembrane proteins, such as protein-3-tetramer, are hydrophilic channels for

the passage of water and water-soluble materials through hydrophobic regions. Some

transmembrane proteins pass multiple times through the bilayer to form channels and receptors.

Cryofracture preparations often split plasma membranes through the hydrophobic region, between

Page 5: Signal&Mito

6/4/2014

5/24

the ends of the phospholipids’ fatty acid tails (Fig. 2–1). Most integral proteins exposed in this way

remain in the side closest to the cytoplasm, termed the P (protoplasmic) face. The membrane half

nearest to the environment, the E (ectoplasmic) face, usually appears smoother.

b. Peripheral membrane proteins (Fig. 2–1, H) are ionically associated with the inner or outer

membrane surface and are released in high-salt solutions; some are globular, some filamentous. In

erythrocytes, examples on the cytoplasmic surface include adapter proteins like spectrin, which

helps maintain membrane integrity, and ankyrin, which links spectrin to protein-3-tetramer.

3. Carbohydrates occur on plasma membranes mainly as oligosaccharide moieties of glycoproteins (Fig. 2–

1, B) and glycolipids. Membrane oligosaccharides have a characteristic branching structure and project

from the cell's outer surface, forming a surface coat called the glycocalyx that participates in cell adhesion

and recognition.

B. Membrane Organization

The fluid mosaic model describes biologic membranes as “protein icebergs in a lipid sea”. Integral proteins

exhibit lateral mobility and may rearrange through their association with peripheral proteins, cytoskeletal

filaments within the cell (III.I), membrane components of adjacent cells, and extracellular matrix components.

Integral proteins may diffuse to and accumulate in one membrane region. Membrane asymmetry refers to

differences in chemical composition between the bilayer's inner and outer halves. Oligosaccharides occur only

on the plasma membrane's outer surface. Phospholipid asymmetries also occur. The outer half has more

phosphatidyl choline and sphingomyelin and the inner half has more phosphatidyl serine and phosphatidyl

ethanolamine.

C. Membrane Functions

1. Selective permeability. Cell membranes separate the internal and external environments of a cell or

organelle, preventing the intrusion of harmful substances, the dispersion of macromolecules, and the

dilution of enzymes and substrates. This selective permeability is essential for maintaining the functional

steady state, or homeostasis, required for cell survival. Homeostatic mechanisms attributable to cell

membranes maintain optimal intracellular concentrations of ions, water, enzymes, and substrates. Three

mechanisms allow selected molecules to cross membranes.

a. Passive diffusion. Some substances (e.g., water and lipids) can cross the membrane in either

direction following a concentration gradient, without the cell expending energy.

b. Facilitated diffusion. Some molecules (e.g., glucose) are helped across the membrane by a

membrane component. This facilitated diffusion is often unidirectional, but it follows a concentration

gradient and requires no energy.

c. Active transport. Some molecules enter or exit a cell against a gradient. This requires energy,

usually as adenosine triphosphate (ATP). One active transport mechanism is the sodium pump

(Na+/K+-ATPase), which expels sodium ions from a cell even when the sodium concentration is

higher outside than inside.

2. Signal transduction. Integral membrane receptor proteins with strong binding affinities for signal

molecules (e.g., neurotransmitters, peptide hormones, and growth factors) are found on cell surfaces. The

signal molecule to which a receptor specifically binds is its ligand. The receptor transduces the signal

across the membrane without the ligand entering the cell. These receptors are critical to intercellular

communication (IV.B). Signal transmission depends on the receptor class involved. There are three major

receptor classes.

Page 6: Signal&Mito

6/4/2014

6/24

a. Ligand-gated (e.g., transmitter-gated) ion channels are long proteins that pass multiple times

through the plasma membrane (Fig. 2–2A). The binding of a neurotransmitter (as ligand) to its

receptor domain at the cell surface induces a conformational change in the rest of the receptor that

opens (or closes) a transmembrane ion channel. Ligand binding and the resulting opening are

brief, permitting regulation. A signal of adequate strength (threshold) allows enough ions to cross

the membrane to open additional channels (voltage-gated ion channels) in response to the

change in potential. Reaching threshold allows a self-perpetuating membrane depolarization to

propagate as a wave along the nerve cell surface (9.VII.B.2).

b. Enzyme-linked receptors (Fig. 2–2B) comprise a heterogeneous group of transmembrane

(typically single-pass) proteins associated with an enzyme (typically a protein kinase) or possessing

kinase activity of their own (e.g., tyrosine kinase). Protein kinases activate additional proteins by

phosphorylating them (III.B.1.b). In this way, a ligand binding to its receptor initiates a cascade of

enzyme activations. Receptor tyrosine kinases (e.g., insulin and growth factor receptors)

often act by activating an adaptor protein, which subsequently activates the guanosine

triphosphate (GTP)-binding protein (G protein) Ras by causing it to exchange the guanosine

diphosphate (GDP) it binds in its inactive state for GTP. Activated Ras then activates a cascade of

cytoplasmic protein kinases, ending with a gene-regulatory protein that changes the pattern of

gene expression. Ras differs from the G proteins described in the next section in that it is

monomeric rather than trimeric. It resembles the α subunit in Figure 2–3 and acts without a β or γ

subunit.

c. G protein–coupled receptors (GPCRs) comprise a family of proteins that make seven passes

through the membrane (Fig. 2–3). Each receptor binds a different ligand, and the effect of signal

binding on the ultimate target protein (either an ion channel [Fig. 2–2A] or an enzyme [Fig. 2–2B])

is indirect and is mediated by a trimeric G protein (Fig. 2–3A). On binding its ligand (Fig. 2–3B), the

receptor interacts with and activates the trimeric G protein (Fig. 2–3C). The inactive G protein

carries a GDP molecule, which it trades for GTP on activation (Fig. 2–3D and E). The activated G

protein complex dissociates from the receptor, allowing the GTP-binding portion to interact with and

activate its target (Fig. 2–3F;), thus inactivating the G protein (i.e., GTP → GDP; Fig. 2–3F).

Ligand binding to GPCRs results in downstream activation of multiple signaling pathways involving

the generation of intracellular second messengers. Two key examples with different second

messengers are described below.

1. Cyclic adenosine monophosphate (cAMP). Some adrenalin receptors are key examples of

GPCRs that lead to cAMP production. Adrenalin binding activates membrane-bound adenylyl

cyclase through a G-protein mediated mechanism, causing the production of cAMP as the

second messenger. Among cAMP's primary downstream targets is protein kinase A (PKA),

which in turn activates a number of other target molecules including the transcription factors

CREB, CREM, and ATF-1.

2. Diacyl glycerol (DAG) and phosphatidylinositol triphosphate (IP3). Some serotonin,

calcitonin, and histamine receptors are GPCRs that lead to DAG and IP3 production. Ligand

binding activates membrane bound phospholipase C (PLC) through a G-protein mediated

mechanism. PLC hydrolyzes the membrane lipid, inositol bisphosphonate (IP2), generating

DAG and IP3.

Page 7: Signal&Mito

6/4/2014

7/24

a. DAG remains in the plasma membrane, facilitating the translocation of protein kinase C

(PKC) from the cytosol to the plasma membrane and serving as an activator of that

enzyme. PKC then activates other downstream target molecules including MARCK.

b. IP3 diffuses into the cytosol to bind to IP3-receptors in the smooth endoplasmic reticulum

(SER) membranes. IP3-receptors are ligand-gated calcium channels that open upon IP3

binding to allow the release of calcium in the SER into the cytosol. The resulting increase in

cytosolic calcium concentration induces the activation of many calcium-dependent

processes such as muscle contraction, membrane polarization, and enzyme activity.

d. Steroid hormone receptor family. Most peptide hormones and growth factors signal the cell

interior through cell-surface receptors and associated second messengers. Steroid hormones

(e.g., hydrocortisone and estrogen), retinoids (vitamin A–related compounds), and vitamin D pass

more readily through the membrane and bind to receptors in the nucleus. The thyroid hormone

receptor also belongs to this group (Because thyroid hormone consists mainly of the aromatic

amino, tyrosine, it passes readily through the plasma membrane). This family of nuclear receptors

is characterized by a specific ligand-binding site and a DNA-binding site. They typically form dimers

and activate or repress gene transcription through their direct association with response

elements in the DNA flanking the gene-coding regions

3. Endocytosis. Cells engulf extracellular substances and bring them into the cytoplasm in membrane-limited

vesicles by mechanisms described collectively as endocytosis.

a. In phagocytosis (cell eating), the cell engulfs insoluble substances, such as large macromolecules

or entire bacteria. The vesicles formed are termed phagosomes.

b. In pinocytosis (cell drinking), the cell engulfs small amounts of fluid, which may contain a variety of

solutes. Pinocytotic vesicles are smaller than phagosomes.

c. In receptor-mediated endocytosis (Fig. 2–4), a cell engulfs ligands along with their surface

receptors. The binding of ligand to receptor causes the ligand–receptor complexes to collect in a

coated pit, a shallow membrane depression whose cytoplasmic surface is covered with the coat

protein, clathrin. After further invagination, the protein, dynamin, coils around the neck of the

budding vesicle and pinches it off to create a coated vesicle, which carries the ligand–receptor

complexes into the cell. The clathrin coat is released from the vesicle, now termed an early

endosome, and the ligands dissociate from the receptors. The late endosome , or compartment

of uncoupling of receptor and ligand (CURL), becomes more tubular and divides into two portions,

segregating the receptors from the ligands. The portion with the receptors pinches off and returns

to fuse with the plasma membrane. The portion with the ligands fuses with a lysosome.

4. Exocytosis ejects substances from the cell. Cells use exocytosis both for secretion and for excretion of

undigested material. A membrane-limited vesicle or secretory granule fuses with the plasma membrane and

releases its contents into the extracellular space, without disrupting the plasma membrane.

5. Compartmentalization. Membranes selectively block the passage of most water-soluble substances. The

cytoplasm has many membrane-limited compartments (organelles), each with different internal solute

concentrations. This compartmentalization prevents the dilution of substrates, metabolic intermediates and

cofactors in multistep biochemical reactions, and protects sensitive reactions from the intrusion of

extraneous substances.

Page 8: Signal&Mito

6/4/2014

8/24

6. Spatial–temporal organization of metabolic processes. Some cell membranes (e.g., inner

mitochondrial membrane and Golgi complex) contain enzymes arranged in series so that intermediates in

multistep metabolic processes are passed from enzyme to enzyme. This arrangement maintains the

chronologic order of such processes and sets rate limits by maintaining local concentrations of

intermediates.

7. Storage, transport, and secretion. Substances in vesicles may be kept for later use (storage), shuttled

from one compartment to another for further processing (transport, II.D), or expelled from the cell

(secretion, II.C.4).

D. Membrane Flow (Traffic)

Membrane moves from one organelle to another. This general feature of organelle function allows homeostatic

regulation of membrane area. Membranes bud as vesicles from one organelle and fuse with another, allowing

the amount of membrane in a particular organelle to change without membrane synthesis or breakdown.

Vesicle budding is energetically facilitated by coat proteins (e.g., β-COP, clathrin) that vesicles acquire on their

cytoplasmic surfaces during budding. These coats help shape the bud and ensure its inclusion of molecules

that direct vesicle transport toward the proper destination. After a coated vesicle is formed, the coat is

released. The resulting transfer vesicle may diffuse to nearby targets (e.g., ER to Golgi) or attach to

microtubules for transport over longer distances (e.g., neuron cell body to axon terminal). Transmembrane

proteins in budding vesicle membranes, called v-SNAREs, remain after the coat is shed. They are critical in

allowing transfer vesicles to recognize, bind, fuse with, and unload their contents, in a GTP-dependent process,

at the correct target organelle. Receptors in the target organelle membranes, called t-SNAREs, selectively bind

the v-SNAREs and recruit other cytoplasmic proteins required for fusion with the target membrane.

FIGURE 2–1.

Schematic diagram of the biochemical components of plasma membranes (II.A). Labeled components include

cholesterol (A), the oligosaccharide moiety (B) of a glycoprotein on the extracellular surface, integral proteins

(C and D), phospholipid molecules (E) with their fatty acid tails (F) and polar head groups (G), and peripheral

protein (H).

FIGURE 2–2.

Page 9: Signal&Mito

6/4/2014

9/24

Schematic diagrams of signal-transducing membrane receptors. A. Ion channel–linked receptor (transmitter-

gated ion channel). Ligand binding causes a conformational change in the receptor, allowing ions to pass

through the membrane. B. Enzyme-linked receptor. Ligand binding causes a conformational change in the

receptor that activates an enzyme (e.g., kinase) that is part of the receptor (shown) or another protein

associated with the cytoplasmic domain of the receptor (not shown).

FIGURE 2–3.

Page 10: Signal&Mito

6/4/2014

10/24

Schematic diagram of G protein–coupled receptor activation. A. System components. B. Ligand-receptor

binding causes a conformational change in the receptor allowing the α subunit of the trimeric G protein complex

to interact with it. C. As the α subunit interacts with the receptor, it undergoes a conformational change,

allowing release of its GDP. D. GDP is traded for GTP. E. This causes a further conformational change of the

α subunit. F. This change causes the subunit to disengage from the receptor, dissociate from the γ and β

subunits, and interact with a G-protein–activated kinase. In this interaction, one high-energy bond in the GTP is

used to activate the kinase, resulting in the transformation of GTP to GDP.

FIGURE 2–4.

Page 11: Signal&Mito

6/4/2014

11/24

Schematic diagram of receptor-mediated endocytosis (II.C.3.c).

III. Cytoplasm

Cytoplasmic structures comprise three groups. Organelles are membrane-bound, enzyme-containing

subcellular compartments (e.g., mitochondria). Each organelle has a distinctive structure and unique functions.

Cytoplasmic inclusions are structures, membrane-limited or not, that are generally transient cargo and

passive in cell metabolism (e.g., lipid droplets). Organelles and inclusions are discussed in sections III.A–H. The

cytoskeleton (III.I) is a proteinaceous supporting network within the cytoplasm, components of which also form

discrete cytoplasmic structures.

A. Mitochondria

These large organelles generate chemical energy (ATP) from nutrients for use by the cell.

1. Structure. Mitochondria are comparable in size to bacteria (typically 2–6 μm long and 0.2 μm wide). They

have various shapes (e.g., spherical, ovoid, filamentous). Each is bounded by two unit membranes.

a. The outer mitochondrial membrane (Fig. 2–5, A) has a smooth contour and forms a continuous

but porous covering. It is permeable to small molecules (<5 k Da) owing to large channel-forming

proteins called porins.

b. The inner mitochondrial membrane (Fig. 2–5, B) is less permeable and has many infoldings, or

cristae (Fig. 2–5, C). The cristae of most mitochondria are shelflike, but in steroid-secreting cells

are tubular. Its inner surface contains inner membrane subunits (Fig. 2–5, H), also called F1

subunits (or lollipops, because of their shape in certain preparations); these are sites of

mitochondrial ATP synthase activity. Intercalated in the inner membrane are components of the

electron transport system, including enzymes and cofactors that have important roles in

mitochondrial function (e.g., cytochromes, proton pumps, dehydrogenases, flavoproteins).

Mitochondrial ribosomes also associate with the inner surface.

c. Membrane-limited spaces. The mitochondrial membranes create two membrane-limited spaces.

The intermembrane space (Fig. 2–5, F) is located between the inner and outer membranes and

is continuous with the intracristal space (Fig. 2–5, E), which extends into the cristae. The

Page 12: Signal&Mito

6/4/2014

12/24

intercristal space, or matrix space (Fig. 2–5, D), is enclosed by the inner membrane and

contains the mitochondrial matrix.

d. Mitochondrial matrix. This matrix contains water, solutes, and large matrix granules (Figure 2–

5, G), which play a role in mitochondrial calcium ion concentration. It also contains circular DNA

and mitochondrial ribosomes similar to those of bacteria. The matrix contains numerous soluble

enzymes involved in such specialized mitochondrial functions as the citric acid (Krebs,

tricarboxylic acid) cycle, lipid oxidation, and mitochondrial protein and DNA synthesis.

2. Function. Mitochondria provide energy for chemical and mechanical work by storing energy from cellular

metabolites in the high-energy bonds of ATP. Energy is generated by oxidative phosphorylation, which

involves four main steps: (1) the citric acid cycle uses acetyl CoA generated from pyruvate (from

carbohydrates) or fatty acids (from lipids) to generate high-energy electrons, which it donates to NADH and

FADH2; (2) these intermediates release the high-energy electrons to the electron transport chain in the

inner mitochondrial membrane, which uses their energy to drive pumps that eject protons from the matrix;

(3) the proton imbalance establishes an electrochemical gradient across the inner mitochondrial

membrane—a form of potential energy; (4) the flow of protons back into the matrix, down the gradient

established by the proton pumps, occurs through a channel in ATP synthase. ATP synthase captures that

energy by converting adenosine diphosphate (ADP) plus inorganic phosphate to ATP. ATP exits the

mitochondrion and releases its stored energy at various intracellular sites. Mitochondria synthesize their

own DNA and some proteins. They grow and reproduce by fission or budding and can undergo rapid

movement and shape changes.

3. Location. Mitochondria occur in nearly all eukaryotic cells, and in most are dispersed throughout the

cytoplasm. They accumulate in cell types and intracellular regions with high energy demands. Cardiac

muscle cells contain abundant mitochondria. Epithelial cells lining kidney tubules have abundant

mitochondria interdigitated between basal plasma membrane infoldings, where active ion and water

transport occurs.

B. Ribosomes

These protein-synthesizing organelles are two main types. Mitochondrial (like prokaryotic) ribosomes are

smaller (20 nm) than cytoplasmic (eukaryotic) ribosomes (25 nm).

1. Structure. Each ribosome type has two unequal ribosomal subunits, named for their ultracentrifugal

sedimentation values (but often called simply “large” and “small”). Mitochondrial ribosomes (70S overall)

have a 50S and a 30S subunit; cytoplasmic ribosomes (80S overall) have a 60S and a 40S subunit.

Cytoplasmic ribosomes are composed of ribosomal RNA (rRNA) synthesized in the nucleolus and many

proteins synthesized in the cytoplasm. Light microscopy reveals cytoplasmic ribosome accumulations as

basophilic patches, formerly termed ergastoplasm in glandular cells and Nissl bodies in neurons. In

EMs, ribosomes appear as small, electron-dense cytoplasmic granules.

2. Location and function. Ribosomal subunits are dispersed in the cytoplasm. Cytoplasmic ribosomes occur

in two forms. A free ribosome forms as a large and small subunit assemble on a strand of messenger

RNA (mRNA). Polyribosomes, or polysomes, are multiple ribosomes attached to a single mRNA,

permitting synthesis of multiple copies of a protein from the same message. Ribosomes translate the mRNA

code and thus play a critical role in assembling amino acids into specific proteins. As each ribosome

reaches the end of the message, its subunits separate and release the mRNA. Polysomes occur free in the

cytoplasm (free polysomes) or attached to RER membranes; free polysomes synthesize intracellular

Page 13: Signal&Mito

6/4/2014

13/24

structural proteins and enzymes. RER polysomes synthesize proteins to be secreted, sequestered, or

inserted into membranes. Various signal sequences are encoded in the 5′ end of mRNAs, helping to

direct the proteins that contain the translated sequence to different organelles. The signal sequences

directing secretory proteins to the RER are described in the next section (III.C.1.b). Other signal sequences

direct proteins to, and help them enter, the nucleus, mitochondria, and peroxisomes. A common

posttranslational modification of cytoplasmic proteins that regulates many cell functions is protein

phosphorylation. Phosphate groups, transferred from ATP or GTP to proteins by protein kinases (e.g.,

tyrosine kinase or serine/threonine kinases), or removed by protein phosphatases, can activate or

inactivate a wide variety of proteins after they have been synthesized. Such reactions are often key steps

in signal transduction (II.C.2).

C. Endoplasmic Reticulum (ER)

This complex organelle participates in synthesizing, packaging, and processing various molecules. It is a freely

anastomosing network (reticulum) of membranes that form vesicles, or cisternae; these may be elongated,

flattened, rounded, or tubular. Transfer vesicles (II.D) bud from the ER and cross the intervening cytoplasm,

delivering their contents to the Golgi complex (III.D) for further processing or packaging. In mature cells, ER

occurs in two forms: rough and smooth.

1. Rough endoplasmic reticulum (RER)

a. Structure. RER is studded with ribosomes in polysomal clusters. RER cisternae are typically

parallel, flattened, elongated, and especially abundant in cells specialized for protein secretion

(e.g., pancreatic acinar cells, plasma cells). Ribosomes render the RER basophilic. RER

membranes and individual ribosomes are visible only with the electron microscope. Proteins unique

to RER membranes include docking protein, ribophorins, and signal peptidase (III.C.1.b).

b. Function. The RER synthesizes proteins for sequestration from the cytoplasm, including secretory

proteins such as collagen, proteins for insertion into cell membranes (integral proteins; II.A.2.a),

and lysosomal enzymes (isolated to prevent autolysis). Ribosomes assemble on and read mRNAs

from 5′ to 3′. The 5′ end of mRNAs for secretory, sequestered and integral membrane proteins

carries the code for a 20- to 25-amino acid signal sequence . The signal sequence is translated

on a free polysome and subsequently binds with a cytoplasmic signal recognition particle (SRP,

six polypeptides plus a 7S RNA molecule). SRP halts translation until the SRP–polyribosome

complex binds to the RER docking protein; the SRP is then released and translation continues.

Ribophorins mediate the attachment of the signal sequence and large ribosomal subunit to the

RER membrane and provide a hydrophilic translocation channel for vectorial discharge

(unidirectional passage) of nascent protein into the RER lumen. Here the signal sequence is

cleaved by signal peptidase and the resulting nascent protein undergoes folding with the aid of

protein chaperones (e.g., calnexin, calreticulin) and assembly. Chaperones also assist in quality

control, retaining misfolded or unassembled protein complexes in the ER cisterna. If modifications

are incomplete or unsuccessful, the new proteins are eventually degraded. Another important

posttranslational modification in the ER is core glycosylation, in which preassembled

oligosaccharides, often high in mannose, are transferred from a lipid carrier (e.g., dolichol

phosphate) to amino acids, especially asparagine. The oligosaccharides help “address” proteins

for transport to intracellular destinations.

c. Location. The RER is suspended in the cytoplasm and continuous with the nuclear envelope's

outer membrane. The RER in protein-secreting epithelial cells often lies in the basal cytoplasm,

Page 14: Signal&Mito

6/4/2014

14/24

between the plasma membrane and the nucleus.

2. Smooth endoplasmic reticulum (SER)

a. Structure. The SER lacks ribosomes and thus appears smooth in electron micrographs. Its

cisternae are tubular or vesicular. SER stains poorly, if at all; thus, with the light microscope, it is

indistinguishable from the rest of the cytoplasm.

b. Function. Because it lacks ribosomes, SER cannot synthesize proteins. It has many enzymes that

are important in lipid metabolism, steroid hormone synthesis, gluconeogenis (glucose-6-

phosphatase), and detoxification. The last occurs by means of enzymatic conjugation, oxidation,

and methylation of potentially toxic substances. It plays a key role in regulating cytosolic calcium

concentrations by sequestering excess calcium and releasing it when IP3 receptors in its

membranes are activated [II.C.2.c.(2)(b)]. Because the steroid hormones it synthesizes diffuse

freely through cell membranes, they need not be transferred to the Golgi complex for packaging

and release.

c. Location. The SER is suspended in the cytoplasm of many cells and is especially abundant in cells

synthesizing steroid hormones (e.g., in the adrenal cortex and gonads). It is also abundant in liver

cells (hepatocytes), where it participates in glucose metabolism and drug detoxification. Specialized

SER termed sarcoplasmic reticulum is found in striated muscle cells, where it regulates muscle

contraction by sequestering and releasing calcium ions (10.II.B.2).

D. Golgi Complex

The Golgi complex (Golgi apparatus) is a focal center of membrane flow (II.D) and vesicle traffic among

organelles and has a key role in secretion.

1. Structure. This membranous organelle comprises three major compartments: (1) a stack of 3–10 slightly

curved, flattened cisternae with peripheral dilations; (2) numerous small vesicles peripheral to the stack;

and (3) a few large condensing vacuoles at the concave surface of the stack. The cis face (convex

face, forming face) of the stack is usually close to adjacent dilated ER cisternae and surrounded by

transfer vesicles. Its cisternae stain with osmium. The trans face (concave face, maturing face) often

harbors several condensing vacuoles and faces away from the nucleus. It connects to a system of tubules

and vesicles called the trans Golgi network (TGN), from which secretory vesicles and lysosomes exit.

2. Functions

a. Polysaccharide synthesis. The Golgi complex contains glycosyltransferases that initiate,

lengthen, or shorten polysaccharide or oligosaccharide chains one sugar at a time.

b. Modifying secretory products. The cis Golgi contains enzymes that glycosylate proteins and

lipids and sulfate glycosaminoglycans (GAGs). It is thus important in synthesizing secretory

glycoproteins, proteoglycans, glycolipids, and sulfated GAGs. It also participates in tagging

lysosomal enzymes with mannose-6-phosphate for segregation by the TGN.

c. Sorting secretory products. Products synthesized by the RER and modified in the cis Golgi are

sorted in the TGN. For example, lysosomal enzymes marked with mannose-6-phosphate and

secretory proteins destined for constitutive versus regulated exocytosis are segregated into

vesicles with different v-SNAREs in their membranes allowing them to directed to different target

sites (II.D).

Page 15: Signal&Mito

6/4/2014

15/24

d. Packaging secretory products. The TGN packages the segregated products into vesicles.

These secretory vesicles, or secretory granules, are transported to the plasma membrane for

exocytosis (II.C.5).

e. Concentrating and storing secretory products. The Golgi complexes of some cells

concentrate and store secretory products prior to secretion. Concentration (e.g., by removing water

from the vesicle lumen) is a major function of the TGN's condensing vacuoles, which also often

serve as precursors to secretory granules.

3. Location. The Golgi complex typically lies near the nucleus (juxtanuclear) and is often found near

centrioles (which organize the Golgi membranes and direct vesicle traffic). Its membranes typically

associate with microtubules by binding the molecular motor, dynein. Because dynein generally moves

vesicles toward the microtubule's minus end, Golgi membranes collect near the microtubule organizing

center (III.H.1.b) adjacent to the nucleus. Golgi complexes are well developed in neurons and secretory

(gland) cells.

4. Flow through the Golgi complex. Secretory materials were long thought to follow a one-way route (cis to

trans) through the Golgi complex. This is an oversimplification. Golgi-associated vesicles differ in their

source, destination, function, contents, and surface composition. Certain nonclathrin, vesicle-coating

proteins (e.g., β-COP, COPI, COPII) and SNAREs are associated with specific Golgi regions, indicating that

various vesicle types fuse with, and bud from, the cis, trans, or intermediate Golgi membranes.

E. Phagosomes

These membrane-limited vesicles of variable size contain material destined for lysosomal digestion.

Heterophagosomes contain extracellular material. Autophagosomes contain intracellular material, such as

worn or damaged organelles). Digestion begins when a phagosome fuses with one or more primary lysosomes

to form a secondary lysosome, as described below. (Note: Some authors use the term “heterophagosome” to

refer to secondary lysosomes [III.F.2].)

F. Lysosomes

These spherical, membrane-limited vesicles function as the cell's digestive system and may contain more than

50 enzymes each. Their enzyme activities distinguish them from other cellular granules. Acid phosphatase

occurs almost exclusively in lysosomes and is often used to identify them. Other common lysosomal enzymes

include ribonucleases, deoxyribonucleases, cathepsins, sulfatases, beta-glucuronidase, phospholipases,

various proteases, glucosidases, and lipases. An inherited lysosomal enzyme deficiency can cause life-

threatening lysosomal storage diseases (substrate accumulations in the cytoplasm). Lysosomal enzymes

usually occur as glycoproteins and are most active at acidic pH. Lysosomes vary in sizes and electron density,

depending on their activity.

1. Primary lysosomes are small (5–8 nm diameter) and electron-dense, appearing as solid black circles in

EMs. Enzymes in these storage lysosomes (released directly from the TGN) are inactive. Lysosomal

enzymes synthesized and core-glycosylated in the RER are transferred to the Golgi complex for further

glycosylation and packaging in vesicles (III.D.2). Primary lysosomes disperse through the cytoplasm and

are abundant in phagocytic cells (e.g., macrophages, neutrophils).

2. Secondary lysosomes are larger, less electron-dense, and have a mottled appearance in EMs. They

form by the fusion of one or more primary lysosomes with a phagosome. Fusion activates a proton pump in

their membrane, which acidifies the contents and activates the enzymes. Their primary function is digesting

products of heterophagy and autophagy. Lysosomal digestion yields metabolites for cell maintenance and

growth (small molecules diffuse into the surrounding cytoplasm) and aids in organelle turnover. Lysosomal

Page 16: Signal&Mito

6/4/2014

16/24

enzymes also digest some cell synthesis products, thus regulating the quality and quantity of secretory

material. Secondary lysosomes occur throughout the cytoplasm in many cells, in numbers that reflect the

cell's lysosomal and phagocytic activity.

3. Residual bodies are membrane-limited inclusions of various sizes and electron densities associated with

the terminal phase of lysosome function. They contain indigestible materials, such as pigments, crystals,

and certain lipids. Some cells (e.g., macrophages) expel residual bodies as waste, but long-lived cells (e.g.,

nerve, muscle) accumulate them. In the latter, waste-containing residual bodies reflect cellular aging and

are termed lipofuscin granules. These granules appear yellow–brown in light microscopy and as

electron-dense particles in EMs.

G. Peroxisomes

Peroxisomes are membrane-limited, enzyme-containing vesicles slightly larger than primary lysosomes. In rats,

they differ from lysosomes because of their electron-dense, granular urate oxidase nucleoid. Peroxisomes

function in hydrogen peroxide metabolism. They contain urate oxidase, hydroxyacid oxidase, and D-amino acid

oxidase, which produce hydrogen peroxide capable of killing bacteria; they also contain catalase, which

oxidizes various substrates and uses the hydrogen removed in the process to convert toxic hydrogen peroxide

to water. Peroxisomes also participate in gluconeogenesis by assisting in fatty acid oxidation. They occur

dispersed in the cytoplasm or in association with SER.

H. Other Cytoplasmic Inclusions

Prominent among storage inclusions are spherical lipid droplets, which differ in appearance depending on

the histologic preparation. Glycogen granules are PAS-positive in light microscopy and appear in EMs as

rosettes of electron-dense particles. Both lipid droplets and glycogen granules lack a limiting membrane.

Melanin is a brown pigment widely distributed in vertebrates and often found in electron-dense, membrane-

limited granules termed melanosomes. It is particularly abundant in epidermal cells, the retina's pigment

epithelium, and the brain's substantia nigra.

I. Cytoskeleton

The cytoskeleton, a mesh of filamentous elements called microtubules, microfilaments, and intermediate

filaments, provides structural stability for maintaining cell shape. It is also important in cell movement and

rearranging cytoplasmic components.

1. Microtubules

a. Structure. Microtubules (Fig. 2–6) are the thickest (24-nm) cytoskeletal components. In EMs these

fine tubules vary in length and have dense walls (5-nm thick) and a clear internal space (14-nm

across). The walls consist of subunits called tubulin heterodimers, each of which comprises one

α-tubulin and one β-tubulin protein molecule. The tubulin heterodimers are arranged in

threadlike chains called protofilaments, 13 of which align parallel to one another to form the wall

of each microtubule. Each microtubule is polarized, with a plus (+) and minus (−) end. They exist in

a state of dynamic instability, undergoing abrupt changes in length through changes in the

balance between polymerization and depolymerization.

b. Function. Microtubules form a network of roadways in the cell, deploy cytoplasmic organelles

(including the ER and Golgi complex), shuttle vesicles from one part of the cell to another, and

move chromosomes during mitosis. Their instability is critical to their function. The drug Taxol

causes permanent stabilization and leads to cell death. Most microtubules anchor by their minus

ends in the γ-tubulin rings of satellite bodies that act as nucleation sites in the microtubule

organizing center (MTOC). The MTOC is a juxtanuclear region containing the centrioles and

Page 17: Signal&Mito

6/4/2014

17/24

pericentriolar (satellite) bodies. Microtubules extend from their nucleation sites by adding GTP-

containing tubulin heterodimers to their plus end. The GTP is hydrolyzed by tubulin to GDP shortly

after heterodimer incorporation into a growing microtubule, causing destabilization,

depolymerization, and retraction as the GDP-containing heterodimers are released. Rapidly

growing tubules develop a GTP cap on their plus end, briefly protecting them from retraction, but

will depolymerize, if they fail to attach to an organelle or a stabilizing microtubule-associated protein

(MAP) in time. Stabilized microtubules acquire other MAPs; large, ATPase-containing molecular

motors (e.g., kinesin and dynein) capable of binding cellular structures and “walking” them along

the microtubules, provide an intracellular transport system. Kinesin carries its cargoes mainly

toward the plus end, and dynein toward the minus end, of stabilized microtubules. Different motors

appear to exist for each organelle or vesicle type. Thus microtubule struts that spread the ER

through the cytoplasm with the aid of kinesin, when exposed to colchicine (a drug causing net

microtubule depolymerization), collapse, allowing the ER to collapse around the nucleus.

Conversely, dynein-assisted aggregation of the Golgi complex toward the nucleus fails, causing

Golgi dispersion, in the presence of the same drug.

c. Location. Microtubules originate repeatedly from the MTOC, growing outward through the

cytoplasm and retracting if they fail to connect. Those that do connect stabilize and provide a

latticework supporting organelle deployment and vesicle traffic. Stabilized microtubule arrays occur

in cilia (III.J), flagella (III.K), basal bodies (III.L), centrioles (III.M), and the mitotic spindle apparatus

(III.N).

2. Microfilaments

a. Structure. Microfilaments are the thinnest cytoskeletal elements (5–7 nm) and are more flexible

than microtubules. They are filamentous polymers of one of several types of globular actin

protein monomers. In striated muscle cells, actin filaments form a stable paracrystalline array in

association with myosin filaments. Actin filaments in other cells are less stable and repeatedly

dissociate and reassemble. These changes are regulated in part by calcium ions, cAMP, and by a

host of actin-binding proteins in the cytoplasm and attached to the plasma membrane's

cytoplasmic surface. In addition to regulating polymerization and depolymerization, actin-binding

proteins arrange microfilaments into the networks and bundles that carry out their many important

functions.

b. Function. Microfilaments are contractile, but to contract they must interact with myosin, the only

actin-associated motor-protein family. In muscle cells, myosin forms thick filaments. In nonmuscle

cells, it exists in soluble form and binds to microfilaments by its globular head, leaving its tail (free

end) to attach to the plasma membrane or other cellular components to move them. Each actin

monomer harbors an ATP molecule that promotes binding during polymerization but hydrolyzes to

ADP shortly after binding to destabilize the microfilament, unless stabilizing actin–binding proteins

are present. Cytochalasins disrupt microfilament organization and interfere with the following

functions: endocytosis; exocytosis; extension and contraction of microvilli; cell movement;

cytoplasmic streaming; maintenance of cell shape; and equatorial constriction of dividing cells.

Phalloidin binds to intact microfilaments, stabilizes them, and interferes with these same functions.

Because it fluoresces, phalloidin may be used to localize actin filaments.

c. Location. In nonmuscle cells, microfilaments form an irregular cytoplasmic mesh. Local

Page 18: Signal&Mito

6/4/2014

18/24

accumulations occur as a thin sheath under the plasma membrane called the terminal web; as

parallel strands in cores of microvilli; in the cytoplasm at the leading edge of pseudopods; in

association with the plasma membrane, organelles, or other cytoplasmic components; or as a belt

(“purse string”) around the equator of dividing cells.

3. Intermediate filaments

a. Structure. Intermediate filaments are ropelike and composed of shorter threadlike protein subunits

twisted around one another to form filaments with a diameter (10–12 nm) intermediate between

microtubules and microfilaments. Their protein subunits are globular at their amino and carboxy

terminals, with an elongated, linear central domain. The individual proteins belong to the same

family as nuclear lamins (3.II.B) and differ depending on the cell type. Examples: cytokeratins in

epithelial cells, vimentin in mesenchyme derived cells (e.g., fibroblasts, chondrocytes), desmin in

muscle cells, glial fibrillary acidic protein in glial cells, and neurofilaments (intermediate

filament bundles) in neurons. The stability and longevity of these proteins, together with their cell-

type specificity, make them particularly useful in immunohistochemical determination of the origin of

neoplastic cells.

b. Function. Intermediate filaments are notable for their tensile strength and durability. Their

abundance in cells subjected to mechanical stress (e.g., cells of skin, connective tissue, and

muscle) indicates a role in stabilizing cell structure and in the many functions that depend on

maintaining cell shape.

c. Location. In most cells, intermediate filaments form a network surrounding the nucleus and extend

throughout the cytoplasm. Their ordered arrangement in certain cells (e.g., neurons and

keratinocytes of the skin) reflects their special role in maintaining cell shape. Cytokeratin-containing

tonofilaments of desmosomes (4.IV.B.3) are examples.

J. Motile Cilia

A ciliated cell typically has hundreds of cilia, which are motile, 5- to 10-μm long, 0.25-μm wide, plasma-

membrane covered, cell-surface extensions. Each contains a core, or axoneme, composed of nine peripheral

microtubule doublets surrounding a pair of discrete microtubules (the “9+2” arrangement). The peripheral

doublets consist of a full A microtubule and a partial B microtubule (Fig. 2–6, B). Attached to each A

microtubule are two dynein arms, which interact with the B microtubule in the adjacent doublet and drive the

movement of the doublets past one another. Other proteins that crosslink the doublets prevent simple sliding

and convert the motion into bending. Ciliary movement occurs in two phases: a forward power stroke, in which

the distal part of the cilium remains straight and rigid, and a return or recovery stroke, in which the cilium is

more flexible and bent.

K. Primary Cilia

Most mammalian cells exhibit a single non-motile cilium extending like an antenna from the cell surface. They

have a microtubule core resembling that of motile cilia, but lacking the two central microtubules (i.e., “9+0”).

These cilia have sensory functions, receiving mechanical and chemical signals from other cells and the

environment. Primary cilia extending from receptor cells in the olfactory epithelium are specialized to detect

odors. The stalk containing the inner and outer segments of visual receptors in the retina contain a primary

cilium-like structure. Inherited defects in the form and function of these structures on the surface of cells lining

the renal tubules result in polycystic kidney disease .

L. Flagella

A flagellum is similar to a motile cilium but it is longer and typically only one or two are present per cell. In

mammals, flagella, typically are 50- to 55-μm long and 0.2- to 0.5-μm thick, occur in the tails of spermatozoa.

Page 19: Signal&Mito

6/4/2014

19/24

The flagellar axoneme is identical to that of a motile cilium but is separated from the surrounding plasma

membrane by large protein masses. Flagellar movement resembles a turning corkscrew more than the whiplike

action of the cilia.

M. Basal Bodies

These structures serve as the anchoring points and microtubule organizers for all cilia or flagella. Centrioles

(III.N) migrate to the plasma membrane and give rise to basal bodies as in centriole self-duplication. Basal

bodies resemble centrioles (Fig. 2–6, C), with nine microtubule triplets.

N. Centrioles

1. Structure. A centriole is a cylinder of microtubules, 150 nm in overall diameter and 350- to 500-nm long,

containing nine microtubule triplets in a pinwheel array (Fig. 2–6, C). Each microtubule in a triplet shares

a portion of its neighbor's wall. An interphase (nondividing) cell has a perpendicular pair of adjacent

centrioles, each surrounded by several electron-dense satellites, or pericentriolar bodies containing γ-

tubulin. Cytoplasmic microtubules radiate from the pericentriolar bodies into the cytoplasm.

2. Function. Centrioles are the cell's structural organizers. Centriole duplication is required for cell division.

During mitosis, the centrioles organize the mitotic spindle (III.O). Even in vitro, isolated centrioles control

microtubule polymerization; centrioles transmit physical organizing forces by means of the microtubules

radiating from the pericentriolar bodies. Through their effects on microtubules, centrioles control organelle,

vesicle, and granule traffic within the cell. Centrioles also give rise to basal bodies (III.M). However,

centrioles are not nucleation sites for cytoplasmic microtubules; the γ-tubulin rings in the pericentriolar

bodies serve this function.

3. Location. Between cell divisions, centrioles lie near the nucleus, often surrounded by Golgi complexes.

The centrioles and associated Golgi complexes constitute the centrosome (cytocenter), which appears as

a clear juxtanuclear zone. During the S phase of interphase (Fig. 3–2), each centriole duplicates, forming a

procentriole perpendicular to the original. During mitosis, new centriole pairs migrate to opposite cell

poles to organize the spindle.

O. Mitotic Spindle Apparatus

In preparation for mitosis, cytoplasmic microtubules depolymerize and repolymerize as the mitotic spindle

apparatus. This spindle-shaped microtubule array assembles between two centriole pairs at opposite poles of

dividing cells (3.VI.A.2.a). Some spindle microtubules (continuous fibers) extend from centriole to centriole.

Others (chromosomal fibers) extend from one centriole to the centromere of a chromosome at the equatorial

plate. The spindle apparatus is crucial for chromosome separation during mitosis.

FIGURE 2–5.

Page 20: Signal&Mito

6/4/2014

20/24

Schematic diagram of a mitochondrion (III.A). Labeled components include the outer mitochondrial membrane

(A), inner mitochondrial membrane (B), cristae (C), mitochondrial matrix in the intercristal space (D), intracristal

extension (E) of the intermembrane space (F), matrix granules (G), and inner membrane subunits (F1 subunits)

(H).

FIGURE 2–6.

Page 21: Signal&Mito

6/4/2014

21/24

Schematic diagrams of microtubules and their contributions to cilia and centrioles. A. Microtubules as seen by

the electron microscope. Cross-sections of tubules show a ring of 13 subunits of dimers arranged in a spiral.

Changes in microtubule length are caused by the addition or loss of individual tubulin subunits. B. A cross-

section through a cilium reveals a microtubule complex, or axoneme, at its core. An axoneme consists of two

Page 22: Signal&Mito

6/4/2014

22/24

central microtubules surrounded by nine microtubule doublets (9 + 2). In the doublets, the A microtubule is

complete and consists of 13 subunits, whereas the B microtubule shares two or three heterodimers with the A.

When activated by ATP, the dynein arms (which harbor ATPase) link adjacent tubules and provide for the

sliding of doublets past each other. C. Centrioles consist of nine microtubule triplets in a pinwheel array. In the

triplets, the A microtubule is complete and consists of 13 subunits, whereas the B and C microtubules share

tubulin subunits. These organelles typically occur in pairs disposed at right angles to each other. (Reproduced,

with permission, from Junqueira LC, Carneiro J, Basic Histology: Text & Atlas, 11th ed. McGraw-Hill, Inc., New

York, 2005.)

IV. Cell Functions

Cells are both empowered and constrained by their available resources. The amounts and types of energy and

raw materials at their disposal, the information encoded in their genes, and intrinsic and extrinsic (epigenetic)

factors controlling access to that information are major determinants of cell function. Cells in tissues

undergoing growth or repair use more resources preparing for, and carrying out, cell division. Fully

differentiated cells concentrate on more specialized functions (e.g., secretion, contraction). Maintaining a

constant internal environment (homeostasis), even in quiescent cells, requires the expenditure of significant

energy and other resources.

A. Cellular Differentiation

Refinements in cell structure and function accompany embryonic and fetal development, maturation, aging, and

death. This cellular differentiation generally results in dividing less often and having fewer, but more efficient,

capabilities than embryonic cells. The functions of a differentiated cell can be gauged by the organelles it

contains. For example, cells specialized for protein secretion contain abundant RER and a well-developed

Golgi complex. Although differentiation can result in dramatic changes, it does not occur suddenly. It occurs in

steps, often separated by one or more passes through the cell cycle, and involves interactions among the cell's

environment, metabolic machinery, and DNA. Among the more obvious changes taking place during

differentiation are the changes in the genes being expressed. Cancer often involves a process called

dedifferentiation in which a cell returns to a more embryonic character. Differentiation-inducing agents (e.g.,

retinoic acid) may be employed to reverse carcinogenesis.

B. Intercellular Communication

Tissues, organs, and organ systems are collections of cells and cell products acting in concert to carry out

complex functions. The embryonic cells that ultimately form a tissue develop communication strategies early in

embryogenesis. Many types of intercellular communication occur—direct and indirect.

1. Direct communication. During cell-cell recognition or contact inhibition of cell division, signal transduction

may require temporary physical contact. In some tissues, especially in epithelia, cells have direct contact

with their neighbors over a large surface area. These areas of contact are often marked by specialized

junctional complexes (4.IV.B). Some components of junctional complexes are specialized for attachment

(physical communication), and others (gap junctions) provide cytoplasmic channels that transmit electric

and chemical signals.

2. Indirect communication. Signals are also transmitted from cell to cell even when they are not in contact.

In paracine communication, the signal traverses a short distance (e.g., hormones, growth factors, or

other signal molecules produced by one cell type have effects on another in the same tissue). In

endocrine and neural communication, the signal travels farther (e.g., when hormone-producing cells in

one tissue elicit responses from targets in others).

C. Cell Adhesion

Many cell functions, especially those involving cell shape and tissue integrity, depend on cell adhesion. Cell-

cell adhesion (e.g., in epithelia), requires physical linkages between the cytoskeletons of neighboring cells.

Page 23: Signal&Mito

6/4/2014

23/24

Intercellular binding is mediated by transmembrane proteins called cadherins. Their intracellular domains are

linked to the cytoskeleton by adaptor proteins that may form plaques on the cytoplasmic surfaces of opposing

membranes. Cell–substrate adhesion involves linkages between the cytoskeleton and collagen fibers of the

extracellular matrix. Transmembrane proteins in these attachments are called integrins. In these junctions,

intracellular adaptor proteins attach integrins to the cytoskeleton and extracellular adaptor proteins (e.g.,

laminin and fibronectin) attach integrins to collagen. Specific examples of cell adhesion are presented in

subsequent chapters (4.IV.B,C,E; 5.II.A.1.b and D.2).

D. Apoptosis

Apoptosis is the term applied to the processes leading to programmed cell death. Because the process

involves activating cellular mechanisms whereby the cell participates in its own destruction, many refer to it as

cellular suicide. Whether a cell does or does not follow an apoptotic pathway depends on the balance between

protective and destructive signals.

1. Structural changes. During apoptosis, the cell shrinks in size. Its plasma membrane undergoes irregular

blebbing and becomes leaky; this exposes phosphatidylserine , normally buried in the membrane's inner

leaflet, to the surface. Internally the cell's organelles, especially the mitochondria, swell and break apart

releasing proteins that are normally sequestered. Nuclear chromatin condenses, fragments, and is

degraded. Ultimately the entire cell breaks into irregular, membrane-limited apoptotic bodies, displaying

phosphatidyl serine on their surface, which attracts and activates phagocytic cells to remove them.

2. Functions of Apoptosis. In adults, apoptosis is primarily a response to cell or tissue damage beyond the

cell's ability to restore its normal structure and function (e.g., sublethal mechanical or chemical trauma, viral

infection, cancer). During development, it is a specialized form of differentiation in response to signals

dictating the normal removal of otherwise healthy cells (e.g., removal of interdigital tissue leading to digit

separation during embryonic development).

3. Mechanisms of Apoptosis. The two primary mechanisms leading to apoptosis are referred to as the

intrinsic and extrinsic pathways. Although the initiation factors differ, both pathways converge on the

same execution mechanism carried out by cysteine-aspartic-acid proteases (caspases), resulting in

apoptotic body formation.

a. Intrinsic (Mitochondrial) Pathway. Mitochondria express a protein, Bcl-2, on their surface that is

associated with protection against apoptosis by preventing the release of mitochondrial

cytochrome C. Mitochondrial damage, as occurs for example in the presence of reactive oxygen

species and resulting lipid peroxidation, allows other mitochondrial proteins, (e.g., Bax, Bad, or

Bak) to reach the mitochondrial surface and inactivate Bcl-2's anti-apoptotic effect. These Bcl-2

inactivators thus initiate perforation of the mitochondrial membrane, releasing cytochrome C into

the cytoplasm, where it forms a complex with apoptotic protease-activating factor-1 (APAF-1) and

ATP. These complexes aggregate to form apoptosomes, which bind and activate the protease

caspase-9 (an initiator caspase). By then activating effector caspases (e.g., caspase 3 and

caspase 7), caspase 9 initiates a cascade of protease activity that leads to destruction of

cytoplasmic proteins, degradation of nuclear DNA, cell fragmentation, and phagocytosis of

apoptotic bodies. Another intrinsic pathway involves the tumor suppressor protein p53, which halts

the cell cycle when sufficient DNA breaks or mutations occur (e.g., after exposure to excess

radiation; 3.VI.B.5.a). Sufficient p53 enhances apoptosis through interactions with multiple pro-

apoptotic mitochondrial factors (e.g., Bax and Bid).

b. Extrinsic Pathway. Several external signals affect apoptosis. Growth factors (e.g., FGF) and

Page 24: Signal&Mito

6/4/2014

24/24

cytokines (e.g., IL-2) can promote cell survival and thus protect against apoptosis. Other signals

(e.g., FAS ligand and tumor necrosis factor (TNF) bind to cell surface receptors that then trigger

apoptosis by activating the initiator caspase, caspase-8, which, like caspase-9, activates a

protease cascade leading to protein and DNA degradation, cell fragmentation, and death.

Copyright © McGraw-Hill Global Education Holdings, LLC.

All rights reserved.

Your IP address is 132.248.9.8