17.01.2015 Views

Cell Proliferation and Its Regulation - UCSF Biochemistry ...

Cell Proliferation and Its Regulation - UCSF Biochemistry ...

Cell Proliferation and Its Regulation - UCSF Biochemistry ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Katherine Hyl<strong>and</strong>, PhD<br />

<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> <strong>Its</strong> <strong>Regulation</strong><br />

(<strong>Biochemistry</strong>/Molecular Biology Lecture)<br />

OBJECTIVES<br />

• Describe the key properties of stem cells.<br />

• List the four phases of the cell cycle <strong>and</strong> describe what happens in each<br />

phase.<br />

• Name the four cyclin-Cdk complexes that drive the human cell cycle <strong>and</strong><br />

explain how the timing of their function is regulated.<br />

• Diagram the pathway by which the G1 Cdk activates the G1/S Cdk. Describe<br />

the molecular events that take place at each step of the pathway, <strong>and</strong> explain<br />

why they are important for the proliferation of normal <strong>and</strong> cancer cells.<br />

• Name the two classes of Cdk inhibitors <strong>and</strong> the cyclin-Cdk complexes they<br />

inhibit.<br />

• Describe the general nature of cell signaling networks that allow cells to<br />

interpret information from numerous extracellular signals.<br />

• Describe three classes of receptor proteins in the plasma membrane, <strong>and</strong><br />

explain how they transmit extracellular signals to the cell interior.<br />

• Diagram the pathway leading from the binding of epidermal growth factor<br />

(EGF) to the EGF receptor to activation of the cyclin D gene. Describe the<br />

molecular events that take place at each step of the pathway.<br />

• Describe the Wnt signaling pathway <strong>and</strong> its effect on cell proliferation.<br />

• Describe in molecular terms how TGFß inhibits cell division.<br />

• Describe how apoptosis can be triggered by either extracellular or<br />

intracellular signals.<br />

• Explain how the balance between pro-apoptotic <strong>and</strong> anti-apoptotic proteins<br />

determines whether a cell will die.<br />

• Explain how p53 causes cell cycle arrest <strong>and</strong> apoptosis.<br />

• Describe the spindle assembly checkpoint <strong>and</strong> the molecular function of the<br />

Mad2 protein.<br />

• Describe how selective proteolysis is achieved by the cell.<br />

KEY WORDS<br />

anaphase<br />

anaphase-promoting complex/cyclosome (APC/C)<br />

antimitogen<br />

APC (adenomatous polyposis coli)<br />

apoptosis<br />

ß-catenin<br />

Bcl-2<br />

caspase<br />

Cdk inhibitor<br />

cell cycle<br />

checkpoint<br />

chromosome segregation<br />

cyclin<br />

Mad2 protein<br />

MEK<br />

metaphase<br />

mitogen<br />

mitogen activated protein kinase (MAPK)<br />

mitotic spindle<br />

Myc<br />

p16<br />

p21<br />

p27<br />

p53<br />

progenitor cell<br />

proteasome<br />

35


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

KEY WORDS (Continued)<br />

cyclin-dependent kinase (Cdk)<br />

cytochrome c<br />

cytokinesis<br />

E2F transcription factor<br />

epidermal growth factor (EGF)<br />

Grb2<br />

growth factor<br />

GTPase-activating protein (GAP)<br />

GTP-binding protein (G-protein)<br />

guanine nucleotide exchange factor (GEF)<br />

Her2/neu<br />

kinetochore<br />

Rb protein<br />

Raf<br />

Ras<br />

receptor tyrosine kinase<br />

SH2 domain<br />

sister chromatid<br />

sister chromatid cohesion<br />

Sos<br />

spindle pole<br />

stem cell<br />

survival factor<br />

terminal differentiation<br />

transforming growth factor ß (TGFß)<br />

Wnt protein<br />

Optional reading:<br />

Alberts et al. Molecular Biology of the <strong>Cell</strong>; 5th Edition, Garl<strong>and</strong> Science, 2008.<br />

Chapter 17: <strong>Cell</strong> Cycle; Chapter 18: Apoptosis.<br />

Kumar, Abbas, Fausto <strong>and</strong> Mitchell. Robbins Basic Pathology; 8th Edition, Elsevier/Saunders,<br />

2007. Chapter 6 - Neoplasia: <strong>Cell</strong> Cycle, pp 188-198.<br />

I. INTRODUCTION<br />

<strong>Cell</strong> proliferation produces two cells from one, <strong>and</strong> it requires cell growth followed by<br />

cell division. Uncontrolled cell proliferation is a hallmark of cancer. As described in the<br />

overview lecture of cancer biology, multiple mutations that accumulate in somatic cells<br />

over many years eventually remove an elaborate set of controls that would otherwise<br />

prevent cancer cells from dividing unchecked. In this lecture, we will focus on the normal<br />

mechanisms that allow nearly all of the billions of cells in our body to proliferate only<br />

when they should. These mechanisms are subverted in cancer cells, <strong>and</strong> it is impossible to<br />

underst<strong>and</strong> cancer without first underst<strong>and</strong>ing the controls that keep the vast majority of<br />

the 10 14 cells (100,000 billion cells) that form the human body from misbehaving.<br />

In normal tissues, cell proliferation is generally restricted to cells that replenish the tissue.<br />

Most tissues are thought to contain stem cells that have this replenishment function<br />

(Figure 1). Stem cells are self-renewing cells that can divide asymmetrically to yield a<br />

new stem cell <strong>and</strong> a progenitor cell. Progenitor cells may or may not undergo further<br />

divisions, ultimately leading to terminal differentiation. Once cells have terminally<br />

differentiated, they have a specialized function <strong>and</strong> are no longer dividing. Most tissues<br />

are made up of such non-dividing cells. Thus proliferation is normally tightly controlled<br />

so that only particular cells in the body are dividing.<br />

<strong>Cell</strong> number is dependent not only on cell proliferation, but also on cell death.<br />

Programmed cell death, or apoptosis, is the process by which excess or damaged cells in<br />

the body are removed. Apoptosis is an extensive, ongoing process in our bodies. It is the<br />

balance between the production of new cells <strong>and</strong> cell death that maintains the appropriate<br />

36


Katherine Hyl<strong>and</strong>, PhD<br />

Stem <strong>Cell</strong> (self-renewing)<br />

Progenitor (Dividing)<br />

Figure 1. Stem <strong>Cell</strong>s. Stem cells are<br />

self-renewing cells. They can divide<br />

asymmetrically to produce a new<br />

stem cell (indicated by a circle) <strong>and</strong> a<br />

progenitor cell. Progenitor cells divide<br />

to produce cells that undergo terminal<br />

differentiation to produce the mature<br />

cells that make up a tissue or organ.<br />

Terminally Differentiated<br />

<strong>Cell</strong>s (Non-Dividing)<br />

number of cells in a tissue (referred to as homeostasis). Apoptosis is also a key<br />

mechanism by which cancer-prone cells are eliminated. Both normal apoptotic processes<br />

<strong>and</strong> normal cell mechanisms that control proliferation usually need to be altered to<br />

produce enough abnormal cell proliferation to cause cancer.<br />

II.<br />

THE CELL CYCLE<br />

<strong>Cell</strong> division occurs in defined stages, which together comprise the cell cycle. In terms<br />

of the genetic material, cells must replicate their chromosomal DNA once every cell<br />

cycle <strong>and</strong> segregate the sister chromatids produced by DNA replication to yield two<br />

genetically identical daughter cells (Figure 2). During DNA replication, cohesion<br />

proteins attach the replicated sister chromatids to each other, holding them together.<br />

This sister chromatid cohesion is critical for the subsequent alignment of each pair of<br />

sister chromatids on the mitotic spindle (see below), <strong>and</strong> it is therefore essential for the<br />

subsequent segregation of one (<strong>and</strong> only one) chromatid of each pair into each of the two<br />

daughter cells.<br />

14<br />

The cell division cycle is broken up into four stages: G 1<br />

, S, G 2<br />

<strong>and</strong> M (Figure3).<br />

DNA replication occurs during S (“synthesis”) phase. DNA packaging, chromosome<br />

segregation <strong>and</strong> cell division (cytokinesis) occur in M (mitosis). S phase <strong>and</strong> M phase<br />

are separated by Gap phases. G<br />

Thompson & Thompson GENETICS IN MEDICINE 1<br />

is the gap between M <strong>and</strong> S. <strong>Cell</strong> growth is one of the<br />

important events of G1. The transition from G1 to S is the critical control point in the cell<br />

M<br />

G 2<br />

(2-4 hr)<br />

G 1<br />

(10-12 hr)<br />

S<br />

(6-8 hr)<br />

Telomere<br />

Telomere<br />

Sister chromatids<br />

Centromere<br />

Figure 2-9 ■ A typical mitotic cell cycle, described in the<br />

text. The telomeres, the centromere, <strong>and</strong> sister chromatids<br />

are indicated.<br />

37<br />

spend a long time, days or years, in G 1 . In fact, some<br />

cell types, such as neurons <strong>and</strong> red blood cells, do not<br />

divide at all once they are fully differentiated; rather,<br />

The two sister chromatids are held together physically<br />

at the centromere, a region of DNA that associates with<br />

a number of specific proteins to form the kinetochore.<br />

Figure<br />

This complex<br />

2. The chromosome<br />

structure serves<br />

cycle.<br />

to attach<br />

A key<br />

each chromosome<br />

to of the cell microtubules division is of the the duplication mitotic spindle <strong>and</strong> to<br />

purpose<br />

of govern the genetic chromosome material movement carried on during mitosis. DNA<br />

chromosomes synthesis during <strong>and</strong> S its phase accurate is not synchronous segregation throughout<br />

such all chromosomes that each daughter or even cell within acquires a single one chromosome;<br />

copy rather, of each along chromatid. each chromosome, (Reproduced it begins at hundreds<br />

from to thous<strong>and</strong>s Thompson of & sites, Thompson called origins Genetics of DNA in replication.<br />

Medicine, Individual 7th chromosome edition, Nusbbaum, segments have et al, their own characteristic<br />

2007, time with of permission replication from during Elsevier.) the 6- to 8-hour<br />

p.14,<br />

S phase.<br />

By the end of S phase, the DNA content of the cell<br />

has doubled, <strong>and</strong> each cell now contains two copies of<br />

the diploid genome. After S phase, the cell enters a brief<br />

stage called G 2 . Throughout the whole cell cycle, ribonucleic<br />

acids <strong>and</strong> proteins are produced <strong>and</strong> the cell<br />

gradually enlarges, eventually doubling its total mass<br />

before the next mitosis. G 2 is ended by mitosis, which<br />

begins when individual chromosomes begin to condense<br />

<strong>and</strong> become visible under the microscope as thin,<br />

extended threads, a process that is considered in greater


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

Figure 3. The cell cycle. Reproduced<br />

with permission from Alberts et<br />

al. Molecular Biology of the <strong>Cell</strong>.<br />

Garl<strong>and</strong> Publishing, 2002.<br />

cycle. G2 is the gap between S <strong>and</strong> M, <strong>and</strong> provides time for proofreading to ensure DNA<br />

is properly replicated <strong>and</strong> packaged prior to cell division. G0 or quiescence occurs when<br />

cells exit the cell cycle due to the absence of growth-promoting signals or presence of<br />

prodifferentiation signals. Ordered progression through each phase is intricately regulated<br />

through both positive <strong>and</strong> negative regulatory signaling molecules. The G 1<br />

, S, <strong>and</strong> G 2<br />

phases comprise interphase, which accounts for most of the time in each cell cycle.<br />

The M phase, mitosis, is relatively short (approximately 1 hour of a 24 hour cell cycle).<br />

Mitosis is itself divided into several steps, described below. (For a review of mitosis, see<br />

the Mitosis <strong>and</strong> Meiosis online module on iROCKET.)<br />

1. Assembly of the mitotic spindle: At the very beginning of M phase (called<br />

prophase), the chromosomes condense while the cytoplasmic microtubules are being<br />

reorganized to build a bipolar mitotic spindle. <strong>Its</strong> purpose is to accurately segregate<br />

the two sister chromatids to opposite poles of the cell.<br />

2. Steps leading to metaphase: The nuclear envelope then breaks down, allowing the<br />

sister chromatids, which are attached to each other through sister chromatid cohesion,<br />

to become linked to the microtubules via attachment sites on each chromatid called<br />

kinetochores. Kinetochores are protein-DNA complexes in which proteins that<br />

can capture microtubules are held tightly by DNA sequences at the centromere<br />

on each sister chromatid pair. The other end of a spindle microtubule is attached<br />

to a centrosome (the major microtubule organizing center in the cell, also called<br />

the spindle pole body), which has duplicated by this time to form the two spindle<br />

poles. Because the two kinetochores on each pair of sister chromatids are attached<br />

to opposite spindle poles, they are under tension due to pulling forces that are<br />

attempting to move them to opposite poles. Eventually, the balance between these<br />

forces causes each chromosome to line up near the center of the spindle, which marks<br />

the metaphase stage of mitosis (Figure 4).<br />

3. Anaphase: After all the chromosomes achieve bipolar attachment to spindle<br />

microtubules in metaphase, sister chromatid cohesion is rapidly dissolved. As a<br />

result, the pulling forces of the microtubules cause the two sister chromatids to move<br />

rapidly to the opposite poles (Figure 5).<br />

4. Cytokinesis: After sister chromatids segregate to opposite poles, cells physically<br />

divide into two daughter cells through a process that involves pinching in of the<br />

plasma membrane (Figure 6).<br />

38


Katherine Hyl<strong>and</strong>, PhD<br />

Figure 4. The mitotic spindle at metaphase. All of the chromosomes are lined up at the equator of<br />

the spindle. Reproduced with permission from Alberts et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong><br />

Publishing, 2008.<br />

Figure 5. Anaphase. Only three pairs of sister chromatids are shown; however, in a diploid cell,<br />

this occurs simultaneously for all 46 human chromosomes (that is, for 46 pairs of sister chromatids).<br />

(Reproduced with permission from Alberts et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong> Publishing, 2008.)<br />

Figure 6. Cytokinesis. After the two sister chromatids are segregated to opposite poles, cells undergo<br />

cytokinesis by an organized pinching in of the plasma membrane. (Reproduced with permission from<br />

Alberts et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong> Publishing, 2008.)<br />

39


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

III.<br />

CELL CYCLE CONTROL: ACTIVATORS <strong>and</strong> BRAKES<br />

How is the cell cycle controlled The mechanisms of regulation can be broken down into<br />

two parts: First, how is the cell cycle regulated so that the different phases occur in the<br />

correct order Second, how do extracellular signals activate or inhibit the cell cycle This<br />

section addresses the first question, the next section (IV) addresses the second.<br />

Not until the 1980s was it discovered that a special regulatory system acts like the<br />

controller on a washing machine to drive the cell through each of its stages. This<br />

regulatory system is more than a billion years old, <strong>and</strong> most of its central components<br />

are essentially the same in single-celled eukaryotes such as yeasts <strong>and</strong> humans. This has<br />

made it possible to use the readily accessible yeast cell to dissect many of the details that<br />

underlie the normal regulatory mechanisms that control the growth of the cells in our<br />

bodies.<br />

A. Cyclin Dependent Kinases: The core activators of the cell cycle control system.<br />

The events that occur in each part of the cell cycle are carried out by specific<br />

proteins, <strong>and</strong> these proteins must be synthesized or activated at the correct time in the<br />

cycle. For example, before DNA synthesis can begin, the enzymes that produce the<br />

nucleotides used in DNA synthesis must be activated. This occurs late in G1 phase.<br />

(Remember Nucleotide Metabolism See lecture from M&N.)<br />

<strong>Cell</strong> cycle progression is positively regulated by a family of protein kinases called<br />

cyclin-dependent kinases (Cdks), which function to turn specific proteins on <strong>and</strong> off<br />

at appropriate times in the cell cycle. Like other protein kinases, Cdks turn proteins<br />

on or off by phosphorylating them. Each cyclin-dependent kinase has two subunits<br />

- a kinase subunit (the Cdk catalytic subunit) <strong>and</strong> a cyclin subunit (Figure 7). As a<br />

monomer, the Cdk has no enzymatic activity; activation requires association with a<br />

cyclin protein, which functions as an allosteric activator.<br />

Cyclins were first identified as key cell-cycle regulators when it was observed that<br />

they undergo a cycle of synthesis <strong>and</strong> regulated destruction during each cell cycle.<br />

There are several different Cdks <strong>and</strong> a number of cyclins. The kinase subunits<br />

are present throughout the cell cycle, while the cyclin subunits are produced <strong>and</strong><br />

degraded at specific times in the cell cycle, thus providing temporal regulation of the<br />

cyclin-Cdk complex. As the cyclin subunit is produced, it binds to the kinase subunit<br />

<strong>and</strong> activates it. The cyclin subunit also targets its kinase partner to specific protein<br />

substrates. The key cyclin-Cdk complexes that drive the human cell cycle are listed in<br />

Table I.<br />

The cell cycle can be viewed as a Cdk cycle (Figure 8). Activation of G1-Cdks<br />

by cyclin D turns on the events that occur in the early phase of G1. One of these is<br />

synthesis of cyclin E. As cyclin E is made, it binds to Cdk2, forming G1/S-Cdk. As<br />

the G1/S-Cdk activity accumulates to a critical threshold, it triggers the transition<br />

from late G1 into S phase. Cyclin A is made in S phase. It also binds to Cdk2,<br />

forming the S-Cdk that is required for DNA synthesis. Cyclin B is made during S<br />

phase <strong>and</strong> G2. As it is made, it binds to Cdk1 forming M-Cdk. When M-Cdk reaches<br />

a threshold activity, it triggers the transition from G2 into the prophase stage of<br />

mitosis.<br />

40


Katherine Hyl<strong>and</strong>, PhD<br />

Figure 7. Cyclin-dependent<br />

kinases (Cdks). Cdks are<br />

the key regulators of the cell<br />

division cycle in organisms as<br />

diverse as baker’s yeast <strong>and</strong><br />

humans. Cyclin-dependent<br />

kinases have two subunits, the<br />

kinase (often simply called the<br />

Cdk) <strong>and</strong> a regulatory protein<br />

called a cyclin.<br />

Table I. The four key cyclin-Cdks that drive the cell cycle<br />

Figure 8. The cell cycle as a Cdk cycle. Different phases of the cell cycle are driven by different cyclin-<br />

Cdk complexes. In this simplified view, only a G1/S-Cdk, a S-Cdk <strong>and</strong> a M-Cdk are shown. These act in<br />

sequence, as each cyclin protein is produced, to program the following critical events: the G1-S transition<br />

known as Start, S phase (DNA synthesis), <strong>and</strong> the start of M phase (mitosis). In addition, as described<br />

in the text, a G1-Cdk activated by cyclin D phosphorylates the Rb protein to produce cyclin E, which is<br />

required for G1/S-Cdk activity. Note that the activity of each Cdk disappears rapidly at a specific time<br />

in the cell cycle (as the specific cyclin protein is degraded). The APC/C is a large protein complex that<br />

controls a proteolytic process required for the separation of sister chromatids at anaphase. (Reproduced<br />

with permission from Alberts et al. Molecular Biology of the <strong>Cell</strong>. 5th Edition, Garl<strong>and</strong> Publishing, 2008;<br />

Fig. 17-16, p. 1062)<br />

41


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

B. G1 regulation: How the G1-Cdk turns on the G1/S Cdk<br />

During G1, cells prepare for DNA replication. They must synthesize proteins<br />

necessary to replicate their genome, <strong>and</strong> then assemble the various components of<br />

the DNA replication machinery onto the origins of replication. This is coordinated<br />

with nutrient <strong>and</strong> growth factor availability to ensure the cell is in an environment<br />

that supports cell division. The G1 phase of the cell cycle is unique in that it<br />

represents the only time where cells are sensitive to signals from their extracellular<br />

environment. <strong>Cell</strong>s require growth factor-dependent signals up to a point in late G1,<br />

referred to as the “restriction point” or Start, after which the transition is made into S<br />

phase. The transition between early G1 <strong>and</strong> late G1 (“Start”) illustrates one way that<br />

cyclin-dependent kinases regulate the progression of the cell around the cell cycle.<br />

This is a crucial control point that is often dysregulated in cancer.<br />

In order to move from early G1 to late G1, the cell must synthesize cyclin E.<br />

Transcription of the cyclin E gene requires a transcription factor called E2F. In<br />

cells that are not proliferating <strong>and</strong> in cells that are in early G1, the E2F transcription<br />

factor is bound to the promoter for the cyclin E gene, but it is inhibited by a protein<br />

that binds it, called Rb. (Rb st<strong>and</strong>s for Retinoblastoma, a childhood tumor of the<br />

retina – more on this in the Tumor Suppressor <strong>and</strong> Oncogene lecture). Rb is a nuclear<br />

phospho-protein that plays a key role in regulating the cell cycle. It exists in an active<br />

underphosphorylated state <strong>and</strong> an inactive hyperphosphorylated state. In its active<br />

state, Rb serves as a brake that prevents advancement of cells from G1 to S phase.<br />

When G1-Cdk activity increases near the middle of G1, G1-Cdk phosphorylates the<br />

Rb protein <strong>and</strong> inactivates it (Figure 9). Inactive phosphorylated Rb releases from<br />

E2F <strong>and</strong> allows transcription of the cyclin E gene to take place. The cyclin E protein<br />

binds to the Cdk2 kinase to form the G1/S-Cdk. E2F also transcribes a number of<br />

other genes important for S phase, including the genes for DNA polymerase <strong>and</strong><br />

thymidylate synthase.<br />

A. Molecular Events B. Control Relationships<br />

G1-Cdk OFF<br />

Cdk4-cyclin D (G1-Cdk)<br />

Rb<br />

E2F<br />

promoter<br />

cyclin E gene<br />

OFF<br />

Rb<br />

E2F<br />

G1-Cdk ON<br />

Rb<br />

P<br />

Cdk2<br />

cyclin E<br />

E2F<br />

promoter<br />

cyclin E gene<br />

ON<br />

Cdk2-cyclin E (G1/S-Cdk)<br />

Figure 9. How the G1-Cdk activates the G1/S-Cdk. The G1-Cdk (Cdk4-cyclin D) phosphorylates<br />

the Rb (Retinoblastoma) protein releasing it from transcription factor E2F. E2F can now activate<br />

transcription of cyclin E, which in turn results in the production of cyclin E protein <strong>and</strong> formation of<br />

the G1/S Cdk. In describing signaling systems, it is common to use an arrow to indicate an activation<br />

<strong>and</strong> the T-shaped symbol to indicate an inhibition.<br />

42


Katherine Hyl<strong>and</strong>, PhD<br />

Importantly, the production a cyclin E, <strong>and</strong> thus CDK2-cyclin E activity, represents<br />

the transition from mitogen-dependent to mitogen–independent cell cycle progression<br />

(or passage through “start”), irreversibly committing the cells to enter S phase. Once<br />

cells enter S phase, they are committed to divide without additional growth factor<br />

stimulation. As we will discuss in later lectures, cells that acquire mutations that<br />

obviate the need for mitogen- dependent signals will bypass this crucial control point.<br />

C. Brakes on the cell cycle: Cdk inhibitors<br />

The Rb protein can be viewed as a “brake” on the cell cycle because it prevents the<br />

transcription of the gene for cyclin E by inhibiting E2F. Three other proteins that act<br />

as “brakes” on the cell cycle are the Cdk inhibitors p16, p21, <strong>and</strong> p27. These act by<br />

binding directly to Cdk-cyclin complexes <strong>and</strong> blocking their protein kinase activity<br />

(Figure 10).<br />

Cdk inhibitors fall into two classes: specific <strong>and</strong> general. The Ink4 (inhibitors of<br />

Cdk4) family of proteins, including p16, bind exclusively to <strong>and</strong> inhibit the G1 Cdks,<br />

Cdk4/6-cylin D. The Cip/Kip family of Cdk inhibitors, including p21 <strong>and</strong> p27, bind<br />

to a broad range of Cdk-cyclin complexes, shutting off the cell cycle at multiple<br />

points. Functionally, p21 <strong>and</strong> p27 appear to mainly inhibit Cdk2 complexes. As will<br />

be discussed in the Tumor Suppressor Gene <strong>and</strong> Oncogene lecture, alterations in<br />

these inhibitor proteins play an important role in cancer.<br />

Why is the cell cycle controlled by both activators (e.g. cyclins) <strong>and</strong> inhibitors (e.g.<br />

Rb, p16, p21, p27) As we will see, it helps each cell to respond to multiple inputs,<br />

so that it enters the cell cycle only when the correct combination of conditions are<br />

present. The control of cycle entry by both growth activating <strong>and</strong> growth inhibiting<br />

signals is part of a “fail-safe” system for insuring that cell proliferation only occurs<br />

when it is useful to a multicellular organism like ourselves. Without a complex<br />

control system of this type, humans could not exist, because we would all get cancer<br />

at a very early age (probably in utero).<br />

Figure 10. How Cdk inhibitors bind to <strong>and</strong> inactivate Cdks. (Reproduced with permission from Alberts<br />

et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong> Publishing, 2008.<br />

43


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

IV.<br />

CONTROLLING PROLIFERATION: MITOGENS, ANTI-MITOGENS, <strong>and</strong><br />

CELL SIGNALING<br />

A. General Principles of <strong>Cell</strong> Signaling<br />

<strong>Cell</strong> signaling processes are central to all of human biology <strong>and</strong> medicine. Although<br />

the details of cell signaling pathways can become very complex, the big picture of<br />

cell signaling (e.g. transmitting information from the extra-cellular environment<br />

into the cell so it can respond appropriately) is straightforward. Signaling<br />

pathways are built from a limited set of molecules <strong>and</strong> molecular mechanisms (e.g.<br />

phosphorylation or proteolysis) that allow for communication within <strong>and</strong> between<br />

cells. The underlying molecular mechanisms used in signaling pathways show a<br />

number of common properties. In particular, they allow signaling proteins to undergo<br />

switch-like activation from an inactive to an active state (for example by receptor<br />

clustering, GTP-binding to Ras proteins, <strong>and</strong> stabilization of β catenin, as described<br />

below) <strong>and</strong> they can also be readily reversed (e.g. by receptor down-regulation,<br />

hydrolysis of bound GTP, <strong>and</strong> β catenin degradation). Dr. Fulton introduced the<br />

subject of signaling last year, <strong>and</strong> it is important to review that material for this block<br />

(see lectures on Protein Function <strong>and</strong> Signaling from Prologue). The introductory<br />

lectures focused on one of the two major classes of cell surface receptor proteins<br />

present in all cells: the G-protein-linked receptor family. The other major class is<br />

referred to as the enzyme-linked receptor family. This class includes receptors linked<br />

to protein kinases, which fall into two subgroups: the receptor tyrosine kinases<br />

(RTKs) <strong>and</strong> the receptor serine/threonine kinases. An example of each is discussed<br />

below.<br />

B. Mitogens <strong>and</strong> Anti-Mitogens<br />

Non-dividing cells exist in phase called G0 (G zero). G0 cells can re-enter the cell<br />

cycle in G1 when stimulated by mitogens, which are extracellular proteins that<br />

stimulate cell proliferation by directly controlling the entry of cells into the cell<br />

cycle. (For historical reasons, mitogens are often loosely referred to as growth<br />

factors. Although it is best to reserve the latter term for those signaling molecules<br />

that actually induce cell growth, i.e. the increase in cell mass, these terms are often<br />

used interchangeably). Conversely, cells can be arrested in G1 via the action of<br />

Lig<strong>and</strong><br />

Binding<br />

Domain<br />

Kinase<br />

Domain<br />

Tyrosine<br />

residue<br />

EGF<br />

P SH2<br />

SH3<br />

Grb2<br />

(adaptor protein)<br />

Sos<br />

Figure 11. Activation of the Epidermal<br />

Growth Factor (EGF) receptor tyrosine<br />

kinase. EGF binds to the EGF receptor through<br />

an extracellular lig<strong>and</strong> binding domain, leading<br />

to dimerization of the receptor. Dimerization<br />

causes one subunit to phosphorylate the other<br />

(transphosphorylation) on specific tyrosine<br />

residues. The SH2 domain of the Grb2 adaptor<br />

protein then binds to the region of the EGF<br />

receptor containing the phosphorylated<br />

tyrosines. Grb2 in turn, uses its second<br />

common protein domain, called SH3, to bind<br />

to another protein called Sos. Grp2 is known<br />

as an adapter protein, since it function to hold<br />

two other proteins together. Sos is a member<br />

of a large family of proteins that regulate G<br />

proteins (GTP-binding proteins) by causing the<br />

exchange of a tightly bound GDP molecule for<br />

GTP (see Figure 14).<br />

44


Katherine Hyl<strong>and</strong>, PhD<br />

anti-mitogens (proteins that inhibit the activity of mitogens). Many mitogens <strong>and</strong> a<br />

smaller number of anti-mitogens are known. We will discuss one example of each:<br />

the mitogen epidermal growth factor (EGF) <strong>and</strong> the anti-mitogen transforming<br />

growth factor ß (TGFß). The receptors for these factors are both enzyme-linked<br />

receptors. The EGF receptor, or EGFR, is an example of a receptor tyrosine kinase<br />

(RTK), <strong>and</strong> the TGFß receptor is a receptor serine/threonine kinase,<br />

What are the normal functions of these factors One function of EGF is to promote<br />

wound healing. After a wound is formed, epidermal <strong>and</strong> inflammatory cells secrete<br />

EGF <strong>and</strong> other growth factors. It signals cells at the margins of the wound to<br />

proliferate so that the wound may be healed. TGFß acts as a brake to this process so<br />

that the proliferation is coordinated with other aspects of wound healing.<br />

C. The EGF Signaling Pathway<br />

The EGF receptor belongs to the ErbB family of RTKs, which has four members<br />

capable of homo– or heterodimerization. Each receptor heterodimer can respond to a<br />

distinct set of extracellular lig<strong>and</strong>s <strong>and</strong> has different intracellular signaling properties.<br />

Interestingly, another member of the ErbB family, the ErbB2 receptor (also called<br />

HER2/neu) lacks intrinsic growth factor-binding activity. Consequently, in normal<br />

cells HER2/neu must function as part of a heterodimer with another ErbB family<br />

member, such as EGF. (More about HER2/neu <strong>and</strong> its role in breast cancer in later<br />

lectures.)<br />

EGF functions by binding to the extracellular domain of EGF receptor, a cell<br />

surface protein with a single transmembrane domain (Figure 11). The cytoplasmic<br />

domain of the receptor is the protein tyrosine kinase. When EGF binds to its<br />

receptor, the receptor forms a dimer in which one subunit phosphorylates the other<br />

(transphosphorylation) on particular tyrosine residues in the cytoplasmic part of the<br />

receptor. These phosphorylated tyrosines serve as binding sites for other cytoplasmic<br />

proteins that contain special domains, called SH2 domains. SH2 domains specifically<br />

recognize phosphorylated tyrosines <strong>and</strong> the adjacent amino acids. One protein that<br />

binds to phosphotyrosine residues in the EGF receptor is an adaptor protein called<br />

Grb2. Grb2, in turn, recruits a protein called Sos. Thus binding of EGF to the EGF<br />

receptor recruits both Grb2 <strong>and</strong> Sos to the intracellular portion of the receptor.<br />

Sos<br />

GDP<br />

GTP<br />

Ras-GDP<br />

Ras-GTP<br />

inactive<br />

active<br />

GAP<br />

GTPase-activating protein<br />

45<br />

Figure 12. Sos is a guanine<br />

nucleotide exchange protein (GEF)<br />

that activates the Ras protein. Ras<br />

is a monomeric GTP-binding protein<br />

that is only active in its GTP-bound<br />

form. In its GDP bound form, Ras<br />

is inactive. When Sos binds to Grb2<br />

at the EGF receptor, it is brought<br />

close to membrane-bound Ras-GDP<br />

molecules, causing the Ras to release<br />

its GDP <strong>and</strong> bind a GTP in its place.<br />

A second common type of protein is<br />

a GTPase-activating protein (GAP),<br />

which inactivates Ras by promoting<br />

its GTP hydrolysis. The cell contains<br />

hundreds of monomeric GTP-binding<br />

proteins that serve to regulate many<br />

different functions. Each is regulated<br />

in a similar way by GEFs <strong>and</strong> GAPs.


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

MEK<br />

inactive<br />

Ras-Raf<br />

MAPK<br />

inactive<br />

MEK-P<br />

active<br />

MAPK-P<br />

active<br />

Target Target-P<br />

inactive<br />

active<br />

Figure 13. Ras activates the MAP<br />

kinase cascade. Ras-GTP binds<br />

directly to Raf, which activates its<br />

kinase activity. Raf phosphorylates<br />

a kinase called MEK (also called<br />

MAP kinase kinase). After it has<br />

been phosphorylated by Raf, MEK<br />

phosphorylates MAP kinase (mitogen<br />

activated protein kinase, MAPK).<br />

Active MAPK then phosphorylates<br />

its target proteins, including<br />

transcription factors, stimulating the<br />

entry of the cell into the cell cycle.<br />

Sos is a guanine nucleotide exchange factor (GEF). It acts on a small monomeric<br />

GTP binding protein, Ras. The Ras protein is bound to the inner surface of the<br />

plasma membrane. Like the G-proteins discussed by Dr. Fulton in the Prologue, Ras<br />

can exist in two states: an inactive state in which GDP is bound, <strong>and</strong> an active state<br />

in which GTP is bound. Sos activates Ras by promoting the release of its GDP <strong>and</strong><br />

binding of GTP (Figure 14). Recruitment of Sos to the plasma membrane where Ras<br />

is located results in the activation of Ras. Ras can be returned to its inactive form<br />

through the hydrolysis of GTP to GDP. This step occurs when a GTPase-activating<br />

protein (GAP) binds Ras <strong>and</strong> induces the hydrolysis of its GTP (see Figure 12).<br />

In its GTP-bound (active) state, Ras turns on a protein kinase cascade, in which<br />

protein kinases sequentially activate each other through phosphorylation (Figure<br />

13). Active Ras binds to <strong>and</strong> activates a protein kinase called Raf. In turn, Raf<br />

phosphorylates <strong>and</strong> activates another kinase called MEK (MAP kinase kinase).<br />

MEK in turn phosphorylates <strong>and</strong> activates mitogen-activated protein kinase, MAP<br />

kinase. This chain of phosphorylation events is called the MAP kinase cascade.<br />

MAP kinase phosphorylates gene-specific transcription factors in the cell nucleus<br />

that bind to the promoters of genes <strong>and</strong> promote cell proliferation. One important<br />

transcription factor that is up-regulated by the MAP kinase cascade is Myc, which is<br />

the product of the c-MYC gene.<br />

One of the targets of transcription factors that are activated by the MAP kinase<br />

cascade is the cyclin D gene. Thus, a multi-tiered pathway connects the presence of<br />

a mitogen (EGF) outside the cell to increased expression of a key component of the<br />

cell cycle control machinery (the cyclin D gene) in the nucleus (Figure 14). Increased<br />

expression of the cyclin D gene leads to the activation of G1-Cdk, pushing the cell to<br />

proliferate, as explained previously.<br />

MAPK<br />

Transcription<br />

Factors<br />

(e.g. Myc)<br />

cyclin D<br />

Figure 14. Activation of MAP kinase leads to the transcription of cyclin D. MAPK phosphorylates<br />

transcription factors. This in turn leads to the transcription of the Myc gene, which itself encodes a<br />

transcription factor for the cyclin D gene.<br />

46


Katherine Hyl<strong>and</strong>, PhD<br />

D. Wnt signaling<br />

The Wnt proteins are mitogens analogous to EGF. They function in a signaling<br />

pathway that regulates cell proliferation by controlling proteolysis of a key signaling<br />

protein (Figure 15). The Wnt signaling pathway plays a central role during embryonic<br />

development, <strong>and</strong> also serves important functions in adults. For example, Wnt<br />

signaling is necessary for the proliferation of stem cells in the proliferative zones<br />

in the gut epithelium (the crypts that lie between the microvilli of the epithelium)<br />

(Figure 16). Colon cancer is almost invariably associated with the hyperactivation of<br />

this pathway in an early step of tumor evolution.<br />

As illustrated in Figure 15, Wnt proteins bind to a cell surface receptor called<br />

Frizzled. Frizzled controls the stability of a protein called ß-catenin, which functions<br />

together with a protein called TCF to form a transcription factor that activates the<br />

promoter of the cyclin D gene.<br />

When Wnt is bound, Frizzled turns off a protein kinase called GSK-3. GSK-3<br />

normally functions to promote the degradation of ß-catenin, thus preventing it from<br />

activating the cyclin D promoter. Phosphorylation of ß-catenin by the protein kinase<br />

GSK-3 results in its degradation. However, GSK-3 can only phosphorylate ß-catenin<br />

when ß-catenin is bound to a protein called APC (adenomatous polyposis coli).<br />

Thus, APC is necessary to hold ß-catenin in check, <strong>and</strong> loss or inactivation of APC<br />

is associated with development of colorectal cancer (as described in later lectures).<br />

(Note: this APC protein is not to be confused with APC/C, the anaphase promoting<br />

complex/cyclosome, to be described later in this lecture.)<br />

Wnt<br />

Frizzled<br />

GSK-3<br />

P<br />

P<br />

b-catenin<br />

APC<br />

b-catenin<br />

APC<br />

degradation<br />

TCF<br />

APC<br />

b-catenin<br />

TCF<br />

cyclin D<br />

Figure 15. The Wnt signaling pathway (see text for details).<br />

47


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

microvillus<br />

microvillus<br />

high<br />

Expression<br />

of APC<br />

crypt<br />

(proliferating<br />

stem cells)<br />

Figure 16. Expression of the APC gene in the gut epithelium. Shown is a<br />

schematic of a microvillus in the gut epithelium showing the zone of proliferation<br />

(crypts) <strong>and</strong> the gradient of expression of the APC gene, whose protein product<br />

inhibits Wnt signaling.<br />

low<br />

Once GSK-3 is inhibited by Frizzled, ß-catenin is no longer degraded, allowing it to<br />

associate with TCF <strong>and</strong> activate the cyclin D promoter <strong>and</strong> promote cell proliferation.<br />

Thus, Wnt signaling promotes cell proliferation through the effect of ß-catenin on<br />

cyclin D production.<br />

While Wnt proteins are the extracellular growth factors that activate this pathway,<br />

cells also control the pathway from within the cell by varying the transcription<br />

of the APC gene, whose protein product inhibits the Wnt signaling pathway. For<br />

example, in the epithelium of the colon, there exists a gradient of APC expression<br />

that is highest in the terminally differentiated nondividing cells in the microvilli <strong>and</strong><br />

lowest in the proliferating stem cells in the crypts (see Figure 16). (More about the<br />

role of APC in colon cancer to come in lectures on Colon Cancer <strong>and</strong> Familial <strong>and</strong><br />

Hereditary Cancer Syndromes.)<br />

E. TGFß-Smad: An anti-mitogenic pathway<br />

Like EGF, TGFß is an extracellular protein that binds a cell surface receptor.<br />

However, instead of causing cell proliferation, this molecule causes cells to<br />

arrest their cell cycle <strong>and</strong> enter G0. How does this occur The TGFß receptor is<br />

a transmembrane serine/threonine kinase. Upon binding to TGFß, the receptor<br />

phosphorylates proteins in the cytoplasm called Smads (Figure 17). Once<br />

phosphorylated, Smad proteins then enter the nucleus <strong>and</strong> function as transcription<br />

factors to turn on specific target genes. A key gene turned on by TGFß is the Cdk<br />

inhibitor p21 discussed above. The activation of p21 blocks G1-S transition by<br />

inhibiting Cdk2-Cyclin E/A, leading to the arrest of the cell cycle. Thus, TGFß<br />

arrests cell division by turning on transcription of the gene for a Cdk inhibitor.<br />

V. APOPTOSIS<br />

As previously explained, the number of cells in a tissue is controlled not only by cell<br />

proliferation, but also by programmed cell death, or apoptosis. For a tissue to stay the<br />

same size, cell proliferation <strong>and</strong> cell death must be perfectly balanced. Apoptosis plays<br />

important roles both during development <strong>and</strong> in mature tissues. For example, during<br />

development of a limb, tissue present between the digits must be removed. This occurs<br />

through localized apoptosis (Figure 18).<br />

48


Katherine Hyl<strong>and</strong>, PhD<br />

TGF<br />

Smad<br />

Smad-P<br />

Smad-P<br />

promoter<br />

kinase<br />

domain<br />

p27 gene<br />

TGF receptor<br />

plasma<br />

membrane<br />

p21<br />

Cdk4-cyclin D<br />

Cdk2-cyclin E<br />

Cdk2-cyclinA<br />

Figure 17. How TGFß<br />

arrests cell division. TGFß<br />

binds to the TGFß receptor.<br />

Binding of TGFß activates<br />

the receptor’s intracellular<br />

protein kinase domain,<br />

leading to phosphorylation of<br />

Smad proteins on serine <strong>and</strong><br />

threonines. Phosphorylated<br />

Smads enter the nucleus <strong>and</strong><br />

bind to promoters of genes to<br />

control transcription. A key<br />

target is the p21 gene. The<br />

p21 protein in turn inhibits<br />

cyclin E/A- cdk2 complexes,<br />

thus leading to cell cycle<br />

arrests.<br />

nucleus<br />

As described in the Prologue block, the process of apoptosis requires the activation of a<br />

special class of proteases inside the cell known as caspases. Caspase molecules normally<br />

exist as inactive procaspase molecules in the cell. Procaspase activation is carefully<br />

controlled, so that the cell only kills itself when this is appropriate for the success of the<br />

organism as a whole.<br />

A. <strong>Cell</strong>-surface death receptors activate an extrinsic apoptotic pathway<br />

Procaspase activation can be initiated from outside the cell, as happens in the immune<br />

system when T cells kill their target cells by producing a signaling protein called<br />

Fas lig<strong>and</strong>. The Fas lig<strong>and</strong> binds to its receptor, Fas, on target cells. The cytoplasmic<br />

domain of a “death receptor” such as Fas is then triggered to bind adaptor proteins<br />

that link the receptor to procaspase-8 molecules. The aggregated procaspase-8<br />

molecules are thereby stimulated to cleave each other, initiating a proteolytic cascade<br />

that leads to apoptosis (Figure 19A).<br />

B. An intrinsic apoptotic pathway depends on mitochondria<br />

When cells are stressed (e.g., hypoxia), damaged (e.g., unrepaired DNA damage), or<br />

become abnormal in other ways, they can activate apoptosis from inside the cell by<br />

triggering a similar process of procaspase aggregation <strong>and</strong> activation. In response to<br />

stress or damage, pro-apoptotic signals induce mitochondria to release cytochrome<br />

c into the cytosol, where it binds <strong>and</strong> activates an adaptor protein called Apaf-1. This<br />

causes Apaf-1 to aggregate into a wheel-like complex called an apoptosome. This<br />

aggregate then recruits a set of procaspase-9 molecules, which become activated to<br />

trigger a caspase cascade causing cell death (Figure 19B).<br />

49


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

Apoptosis<br />

Figure 18. Programmed cell death. Fas During lig<strong>and</strong> development of limb, tissue present<br />

Fas<br />

between the digits is removed by apoptosis.<br />

Apoptosis<br />

Killer T-cell<br />

Target <strong>Cell</strong><br />

Figure 19. Induction of apoptosis by either extracellular or intracellular signals. (A) Extracellular<br />

activation. Adaptor proteins bind the intracellular region of aggregated Fas proteins, causing the<br />

aggregation of procaspase-8 molecules. These then cleave one another to initiate the caspase cascade. (B)<br />

Intracellular activation. Mitochondria release cytochrome c, which binds to <strong>and</strong> causes the aggregation of<br />

the adaptor protein Apaf-1. Apaf-1 binds <strong>and</strong> aggregates procaspase-9 molecules, which are activated to<br />

trigger a caspase cascade, leading to apoptotic cell death (From Alberts et al., Molecular Biology of the<br />

<strong>Cell</strong>, 2002)<br />

50


Katherine Hyl<strong>and</strong>, PhD<br />

Domains BH 1, 2, 3 BH 1, 2, 3, 4 BH 3 only<br />

Function Pro-apoptotic Anti-apoptotic Pro-apoptotic<br />

Examples Bak, Bax Bcl-2 Bad, Bid, Puma<br />

Table 2. Three subclasses of proteins in the Bcl-2 family that control apoptosis by the<br />

intracellular (intrinsic) pathway.<br />

The release of cytochrome c from the mitochondria is tightly controlled by members<br />

of the Bcl-2 family of proteins, all of which contain at least one BH protein domain.<br />

Within this family of proteins, there are three sub-classes (Table 2): two subclasses<br />

promote apoptosis (the “pro-apoptotic” BH123 proteins, which contain 3 different<br />

BH protein domains, <strong>and</strong> the BH3-only proteins), <strong>and</strong> one subclass antagonizes<br />

apoptosis (the “anti-apoptotic” Bcl-2 proteins). The BH3 domain is the only domain<br />

shared by all three subclasses of proteins, <strong>and</strong> it can mediate a direct binding<br />

interaction between one pro-apoptotic protein <strong>and</strong> one anti-apoptotic protein to form<br />

heterodimers. The central players are the BH123 family members, Bak <strong>and</strong> Bax,<br />

which can form channels in the mitochondrial outer membrane that cause cytochrome<br />

c <strong>and</strong> other proteins in the mitochondrion’s intermembrane space to be released into<br />

the cytoplasm, thereby activating procaspase-9 via Apaf-1. The anti-apoptotic Bcl-2<br />

proteins appear to bind directly to Bak <strong>and</strong> Bax to inhibit them, thereby serving to<br />

keep the cell alive. The remaining BH3-only pro-apoptotic subclass is composed<br />

of a large number of proteins that bind to various subsets of the anti-apoptotic<br />

Bcl-2 proteins, forming heterodimers with them. If large enough amounts of the<br />

BH3 proteins are present in the right combinations, they will dissociate all of these<br />

inhibitors from Bak <strong>and</strong> Bax, thereby permitting the channel formation <strong>and</strong> inducing<br />

cell death (Figure 20).<br />

In summary, it is the balance between the activities of the set of anti-apoptotic Bcl-2<br />

proteins <strong>and</strong> the two subclasses of pro-apoptotic proteins that determines whether<br />

a mammalian cell lives or dies by the intrinsic pathway of apoptosis. This balance<br />

is determined through a complex <strong>and</strong> poorly understood signaling network that<br />

continually monitors the state of the cell. For example, only if a cell is in its expected<br />

location in the organism will it receive the specific survival signals that it requires to<br />

prevent apoptosis. Thus it is not surprising that cancer cells often acquire mutations<br />

that allow them to alter the balance between pro- <strong>and</strong> anti-apoptotic proteins, making<br />

it less likely for them to commit suicide even under conditions when normal cells<br />

would.<br />

VI.<br />

p53, THE CELL CYCLE, <strong>and</strong> APOPTOSIS<br />

The cell cycle is controlled at certain stages by checkpoints. These are biochemical<br />

mechanisms that stop the cell cycle if certain conditions are not met.<br />

One checkpoint is the G1 DNA damage checkpoint. If cells contain unrepaired damage to<br />

their DNA, the cell cycle is arrested in G1. This arrest requires a key transcription factor,<br />

p53, which is activated by DNA damage (Figure 21). There are three components to the<br />

system: 1) a DNA damage sensor, 2) the Mdm2 protein that normally causes p53 to be<br />

degraded, <strong>and</strong> 3) the p53 protein itself. DNA damage causes phosphorylation of p53 <strong>and</strong><br />

blocks the binding of Mdm2. This leads to the stabilization <strong>and</strong> accumulation of p53. p53<br />

can then bind to the promoter of the p21 Cdk inhibitor described earlier <strong>and</strong> activate its<br />

transcription, causing p21 to accumulate. The resulting inhibition of Cdks leads to cell<br />

cycle arrest.<br />

51


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

Figure 20. How pro-apoptotic BH3-<br />

only <strong>and</strong> anti-apoptotic Bcl2 proteins<br />

regulate the intrinsic pathway of<br />

apoptosis. (A) In the absence of an<br />

apoptotic stimulus, anti-apoptotic<br />

Bcl2 proteins bind to <strong>and</strong> inhibit the<br />

BH123 proteins on the mitochondrial<br />

outer membrane (<strong>and</strong> in the cytosol -<br />

not shown). (B) In the presence of an<br />

apoptotic stimulus, BH3-only proteins are<br />

activated <strong>and</strong> bind to the anti-apoptotic<br />

Bcl2 proteins so that they can no longer<br />

inhibit the BH123 proteins, which no<br />

become activated an aggregate in the<br />

outer mitochondrial membrane <strong>and</strong><br />

promote the release of intermembrane<br />

mitochondrial proteins into the cytosol.<br />

Some activated BH3-only proteins may<br />

stimulate mitochondrial protein release<br />

more directly by binding to <strong>and</strong> activating<br />

the BH123 proteins. Although not shown,<br />

the anti-apoptotic Bcl2 protins are bound<br />

to the mitochondrial surface. (Reproduced<br />

with permission from Alberts et al.<br />

Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong><br />

Publishing, 2008. Figure e18-11, p. 1124.)<br />

If p53 activation continues for a prolonged period of time, apoptosis ensues. This<br />

process kills cells with damaged DNA that remain unrepaired, <strong>and</strong> serves to remove cells<br />

from tissues that may otherwise accumulate mutations that would be passed on to their<br />

daughter cells. High levels of p53 are thought to activate apoptosis by increasing the<br />

transcription of several genes. One target gene is the BH123 protein Bax, whose gene is<br />

directly activated by p53 (Figure 22).<br />

In light of the important role p53 plays in preventing unrepaired DNA damage to be<br />

passed on to daughter cells, it is not surprising that p53 is found to play a central role in<br />

cancer development. In fact, the p53 pathway is mutated in nearly all cancers, thereby<br />

allowing damaged DNA to remain in cells as they proliferate (more in the lecture on<br />

Tumor Suppressor Genes <strong>and</strong> Oncogenes).<br />

52


Katherine Hyl<strong>and</strong>, PhD<br />

Figure 21. How DNA damage<br />

activates p53 <strong>and</strong> causes cellcycle<br />

arrest. DNA damage<br />

activates a protein kinase that<br />

phosphorylates p53, preventing<br />

its degradation. This leads to the<br />

production of high levels of the<br />

Cdk inhibitor p21. (reproduced<br />

with permission from Alberts et<br />

al. Molecular Biology of the cell.<br />

5th Edition, Garl<strong>and</strong> Publishing,<br />

2008: Fig 17-63.)<br />

Prolonged<br />

p53<br />

Activation<br />

p53<br />

promoter<br />

BAX gene<br />

Figure 22. DNA damage can lead to<br />

apoptosis. Prolonged activation of p53 in<br />

response to DNA damage results in apoptosis.<br />

p53 activates the transcription of<br />

several genes involved in apoptosis including<br />

that for the pro-apoptotic BH123 protein<br />

Bax shown here.<br />

cytochrome C<br />

apoptosis<br />

Bax channel<br />

53


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

VII.<br />

THE SPINDLE ASSEMBLY CHECKPOINT: THE IMPORTANCE OF<br />

REGULATED PROTEOLYSIS IN THE CELL<br />

In addition to monitoring the state of DNA in G1 before entering S phase, cells also<br />

monitor the state of the cell at several other checkpoints. One, called spindle assembly<br />

checkpoint, ensures that mitosis does not proceed beyond metaphase until the spindle is<br />

properly assembled. This checkpoint monitors the attachment of spindle microtubules<br />

to each kinetochore through the action of the Mad2 protein (Figure 23). There are two<br />

key features of the checkpoint: 1) Mad2 associates with kinetochores only when they<br />

are not attached to microtubules, <strong>and</strong> 2) Mad2 becomes activated for arresting mitosis<br />

only when bound to such kinetochores. If even one of the 46 human chromosomes<br />

is not attached correctly to microtubules, enough Mad2 is activated to keep the cell<br />

in metaphase. Only when the spindle has been properly assembled with all of the<br />

kinetochores bound to microtubules does Mad2 becomes inactive <strong>and</strong> allow anaphase to<br />

proceed. If there is a problem with spindle assembly, Mad2 will arrest the cell cycle until<br />

the problem is resolved.<br />

Active Mad2 exerts its effects by blocking the key regulator of the metaphase-toanaphase<br />

transition, the anaphase-promoting complex/cyclosome (APC/C). The APC/C<br />

is a member of a large family of important enzymes, called ubiquitin ligases, that trigger<br />

the regulated destruction of target proteins in the cell. The actual proteolysis is carried out<br />

by proteasomes, large protein complexes that pump selective proteins into their interior<br />

in order to cleave them into small fragments.<br />

As a ubiquitin ligase, the APC/C marks proteins for uptake into proteasomes by<br />

covalently adding multiple copies of the small protein called ubiquitin to them. The<br />

polyubiquitin chain added to a protein is recognized by the proteasome, causing the<br />

protein to be destroyed. One of the destroyed proteins is an inhibitor of the protein that<br />

centrosome<br />

microtubules<br />

Mad2<br />

cell cycle arrest<br />

kinetochore<br />

Mad2<br />

inactive<br />

Sister<br />

Chromatids<br />

Figure 23. Spindle assembly checkpoint. This checkpoint functions through the action of the Mad2<br />

protein, which binds to kinetochores that have not attached to microtubules. When bound to kinetochores,<br />

Mad2 triggers cell cycle arrest. Once microtubules are attached to all of the kinetichores, Mad2 is no<br />

longer active <strong>and</strong> the cell cycle proceeds.<br />

54


Katherine Hyl<strong>and</strong>, PhD<br />

cuts the linkages holding the sister chromatid pairs together. The removal of the inhibitor<br />

allows the separation of sisters <strong>and</strong> unleashes anaphase. The S- <strong>and</strong> M-cyclins are the<br />

second major targets of the APC/C. The destruction of these cyclins inactivates the<br />

corresponding Cdks (see Figure 8). As a result, the many proteins phosphorylated by<br />

Cdks from S phase to early mitosis are dephosphorylated by various protein phosphatases<br />

that are present in the anaphase cell. This dephosphorylation of Cdk targets is required<br />

for the completion of M phase, including the final steps in mitosis <strong>and</strong> the process of<br />

cytokinesis. Not surprisingly, cells defective in the spindle assembly checkpoint show<br />

high rates of aneuploidy because of errors in chromosome segregation during mitosis.<br />

Defects in the spindle-assembly checkpoint, <strong>and</strong> specifically in Mad2, have been<br />

associated with tumorigenesis.<br />

VIII.<br />

THE EVOLUTION OF CELL SIGNALING AND CANCER<br />

Now that we have explored the key aspects of normal cell proliferation, we can begin<br />

to consider what goes awry in cancer cells. The following section considers theoretical<br />

aspects of the evolution of cell signaling <strong>and</strong> cancer to provide you with context for<br />

thinking about cancer development <strong>and</strong> treatment. (From Dr. Bruce Alberts)<br />

A. Elaborate cell signaling mechanisms had to evolve in multicellular organisms to<br />

prevent cancer. Various types of evidence suggest that single-celled life was present on<br />

the earth 3.5 billion years ago, about a billion years after the earth formed (prokaryotic<br />

cells such as bacteria). However, it appears to have required another two billion years<br />

to evolve the first multicellular organisms. Initially these were very small aggregates of<br />

eukaryotic cells that had learned how to cooperate, with each cell restraining its own<br />

growth for the good of the entire aggregate. Although this had the advantage of allowing<br />

each type of cell to specialize, it meant that each cell had to send <strong>and</strong> receive an elaborate<br />

set of signals to determine its appropriate behavior, <strong>and</strong> that fail-safe controls had to<br />

evolve to prevent the type of selfish cell behavior that we call cancer.<br />

As larger <strong>and</strong> larger organisms evolved, major improvements to these fail-safe controls<br />

had to develop in the form of multiple, largely redundant systems that prevent aberrant<br />

cell proliferation. Why Even with the overlapping set of proofreading mechanisms that<br />

allow us to replicate the three billion (3 x 10 9 ) nucleotide pairs in the human genome with<br />

an error rate of only about one in a billion (10 -9 ), the fact that humans are formed from<br />

about 10 14 cells means that billions of cells experience mutations every day, potentially<br />

disrupting the normal controls on cell growth. Viewed from this perspective, the<br />

surprising thing about large multicellular organisms is how infrequently cells misbehave<br />

to create a tumor. As we shall see, the reason we do not all die of cancer is that, in<br />

general, many different mutations need to accumulate in a single line of cells to cause<br />

this disease – perhaps 10 to 20. Obtaining a better underst<strong>and</strong>ing the multiple layers of<br />

control that are circumvented during tumorigenesis will be key to controlling cancer.<br />

Unfortunately, there is still much to learn in this critical area of research.<br />

B. <strong>Cell</strong>s integrate the many signals that they receive in deciding whether to survive,<br />

grow <strong>and</strong> divide (proliferate), differentiate, or die (apoptosis). Every cell contains<br />

many different cell surface receptor molecules, each of which recognizes a particular<br />

molecule at the cell exterior. Some of these bind to signaling molecules that have been<br />

secreted by neighboring cells, others bind to protein molecules held in the plasma<br />

membrane of tightly opposed adjacent cells, while others bind to the extracellular matrix.<br />

All of these signal molecules work in combinations to regulate the behavior of the cell,<br />

55


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

with each of the hundreds or thous<strong>and</strong>s of different cell types in our bodies responding<br />

to this babble of signals differently. As shown in Figure 24, an individual cell generally<br />

requires multiple signals just to survive. It requires additional signals to grow <strong>and</strong> divide,<br />

<strong>and</strong> a different set of additional signals to differentiate. If deprived of its required survival<br />

signals, a cell will undergo cell suicide (apoptosis). The actual situation is even more<br />

complex than illustrated in Figure 24, since some extracellular signal molecules act to<br />

inhibit these <strong>and</strong> other cell behaviors, or even to induce apoptosis.<br />

How exactly a cell makes each of the all-or-none decisions illustrated in Figure 24 is<br />

not understood in detail. Speaking metaphorically, the decision is analogous to “cell<br />

thinking”. <strong>Cell</strong>s integrate the many signals they receive through a “cross-talk” between<br />

different intracellular events triggered by different cell surface receptors. Some of the<br />

cross talk depends on simple “coincidence detectors”, as in the example shown in Figure<br />

25. Here two different signaling events are needed to activate a single protein inside the<br />

cell, because the protein needs to be phosphorylated at two different sites to become<br />

active. Thus, the activation of this protein occurs if, <strong>and</strong> only if, two specific extracellular<br />

signals are present simultaneously. But much of the cross talk is more complex <strong>and</strong> not<br />

yet decipherable.<br />

Figure 24. How an animal cell depends on multiple extracellular signal molecules.<br />

Reproduced with permission from Alberts et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong><br />

Publishing, 2008.<br />

56


Katherine Hyl<strong>and</strong>, PhD<br />

Figure 25. Signal integration inside the cell. Here signals A <strong>and</strong> B each trigger a different intracellular<br />

signaling pathway. Both pathways involve the activation of a protein kinase that phosphorylates protein<br />

Y, but at a different site. Because both sites must be phosphorylated for protein Y to become activated,<br />

protein Y serves as a coincidence detector that indicates that both extracellular molecules A <strong>and</strong> B<br />

are present. (Reproduced with permission from Alberts et al. Molecular Biology of the <strong>Cell</strong>. Garl<strong>and</strong><br />

Publishing, 2008.)<br />

Why is it so important to underst<strong>and</strong> how cells “think” Cancer can be viewed as a<br />

disease in which a cell has accumulated so many changes in its intracellular processes<br />

that it has escaped from all of its normal requirements, thinking that it should proliferate<br />

<strong>and</strong> survive independent of its environment. The ideal cancer therapy would be based<br />

on an underst<strong>and</strong>ing of the exact, highly abnormal intracellular state of the cells in<br />

a particular tumor. One might then be able to induce apoptosis in the cancer cells by<br />

exposing them to a mixture of two or three specific signaling molecules (or inhibitors<br />

of such molecules), with no deleterious effect on normal cells. It is important to keep<br />

in mind, however, that each tumor has its own unique set of mutations <strong>and</strong> aberrant<br />

signaling pathways, resulting from a long evolutionary process of r<strong>and</strong>om mutation <strong>and</strong><br />

natural selection during tumor progression. Thus, we should view cancer as a collection<br />

of different but related diseases, each of which may require its own specific combination<br />

of therapies to treat it.<br />

Acknowledgements:<br />

Significant contributions to this lecture were made by<br />

Hiten Madhani, PhD, Department of <strong>Biochemistry</strong> <strong>and</strong> Biophysics, who originally<br />

developed this lecture (lecturer 2002 – 2006); <strong>and</strong><br />

Bruce Alberts, PhD, Department of <strong>Biochemistry</strong> <strong>and</strong> Biophysics (lecturer 2007).<br />

57


<strong>Cell</strong> <strong>Proliferation</strong> <strong>and</strong> its <strong>Regulation</strong><br />

58

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!