Enzymatic Reactions

Chapter 4 Enzymatic Reactions


The living cell is a cauldron in which thousands of chemical reactions proceed at the same time. Hardly any of these reactions would proceed at any noticeable rate if their starting materials, or substrates, were simply mixed in a test tube by an overoptimistic chemist. The chemist could possibly force the reactions by increasing the temperature or by using a nonselective catalyst, such as a strong acid or a strong base. But the human body is not in this lucky position because body temperature and pH must be kept within narrow limits.


Therefore living things depend on highly selective catalysts called enzymes. By definition, a catalyst is a substance that accelerates a chemical reaction without being consumed in the process. Because it is regenerated at the end of each catalytic cycle (Fig. 4.1), a single molecule of the catalyst can convert many substrate molecules into product. Only a tiny amount of the catalyst is needed.



The thermodynamic properties of a reaction are related to energy balance and equilibrium, whereas kinetic properties are related to the speed (velocity, or rate) of the reaction. Enzymes do not change the equilibrium of a reaction or its energy balance; they only make the reaction go faster. Enzymes change the kinetic but not the thermodynamic characteristics of the reaction.



The equilibrium constant describes the equilibrium of the reaction


In theory, all chemical reactions are reversible. The reaction equilibrium can be determined experimentally by mixing substrates (or products) with a suitable catalyst and allowing the reaction to proceed to completion. At this point, the concentrations of substrates and products can be measured to determine the equilibrium constant Kequ, which is defined as the ratio of product concentration to substrate concentration at equilibrium. For a simple reaction



the equilibrium constant is



[B] and [A] are the molar concentrations of product B and substrate A at equilibrium.


When more than one substrate or product participate, their concentrations have to be multiplied. For the reaction



the equilibrium constant is



The alcohol dehydrogenase (ADH) reaction provides an example:



   (1)



NAD+ (nicotinamide adenine dinucleotide) is a coenzyme that accepts hydrogen in this reaction (see Chapter 5). The equilibrium constant of the reaction is



   (2)



From Equation (2) we can calculate the relative concentrations of acetaldehyde and ethanol at equilibrium when [NADH] = [NAD+] and pH = 7.0:



   (3)



There is 10,000 times more ethanol than acetaldehyde at equilibrium!


Under aerobic conditions, however, NAD+ is far more abundant than NADH in the cell. When [NAD+] is 1000 times higher than [NADH], Equation (3) assumes the numerical values of



The pH also is important. At pH of 8.0 and [NAD+]/[NADH] ratio of 1000, for example, Equation (3) yields



This example shows that a reaction can be driven toward product formation by raising the concentration of a substrate or lowering the concentration of a product.


To adapt the equilibrium constant to physiological conditions, a “biological equilibrium constant,” K′equ, is used. In the definition of K′equ, a value of 1.0 is assigned to the water concentration if water participates in the reaction, and a value of 1.0 to a proton concentration of 10−7 mol/L (pH = 7.0) if protons participate in the reaction. The K′equ of the alcohol dehydrogenase reaction, for example, is not 10−11 mol/L but 10−4 mol/L. At a pH of 8.0 in the preceding example, the proton concentration would be given a numerical value of 10−1.



The free energy change is the driving force for chemical reactions


During chemical reactions, energy is either released or absorbed. This is described as the enthalpy change ΔH:



(4) image



ΔE is the heat that is released or absorbed. It is measured in either kilocalories per mole (kcal/mol) or kilojoules per mole (kJ/mol, where 1 kcal = 4.184 kJ). By convention, a negative sign of ΔE means that heat is released; a positive sign indicates that heat is absorbed. P is the pressure, and ΔV is the volume change. P × ΔV is the work done by the system. It can be substantial in a car motor when the volume in the cylinder expands against the pressure of the piston, but volume changes are negligible in the human body. Therefore ΔH ≈ ΔE. The enthalpy change describes the difference in the total chemical bond energies between the substrates and products.


Reactions are driven not only by ΔH but by the entropy changeS) as well. Entropy is a measure of the randomness or disorderliness of the system. A cluttered desk is often cited as an example of a high-entropy system. A positive ΔS means that the system becomes more disordered during the reaction. Entropy change and enthalpy change are combined in the free energy change ΔG:



(5) image



where T = absolute temperature measured in Kelvin.


ΔG is the driving force of the reaction. Like ΔE and ΔH in Equation (4), it is measured in kilocalories per mole (kcal/mol) or kilojoules per mole (kJ/mol). A negative sign of ΔG defines an exergonic reaction. It can proceed only in the forward direction. A positive sign of ΔG signifies an endergonic reaction. It can proceed only in the backward direction. At equilibrium, ΔG equals zero.


Equation (5) shows that a reaction can be driven by either a decrease in the chemical bond energies of the reactants (negative ΔH) or an increase in their randomness (positive T × ΔS). Low energy content and high randomness are the preferred states. Like most students, Nature tends to slip from energized order into energy-depleted chaos.


Entropy changes are small in most biochemical reactions, but diffusion is an entropy-driven process (Fig. 4.2, A). There is no making and breaking of chemical bonds during diffusion. Therefore the enthalpy change ΔH is zero. This leaves the T × ΔS part of Equation (5) as the only driving force. Thus diffusion can produce only a random distribution of the dissolved molecules.



The human body is a very orderly system. To maintain this improbable and therefore thermodynamically disfavored state of affairs, biochemical reactions must antagonize the spontaneous increase in entropy.Equation (5) shows that a reaction can reduce entropy (negative T × ΔS) only when it consumes chemical bond energy (negative ΔH). In other words, the human body must consume chemical bond energy to maintain its low-entropy state.


In the example of Figure 4.2, B and C, the cell maintains a sodium gradient across the membrane by pumping sodium out of the cell (negative T × ΔS). Sodium pumping is driven by the hydrolysis of adenosine triphosphate (ATP) to adenosine diphosphate (ADP) and inorganic phosphate (Pi) (negative ΔH). Without ATP the gradient dissipates, the entropy of the system increases, and the cell dies. That is what death and dying are all about: a sharp rise in the entropy of the body.



The standard free energy change determines the equilibrium


The free energy change ΔG is affected by the relative reactant concentrations. It is not a property of the reaction as such. To describe the energy balance of a reaction, the standard free energy change, ΔG0′, must be defined: ΔG0is the free energy change under standard conditions. Standard conditions are defined by a concentration of 1 mol/L for all reactants (except protons and water) at a pH of 7.0. As in the definition of K′equ, values of 1 are assigned both to the water concentration and to the proton concentration at pH 7.


For the reaction



image



the standard free energy change ΔG0′ is related to the real free energy change ΔG by Equation (6):



   (6)



where R = gas constant, and T = absolute temperature measured in Kelvin. The numerical value of R is 1.987 × 10−3 kcal × mol−1 × K−1. At a “standard temperature” of 25°C (298K), Equation (6) assumes the form of



   (7)



At equilibrium, ΔG = 0, and Equation (7) therefore yields



   (8)



The reactant concentrations under the logarithm are now the equilibrium concentrations. Their ratio defines the biological equilibrium constant K′equ



   (9)



Substituting Equation (9) into Equation (8) yields



(10) image



There is a negative logarithmic relationship between ΔG0and the equilibrium constant K′equ (Table 4.1). When ΔG0′ is negative, product concentrations are higher than substrate concentrations at equilibrium; when it is positive, substrate concentrations are higher.


Table 4.1 Relationship between the Equilibrium Constant Kequ and the Standard Free Energy Change ΔG0







































K′equ ΔG0′ (kcal/mol)
10−5 6.82
10−4 5.46
10−3 4.09
10−2 2.73
10−1 1.36
1 0
10 −1.36
102 −2.73
103 −4.09
104 −5.46
105 −6.82



The substrate must bind to its enzyme before the reaction can proceed


Enzymatic catalysis, like sex, requires intimate physical contact. It starts with the formation of an enzyme-substrate complex:



where E = free enzyme, S = free substrate, and E·S = enzyme-substrate complex.


In the enzyme-substrate complex, the substrate is bound noncovalently to the active site on the surface of the enzyme protein. The active site contains the functional groups for substrate binding and catalysis. If a prosthetic group participates in the reaction as a coenzyme, it is present in the active site.


According to the lock-and-key model, substrate and active site bind each other because their surfaces are complementary. In many cases, however, substrate binding induces a conformational change in the active site that leads to further enzyme-substrate interactions and brings catalytically active groups to the substrate. This is called induced fit (Fig. 4.3).



The enzyme’s substrate specificity is determined by the geometry of enzyme-substrate binding. If the substrate is optically active, generally only one of the isomers is admitted. This is to be expected because the enzyme, being formed from optically active amino acids, is optically active itself. A three-point attachment (shown schematically in Fig. 4.4) is the minimal requirement for stereoselectivity.




Rate constants are useful for describing reaction rates


The rate (velocity) of a chemical reaction can be described by a rate constant k:



In this one-substrate reaction, the reaction rate is defined by



(11) image



The rate constant has the dimension s−1 (per second), and the velocity V is the change in substrate concentration per second.


For a reversible reaction, the forward and backward reactions must be considered separately:



   (12)




(13) image



At equilibrium, Vforward = Vbackward. Therefore the net reaction is zero:



(14) image




   (15)



Equation (15) shows that the equilibrium constant Kequ, previously defined as [B]/[A] at equilibrium, is also the ratio of the two rate constants.


The forward reaction in Equation (12) is a first-order reaction. In a first-order reaction, the reaction rate is directly proportional to the substrate concentration. When the substrate concentration [A] is doubled, the reaction rate V is doubled as well. Uncatalyzed one-substrate reactions follow first-order kinetics. A classic example is the decay of a radioactive isotope.


When two substrates participate, the reaction rate is likely to depend on the concentrations of both substrates. This is called a second-order reaction. For



the following is obtained:



(16) image



Doubling the concentration of one substrate doubles the reaction rate; doubling the concentrations of both raises it fourfold.


A zero-order reaction is independent of the substrate concentration. No matter how many substrate molecules are present in the test tube, only a fixed number is converted to product per second:



(17) image



Zero-order kinetics are observed only in catalyzed reactions when the substrate concentration is high, and the amount and turnover number of the catalyst, rather than the substrate availability, is the limiting factor.


Jun 18, 2016 | Posted by in BIOCHEMISTRY | Comments Off on Enzymatic Reactions

Full access? Get Clinical Tree

Get Clinical Tree app for offline access