Biophysical Principles

Published on 01/03/2015 by admin

Filed under Basic Science

Last modified 22/04/2025

Print this page

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

This article have been viewed 3288 times

CHAPTER 4 Biophysical Principles*

The concepts in this chapter form the basis for understanding all the molecular interactions in chemistry and biology. To illustrate some of these concepts with a practical example, the chapter concludes with a section on an exceptionally important family of enzymes that bind and hydrolyze the nucleotide GTP. This example provides the background knowledge to understand how GTPases participate in numerous processes covered in later chapters.

Most molecular interactions are driven by diffusion of reactants that simply collide with each other on a random basis. Similarly, dissociation of molecular complexes is a random process that occurs with a probability determined by the strength of the chemical bonds holding the molecules together. Many other reactions occur within molecules or molecular complexes. The aim of biophysical chemistry is to explain life processes in terms of such molecular interactions.

The extent of chemical reactions is characterized by the equilibrium constant; the rates of these reactions are described by rate constants. This chapter reviews the physical basis for rate constants and how they are related to the thermodynamic parameter, the equilibrium constant. These simple but powerful principles permit a deeper appreciation of molecular interactions in cells. On the basis of many examples presented in this book, it will become clear to the reader that rate constants are at least as important as equilibrium constants, since the rates of reactions govern the dynamics of the cell. The chapter includes discussion of the chemical bonds important in biochemistry. Box 4-1 lists key terms used in this chapter.

BOX 4-1 Key Biophysical Terms

Rate constants, designated by lowercase ks, relate the concentrations of reactants to the rate of a reaction.

Equilibrium constants are designated by uppercase Ks. One important and useful concept to remember is that the equilibrium constant for a reaction is related directly to the rate constants for the forward and reverse reactions, as well as the equilibrium concentrations of reactants and products.

The rate of a reaction is usually measured as the rate of change of concentration of a reactant (R) or product (P). As reactants disappear, products are formed, so the rate of reactant loss is directly related to the rate of product formation in a manner determined by the stoichiometry of the mechanism. In all the reaction mechanisms in this book, the arrows indicate the direction of a reaction. In the general case, the reaction mechanism is expressed as

image

Reaction rates are expressed as follows:

image

At equilibrium, the forward rate equals the reverse rate:

image

and concentrations of reactants Req and products Peq do not change with time.

The equilibrium constant K is defined as the ratio of the concentrations of products and reactants at equilibrium:

image

so it follows that

image

In specific cases, these relationships depend on the reaction mechanism, particularly on whether one or more than one chemical species constitute the reactants and products. The equilibrium constant will be derived from a consideration of the reaction rates, beginning with the simplest case in which there is one reactant.

First-Order Reactions

First-order reactions have one reactant (R) and produce a product (P). The general case is simply

image

Some common examples of first-order reactions (Fig. 4-1) include conformational changes, such as a change in shape of protein A to shape A*:

image

and the dissociation of complexes, such as

image

The rate of a first-order reaction is directly proportional to the concentration of the reactant (R, A, or AB in these examples). The rate of a first-order reaction, expressed as a differential equation (rate of change of reactant or product as a function of time [t]), is simply the concentration of the reactant times a constant, the rate constant k, with units of s−1 (pronounced “per second”):

image

The rate of the reaction has units of M s−1, where M is moles per liter and s is seconds (pronounced “molar per second”). As the reactant is depleted, the rate slows proportionally.

A first-order rate constant can be viewed as a probability per unit of time. For a conformational change, it is the probability that any A will change to * in a unit of time. For dissociation of complex AB, the first-order rate constant is determined by the strength of the bonds holding the complex together. This “dissociation rate constant” can be viewed as the probability that the complex will fall apart in a unit of time. The probability of the conformational change of any particular A to * or of the dissociation of any particular AB is independent of its concentration. The concentra-tions of A and AB are important only in determining the rate of the reaction observed in a bulk sample (Box 4-2).

To review, the rate of a first-order reaction is simply the product of a constant that is characteristic of the reaction and the concentration of the single reactant. The constant can be calculated from the half-time of a reaction (Box 4-2).

Second-Order Reactions

Second-order reactions have two reactants (Fig. 4-2). The general case is

image

A common example in biology is a bimolecular association reaction, such as

image

where A and B are two molecules that bind together. Some examples are binding of substrates to enzymes, binding of ligands to receptors, and binding of proteins to other proteins or nucleic acids.

The rate of a second-order reaction is the product of the concentrations of the two reactants, R1 and R2, and the second-order rate constant, k:

image

The second-order rate constant, k, has units of M−1 s−1 (pronounced “per molar per second”). The units for the reaction rate are

image

the same as a first-order reaction.

The value of a second-order “association” rate constant, k+, is determined mainly by the rate at which the molecules collide. This collision rate depends on the rate of diffusion of the molecules (Fig. 4-2), which is determined by the size and shape of the molecule, the viscosity of the medium, and the temperature. These factors are summarized in a parameter called the diffusion coefficient, D, with units of m2 s−1. D is a measure of how fast a molecule moves in a given medium. The rate constant for collisions is described by the Debye-Smoluchowski equation, a relationship that depends only on the diffusion coefficients and the area of interaction between the molecules:

image

where b is the interaction radius of the two particles (in meters), the Ds are the diffusion coefficients of the reactants, and N o is Avogadro’s number. The factor of 103 converts the value into units of M−1 s−1.

For particles the size of proteins, D is approximately 10−11 m2 s−1 and b is approximately 2 × 10−9 μ, so the rate constants for collisions of two proteins are in the range of 3 × 108 M−1 s−1. For small molecules such as sugars, D is approximately 10−9 m2 s−1 and b is approximately 10−9 μ, so the rate constants for collisions of a protein and a small molecule are about 20 times larger than collisions of two proteins, in the range of 7 × 109 M−1 s−1. On the other hand, experimentally observed rate constants for the association of proteins are 20 to 1000 times smaller than the collision rate constant, on the order of 106 to 107 M−1 s−1. The difference is attributed to a steric factor that accounts for the fact that macromolecules must be correctly oriented relative to each other to bind together when they collide. Thus, the complementary binding sites are aligned correctly only 0.1% to 5% of the times that the molecules collide.

Many binding reactions between two proteins, between enzymes and substrates, and between proteins and larger molecules (e.g., DNA) are said to be “diffusion limited” in the sense that the rate constant is determined by diffusion-driven collisions between the reactants. Thus, many association rate constants are in the range of 106 to 107 M−1 s−1.

To review, the rate of a second-order reaction is simply the product of a constant that is characteristic of the reaction and the concentrations of the two reactants. In biology, the rates of many bimolecular association reactions are determined by the rates of diffusion-limited collisions between the reactants.

Reversible Reactions

Most reactions are reversible, so the net rate of a reaction is equal to the difference between the forward and reverse reaction rates. The forward and reverse reactions can be any combination of first- or second-order reactions. A reversible conformational change of a protein from A to * is an example of a pair of simple first-order reactions:

image

The forward reaction rate is k+A with units of M s−1, and the reverse reaction rate is k* with the same units. At equilibrium, when the net concentrations of A and * no longer change,

image

and

image

This equilibrium constant is unitless, since the units of concentration and the rate constants cancel out.

The same reasoning with respect to the equilibrium constant applies to a simple bimolecular binding reaction:

image

where A and B are any molecule (e.g., enzyme, receptor, substrate, cofactor, or drug). The forward (binding) reaction is a second-order reaction, whereas the reverse (dissociation) reaction is first-order. The opposing reactions are

image

The overall rate of the reaction is the forward rate minus the reverse rate:

image

Depending on the values of the rate constants and the concentrations of A,B, and AB, the reaction can go forward, backward, or nowhere.

At equilibrium, the forward and reverse rates are (by definition) the same:

image

The equilibrium constant for such a bimolecular reaction can be written in two ways:

image

This is the classical equilibrium constant used in chemistry, where the strength of the reaction is proportional to the numerical value. For bimolecular reactions, the units of reciprocal molar are difficult to relate to, so biochemists frequently use the reciprocal relationship:

image

When half of the total A is bound to B, the concentration of free B is simply equal to the dissociation equilibrium constant.

Thermodynamic Considerations

The driving force for chemical reactions is the lowering of the free energy of the system when reactants are converted into products. The larger the reduction in free energy, the more completely reactants will be converted to products at equilibrium. A thorough consideration of thermodynamics is beyond the scope of this text, but an overview of this subject is presented to allow the reader to gain a basic understanding of its power and simplicity.

The change in Gibbs free energy, δG, is simply the difference in the chemical potential, μ, of the reactants (R) and products (P):

image

The chemical potential of a particular chemical species depends on its intrinsic properties and its concentration, expressed as the equation

image

where μ0 is the chemical potential in the standard state (1 M in biochemistry), R is the gas constant (8.3 J mol−1 degree−1), T is the absolute temperature in degrees Kelvin, and C is the ratio of the concentra-tion of the chemical species to the standard concentration. Because the standard state is defined as 1 μ, the parameter C has the same numerical value as the molar concentration, but is, in fact, unitless. The term RT ln C adjusts for the concentration. When C = 1, μ= μ0.

Under standard conditions in which one mole of reactant is converted to one mole of product, the standard free energy change, δG0, is

image

However, because most reactions do not take place under these standard conditions, the chemical potential must be adjusted for the actual concentrations. This can be done by including the concentration term from the definition of the chemical potential. An equation for the free energy change that takes concentrations into account is

image

Substituting the definition of δG0, we have

image

This relationship tells us that the free energy change for the conversion of reactants to products is simply the free energy change under standard conditions corrected for the actual concentrations of reactant and products.

At equilibrium, the concentrations of reactants and products do not change and the free energy change is zero, so

image

or

image

The reader is already familiar with the fact that the equilibrium constant for a reaction is the ratio of the equilibrium concentrations of products and reactants. Thus, that relationship can be substituted in this thermodynamic equation:

image

or

image

This profound relationship shows how the free energy change is related to the equilibrium constant. The change in the standard Gibbs free energy, δG0, specifies the ratio of products and reactants when the reaction reaches equilibrium, regardless of the rate or path of the reaction. The free energy change provides no information about whether or not a given reaction will proceed on a time scale relevant to cellular activities. Nevertheless, because the equilibrium constant depends on the ratio of the rate constants, knowledge of the rate constants reveals the equilibrium constant and the free energy change for a reaction. Consider the consequences of various values of δG0:

It is sometimes said that a reaction with a positive δG0 will not proceed spontaneously. This is not strictly true. Reactants will still be converted to products, although relative to the concentration of reactants, the concentration of products will be small. The size and sign of the free energy change tell nothing about the rate of a reaction. For example, the oxidation of sucrose by oxygen is highly favored with a δG0 of −5693 kJ/mol, but “a flash fire in a sugar bowl is an event rarely, if ever, seen.”*

The free energy change is additionally related to two thermodynamic parameters that are important to the subsequent discussion of molecular interactions. The Gibbs-Helmholtz equation is the key relationship:

image

where δH is the change in enthalpy, an approximation (with a small correction for pressure-volume work) of the bond energies of the molecules. Thus, δH is the heat given off when a bond is made or the heat taken up when a bond is broken. The change in enthalpy is simply the difference in enthalpy of reactants and products. In biochemical reactions, the enthalpy term principally reflects energies of the strong covalent bonds and of the weaker hydrogen and electrostatic bonds. If no covalent bonds change, as in a binding reaction or a conformational change, δH is determined by the difference in the energy of the weak bonds of the products and reactants.

The change in entropy, expressed as δS is a measure of the change in the order of the products and reactants. The value of the entropy is a function of the number of microscopic arrangements of the system, including the solvent molecules. Note the minus sign in front of the TδS term. Reactions are favored if the change in entropy is positive, that is, if the products are less well ordered than the reactants. Increases in entropy drive reactions by increasing the negative free energy change. For example, the hydrophobic effect, which is discussed later in this chapter, depends on an increase in entropy. Increases in entropy provide the free energy change for many biologic reactions, especially macromolecular folding (see Chapters 3 and 17) and assembly (see Chapter 5).

As was emphasized in the case of δG, neither the rate of the reaction nor the path between reactants and products is relevant to the difference in enthalpy or entropy of reactants and products. The reader may consult a physical chemistry book for a fuller explanation of these basic principles of thermodynamics.

Linked Reactions

Many important processes in the cell consist of a single reaction, but most of cellular biochemistry involves a series of linked reactions (Fig. 4-3). For example, when two macromolecules bind together, the complex often undergoes some type of internal rearrangement or conformational change, linking a first-order reaction to a second-order reaction.

image

One of thousands of such examples is GTP binding to a G protein, causing it to undergo a conformational change from the inactive to the active state (Figs. 4-6 and 4-7 ahead).

image

Figure 4-7 Kinetic dissection of the Ras gtpase cycle using a series of “single turnover” experiments, in which each enzyme molecule carries out a reaction only once. A, GTP binding. Nucleotide-free Ras is mixed rapidly with a fluorescent derivative of GTP (mGTP), and fluorescence is followed on a millisecond time scale. With 100 mM mGTP (approximately 10% of the cellular concentration), binding is fast (half-time less than 5 ms), but the change in fluorescence is slower, about 30 s−1, since it depends on a subsequent, slower conformational change. Linking the association reaction to this highly favorable (K = 106) first-order conformational change accounts for the exceedingly high affinity (Kd = ˜10−11 M) of Ras for GTP. Binding and dissociation of GDP are similar.B, GTP hydrolysis and γ-phosphate dissociation. GTP is mixed with Ras, and hydrolysis is followed by collecting samples on a millisecond time scale with a “quench-flow” device, dissociating the products from the enzyme and measuring the fraction of GTP converted to GDP. The Ras-GDP-P intermediate releases γ-phosphate spontaneously in a first-order reaction. A fluorescent phosphate-binding protein is used to measure free phosphate. On this time scale in this figure, Ras alone does not hydrolyze GTP or dissociated phosphate, since the hydrolysis rate constant is 5 × 10−5 s−1, corresponding to a half-time of 1400 seconds. The GTPase activating protein (GAP) neurofibromin 1 (NF1) at a concentration of 10 mM increases the rate of hydrolysis to 20 s−1 and allows observation of the time course of phosphate dissociation at 8 s−1. C, GDP dissociation. Ras with bound fluorescent mGDP is mixed with GTP, which replaces the mGDP as it dissociates. The loss of fluorescence over time gives a rate constant for mGDP dissociation of 0.00002 s−1. The guanine nucleotide exchange factor Cdc24Mn at a concentration of 1 mM increases the rate of mGDP dissociation 500-fold to 0.01 s−1.

(Compiled from experiments reported by Lenzen C, Cool RH, Prinz H, et al: Kinetic analysis by fluorescence of the interaction between Ras and the catalytic domain of the guanine nucleotide exchange factor Cdc24Mn. Biochemistry 37:7420–7430, 1998; and by Phillips RA, Hunter JL, Eccleston JF, Webb MR: Mechanism of Ras GTPase activation by neurofibromin. Biochemistry 42:3956–3965, 2003.)

Similarly, the basic enzyme reaction considered in most biochemistry books is simply a series of reversible second- and first-order reactions:

image

where E is enzyme, S is substrate, and P is product. These and more complicated reactions can be described rigorously by a series of rate equations like those explained previously. For example, enzyme reactions nearly always involve one or more additional intermediates between ES and EP, coupled by first-order reactions, in which the molecules undergo conformational changes.

Linking reactions together is the secret of how the cell carries out unfavorable reactions. All that matters is that the total free energy change for all coupled reactions is negative. An unfavorable reaction is driven forward by a favorable reaction upstream or downstream. For example, the unfavorable reaction producing adenosine triphosphate (ATP) from adenosine diphosphate (ADP) and inorganic phosphate is driven by being coupled to an energy source in the form of a proton gradient across the mitochondrial membrane (see Fig. 8-5). This proton gradient is derived, in turn, from the oxidation of chemical bonds of nutrients. To use a macroscopic analogy, a siphon can initially move a liquid uphill against gravity provided that the outflow is placed below the inflow, so that the overall change in energy is favorable.

An appreciation of linked reactions makes it possible to understand how catalysts, including biochemical catalysts—protein enzymes and ribozymes—influence reactions. They do not alter the free energy change for reactions, but they enhance the rates of reactions by speeding up the forward and reverse rates of unfavorable intermediate reactions along pathways of coupled reactions. Given that the rates of both first- and second-order reactions depend on the concentrations of the reactants, the overall reaction is commonly limited by the concentration of the least favored, highest-energy intermediate, called a transition state. This might be a strained conformation of substrate in a biochemical pathway. Interaction of this transition state with an enzyme can lower its free energy, increasing its probability (concentration) and thus the rate of the limiting reaction. Acceleration of biochemical reactions by enzymes is impressive. Enhancement of reaction rates by 10 orders of magnitude is common.

Chemical Bonds

Covalent bonds are responsible for the stable architecture of the organic molecules in cells (Fig. 4-4). They are very strong. C—C and C—H bonds have energies of about 400 kJ mol−1. Bonds this strong do not dissociate spontaneously at body temperatures and pressures, nor are the reactive intermediates required to form these bonds present in finite concentrations in cells. To overcome this problem, living systems use enzymes, which stabilize high-energy transition states, to catalyze formation and dissolution of covalent bonds. Energy for making strong covalent bonds is obtained indirectly by coupling to energy-yielding reactions. For example, metabolic enzymes convert energy released by breaking covalent bonds of nutrients, such as carbohydrates, lipids, and proteins, into ATP (see Fig. 19-4), which supplies energy required to form new covalent bonds during the synthesis of polypeptides. Metabolic pathways relating the covalent chemistry of the molecules of life are covered in depth in many excellent biochemistry books.

For cell biologists, four types of relatively weak interactions (Fig. 4-5) are as important as covalent bonds because they are responsible for folding macromolecules into their active conformations and for holding molecules together in the structures of the cell. These weak interactions are (1) hydrogen bonds, (2) electrostatic interactions, (3) the hydrophobic effect, and (4) van der Waals interactions. None of these interactions is particularly strong on its own. Stable bonding between subunits of many macromolecular structures, between ligands and receptors, and between substrates and enzymes is a result of the additive effect of many weak interactions working in concert.

Hydrogen and Electrostatic Bonds

Hydrogen bonds (Fig. 4-5) occur between a covalently bound donor H atom with a partial positive charge, δ+ (due to electron withdrawal by a covalently bonded O or N), and an acceptor atom (usually O or N) with a partial negative charge, δ−. These bonds are highly directional, with optimal bond energy (12 to 29 kJ mol−1) when the H atom points directly at the acceptor atom. Hydrogen bonds are extremely important in the stabilization of secondary structures of proteins, such as α-helices and β-sheets (see Fig. 3-8) and in the base pairing of DNA and RNA (see Fig. 3-14).

Electrostatic (or ionic) bonds occur between charged groups that have either lost or gained a proton (e.g., —COO and —NH3+). Although these bonds are potentially about as strong as an average hydrogen bond (20 kJ mol−1), it has been argued that they contribute little to biological structure. This is because a charged group is usually neutralized by an inorganic counterion (such as Na+ or Cl) that is itself surrounded by a cloud of water molecules. The effect of having the cloud of water molecules is that the counterion does not occupy a single position with respect to the charged group on the macromolecule; so these interactions lack structural specificity.

The Hydrophobic Effect

Self-assembly and other association reactions that involve the joining together of separate molecules to form more ordered structures might seem unlikely when examined from the point of view of thermodynamics. Nonetheless, many binding reactions are highly favored, and when such processes are monitored in the laboratory, it can be shown that Ds actually increases.

How can association of molecules lead to increased disorder? The answer is that the entropy of the system—including macromolecules and solvent—increases ow-ing to the loss of order in the water surrounding the mac-romolecules (Fig. 4-5). This increase in the entropy of the water more than offsets the increased order and decreased entropy of the associated macromolecules. Bulk water is a semistructured solvent maintained by a loose network of hydrogen bonds (see Fig. 3-1). Water cannot form hydrogen bonds with nonpolar (hydrophobic) parts of lipids and proteins. Instead, water molecules form “cages” or “clathrates” of extensively H-bonded water molecules near these hydrophobic surfaces. These clathrates are more ordered than is bulk water or water interacting with charged or polar amino acids.

When proteins fold (see Fig. 17-12), macromolecules bind together (see Chapter 5), and phospholipids associate to form bilayers (see Fig. 7-5), hydrophobic groups are buried in pockets or between interfaces that exclude water. The highly ordered water formerly associated with these surfaces disperses into the less ordered bulk phase, and the entropy of the system increases.

The increase in the disorder of water that results when hydrophobic regions of macromolecules are buried is called the hydrophobic effect. Hydrophobic interactions are a major driving force, but they would not confer specificity on an intermolecular interaction except for the fact that the molecular surfaces must be complementary to exclude water. The hydrophobic effect is not a bond per se, but a thermodynamic factor that favors macromolecular interactions.

Analysis of an Enzyme Mechanism: The Ras GTPase

This section uses a vitally important family of enzymes called GTPases to illustrate how enzymes work. The example is Ras, a small GTPase that serves as part of a biochemical pathway linking growth factor receptors in the plasma membrane of animal cells to regulation of the cell cycle. The example shows how to dissect an enzyme reaction by kinetic analysis and how crystal structures can reveal conformational changes related to function. GTPases related to Ras regulate a host of systems (see Table 25-3) including nuclear transport (see Fig. 14-17), protein synthesis (see Figs. 17-9 and 17-10), vesicular trafficking (see Fig. 21-6), signaling pathways coupled to seven-helix receptors including vision and olfaction (see Figs. 25-8 and 25-9), the actin cytoskeleton (see Figs. 33-17 and 33-20), and assembly of the mitotic spindle (see Fig. 44-8). This section gives the reader the background required to understand the contributions of GTPases to all of these processes as they are presented in the following sections of the book.

Having evolved from a common ancestor, Ras and its related GTPases share a homologous core domain that binds a guanine nucleotide and use a common enzymatic cycle of GTP binding, hydrolysis, and product dissociation to switch the protein on and off (Fig. 4-6). The GTP-binding domain consists of about 200 residues folded into a six-stranded β-sheet sandwiched between five α-helices. GTP binds in a shallow groove formed largely by loops at the ends of elements of secondary structure. A network of hydrogen bonds between the protein and guanine base, ribose, triphosphate, and Mg2+ anchor the nucleotide. Larger GTPases have a core GTPase domain plus domains required for coupling to seven-helix receptors (see Fig. 25-9) or regulating protein synthesis (see Figs. 17-10 and 25-7).

The bound nucleotide determines the conformation and activity of each GTPase. The GTP-bound conformation is active, as it interacts with and stimulates effector proteins. In the example considered here, the Ras-GTP binds and stimulates a protein kinase, Raf, which relays signals from growth factor receptors to the nucleus (see Fig. 27-6). The GDP-bound conformation of Ras is inactive because it does not bind effectors. Thus, GTP hydrolysis and phosphate dissociation switch Ras and related GTPases from the active to the inactive state.

All GTPases use the same enzyme cycle, which involves four simple steps (Fig. 4-6). GTP binding favors the active conformation that binds effector proteins. GTPases remain active until they hydrolyze the bound GTP. Hydrolysis is intrinsically slow, but binding to effector proteins or regulatory proteins can accelerate this inactivation step. GTPases tend to accumulate in the inactive GDP state, because GDP dissociation is very slow. Specific proteins catalyze dissociation of GDP, making it possible for GTP to rebind and activate the GTPase. Seven-helix receptors activate their associat-ed γ-proteins. Guanine nucleotide exchange proteins (GEFs) activate small GTPases.

Figure 4-7 illustrates the experimental strategy used to establish the mechanism of the Ras GTPase cycle.

Step 1: GTP binding.

GTP binds rapidly to nucleotide-free Ras in two linked reactions (Fig. 4-7A). The first is rapid but reversible association of GTP with Ras. Second is a slower but highly favorable first-order conformational change, which produces the fluorescence signal in the experiment and accounts for the high affinity (Kd typically in the range of 10−11 M). The conformation change involves three segments of the polypeptide chain called switch I, switch II, and switch III. Folding of these three loops around the γ-phosphate of GTP traps the nucleotide and creates a binding site for the Raf kinase, the downstream effector (see Fig. 29-6).

Step 2: GTP hydrolysis.

Hydrolysis is essentially irreversible and slow with a half-time of about 4 hours (Fig. 4-7B). Although slow, GTP hydrolysis on the enzyme is many orders of magnitude faster than in solution. Like other enzymes, interactions of the protein with the substrate stabilizes the “transition state,” a high-energy chemical intermediate be-tween GTP and GDP. In this transition state, the γ-phosphate is partially bonded to both the β-phosphate and an attacking water. Hydrogen bonds between protein backbone amides and oxygens bridging the β- and γ-phosphates and on the γ- and β-phosphates stabilize negative charges that build up on these atoms in the transition state. Hydrolysis is slow in comparison with most enzyme reactions, because none of these hydrogen bonds is particularly strong. Another hydrogen bond from a glutamine side chain helps to position a water for nucleophilic attack on the γ-phosphate. The importance of this interaction is illustrated by mutations that replace glutamine 61 with leucine. This mutation reduces the rate of hydrolysis by orders of magnitude and predisposes to the development of many human cancers by prolonging the active state and thus amplifying growth-promoting signals from growth factor receptors.

Step 3: Dissociation of inorganic phosphate.

Step 4: Dissociation of GDP.

Ras and most other small GTPases depend on regulatory proteins to stimulate the two slow steps in the GTPase cycle: GDP dissociation and GTP hydrolysis. For example, when growth factors stimulate their receptors, a series of reactions (see Fig. 27-6) brings a guanine nucleotide exchange factor (GEF) to the plasma membrane to activate Ras by accelerating dissociation of GDP. First the GEF binds Ras-GDP and then favors a slow conformational change that distorts a part of Ras that interacts with the β-phosphate. This allows GDP to dissociate on a time scale of seconds to minutes rather than 10 hours (Fig. 4-7C). Once GDP has dissociated, nucleotide-free Ras can bind either GDP or GTP. Binding GTP is more likely in cells, because the cytoplasmic concentration of GTP (about 1 mM) is 10 times that of GDP. GTP binding activates Ras, allowing transmission of the signal to the nucleus.

GTPase-activating proteins (GAPs) turn off Ras and related GTPases, by binding Ras-GTP and stimulating GTP hydrolysis, thereby terminating GTPase activation (Fig. 4-7B). Ras GAPs stabilize the transition state, by contributing a positively charged arginine side chain that stabilizes the negative charges on the oxygen bridging the β- and γ-phosphates and on the γ-phosphate. GAPs also help to position Gln61 and its attacking water. In the experiment in the figure, a GAP called neurofibromin (NF1) binds Ras with a half-time of 3 ms (not illustrated) and stimulates rapid hydrolysis of GTP at 20 s−1. This is followed by rate-limiting dissociation of γ-phosphate from the Ras-GDP-P intermediate at 8 s−1 and rapid dissociation of NF1 from Ras at 50 s−1. NF1 is the product of a human gene that is inactivated in the disease called neurofibromatosis. Lacking the NF1 GAP activity to keep Ras in check, affected individuals develop numerous neural tumors that disfigure the skin and may compromise the function of the nervous system.