5.2 Enzymes: The Reaction Catalysts of Biological Systems

Rare indeed is the organic chemical reaction that proceeds unaided at a rate sufficient to support living systems. Enzymes have extraordinary catalytic power, often far greater than that of synthetic or inorganic catalysts. They have a high degree of specificity for their substrates, they substantially accelerate chemical reactions, and they function in aqueous solutions under very mild conditions of temperature and pH. Few nonbiological catalysts have all these properties.

Enzymes are central to every cellular process. Acting in organized sequences, they catalyze the hundreds of stepwise reactions that degrade nutrient molecules, conserve and transform chemical energy, make biological macromolecules from simple precursors, and carry out the various processes of DNA and RNA metabolism.

In molecular biology, the study of enzymes has immense practical importance. In some diseases, especially hereditary genetic disorders related to DNA or RNA metabolism, there may be a deficiency or even a total absence of one or more enzymes. Other disease conditions may be caused by excessive activity of an enzyme. Many medicines act through interactions with enzymes. Furthermore, researchers can isolate and harness enzyme functions to suit their purposes in the laboratory. The set of methods collectively described as “biotechnology” (many of which are described in Chapter 7 and elsewhere throughout this book) is made possible by our understanding of the enzymes of DNA and RNA metabolism.

Enzymes Catalyze Specific Biological Reactions

With the exception of a small group of catalytic RNA molecules (described in Chapter 16), all enzymes are proteins. Their catalytic activity depends on the integrity of their native protein conformation. Enzymes, like other proteins, have molecular weights ranging from about 12,000 to more than 1 million. Some enzymes require for their activity no additional chemical groups other than their amino acid residues. Others require an additional chemical component called a cofactor, either one or more inorganic metal ions (Table 5-3) or a complex organic or metallo-organic molecule called a coenzyme, which acts as a transient carrier of specific functional groups (Table 5-4). Most coenzymes are derived from vitamins, organic nutrients required in small amounts in the human diet. Some enzymes require both a coenzyme and one or more metal ion cofactors for activity. A coenzyme or inorganic cofactor that is very tightly or even covalently bound to the enzyme protein is known as a prosthetic group. A complete, catalytically active enzyme, with its bound coenzyme and/or inorganic cofactor, is referred to as a holoenzyme. The protein part of such an enzyme is called the apoenzyme or apoprotein.

Figure 5-3: Inorganic Elements as Cofactors for Enzymes and Regulatory Proteins
Figure 5-4: Coenzymes: Transient Carriers of Specific Atoms or Functional Groups

143

An enzyme catalyzes a reaction by providing a specific environment in which the reaction can occur more rapidly. Here are some key principles of enzyme-catalyzed reactions:

  1. A molecule that undergoes an enzyme-catalyzed reaction is referred to as a substrate. A substrate differs from a ligand in that it can undergo a chemical transformation while bound to the enzyme, whereas a ligand does not.

  2. The substrate interacts with the enzyme in a pocket known as the active site (Figure 5-8). The active site is typically lined with multiple chemical groups—amino acid side chains, metal ion cofactors, and/or coenzymes—all oriented to facilitate the reaction.

    Figure 5-8: Binding of a substrate to an enzyme at the active site. This is the enzyme T4 RNA ligase, with a bound substrate, ATP (stick representation). Active sites are typically pockets in the surface of enzymes.
  3. An enzyme-catalyzed reaction is highly specific for that particular reaction. An active site that is set up to catalyze one reaction with one substrate will not interact well with other substrates. Specificity is an important property of every enzyme. The catalysis of a different reaction requires a different enzyme.

  4. Catalysis often requires conformational flexibility. As we have seen for proteins that reversibly bind ligands, conformational changes have an essential role in enzyme function. Induced fit and cooperativity also play roles in enzyme catalysis.

  5. Many enzymes are regulated. The panoply of enzymes available to a given cell confers the opportunity not just to accelerate reactions but to control them. In this way, cellular metabolism can be modulated as resources and circumstances demand.

The enormous and highly selective rate enhancements achieved by enzymes can be explained by the many types of covalent and noncovalent interactions between enzyme and substrate. Chemical reactions of many types may take place between substrates and the functional groups (specific amino acid side chains, metal ions, and coenzymes) on enzymes. The particular reactions that occur depend on the requirements of the overall reaction to be catalyzed. An enzyme’s catalytic functional groups may form a transient covalent bond with the substrate and activate it for reaction. Or a group may be transiently transferred from the substrate to the enzyme. The most common type of group transfer involves the transfer of protons between ionizable amino acid side chains in the active site and groups on the substrate molecule, a process called general acid and base catalysis (Figure 5-9). In the enzymes important to molecular biology, phosphoryl group transfers are also common. In many cases, these group transfer reactions occur only in the enzyme active site. The capacity to facilitate multiple interactions and transfers of this type, sometimes all at once, is one of the factors contributing to the rate enhancements provided by enzymes.

Figure 5-9: An enzyme-catalyzed reaction. Shown here is the key step in the formation of a new phosphodiester bond in the active site of a DNA or RNA polymerase. The end of a growing chain of nucleic acid (the primer chain) is at the top left, and an incoming (deoxy)nucleoside triphosphate is at the lower left. The reaction begins with general base catalysis by an active-site residue (B) that abstracts a proton (Ha) from the attacking 3′ hydroxyl at the end of the primer chain. The oxygen of the hydroxyl group concurrently attacks the phosphorus of the a-phosphoryl group of the nucleoside triphosphate, displacing pyrophosphate (PPi). The pyrophosphate is protonated by another active-site residue (usually a Lys, shown here as A), an example of general acid catalysis, which facilitates ejection of the PPi. Two metal ions, usually two Mg2+, are in the active site. One metal ion lowers the pKa of the primer 3′ hydroxyl to facilitate the general base catalysis. The other metal ion coordinates with and orients oxygens of the triphosphate and also aids catalysis by stabilizing the transition state of the reaction.

144

Covalent interactions are only part of the story, however. Much of the energy required to increase the reaction rate is derived from weak, noncovalent interactions between substrate and enzyme, including hydrogen bonds and hydrophobic and ionic interactions. The formation of each weak interaction is accompanied by the release of a small amount of free energy that stabilizes the interaction. The energy derived from enzyme-substrate interaction is called binding energy, ΔGB. Its significance extends beyond a simple stabilization of the enzyme-substrate interaction. Binding energy is a major source of the free energy used by enzymes to increase the rates of reactions.

In the context of nucleic acid metabolism, one type of weak interaction merits special mention. Ionic interactions can include interactions between bound metals (such as Mg2+, Mn2+, and Fe2+ or Fe3+ ions) and substrates. About one-third of all enzymes utilize metals in their catalytic mechanisms, and that proportion is much higher for the enzymes that act on DNA and RNA. For example, the active sites of DNA and RNA polymerases universally feature two metal ions, usually two Mg2+ions, that help orient substrates and facilitate the overall reaction in multiple ways (see Figure 5-9). Mg2+ ions play key roles at the active sites of a wide range of enzymes discussed in later chapters.

145

Enzymes Increase the Rate of a Reaction by Lowering the Activation Energy

A simple enzyme reaction might be written like this:

where E, S, and P represent the enzyme, substrate, and product, and ES and EP are transient complexes of the enzyme with the substrate and with the product.

To understand catalysis, we must recall the important distinction between reaction equilibria and reaction rates. The equilibrium between S and P reflects the difference in the free energies of their ground states. Any reaction, such as S ⇌ P, can be described by a reaction coordinate diagram, a picture of the energy changes during the reaction (Figure 5-10a). Energy in biological systems is described in terms of free energy, G (see Chapter 3). In the diagram, the free energy of the system is plotted against the progress of the reaction (the reaction coordinate). In this example, the free energy of the ground state of P is lower than that of S, so the biochemical standard free-energy change, or ΔG′°, for the reaction is negative and the equilibrium favors P.

Figure 5-10: The use of noncovalent binding energy to accelerate an enzyme-catalyzed reaction. (a) A reaction coordinate diagram. The free energy of a system is plotted against the progress of the reaction S → P. This kind of diagram describes the energy changes during the reaction; the horizontal axis (reaction coordinate) reflects the progressive chemical changes (e.g., bond breakage or formation) as S is converted to P. The activation energies, ΔG, for the S → P and P → S reactions are indicated. ΔG′° is the overall biochemical standard free-energy change in the direction S → P. The ES intermediate occupies a minimum in the energy progress curve of the enzyme-catalyzed reaction. The terms ΔGuncat and ΔGcat correspond to the activation energy for the uncatalyzed reaction (black, dashed curve) and the overall activation energy for the catalyzed reaction (blue, solid curve), respectively. The activation energy is lowered by the amount ΔGcat when the enzyme catalyzes the reaction. (b) An imaginary enzyme (stickase) designed to catalyze the breaking of a metal stick. Before the stick is broken, it must be bent (transition state). Magnetic interactions replace weak enzyme-substrate bonding interactions. A stickase with a magnet-lined pocket that is structurally complementary to the stick (substrate) stabilizes the substrate (middle). Bending is impeded by the magnetic attraction between stick and stickase. An enzyme with a pocket complementary to the reaction transition state helps destabilize the stick (bottom), contributing to catalysis. The binding energy of the magnetic interactions compensates for the increase in free energy needed to bend the stick. In enzyme active sites, weak interactions that occur only in the transition state aid in catalysis.

KEY CONVENTION

To describe the free-energy changes for reactions, chemists define a standard set of conditions (temperature 298 K; partial pressure of each gas 1 atm, or 101.3 kPa; concentration of each solute 1 m) and express the free-energy change for a reacting system under these conditions as ΔG°, the standard free-energy change. Because living systems commonly involve H+ concentrations far below 1 m, biochemists and molecular biologists define a biochemical standard free-energy change, ΔG′°, the standard free-energy change at pH 7.0; we use this definition throughout the book.

146

A favorable equilibrium does not mean that the S → P conversion will occur at a fast or even detectable rate. An unfavorable equilibrium does not mean that the reaction will be slow. Instead, the rate of a reaction depends on the height of the energy hill that separates the product from the substrate. At the top of this hill lies the transition state (denoted by ‡ in Figure 5-10a). The transition state is not a stable species but a transient moment when the alteration in the substrate has reached a point corresponding to the highest energy in the reaction coordinate diagram. The difference between the energy levels of the ground state and the transition state is the activation energy, ΔG. A higher activation energy corresponds to a slower reaction.

The function of a catalyst is to increase the rate of a reaction. Catalysts do not affect reaction equilibria, and enzymes are no exception. The bidirectional arrows in Equation 5-6 make this point: any enzyme that catalyzes the reaction S → P also catalyzes the reaction P → S. The role of enzymes is to accelerate the interconversion of S and P. The enzyme is not used up in the process, and the equilibrium point is unaffected. However, the reaction reaches equilibrium much faster when the appropriate enzyme is present, because the rate of the reaction is increased. Enzymes increase reaction rates by lowering the activation energy of the reaction. To achieve this, enzymes utilize noncovalent and covalent interactions in somewhat different ways.

Two fundamental and interrelated principles provide a general explanation for how enzymes use noncovalent binding energy to accelerate a reaction:

  1. Much of the catalytic power of an enzyme is ultimately derived from the free energy released in forming many weak bonds and interactions between the enzyme and its substrate. This binding energy contributes to specificity as well as to catalysis.

  2. Weak interactions are optimized in the reaction transition state; enzyme active sites are complementary not to substrates per se but to the transition states through which substrates pass as they are converted to products during the reaction (Figure 5-10b).

When the enzyme active site is complementary to the reaction transition state, some of the noncovalent interactions between enzyme and substrate occur only in the transition state. The free energy (binding energy) released by the formation of these interactions partially offsets the energy required to reach the top of the energy hill. The summation of the unfavorable (positive) activation energy ΔG and the favorable (negative) binding energy ΔGB results in a lower net activation energy (see Figure 5-10a). Even on the enzyme, the transition state is not a stable species but a brief point in time that the substrate spends atop an energy hill. The enzyme-catalyzed reaction is much faster than the uncatalyzed process, however, because the hill is much smaller. The groups on the substrate that are involved in the weak interactions between the enzyme and transition state can be at some distance from the substrate bonds that are broken or changed. The weak interactions formed only in the transition state are those that make the primary contribution to catalysis (see Figure 5-10b).

Covalent interactions can accelerate some enzyme-catalyzed reactions by creating a different, lower-energy reaction pathway. When the reaction occurs in solution in the absence of the enzyme, the reaction takes a particular (and usually very slow) path. In the enzyme-catalyzed reaction, if a group on the enzyme is transferred to or from the substrate during the reaction, the reaction path is altered. The new pathway results in acceleration of the reaction only if its overall activation energy is lower than that of the uncatalyzed reaction.

The Rates of Enzyme-Catalyzed Reactions Can Be Quantified

The oldest approach to understanding enzyme mechanisms, and the one that remains most important, is to determine the rate of a reaction and how it changes in response to changes in experimental parameters, a discipline known as enzyme kinetics. We provide here a brief review of key concepts related to the kinetics of enzyme-catalyzed reactions (for more advanced treatments, see the Additional Reading list at the end of the chapter.)

Substrate concentration affects the rate of enzyme-catalyzed reactions. Studying the effects of substrate concentration [S] in vitro is complicated by the fact that it changes during the course of the reaction, as substrate is converted to product. One simplifying approach in kinetics experiments is to measure the initial velocity, designated V0 (Figure 5-11a). In a typical reaction, the enzyme may be present in nanomolar quantities, whereas [S] may be five or six orders of magnitude higher. If just the beginning of the reaction is monitored (often no more than the first few seconds), changes in [S] can be limited to a small percentage, and [S] can be regarded as constant. V0 can then be explored as a function of [S], which is adjusted by the investigator. The effect on V0 of varying [S] when the enzyme concentration is held constant is shown in Figure 5-11b. At relatively low concentrations of substrate, V0 increases almost linearly with an increase in [S]. At higher substrate concentrations, V0 increases by smaller and smaller amounts in response to increases in [S]. Finally, a point is reached beyond which increases in V0 are vanishingly small as [S] increases. In this plateau-like V0 region, the reaction approaches its maximum velocity, Vmax.

Figure 5-11: The initial velocity of an enzyme-catalyzed reaction. (a) A theoretical enzyme catalyzes the reaction S ⇌ P. Progress curves for the reaction (product concentration, [P], vs. time) measured at three different initial substrate concentrations ([S]) show that the rate of the reaction declines as substrate is converted to product. A tangent to each curve taken at time zero (dashed lines) defines the initial velocity, V0, of the reaction. (b) The maximum velocity, Vmax, is indicated as a horizontal dashed line. The straight solid line describes the linear dependence of the initial velocity, V0, on [S] at low substrate concentrations. However, as [S] increases, the line in reality becomes nonlinear, as depicted by the curved line. V0 approaches but never quite reaches Vmax. The substrate concentration at which V0 is half maximal is Km, the Michaelis constant. The concentration of enzyme in an experiment such as this is generally so low that [S] ≫ [E] even when [S] is described as low or relatively low. At low [S], the slope of the line is defined by V0 = Vmax[S]/Km, and this is where V0 exhibits a linear dependence on [S]. The units shown here are typical for enzyme-catalyzed reactions and help illustrate the meaning of V0 and [S]. (Note that the curved line describes part of a rectangular hyperbola, with one asymptote at Vmax. If the curve were continued below [S] = 0, it would approach a vertical asymptote at [S] = −Km.)

147

J. B. S. Haldane, 1892 –1964

When the enzyme is first mixed with a large excess of substrate, there is an initial period called the pre-steady state, when the concentration of ES (enzyme-substrate) builds up. This period is usually too short to be observed easily, lasting just microseconds, and is not evident in Figure 5-11a. The reaction quickly achieves a steady state in which [ES] (and the concentration of any other intermediates) remains approximately constant over time. The concept of a steady state was introduced by G. E. Briggs and J. B. S. Haldane in 1925. The measured V0 generally reflects the steady state, even though V0 is limited to the early part of the reaction, and analysis of these initial rates is referred to as steady-state kinetics.

Leonor Michaelis, 1875 –1949
Maud Menten, 1879 –1960

The curve expressing the relationship between [S] and V0 (see Figure 5-11b) has the same general shape for most enzymes (approaching a rectangular hyperbola). It can be expressed algebraically by an equation developed by Leonor Michaelis and Maud Menten, called the Michaelis-Menten equation:

The important terms are [S], V0, Vmax, and a constant called the Michaelis constant, Km. All these terms are readily measured experimentally.

The Michaelis-Menten equation is the rate equation for a one-substrate enzyme-catalyzed reaction. It states the quantitative relationship between the initial velocity V0, the maximum velocity Vmax, and the initial substrate concentration [S], all related through the Michaelis constant Km. Note that Km has units of concentration. Does the equation fit experimental observations? Yes; we can confirm this by considering the limiting situations where [S] is very high or very low, as shown in Figure 5-11b. The Km is functionally equivalent to the [S] at which V0 is one-half Vmax. At low [S], Km ≫ [S], and the [S] term in the denominator of the Michaelis-Menten equation (Equation 5-7) becomes insignificant. The equation simplifies to V0 = Vmax[S]/Km, and V0 exhibits a linear dependence on [S]. At high [S], where [S] ≫ Km, the Km term in the denominator of the Michaelis-Menten equation becomes insignificant and the equation simplifies to V0 = Vmax; this is consistent with the plateau observed at high [S]. The Michaelis-Menten equation is therefore consistent with the observed dependence of V0 on [S], and the shape of the curve is defined by the terms Vmax/Km at low [S] and Vmax at high [S].

148

Kinetic parameters are used to compare enzyme activities. Many enzymes that follow Michaelis-Menten kinetics have different reaction mechanisms, and enzymes that catalyze reactions with six or eight identifiable intermediate steps within the enzyme active site often exhibit the same steady-state kinetic behavior. Even though Equation 5-7 holds true for many enzymes, both the magnitude and the real meaning of Vmax and Km can differ from one enzyme to the next. This is an important limitation of the steady-state approach to enzyme kinetics. The parameters Vmax and Km can be obtained experimentally for any given enzyme, but by themselves they provide little information about the number, rates, or chemical nature of discrete steps in the reaction. Nevertheless, steady-state kinetics is the standard language with which biochemists compare and characterize the catalytic efficiencies of enzymes.

Figure 5-11b shows a simple graphical method for obtaining an approximate value for Km. The Km, as noted, can vary greatly from enzyme to enzyme, and even for different substrates of the same enzyme. The magnitude of Km is sometimes interpreted (often inappropriately) as an indicator of the affinity of an enzyme for its substrate. However, Km is not equivalent to the substrate Kd (the simple binding equilibrium for the enzyme substrate) for many enzymes.

149

The term Vmax depends on both the concentration of enzyme and the rate of the rate-limiting step in the reaction pathway. Because the number of steps in a reaction and the identity of the rate-limiting step can vary, it is useful to define a general rate constant, kcat, to describe the limiting rate of any enzyme-catalyzed reaction at saturation. If the reaction has several steps and one is clearly rate-limiting, kcat is equivalent to the rate constant for that limiting step. When several steps are partially rate-limiting, kcat can become a complex function of several of the rate constants that define each individual reaction step. In the Michaelis-Menten equation, Vmax = kcat[Et], where [Et] is the total concentration of enzyme, and Equation 5-7 becomes:

The constant kcat is a first-order rate constant and hence has units of reciprocal time. It is equivalent to the number of substrate molecules converted to product in a given unit of time on a single enzyme molecule when the enzyme is saturated with substrate. Hence, this rate constant is also called the turnover number. A kcat may be 0.01 s−1 for an enzyme with an intrinsically slow function (some enzymes with regulatory functions act very slowly) or as high as 10,000 s−1 for an enzyme catalyzing some aspect of intermediary metabolism.

In cells, there are many situations in which enzyme activity is inhibited by specific molecules, including other proteins. From a practical standpoint, the development of pharmaceutical and agricultural agents almost always involves the development of inhibitors for particular enzymes. Some aspects of enzyme inhibition are reviewed in Highlight 5-1.

HIGHLIGHT 5-1 A CLOSER LOOK: Reversible and Irreversible Inhibition

Enzyme inhibitors are molecules that interfere with catalysis by slowing or halting enzyme reactions. The two general categories of enzyme inhibition are reversible and irreversible. There are three types of reversible inhibition: competitive, uncompetitive, and mixed.

A competitive inhibitor competes with the substrate for the active site of an enzyme (Figure 1a). While the inhibitor (I) occupies the active site, it prevents binding of the substrate to the enzyme. Many competitive inhibitors are structurally similar to the substrate and combine with the enzyme to form an EI complex, but without leading to catalysis. Even fleeting combinations of this type will reduce the efficiency of the enzyme. The two other types of reversible inhibition, though often defined in terms of one-substrate enzymes, are in practice observed only with enzymes having two or more substrates. An uncompetitive inhibitor binds at a site distinct from the substrate active site and, unlike a competitive inhibitor, binds only to the ES complex (Figure 1b). A mixed inhibitor also binds at a site distinct from the substrate active site, but it binds to either E or ES (Figure 1c).

All of these inhibition patterns can be analyzed with the aid of a single equation derived from the Michaelis-Menten equation:

where α and α′ reflect the interaction of an inhibitor with the free enzyme (through KI) and with the ES complex (through ), respectively. These terms are defined as:

and

FIGURE 1 The three types of reversible inhibition. (a) Competitive inhibitors bind to the enzyme’s active site. (b) Uncompetitive inhibitors bind at a separate site, but bind only to the ES complex. (c) Mixed inhibitors bind at a separate site, but may bind to either E or ES.

For a competitive inhibitor, there is no binding to the ES complex, and α′ = 1. For an uncompetitive inhibitor, there is no binding to the free enzyme (E), and α = 1. For a mixed inhibitor, both α and α′ are greater than 1. Each class of inhibitor has characteristic effects on the key kinetic parameters in the Michaelis-Menten equation, as summarized in Table 1. The altered Km or Vmax measured in the presence of an inhibitor is often referred to as an apparent Km or Vmax.

Figure 1: Effects of Reversible Inhibitors on Apparent Vmax and Apparent Km

Many reversible enzyme inhibitors are used as pharmaceutical drugs; two examples are shown in Figure 2. The human immunodeficiency virus (HIV) encodes a DNA polymerase that can use either RNA or DNA as template; this enzyme is a reverse transcriptase (discussed in Chapter 14). It uses deoxynucleoside triphosphates as substrates, and it is competitively inhibited by the drug AZT—the first drug to be used in treating HIV infections. Similarly, quinolone antibiotics widely used to treat bacterial infections are uncompetitive inhibitors of enzymes called topoisomerases (described in Chapter 9).

FIGURE 2 Examples of inhibitors with medical applications. AZT (3′-azido-3′-deoxythymidine) is used in the treatment of HIV/AIDS; ciprofloxacin is a quinolone antibiotic.

An irreversible inhibitor can bind covalently with or destroy a functional group on an enzyme that is essential for the enzyme’s activity, or it can form a particularly stable noncovalent association. The formation of a covalent link between an irreversible inhibitor and an enzyme is common. Because the enzyme is effectively inactivated, irreversible inhibitors affect both Vmax and Km. An inhibitor that does not form a covalent link but binds so tightly to the enzyme active site that it does not dissociate within hours or days is also effectively an irreversible inhibitor. Given that enzyme active sites bind most tightly to the transition state of the reactions they catalyze, a molecule that mimics the transition state can be a tight-binding inhibitor. Inhibitors designed in this way are called transition state analogs. Many drugs used to treat people with HIV/AIDS are designed in part as transition state analogs that bind tightly to the HIV protease.

Note that uncompetitive and mixed inhibitors should not be confused with allosteric modulators (see Section 5.4). Although the inhibitors bind at a second site on the enzyme, they do not necessarily mediate conformational changes between active and inactive forms, and the kinetic effects are distinct.

150

DNA Ligase Activity Illustrates Some Principles of Catalysis

An understanding of the complete mechanism of action of a purified enzyme requires the identification of all substrates, cofactors, products, and regulators. Moreover, it requires a knowledge of (1) the temporal sequence in which enzyme-bound reaction intermediates form, (2) the structure of each intermediate and each transition state, (3) the rates of interconversion between intermediates, (4) the structural relationship of the enzyme to each intermediate, and (5) the energy contributed by all reacting and interacting groups to intermediate complexes and transition states. As yet, there is probably no enzyme for which we have an understanding that meets all these requirements.

It is impractical, of course, to cover all possible classes of enzyme chemistry, and we focus here on an enzyme reaction important to molecular biology: the reaction catalyzed by DNA ligases. The discussion concentrates on selected principles, along with some key discoveries that have helped bring these principles into focus. We also use the DNA ligase example to review some of the conventions used to depict enzyme mechanisms. Many mechanistic details and pieces of experimental evidence are necessarily omitted; an entire book would be needed to document the rich experimental history of enzyme research.

DNA ligases were discovered in 1967, and reports were published from four different research groups in that year. These enzymes catalyze the joining of DNA ends at strand breaks (also called nicks) that have a phosphorylated 5′ terminus and a 3′ terminus with a free hydroxyl group. In cells, these enzymes provide the critical links between discontinuous segments of replicated DNA (called Okazaki fragments; see Chapter 11) and carry out the final step in most DNA repair reactions (see Chapter 12). Since their discovery, many details of the reaction mechanism of these enzymes have been elucidated. DNA ligases have become essential tools of biotechnology, used by laboratories around the world to covalently join DNA segments to create recombinant DNA (see Chapter 7). RNA ligases have also been characterized; they use a similar reaction mechanism.

DNA ligases make use of two cofactors: Mg2+ ions and either ATP or nicotinamide adenine dinucleotide (NAD+). ATP-dependent DNA ligases are found in eukaryotes, viruses, and some bacteria and archaea. NAD+-dependent ligases are found in most bacteria, as well as in some viruses and archaea, but are not found in eukaryotes. Commonly, NAD+ is a cofactor participating in oxidation-reduction reactions. Its role in DNA ligase reactions is quite different, however, and it parallels the role of ATP in the ATP-dependent ligases.

All DNA ligases promote a reaction that involves three chemical steps (Figure 5-12). In step 1, an adenylate group, adenosine 5′-monophosphate (AMP), is transferred from either ATP or NAD+ to a Lys residue in the enzyme active site. This process occurs readily in the absence of DNA. In step 2, the enzyme binds to DNA at the site of a strand break and transfers the AMP to the 5′ phosphate of the DNA strand. This activates the 5′ phosphate for nucleophilic attack by the 3′-hydroxyl group of the DNA, in step 3, leading to displacement of the AMP and formation of a new phosphodiester bond in the DNA that seals the nick. Each of the three steps has a highly favorable reaction equilibrium that renders it effectively irreversible. In the absence of DNA, the adenylated enzyme formed in step 1 is quite stable, and it is likely that most DNA ligases in a cell are already adenylated and ready to react with DNA. In addition to illustrating the reaction pathway, Figure 5-12 introduces the conventions commonly used to describe enzyme-catalyzed reactions.

Figure 5-12: The DNA ligase reaction. The reaction creates a new phosphodiester bond at the site of a break, or nick, in the DNA. The same series of three chemical steps is used by every RNA or DNA ligase. In each of the three steps, one phosphodiester bond is formed at the expense of another. Steps 1 and 2 lead to activation of the 5′ phosphate in the nick. In the E. coli DNA ligase reaction, AMP is derived from NAD+ rather than ATP, and the reaction releases nicotinamide mononucleotide (NMN) rather than pyrophosphate.

The overall picture of the DNA ligase reaction mechanism is a composite derived from kinetic and structural studies of many closely related enzymes, including those isolated from bacteriophages T4 and T7, bacteria, eukaryotic viruses (that is, viruses with eukaryotic host cells), and mammals. The ligases typically consist of a DNA-binding domain, a nucleotidyltransferase (NTase) domain, and an OB-fold domain. In DNA ligases, the OB fold, normally associated with binding to single-stranded DNA (see Section 5.1), interacts with the minor groove of double-stranded DNA. These domains provide a flexible structure that closes to completely encircle a nicked DNA molecule, with all three domains in contact with the DNA (see Figure 5-12). During step 1, some residues in the OB fold become part of the active site for transfer of AMP to the active-site Lys residue. As the covalent link between the enzyme and AMP is formed, the adenine base of AMP is fixed in a binding site in the NTase domain, where it stays throughout the remaining steps. In steps 2 and 3, a conformational change rearranges this part of the OB fold so that the same residues now face the solvent, while other parts of the OB fold bind to the DNA and interact primarily with the strand adjacent to the 5′-phosphate end of the strand break. At the same time, the conformational change closes the enzyme around the nicked DNA, and the N-glycosyl bond between the adenine base and the ribose moiety of AMP is rotated, thereby realigning the phosphate of AMP for reaction, in step 2, with the 5′ phosphate of DNA. Step 3 follows closely behind step 2 within the same complex.

151

This one example cannot provide a complete overview of the broad range of strategies that enzymes use, but it does serve as a good introduction to the complexity of enzyme reaction mechanisms. In addition, it illustrates the transfer of phosphoryl groups, a reaction catalyzed by protein enzymes and RNA enzymes (ribozymes) involved in almost every aspect of molecular biology—from DNA polymerases and RNA polymerases to nucleases, topoisomerases, spliceosomes, and ligases.

SECTION 5.2 SUMMARY

  • Most enzymes are proteins. They facilitate the reactions of substrate molecules. The catalyzed reaction occurs in an active site, a pocket on the enzyme that is lined with amino acid side chains and, in many cases, bound cofactors that participate in the reaction.

  • Enzymes are catalysts. Catalysts do not affect reaction equilibria; they enhance reaction rates by lowering activation energies.

  • Enzyme catalysis involves both covalent enzyme-substrate interactions and noncovalent interactions. Enzyme active sites bind most tightly to the transition states of the reactions they catalyze.

  • The rates of most enzyme-catalyzed reactions are described by the Michaelis-Menten equation, which relates the key kinetic parameters Vmax, Km, and kcat.

  • A DNA ligase catalyzes a series of phosphoryl transfer reactions to seal nicks in the DNA backbone; its reaction mechanism illustrates several general principles of enzyme-catalyzed reactions.