You are on page 1of 25
Chemical Kinetics The aims of studying chemical kinetics are to determine experimentally the rate of f reaction and its dependence on parameters such as concentration, temperature and catalysts, and to understand the mechanism of a reaction—that is, the number of steps involved and the nature of intermediates formed The subject of chemical kinetics is conceptually easier to understand than some other topics in physical chemistry, such as thermodynamics and quantum mechanics, although rigorous theoretical treatment of the energetics involved is possible only for very simple systems in the gas phase. Nevertheless, the macroscopic, empirical approach to the subject can provide much useful information In this chapter we discuss general topics in chemical kinetics and consider some important examples, including fast reactions. Enzyme kinetics will be treated in Chapter 10. 9.1 Reaction Rates { The rate of a reaction is expressed as the change in reactant concentration with time. Consider the stoichiometrically simple reaction RP Let the concentrations (in mol L~!) of R at times fy and 13 (t2 > 1) be [R]: and [Rh ‘The rate ofthe reaction over the time interval (f2 ~ 1) is given by Because [R]p < [R];, we introduce a minus sign so that the rate will be a positive quantity rate ‘The rate can be expressed also in terms of the appearance of a product, P},~(P], _ API ae 4 At 311 312 Chapter 9: Chemical Kinetics In this case, we have (Pl) > [Ph In practice, we find that the quantity of interest is not the rate over a certain time interval (because this is only an average quantity ‘whose value depends on the particular value of Ad); rather, we are interested in the instantaneous rate. In the language of calculus, as Af becomes smaller and eventually approaches zeo, the rate ofthe foregoing reaction at a specific time fis given by 3 d{R) _ ale) “de dt ‘The units of reaction rates are usually M s~! or M min! For stoichiometrically more complicated reactions, the rate must be expressed in an unambiguous manner. Suppose that the reaction of interest is 2R+P The ratios —d[R]/dt and d{P|/dr still express the rate of change of the reactant and the product, respectively, but they are no longer equal to each other because the reactant is disappearing twice as fast as the product is appearing. For this reason, we write the rate of this reaction as, 1 aR) _ a) eee lamar In general, for the reaction aA +bB + cC+dD the rate is given by LD) “a 1djA)__1.4{B] _1.4(C) ad b dtc dt rate = — (9.1) where the expressions in brackets refer to the concentrations of the reactants and ] products at time r after the start of the reaction, 9.2 Reaction Order ‘The relationship between the rate of a chemical reaction and the concentrations of the reactants is a complex one that must be determined experimentally. Referring to the general equation above, however, we find that usually (but by no means always) the reaction rate can be expressed as rate x [AJ"[B)” = KA}*(B) (9.2) This equation, known as the rave law, tells us that the rate of a reaction is not con- stant; its value at any time, ¢, is proportional to the concentrations of A and B raised to some powers, The proportionality constant, k, is called the rate constant. The rate law is defined in terms of the reactant concentrations, but the rate constant for a given reaction does not depend on the concentrations of the reactants. The rate con- stant is affected only by temperature, as we shall see later. 92 Reaction Order 313 Expressing the rate of a reaction as shown in Equation 9.2 enables us to define the order of @ reaction, We say that the reaction is x order with respect to A and y order with respect to B. Thus, the reaction has an overall order of (x-+ y). It is im- portant to understand that, in general, there is no connection between the order of ‘This must be the case because fa reactant in the rate expression and its stoichiometric coefficient in the balanced _@ chemical equation can be chemical equation. For example, the rate ofthe reaction balanced in many different ways 2N;0s(g) + 4NOx(g) + Ox(9) is given by rate = k[N:0s ‘The reaction is first order in NoOs—not second order as we might have inferred from the balanced equation The order of a reaction specifies the empirical dependence of the rate on con- centrations, It may be zero, an integer, or even a noninteger. We can use the rate law to determine the concentrations of reactants at any time during the course of a rea tion. To do so, we need to integrate the rate law expressions. For simplicity, we shall focus only on reactions that have integral orders Zero-Order Reactions The rate law for a zero-order reaction of the type A= product is given by 4A) _ yay =-4A_ ap = 3) g rate = — 1 - gay? = ) og “The quantity k (M s~!) isthe zero-order rate constant. As you can see, the rate of the reaction is independent of the reactant concentration (Figure 9.1). Rearranging — Equation 9.3, we obtain «“ Figure 91 Cae Pio of rate versus concentration for a zero-order reaction, Integration between 1 = 0 and ¢ =f at concentrations [Alp and (Aj gives [i asi = ial (4) Note that Equation 9.4 gives the time dependence of [A], but cannot be the full description of the factors affecting the rate, To illustrate this point, consider the de- ‘composition of gaseous ammonia on a tungsten surface: (Alp — ke 04) NHs(9) — JNo(q) + 3H2(9) 314 ‘Chapter 9: Chemical Kinetics Under certain conditions, this reaction obeys a zero-order rate law. Such a zero-order reaction can occur if the rate is limited, for example, by the concentration of a cata- lyst. The rate of the reaction is given by rate = k'0A where k’ is a constant, 0 is the fraction of metal surface covered by the adsorbed ammonia molecules, and A is the total catalyst surface area. If the pressure of am- ‘monia is large enough, @ = 1, and the reaction is zero order in ammonia. At sufli- ciently low pressures, however, 0 is proportional to [NH] in the gas phase and the reaction becomes first order in ammonia. Note that the rate will also depend on the amount of catalyst (that is, on the area A), First-Order Reactions A first-order reaction is one in which the rate of the reaction depends only on the concentration of the reactant raised to the first power: Al _ gay rate 5) te = 5) where k (5~!) is the first-order rate constant. Rearranging Equation 9.5, we get dial aN = ka aD ket Integrating between ¢ = 0 and + at concentrations [Alp and [A], we obtain { Ble lea [Al nag 7) [A= (Abe A plot of In({A]/[Alp) versus ¢ gives a straight line whose slope, which is negative, is given by -k (Figure 9.2a). Equation 9.7 shows that in first-order reactions, the de- crease in reactant concentration with time is exponential (Figure 9.2b) jal inh @ o Figure 9.2 Firstorder reaction. (a) Plot based on Equation 9.6 with a slope of ~X. (b) Exponential decay of [A] with time according to Equation 9.7 ) ; i , 92 Rescton ner 315, Radioactive decays fit first-order kinetics. An example is 2BRn + kPo +a where a represents the helium nucleus (He?*), The thermal decomposition of N2Os mentioned earlier is first order in NOs, Another example is the rearrangement of methyl isonitrile to acetonitrile: CHsNC(g) + CHsCN(g) Half-Life of a Reaction. A measure of considerable practical importance in kinetic studies is the half-life (t,;2) of a reaction. The half-life of a reaction is defined as the time it takes for the concentration of the reactant to decrease by half of its original value. For example, in a first-order reaction, as [A] = [Alo/2, t= &1/2 and Equation 3.6 becomes in Ale/? = pap In2 _ 0.69: fun = 92 = O2 os) Thus, the half-life of a first-order reaction is independent of the initial concentration (Figure 9.3). For A to decrease from 1 M to 0.5 M takes just as much time as it does for A to decrease from 0.1 M to 0.05 M. Table 9.1 lists the half-lives of radioactive isotopes that are used extensively in biochemical research and medicine. In contrast to first-order reactions, the half-lives of other types of reaction all depend on the initial concentration. In general, we can show the dependence of halt life on the initial concentration as follows: b tae 09) where 1 is the order of the reaction. 10 08 os wm 0 oak oo( 1 _1__,_ + ‘060100 150200 250800 Figure 9.3 ‘The halfives ofa first-order reaction (A ~>+ product). The initial concentration is arbitrarily setat | M and A reacts with a constant halflife of 50s. 316 Chapter 8: Chemica Kinetics Table 9.1 Hatt.ives of Common Radioisotopes isotope Decay Proves ta iW if B TS ye Ho 4N4+ 1p 5.73 x 10° yr NAHM + 98 15h Be yP- M34 ys us— Nei+ tp 88d @Co Emission of y rays 526 yr ™BTet Emission of 7 rays oh i a WE UXe + 805.4 “The superscript m denotes the excited nuclear energy sate. Example 914 ‘The thermal decomposition of 2,2'-azobisisobutyronitrile (AIBN) ii Ks } Rese ho oot + ™ CH, Hy has been studied in an inert organic solvent at room temperature. The progress of the reaction ean be monitored by the optical absorption of AIBN at 350 nm, The following data are obtained: Us 4 0. 130 2,000 1.26 4000107 6000092 8.000 08 10,000 0.72 12,000 0.65 2 040 where A is the absorbance, Assume that the reaction is first order in AIBN, and calculate the rate constant. ANSWER From Equation 9.6, we have in {AIBN} [AIBN], ‘The difference in absorbance at ¢ = 0 and at f = <0, (Ay ~ Av), is proportional to the initial concentration of AIBN in the solution. Similarly, the difference (4; ~ uc), where CSE La 9.2 Reaction Order = 317. cc Lame etd ee ‘02000 40006000 6000 1.0x10" 12x10" Figure 9.4 ‘The fit ofthe linear equation is y constant, which is equal to the negative slope, is 1.24 x 10~ = 0.000124x ~ 0.00167. Therefore, the first-order rate A; is the absorbance of AIBN at time 1, is proportional to the instantaneous con- ‘centration [AIBN],* The rate equation can now be expressed as 2 ‘ Because Ay = 1.50 nd A. = 040, we have Ande fs a= Awe “3000286 400-0496 6000-0749 000-0987 1000 1.240 100 1482 “The firs-order rate constant can be obtained by plotting the natural logarithmic term versus #, as shown in Figure 9,4, It is given by the slope of the straight line, which is 1.24 x10 These statements hold true if ithe oF no ATBN remains unreacted as ¢ approaches infinity and the absorbance of products does not interfere with that of AIBN at 350 nm. sta 318 Chapter 9: Chemical Kinetics Second-Order Reactions We consider two types of second-order reactions here. In one type, there is just cone reactant, The second type involves two different reactants. The first type is rep- resented by the general reaction : ‘A= products and that rate is | dA] 2 =-4A. ou } rate = TE = aA (9.10) “That is, the rate is proportional to the concentration of A raised to the second power, and k (M-! 5-1) is the second-order rate constant. Separating the variables and in- tegrating, we obtain — | a f d{ Oo) | Seve = KiB 1A) oe.) (BIA ix. ‘where (Alp isthe initial concentration, Thus, a plot of 1/[A] versus ¢ gives a straight fine with a slope equal to k (Figure 9.5a). To derive the half-life of a second-order ‘ equation, we set [A] = [Alp/2 in Equation 9.11 and write : © Figure 9.5 Second-order reaction (@) Plot based on Equation 9.11 () Plot based on Equation 9.14. “The slope gives the rate constant 1 kin +a (Alo. (9.12) ‘As mentioned earlier, except for first-order reactions, the half-lives of all other reac: tions are concentration dependent. “The second type of second-order reactions is represented by A+B — products and dial. _ a) mie dh KAIB] (9.13) “This reaction is first order in A, first order in B, and second order overall. Let gn gz i 2 Rescon Over 318 [A] = [Aly —* [B] = [Bl -* where x (in mol L~!) is the amount of A and B consumed in time t. From Equation ons dial x Atal a_i dt dt MAI = k([Aly — X)(IBlo — Rearranging, we obtain a Toa bythe somewhat eious but straightforward method of integration by part fue tins, we ean obtain the inal resul 1 (Be = Ale Bp —TAl (Alo = »)1Blo BIA . ou) Bh -TAl [ANBlo : nuaton 9.14 was derived by assuming tha [Alp < [Bly [Al = [Bl the solution ; je the same as that in Equation 9.11. (Note that Equation 9.11 cannot be obtained from Equation 9.14 by setting [Alp = [Blo.) A plot of Equation 9.14 is shown in Figure 9.5b. Below are a few examples of second-order reactions Clg) + Hala) + HCKa) + H@) 2NO2(a) ~ 2NO(g) + Or(9) HOH (ag) + Br” (a) |sBr(ag) + OH” (ag) > ‘An interesting special ease of second-order reactions occurs when one of the reactants is present in great excess. An example is the hydrolysis of acetyl chloride: CH;COCI(ag) + H,0(!) + CHyCOOH (ag) + HCI(ag) Because the concentration of water in the acetyl chloride solution is quite high (about pene the conventration of pure water) and the concentration of acetyl chloride is In one liter of wate, Weve are sine onder of | M of les, the amount of water consumed is negligible compared J000 & 18.02 g mol”) or Sith the amount of water originally present. Thus, we can express the rate as 35.5 moles of water. = k{CH;COC!| 9320 Chapter 8: Chemical Kinatios | Table 92 ‘Summary of Rate Equations for A —- Products i _—_— —e———e—e—ee 1 Units of the Order Differential Form Integrated Form Half-Life Rate Constant (Alp ~ (Al = kt [A] =[Aloe™* “For A+ B = product. where k = k'{H,0]. The reaction therefore appears to follow first-order kineties and is called a pseudo-first-order reaction. ‘Table 9.2 summarizes the rate laws and half-life expressions for zero-, first, and second-order reactions. Third-order reactions are known but are uncommon, and so we shall not discuss them. Renaturation of DNA—A Case Study. A wel-known example of a second-order re- action is the renaturation of DNA in solution. Kinetic studies of this process provide See Chapter 16 for the struc- information regarding the sequence of the DNA molecule. In atypical experiment, a ture and composition of DNA. large piece of DNA is broken down into smaller fragments of about the same size by sonication (agitation by ultrasonic vibrations). Heating the solution briefly to 90°C ddenatures the fragments into single strands. As they cool, the single strands recom- bine (renature) to form double-stranded fragments (Figure 9.6). The rate of renatu- i ration depends on the makeup of the DNA molecule. If the molecule has a unique base-pair sequence, then the probability of a single fragment strand meeting its complementary strand in solution is small, and the rate of renaturation is slow. On the other hand, if a DNA molecule has many repeat, or “redundant,” sequences, the concentration of similar strands will be high and so will the rate of renaturation. In the extreme case of synthetic DNA containing, say, only adenine (A) and thymine (T) complementary base pairs, the renaturation rate would be faster than the rates of the other two examples. "The kinetic analysis of DNA renaturation begins with the combination of two complementary strands, A and B, to form a double helix: A+BOAB Because [A] = [B], the rate ofthis second-order reaction is given by rate = &(AI/B) = KAP where kis the second-order rate constant. From Equation 9.11 we write (Al "l=T Tae (9.15) Fe 92 Reacton Ordor 321 Figure 8.6 Figure 96 sation experiment, A native DNA molecule is broken ito double stranded DNA rena about th sue ze by sonication The small fragments ae denature Ye8t fragments of apo yom of the complementary single strands follows second-order kinetics: Arete I DNA molceule has litle or no repeat sequence, then the concentration ot the od agents is very stall compared othe total concentration of he Fagments finde de rate of renatraton wil be ow. In contrast, fora synthetic DY coe ee Ie same basepair (ay A and), the concentration ofthe singlestrand fragments j Seite high and the rate of renaturation will be much faster. ta} st [Alo “1+ [Alok (9.16) where f is the fraction of strands that dissociate. Note that (A]p is the initial con: ‘ ereigon (in moles of nucleotide per liter) of fragment A that is complementary © Frapment B (before denaturation) and [| isthe concentration of the single stn fragment labo in moles of nucleotide per liter. Let Cy be the total concentration (2 at UM gucleotide per liter) of all the single strands before renaturation, Therefore, [Alp is related to Co by the relation & (Alo (0.17) 322 _Chapler 9: Chemical Kinetics where NV is the smallest repeating sequence, called the complexity of the nucleotide pairs (the larger WV is, the more complex the sequence). For example, for the synthetic poly A- poly T DNA, we have NV = 1, because the repeating sequence is one (every 'A-T base pair is the same as the next), and [Aly = Co/2. On the other hand, if a DNA extracted from some organism has no repeating sequence, then 1 is the same as the total number of base pairs present, which may be on the order of 1 x 10° ot dreater, In this case, [Alp is @ very small number because each fragment has a unique sequence Equation 9.16 can now be written as (9.18) T+ GikpN “The halflife of the reaction isthe amount of time it takes for half ofthe single strands ALf=4, the faction diss to renature, so f=} and sociated is equal to the frac- tion reassociated. or (9.19) k/2. Figure 9.7 shows plots of f versus Cot for different DNA samples. , Complexity (nucleotide pairs) ho Le | noe rot se br or rol eo cat < a ronrepettive — fraction Fraction D joe 107 tO 010 t 101001000 10,000 otis Figure 9.7 Plot of fraction reassociated versus Cot (logarithmic scale) for DNA from different sources “The top seale gives the complexity of the base pairs (N). The smallest value of N is unity. which corresponds to the synthetic poly U- poly A. From the Cyt values at f = 4 and the Second-order rate constant, we can calculate N for other DNA, [From R. J. Britten and D.E_ Kobe, Science 161, 529 (1968). Copyright 1968 by the American Association for the ‘Advancement of Science | For poly U- poly A (a synthetic double-stranded RNA molecule that bets larly to DNA), the Coty» value at f=} is 2x 10"6 M s. Because N Equation 9.19 we write simi- , from 2x10 Ms =sx10 ats! We can now use the’ second-order rate constant to determine the value of N for DNA Wea uher sourees. For example, at f=} for calf DNA, Catia is 3 10° M s Therefore, RCo = (5x 108 M7! s1)(3 x 10° M 5) =2x10 We can conclude that this DNA sample has a very high order of complexity. Determination of Reaction Order In the study of chemical kinetics, one of the fis tasks isto determine the order of the seaction, Several methods are available for détermining the order of @ reaction, and we shall briefly discuss four common approaches. 4. Antegration Method. An obvious procedure is 19 measure the concentration of the te Atlemr() at various time intervals of a reaction and to substitute the data into the Caquations listed in Table 9.2. The equation giving the most constant value of the Fae canstant for a series of time intervals is the one that corresponds best to the cortect some of the reaction. In practice, this method is not precise enough to do more than to distinguish between, say, first- and second-order reactions, 2, Differential Method. This method was developed by van't Hoff in 1884. Because the rate of an nth-order reaction (0) is proportional to the mth power of the concen tration of the reactant, we write vial “The methods described above apply only in ideal cases. In practice, determining reaction order can be very difficult because of uncertainty in concentration measure” Tronts (for example, when there are small concentration changes in inital rate deter- rninations) as well asthe complexity of the reactions (for example, when reactions aft Feversible and reactions occur between reactants and products). To a certain extent, the procedure is one of trial and error. The use of computers has significanty facili tated the analysis of kinetic data. ‘Once the order of the reaction has been determined, the rate constant at @ par ticular temperature can be ealeulated from the ratio of the rate and the concentta: reef ihe reactants, cach raised to the power of its order. Knowledge of the order tt rate constant then enables us to write the rate law for the reaction. Taking common logarithms of both sides, we obtain loge = nlogiA] + logk 92 Reacton Order 323, We assume that the second- order rate constant is the same for different single strands 924 Chapter 9: Chemical Kinetics © CO) Figure 98 a) Measurement of the initial rates, rp, of a reaction at different concentrations (b) Plot of logey versus Logi Ale, | ‘Thus, by measuring v at several different concentrations of A, we can obtain the : value of n from a plot of logy versus log(A]. A satisfactory procedure is to measure the initial rates (10) of the reaction for several different starting concentrations of A, fas shown in Figure 9.8. The advantages of using initial rates are (1) that it avoids possible complications due to the presence of products that might affect the order of the reaction and (2) that the reactant concentrations are known most accurately at this time. 3, Half-Life Method. Another simple method of determining reaction order is to find the dependence of the half-life of a reaction on the initial concentration, again using the equations in Table 9.2 or Equation 9.9. Thus, measuring the halflife ofa reaction will help us determine the order of the reaction. This procedure is particularly useful Tor first-order reactions because their half-lives are independent of concentration. ' 4, Isolation Method. If a reaction involves more than one type of reactant, we can keep the concentrations of all but one reactant constant and measure the rate as a function of its concentration, Any change in rate must be due to that reactant alone. ‘Once we have determined the order with respect to this reactant, we repeat the pro- cedure for the second reactant, and so on. In this way, we can obtain the overall order of the reaction. 9.3 Molecularity of a Reaction Knowledge of the order of a reaction is but one step toward a detailed understanding of how a reaction occurs. A reaction seldom takes place in the manner suggested by balanced chemical equation, Typically the overall reaction is the sum of several steps: the sequence of steps by which a reaction occurs is called the mechanism of the reac- tion, To know the mechanism of a reaction is to know how molecules approach one ane eT) a 9.3 Moleculanty of Reaction 325 another during a collision, and how chemical bonds are broken and formed, charges transferred, and so on when the reactant molecules are in close proximity. The mechanism proposed for a given reaction must account for the overall stoichiometry, the rate law, and other known facts. Consider the reaction between nitrogen dioxide and carbon monoxide: NO2(g) + CO(g) + NO(g) + CO2(a) Experimentally, the rate law is found to be rate = k{NO:) ‘The reaction is more complicated than the balanced equation shows because one of the reactants, CO, does not appear in the rate law expression. ‘Whereas the word order reflects the overall change in going from the reactants to products, the molecularty of a reaction refers to a single, definite Kinetic process that ay be only one step in the overall reaction. Evidence shows that the reaction takes place in two steps, as follows: Step (1) NOa(g) +NO2(g) “+ NOx(@) + NOW) Step (2) NOs(g) + CO(g) * NO2{g) + COx(9) Each of these so-called elementary steps describes what actually happens at the mo- Jecular level, How do we account for the observed rate dependence in terms of these two steps? We simply assume that the rate for the first step is much slower than that for the second step (that is, « ka), The overall rate of reaction then is completely Controlled by the rate ofthe first step, which is aptly called the rate-determining step, so rate = ki{NOz]?, where k, =k. Note that the sum of steps (1) and (2) gives us the overall reaction, because the species NO3 cancels out. Such a species is called an imermediate because it appears in the mechanism of the reaction but is not in the Gverall balanced equation. Keep in mind that an intermediate is always formed in an early elementary step and consumed in a later elementary ste. ‘ “The preceding discussion shows that our insight into a reaction comes from an understanding of molecularity, not of order. Once we know the mechanism and the fate-determining step, we can write the rate law for the reaction, which must agree ‘with the experimentally determined rate law. Although most reactions are kinetically ‘ ‘complex, the mechanisms for a number of them are sufficiently well understood to Heer wd in molecular terms. In general, however, proving the uniqueness of a As in a court of law we ask yeshanism is very dificult, or impossible, especially for a complex reaction only for proof beyond a rea- Hee now cxamine three different types of molecularity. Unlike the order ofa sonable doubt reaction, molecularity cannot be zero or @ noninteger. Unimolecular Reat ns. Reactions such as cis-trans isomerization, thermal decomposition, ring opening, q ‘and racemization are usually wnimolecular—that is, they involve only one reactant molecule in the elementary step. For example, the following gas-phase elementary steps are unimolecular: N,Ox(g) —> 2NOxo) ote fe + cH,cH=cH, Hc—on, cyclopropane Propene 326 Chapter 9: Chemical Kinatios Unimolecular reactions often follow a first-order rate law. Because these reactions presumably occur as the result of a binary collision through which the reactant mol- ‘ecules acquire the necessary energy to change forms, we would expect them to be bimolecular processes and hence second-order reactions. How do we account for the discrepancy between the predicted and observed rate laws? To answer this question, let us consider the treatment put forward by the British chemist Frederick Alexander Lindemann (1886-1957) in 1922." Every now and then a reactant molecule, A, col- lides with another A molecule, and one becomes energetically excited at the expense of the other: AtASA+A where the asterisk denotes the activated molecule, The activated molecule can form the desired product according to the elementary step AY product, | Another process that may also be going on is the deactivation of the A’ molecule: A +ASSALA } ‘The rate of product formation is given by d{product] " Ss hla) Al that remains for us to do is to derive an expression for (A'). Because A" is an energetically excited species, it has litle stability and a short lifetime. Its concen- | tration in the gas phase is not only low but probably fairly constant as well. Using this assumption, we can apply the steady-state approximation as follows. The rate of change of {A’] is given by the steps leading to the formation of A‘ minus the steps leading to the removal of A’. According to the steady-state approximation, however, | this rate of change can be treated as zero. Mathematically, we havet da] dt (AL? — Ka[AI[A"] = halt | Solving for (A"], we obtain } Two important limiting cases may be applied to the above equation. At higher A similar treatment was proposed independently and almost simultaneously by the Danish 1 chemist Jens Anton Christiansen (1888-1969). Note thatthe steady-state approximation doesnot always apply to intermediates. ts use must be justified by ether experimental evidence or theoretical considerations, 9.9 Moleculary ofa Reacion 327 pressures (21 atm), most A’ molecules wll be deactivated instead of forming prod- uct, and we have KAA’) > BIA] KalAl> ke ‘The rate in this ease is given by dfproduct| _ kiks Alprodvel _ fifa (a) ak ‘and the reaction is first order in A. On the other hand, if the reaction is run at low pressures (<0.5 atm) so that most A’ molecules form the product instead of being deactivated, the following inequality will hold: KAA) < ela kala] eke ‘The rate now becomes {produc _y, iy dt which is second order in A. Lindemann's theory has been tested for a number of systems and is found to be essentially correct. The analysis for the intermediate case (that is, k-.[AJ[A"] = Ka[A*]) is more complex and will not be discussed here, Bimolecular Reactions ‘Any elementary step that involves two reactant molecules is a bimolecular reac- tion. Some of the examples in the gas phase are H+H:— Ha +H NO; + CO + NO + 2NOCI— 2NO + Ch In the solution phase we have 2CH,COOH — (CHsCOOH), _ (in nonpolar solvents) Fe? + Fe’ + Fe + Fe* Termolecular Reactions Finally, we note that an elementary step that involves the simultaneous encoun- ter of three reactant molecules is called a termolecular reaction. The probability of @ 328 Chapter 9: Chemical Kinetics three-body collision is usually quite small and only a few such reactions are known. Interestingly, they all involve nitric oxide as one of the reactants: 2NO(g) + Xa(g) > 2NOX(a) where X= Cl, Br, or L Another type of termolecular “reaction” involves atomic recombinations in the gas phase; for example, H+H+M—ib+M | 1+14M>h+M where M is usually some inert gas such as Np or Ar. When atoms combine to form Tome moleculy they possess an excess of Kinetic energy, which is converted to Cibvational motion, resulting in bond dissociation. Through three-body collisions; the | M opesies can take away some of this excess encray to prevent the breakup of the diatom molecules. ‘No clementary steps with a molecularity greater than three are known, 9.4 More Complex Reactions \ All the reactions discussed so far are simple in the sense that only one reaction is | taking place in each case. Unfortunately, this condition is often not satisfied in actual practice. Three examples of more complex reactions will now be discussed. ! Reversible Reactions ‘Most reactions are reversible to a certain extent, and we must consider both the forward and reverse rates. For the reversible reaction that proceeds by two elemen- tary steps: | 4 a | le wwe represent the net rate of change in [A] as dial oe = Rill + AlB) | At equilibrium, there is no net change in the concentration of A with time; that is, djAl/de = 0, so that Ay{Al = kB) This expression leads to where X is the equilibrium constant The discussion of the relationship between reaction rates and equilibria is rooted in a principle of great importance in chemical kinetics. The principle of microscopic reversibility states that at equilibrium, the rates of the forward and reverse processes Va ; 1 7 » © owe 98 Fee tn count ofA 3, nies ern esa recias sI8 0 G (@) ka = ke; (b) ka = Ski; (c) ka = 25k Equation (9.25) holds if ky > ki. Under this condition, B molecules are converted to Cas soon as they are formed so [B] is kept constant and low compared to [A] To get an expression for (Cl, we note that at any instant we have (Alp B| + [C). Therefore, from Equations 9.23 and 9.25 we obtain [C= [Al — [A] - B) = [Al = [Ab(t-e*) (9.26) ‘The (k1/kx)exp(~Kit) term is eliminated because it is much smaller than 1 Figure 9.9 shows plots of (Al, (B], and (C] with time for different rate constant ratios. In all cases, [A] falls steadily from [Ap to zero while [C] rises from zero to [Alp. The concentration of B rises from zero to a maximum and then falls to zero. Note that as kz becomes much larger than ky, the steady-state approximation becomes valid over the time period when [B] remains constant (Figure 9.9c). ‘A more complicated but common consecutive reaction is shown below: ne A+BSCSP where P denotes product. This scheme involves pre-equilibrium, in which an inter~ mediate is in equilibrium with the reactants. A pre-equilibrium arises when the rates of formation of the intermediate and of its decay back into reactants are much faster than its rate of formation of products—that is, when k > ke. Because A, B, and C are assumed to be in equilibrium, we can write ky x ki TAB! and the rate of formation of P is given by alr a alC] =bK\AIB) 932 Chapter 9: Chemical Kinatios In Chapter 10 we shall see applications of both the steady-state approximation and the pre-equilibrium treatment to enzyme kinetics. Chain Reactions ‘One of the best-known gas-phase chain reactions involves the formation of hydrogen bromide from molecular hydrogen and bromine between 230°C and 300" H,(g) + Bro(g) + 2HBr(g) “The complexity of this reaction is indicated by the rate equation | {HB a aHa][Br2]!7 T+ {HB s/1Br] (9.27) where a and f are some constants. Thus, the reaction does not have an integral re- detion order. It has taken many experiments and considerable chemical intuition to ‘come up with Equation 9.27. We assume that a chain of reactions proceeds as follows: Br; “5 2Br chain initiation chain propagation chain propagation chain inhibition chain termination The following reactions play only a minor role in determining the rate: H > 2H chain initiation Br+HBr— Br; +H _ chain inhibition H+H— Hb chain termination H4+Br—HBr chain termination For this reason, they are not included in the kinetic analysis. By applying the steady- state approximation to the intermediates H and Br, we can derive Equation 9.27 using the fist five elementary steps (see Problem 9.20), 9.5 The Effect of Temperature on Reaction Rate \ Figure 9.10 shows four types of temperature dependence for reaction rate constants | ‘Tene (a) tepresents normal reactions whose rates increase with increasing tempera- Taos Toe (6) shows a rate that initially increases with temperature, reaches @ max ‘mum, and finally decreases with further temperature rise. ‘Type (c) shows a steady ‘ecrease of rate with temperature. The behavior outlined in (b) and (c) may be sur- prising, because we might expect the rate ofa reaction to depend on two quantities: ‘the number of collisions per second and the fraction of collisions that activate mole- {les for the reaction. Both quantities should increase with increasing temperature.

You might also like