Vous êtes sur la page 1sur 6

70 Ind. Eng. Chem. Fundam.

1986, 25, 70-75

Some Applications of the Generalized De Donder Equation to


Industrial Reactions

Mlchel Boudart
Department of Chemlcal Engineering, Stanford Unlversify, Stanford, California 94305

The generalized De Donder equation relates the ratio of forward and reverse reaction rates to the exponential of
the thermodynamic driving force for reaction. The latter is the affinity divided by 8RT where 8 is the average
stoichiometric number for the reaction, equal to the stoichiometric number for the rate-determining step, if there
is one, and always equal to unity for an elementary step. I n the latter case,the equation provides useful i n f m t l o n
on the kinetic coupling between steps in a chain or catalytic sequence. I f there exists a rate-determining step,
the equation may help in revealing its identity.

Introduction powerful links between the kinetic of elementary steps and


Thousands of rate constants are tabulated for elemen- the rate of catalytic cycles. In turn, this would accelerate
tary reactions involving free radicals (Kondrat’ev, 1970). the transfer of data from industrial catalysis to surface
Many more can be estimated (Benson, 1976). Thus, rates science, and vice versa. Let us now present the main useful
of chain reactions proceeding singly or in networks can be relations connecting micro- and macrokinetics or kinetics
calculated or estimated from kinetic information on ele- and mechanism.
mentary steps, Le., a microkinetic database. In this way The Quasi-Steady-State Approximation
macrokinetic behavior of many free radical chain reactions Any catalytic cycle consists of a sequence of elementary
can be described from microkinetic information. Many processes. Each one proceeds at a net rate:
examples can be found in pyrolysis of hydrocarbons
V i = Ci - gi
(Baronnet and Niclause, 1985), vinyl polymerization
(North, 1966), oxidation (Van Tiggelen, 1968), combustion Steps must be written as they are believed to occur at the
(Glassman, 1977), and chemistry of the troposphere molecular level. Thus, for dissociative chemisorption of
(Hampton, 1973) or the stratosphere (Boudart and N2 on a catalytic site denoted by *, one may write two
DjBga-Mariadasou, 1984). Progress in the numerical so- successive steps
lution of stiff differential equations now makes the qua-
si-steady-state approximation unnecessary in the kinetic +
Nz * -i N*N
analysis of these reactions (Allara and Edelson, 1975). The N*N + * & 2N*
interplay of micro- and macrokinetics in free radical re-
actions provides a prime example of the feedback loop or perhaps in a single step
between industrial chemistry dealing with, e.g., steam N2 + 2* s 2N*
cracking and academic chemical kinetics dealing with rate
constants of elementary steps, e.g., of free radical reactions. but certainly not
By contrast, in the case of catalytic kinetics, the old 1/2N2+ * e N*
tradition of inferring microkinetic information from ma-
crokinetic observation continues tQ flourish in spite of the If the stoichiometric equations for each step in a catalytic
well-known ambiguities of mechanisms derived from ki- sequence are summed up side by side, the sum reproduces
netics, the so-called kinetic mechanisms (Boudart and the stoichiometric equation for the overall reaction, pro-
DjBga-Mariadassou,1984). The reason is clear. Until very vided that each step is taken u, times where ui is the
recently, there was very little known about the microki- stoichiometric number of that step. The value of cri de-
netics of catalytic reactions. The situation is now changing pends on the arbitrary way in which the stoichiometric
in the case of homogeneous catalysis (Hjortkjaer, 1982). equation for the overall reaction has been written, e.g.
As to heterogeneous catalysis, recent advances in surface N2 + 3H2 = 2NH, or 1/2N2+ 3/2H2= NH,
science have made available a small but rapidly increasing
number of rate constants of elementary steps on so-called The quasi-steady-state approximation (QSSA), in the
well-defined surfaces, mostly metallic single crystals form first clearly enunciated by Christiansen (1953), states
(Madix, 1979). Such data are now available for adsorption, that at the kinetic steady state
surface reaction, and desorption (Boudart and DjBga- UiU = u’i - fii
Mariadassou, 1984). As these data accumulate, it seems (1)
opportune to review the available relations linking micro- for all steps in the cycle proceeding at a net overall rate:
and macrokinetics and to examine the general kinetic v=v‘-v’
structure of catalytic cycles. In particular, some special
kinetic features of cycles in heterogeneous catalysis will If the rate, u, is referred to the number of catalytic sites,
be discussed. Indeed, it is hoped that the kinetic it becomes the turnover rate, ut. The inverse of ut is a time,
knowledge made available by studies on single crystals will the turnover time, T. It seems intuitively clear that the
quickly find its application in the mainstream of catalytic kinetic steady state will not be established before a re-
research as a result of a keener awareness of the links laxation time of the order of T. In any event, a substance
between micro- and macrokinetics. It is also hoped that is not a catalyst unless it turns over more than once. An
this review may stimulate the discovery of new, more example of relaxation time (Figure 1)to reach the steady
0 1986 American Chemical Society
Ind. Eng. Chem. Fundam., Vol. 25, No. 1, 1986 71

6
that, for any elementary process, the rate constants, and
ki,are related to the equilibrium constant, K ; ,by means
of the relation
&/Ei = Ki (4)
As is well-known, this relation was anticipated by Ar-
rhenius and others, in less fundamental terms. Relation
4 is very useful as it gives access to a rate constant that
w
o
0.6 t- 1 may be difficult to measure, provided that the rate con-
stant for the step in the reverse direction is available to-
$ 2
gether with the equilibrium constant. The unexpected
51
lJ?k validity of (4) over 40 orders of magnitude of rate was
u. 3 0.4
8 pointed out by Johnston (1968) for an elementary process
Zk!
O W O,+M+ O + O + M
6a u6 O.* which may not seem amenable at first glance to the for-
[L
LL
malism of transition-state theory.
o 20 40 60 ao 100 120
TIME/S The De Donder Irreversibility Relation
Figure 1. Relaxation to the steady-state decomposition of ammonia De Donder's concern with irreversibility and affinity, A,
(pressure 2.4 X Pa) on a molybdenum foil at 1060 K (Boudart which is identical with -AG, the Gibbs free energy of re-
et al., 1982): upper curve, rate vs. time; lower curve, fraction of action with the minus sign, led him (De Donder, 1927) to
surface covered with nitrogen vs. time. a very useful relation that is similar in form to Eyring's
relation above.
state relates to the decomposition of ammonia at low
pressure (-lo+ Pa) and high temperature on a molyb- Ci/Ci = exp(Ai/Rr) (5)
denum foil in an ultrahigh-vacuum chamber used as a Indeed, to obtain Eyring's relation (4) from the relation
continuous stirred tank reactor (Boudart et al., 1982). The of De Donder (5), all that is necessary is to specify
reaction is run at the steady state. Then, the foil is standard-state conditions for which rates of elementary
"flashed" and N* leaves the foil. The upper curve shows process become equal to rate constants and affinities be-
the return to the steady-state rate as measured by a mass come standard affinities so that exp(Aio/RT) = Ki with
spectrometer. The relaxation time is of the same order the superscript denoting standard state. In actual fact,
as the value of the turnover time at the steady state. As it is ( 5 ) which is obtained most simply from Eyring's re-
the steady state is approached, the fraction of surface lation (4) and transition-state theory (Boudart and
covered with nitrogen, as measured by Auger electron DjBga-Mariadassou, 1984). The De Donder relation ex-
spectroscopy, also reaches its steady-state value (lower presses the kinetic irreversibility of reaction in terms of
curve of Figure 1)with the same relaxation time as that its thermodynamic driving force. In the form of eq 5, it
for the rate of reaction. is valid only for elementary steps.
The Product Rule A relation similar to (5)-relat$s forward and reverse
Following Temkin (Horiuti, 1973),the following identity currents for an electrode, i and i, to the overvoltage, 7
t -
can always be written for some or for all of the steps in i / i = exp(aq/FRT)
a catalytic sequence
where F is the Faraday constant and a is an empirical
(ij1 - v'l)i;2v'3...v', +
v'1(4 - ii,)v'3 ...v', + ... +
coefficient. This expression gives a measure of the irre-
v'&73*..(v'n - fin) = hi- hi(2)
Pl i=l
versibility of the electrode reaction in terms of its driving
force. It was first derived by Volmer and Butler (Bockris
Substitution of (1) for each step into (2) yields and Reddy, 1970).
Let us note a simple application of De Donder's relation.
u= [fi& - fiCi]/D Since in the expression (1) of the QSSA Ci > Si as long as
i=l i=l the reaction moves forward with u > 0, it follows from (5)
that, for a sequence of elementary steps at the steady state,
+
where D = 1 ~ ~ 4 v ' ~ . . . vv'l~2v'3...Cn
'~ + ... +
4C2&.,~,,. If the the affinity, Ai, is positive for all steps. This statement
catalytic sequence consists of n steps, u = v' - u is its net may sound evident, but it is useful to stress it (Boudart,
rate and, from the above 1983) as a result of statements in the literature in which
no distinction is made between Ai and its standard value,
Ai". Of course, a negative value of Aio does not prevent
a step from proceeding forward until it reaches its unfa-
Hence vorable equilibrium. We shall return to this point later
as we talk about kinetic coupling of catalytic cycles.
The Generalized De Donder Irreversibility
(3) Relation
For any catalytic chain sequence, let us introduce Tem-
kin's (Temkin, 1971) average stoichiometric number, 8,
This product rule will be used in what follows. defined as
Arrhenius' Relation between Rate Constants and 5 = CaiAi/ZAi = A / C A i (6)
i i i
the Equilibrium Constant
It is a straightforward consequence of Eyring's ther- where the summation extends to all steps and A is the
modynamic formulation of reaction rates (Eyring, 1935) affinity of the overall reaction.
72 Ind. Eng. Chem. Fundam., Vol. 25, No. 1, 1986

Substitution of the De Donder relation (5) into Temkin’s I I I


3
product rule (3) followed by the use of definition (6) leads
to the generalized De Donder relation
i3/C = exp(A/@Rn (7) 2

which expresses the kinetic irreversibility of the overall


reaction proceeding at a net rate
- 1
Im
u=u’-6 (8) c

N
L
in terms of its thermodynamic driving force. The affinity -0
is modified by the average stoichiometric number of the
reaction.
.
$
>
0

Rearranging (7) with the use of (8) gives simply


-I
u = u’[l - exp(-A/@RT)] (9)
Thus, the net rate of a reaction is given by its forward rate
multiplied by a thermodynamic potential (Happel, 1972) -2
given by the expression between brackets in (9).
Linear Relation between Rate and Affinity near -0 4 0 04 08
Equilibrium A IRT

Sufficiently near equilibrium, where AIaRT << 1so that Figure 2. Linear relation between net rate u and affinity A for the
the exponential in (9) can be expanded in series with only reaction CsHl2 = CsHs + 3Hz on both sides of equilibrium. Data are
from Prigogine et al. (1948), and the least-squares line is from Bou-
the first term retained in the expansion, eq 9 becomes dart et al. (1985).
u = u’(A/sRT) (10)
is not evident that the linear relation (11)will be obeyed
Further, when AIaRT tends to zero, ?I tends to u,, the rate except perhaps so near equilibrium that the h e a r law may
at equilibrium or the exchange rate (Wagner, 1970) be a trivial one (Garfinkle, 1983). It is therefore interesting
u = u,(A/@RT) (11) that linearity between u and A was established experi-
mentally by Prigogine et al. (1948) on both sides of the
The linear relation (11)between rate and affinity was first equilibrium of the reaction
formulated by Horiuti (1953). It was then used by Horiuti
and Nakamura (1967) in the case of ammonia synthesis CCH12 = C&6 + 3H2 (15)
Nz + 3H2 = 2NH3 (12) taking place on a nickel catalyst. In fact, it has been
with the idea of obtaining the stoichiometric number, g d , noticed (Boudart et al., 1985) that the small standard
of the rate-determining step (rds) of that reaction. The deviation of the data from the best least-squares straight
existence of an rds predicates that the affinity of all steps line (Figure 2) suggests, as follows from (ll),a rather high
is zero (or essentially zero) except that of the rds. In such value of a, equal to at least 3. Let us assume that it is 3
a case, it follows from (6) that and that there exists an rds. As suggested by Herbo (1942),
the rds involved is the dehydrogenation of cyclohexane
?? = g d (13) (C6H12)to cyclohexene (C6H10):
Thus, if there exists an rds, (11)becomes
u = u,(A/gdRT) (14)
Then, if the reaction proceeds further by the equilibrated
Thus, it is possible to determine g d by measuring the disproportionation of C6H10
net rate, u , near equilibrium and the exchange rate, u,, at
equilibrium by means of a suitable tracer. It was first 3C6HIO C6H6 t 2C6HIZ (17)
reported (Horiuti and Nakamura, 1967) that g d = 2 for
ammonia synthesis when the reaction is written as in (12). where the symbol =8= denotes zero net rate, or an
However, later work suggested that g d = 1 (Mars et al., equilibrated reaction. This reaction proceeds very rapidly
1960). If the latter value is correct, it follows from any on nickel. The value of ad should be equal to 3 since the
dissociative mechanism of catalytic ammonia synthesis rds implied in (16) followed by (17) must take place three
that the rds is the chemisorption of N2 at the surface of times for reaction 15 to turn over once. This conclusion
the catalyst. By the words “dissociative mechanism” is is supported by the excellent fit of the data up to high
meant one in which both N2 and H2 first dissociate at the values of A. Details are given in Boudart et al. (1985).
surface; then successive recombination steps take place Thus, not only is the linear relationship (11)of Horiuti
between adsorbed species until NH, is released from the valid near equilibrium but it can provide useful mecha-
surface (Boudart, 1981). nistic information without the need of rate measurements
This example is just one of the many nice illustrations at equilibrium.
of the fact that work on ammonia synthesis, the epitome Virtual Pressure [N,],in Ammonia
of an industrial catalytic reaction, has fed back many new Decomposition
general concepts to fundamental catalysis science. This
point has been discussed in detail elsewhere (Boudart, This is another interesting application of the De Donder
1978; Timm, 1974). relation. The overall reaction with equilibrium constant
K is
Validity and Further Use of the Linear Relation
K
between v and A 2NH3 = N2 + 3H2
Since u’, a variable in (IO), can be replaced by u,, a
constant in ( l l ) , only when the affinity tends to zero, it On many catalysts, at not too high temperatures and not
Ind. Eng. Chem. Fundam., Vol. 25, No. 1, 1986 73

too low pressures, the accepted mechanism is Kinetic Coupling of Elementary Steps in a
Catalytic Sequence
2NH3 + 2" -9- 2N" + 3H2
To drive a thermodynamically unfavorable step, (Aio <
0), two possibilities exist: either the concentration of a
2N* 2" + N,
reactant should be made large enough, as was the case
above in ammonia decomposition, or a product must be
x2
kept at a sufficiently low concentration level by removing
where the first reaction is equilibrated and the second is it is a subsequent step faster than it can return to the
the rds on a surface where N is the most abundant reaction original reactants. This can be done by kinetic coupling
intermediate. (Boudart, 1976). Let us explain briefly the nature of this
Define a virtual pressure [N2], as the value of [N,] re- kinetic coupling in the free radical reaction, H2 Br2 = +
quired to obtain at virtual equilibrium the value [N],, 2HBr, that takes place according to the classical mecha-
pertaining to the steady state of the reaction nism following an equilibrium between Br and Br,, which,
by the way, does not take place through an elementary step
v'2 = Ldd(aN*,ss) = 62 = k'ja(a*,ss)[Nzle (18)
where we have carefully used a thermodynamic activity, v2Br2 =& Br
a, for the surface species and have not specified the form
+ -
kI
+
1
of the functions, fa and fd, that depend on surface activity Br H, HBr H (k,/k-,=K,)
T
at the steady state, uN*,,,. Thus, in all generality, we can propagation
write with the help of (18) H -I- Br,
&
2 HBr + Br

The first step in the propagation sequence should be se-


verely limited by equilibrium since it is sizably endo-
thermic and essentially isentropic. At equilibrium the
Applying the De Donder relation and noting that A2 = A, concentration of hydrogen, [HI,, is obtained easily by
since step 2 is the rds, and expressing A in terms of K and noting that
steady-state values of gaseous concentrations or fugacities,
we get from (19) Kl[Br], = K1K[Br2]1/2= Ko[Br2]1/2
where KOis the equilibrium constant of
"
- =-= exp(A,/RT) = exp(A/RT) =
fi2 [N21,, '/2Br2 + H, HBr + H

~~NH31,,2/[N2lss~H21ss3
so that
Finally, for example
[HI, = Ko[Br2]1/2
[N2Ie= K[NH3],,2/[H2],s3 N 6400 bar at 673 K
Now, if we start from a stoichiometric mixture, we have
with ammonia and dihydrogen at 1 bar each. at half-reaction [H,] = [Br,] = [HBr]. The De Donder
The concept of virtual pressure, presented here as a relation gives, at half-reaction
consequence of the De Donder relation, originated with Cl/Cl = exp(Al/RT) = [H],/[H],, = 1 + k z / k l (20)
Temkin and Pyzhev (1940), who introduced the related
concept of fugacity of adsorbed species. The concept of since
virtual pressure was further developed by Kemball(l966).
Although it is yet another concept that came about as a [HI,, = k,[BrI,/(k-, + k2)
result of fundamental studies of ammonia synthesis, it is as readily obtained from the steady-state condition (1)
general and useful. u = v'l - v'l = v'p
Thus, by using ammonia as a nitriding agent, it is pos-
sible to form iron nitride at low pressure and at the high The meaning of eq 20 is that the thermodynamically un-
temperatures required so that diffusion of nitrogen into favorable step proceeds forward with a decided irrevers-
the bulk of the metal proceeds at reasonable speed. Yet, ibility even at half-reaction, since
without the high virtual pressure of nitrogen, the nitride kZ/k-1 10 N v'l/C1
would not be stable at those high temperatures. Another
consequence of eq 19 is that the irreversibility of the as- because, at steady state, H atoms are "pumped down" from
-
sociative desorption of nitrogen is very large under the
conditions specified, since 4/17, 6400 during the
steady-state decomposition of ammonia. This marked
irreversibility of the desorption step means that the
H + Br, HBr + Br -
their equilibrium concentration by the subsequent step

As a result, the affmity for the difficult step at steady state,


steady-state concentration of surface nitrogen must be Al,,,, is positive although the standard affinity, AlO,is
considerably larger than the concentration of surface ni- negative. Thus, in calculations of the standard affinity,
trogen that would be in equilibrium with the steady-state Ai', of steps in a catalytic cycle (Goddard, 1985), the oc-
concentration of gaseous dinitrogen. This conclusion was currence of negative values for Aio should not prevent the
checked experimentally in the case of decomposition of cycle from turning over at an acceptable rate, if kinetic
ammonia on a molybdenum foil at high temperatures but coupling is favorable.
low pressures, at which the steady-state concentration of Yet, in the above example, because k 2 / k - , is known to
surface nitrogen could be monitored by Auger electron be only -10, the step Br + H2 F? HBr + H is nonetheless
spectroscopy (Boudart et al., 1982). Another consequence reversible and HBr is an inhibitor of the overall rate, al-
of the irreversibility of the desorption step in ammonia though the overall reaction itself is not limited by equi-
decomposition is that the rate is not inhibited by nitrogen librium. The inhibiting effect of the product HBr on the
as might be the case if the desorption step was limited by rate of a reaction which is run under irreversible conditions
equilibrium. We will return to this point later. was indeed a formidable puzzle for Bodenstein and Lind
74 Ind. Eng. Chem. Fundam., Vol. 25, No. 1, 1986

(1907), who first observed it but did not know the mech- mediate is quite valuable. Note that the lack of inhibition
anism of the reaction. Let us now return to the inhibition of the reaction rate by the other reaction product, di-
or lack of inhibition of the rate of a catalytic reaction by hydrogen, follows ipso facto from its inability to compete
a reaction product. with the abundant reactive intermediate.
In yet another example, when an order of reaction is zero
Mechanistic Information Obtainable from the with respect to a certain component, the lack of inhibition
Presence or Absence of Product Inhibition of the rate by the product issuing from the species related
In a catalytic reaction, the desorption step or steps that to the zero-order component is not unexpected. Thus, for
release products are moderately or strongly endothermic. ethylene hydrogenation on platinum (Emmett, 1962),the
Although desorption proceeds with an increase in standard zero order with respect to ethane appears related to the
entropy, a desorption step is normally expected to be lim- zero order with respect to ethylene. Although the nature
ited by equilibrium. Thus, it is likely that, at the steady of the surface species that saturates the catalytic sites is
state, the desorption steps will be in quasi-equilibrium. If not known (de Boer, 1957; Cimino et al., 1954),there is no
the surface species that desorbs occupies a significant doubt that such a species exists.
fraction of the catalytic sites, it follows that the product A third explanation of the lack of inhibition of the
of desorption will inhibit the rate of the overall reaction. forward rate of reaction by a product is that the desorption
This is the case of ammonia synthesis on iron that is in- of the product from the surface of the catalyst is indeed
hibited by ammonia (Ozaki and Aika, 1981), of oxidation exothermic. This implies, of course, that adsorption of that
of sulfur dioxide on vanadium oxide that is inhibited by species would be endothermic and therefore very unfa-
sulfur trioxide (Bodenstein and Fink, 19071, and of the vorable (Aio< 0) since adsorption is normally accompanied
hydrodesulfurization of sulfur-containing hydrocarbons on by a loss of entropy (de Boer, 1957). This explanation
cobalt-molybdenum sulfide that is inhibited by hydrogen suggests itself in the case of the hydrogenolysis of ethane
sulfide (Satterfield and Roberts, 1968). (Cimino et al., 1954)
When a reaction product is not an inhibitor, three ex-
CZH, + Hz = 2CH4
planations are possible. The first is that the surface species
that desorbs occupies a kinetically insignificant fraction and the methanation reaction (Vannice, 1982)
of the catalytic sites. This is true for ammonia decom- CO + 3Hz = CH4 + HzO
position at low pressures and high temperatures (Boudart
et al., 1982) or at high pressures and low temperatures on transition metals. In both cases, methane is not an
(Emmett, 1962), where dihydrogen goes not inhibit the rate inhibitor of the rate. That adsorption of methane on
of ammonia decomposition. Correspondingly, nitrogen is metals is endothermic is not surprising in view of the
the most abundant reactive intermediate and hydrogen known difficulty of methane activation. It may be noted
does not compete effectively for catalytic sites. that methane activation is not assessed correctly by
A second explanation applies to dinitrogen in ammonia methane-deuterium exchange.
decomposition. In this case, dinitrogen is not an inhibitor Thus, mere inspection of a rudimentary rate equation
of the rate. As noted above, this is attributed to the fact or even the elementary knowledge whether a product in-
that, as a result of the high virtual pressure of nitrogen, hibits the rate or not contains nontrivial mechanistic in-
the surface concentration of adsorbed nitrogen consider- formation. Ultimately, the qualitative or quantitative
ably exceeds the equilibrium value corresponding to the meaning of this information is contained in the De Donder
steady-state pressure of dinitrogen. Hence, the desorption relation that measures the irreversibility of a step.
of nitrogen proceeds essentially irreversibly and no in- Conclusions
hibition by nitrogen is observed. The De Donder relation for an elementary step and its
A similar situation applks to the dehydrogenation of generalization for a catalystic or chain sequence contain
methylcyclohexane, M, to toluene, T, on a platinum-re- useful mechanistic information. Thus, the relation pro-
forming catalyst (Sinfelt et al., 1960). I t is found that vides a useful bridge between macro- and microkinetics.
toluene does not inhibit the rate. The preferred mecha- Its measure of irreversibility, far from or near equilibrium,
nism (Sinfelt et al., 1960) is one in which M adsorbs ir- can be used to understand or determine kinetic coupling,
reversibly in an opening step and T*, the most abundant virtual pressure, the difference between equilibrium and
reactive intermediate, desorbs irreversibly in a final step. steady-state concentrations, the inhibition of a catalytic
Again, why is the desorption step essentially irreversible? reaction by a reaction product or the lack of it, and the
Our explanation is that, at the steady state, the surface stoichiometric number of the rate-determining step, if
concentration of toluene is substantially greater than that there is one. The De Donder relation is obviously a useful
corresponding to the equilibrium concentration that could bridge between thermodynamics and kinetics. The present
be reached at equilibrium between the surface and toluene paper underscores its usefulness in building bridges be-
at the steady-state pressure of the reaction. In other words, tween macro- and microkinetics, i.e., between the rates of
the virtual pressure of toluene resulting from kinetic the overall reaction and the rates of its component steps,
coupling is much higher than its steady-state pressure. or finally, between industrial data and fundamental in-
This explains clearly why addition of benzene to the feed vestigations.
(Sinfelt et al., 1960) hardly inhibited the rate of M * T
+ 3Hz. To compete effectivelywith toluene for the surface Acknowledgment
benzene should be added at a pressure corresponding to This work was carried out as part of a continuing NSF
the virtual pressure of toluene, not its real pressure. These program, currently under Grant NSF-CBT 8219066.
remarks are qualitative but could be quantified by
measuring the steady-state and equilibrium concentrations Literature Cited
of adsorbed toluene. Then the irreversibility of the de- A k a , D. L.; Edelson, D. Int. J . Chem. Kinet. 1975, 7, 479.
Baronnet, F.; Niclause, M. Ind. fng. Chem. Fundam.. this issue.
sorption step could be calculated. But even without this Benson, S. W. "Thermochemical Kinetics"; Wlley: New York, 1976.
knowledge, the mechanistic information inferred from the Bockris, J. O.'M.; Reddy, A. K. N. "Modern Electrochemistry"; Plenum Press:
New York, 1970; Voi. 2, Chapter 8.
mere lack of inhibition of the rate of reaction by the Bodenstein, M.; Fink, C. G. 2.fhys. Chem. 1907, 60, 1.
product issuing from the most abundant reactive inter- Bodenstein, M.; Lind. S. C. Z .fhys. Chem. 1907, 57, 168.
Ind. Eng. Chem. Fundam. 1986, 25, 75-84 75

Boer, J. H. de A&. Catal. Re/. Subj. 1957, 9 , 472. Kembali, C. Discuss. F a r a h y SOC. 1066, 41, 190.
Boudart, M. CHEMTECH 1978. 8 , 231. Kondrat’ev, V. N. “Rate Constants of Gas Phase Reactions”; Nauka: Mos-
Boudart, M. Catal. Rev.-Sci. Eng. 1981, 23, 1. cow, 1970.
Boudart, M. J. Phys. Chem. 1983, 8 7 , 2786. Madix, R. J. “Chemistry and Physics of Solid Surfaces”; Vanseiow, R., Ed.;
Boudart, M.; Dj6ga-Mariadassou, G. “Kinetics of Heterogeneous Catalytic CRC Press: Boca Raton, FL, 1979; Vol. 2, p 63.
Reactions”; Princeton University Press: Princeton, NJ, 1984. Mars, P.; Schoten, J. J. F.; Zwietering, P. “The Mechanism of Heterogeneous
Boudart, M.; Egawa, S.; Oyama, S. T.; Tamaru, K. J. Phys. Chem. 1982, Catalysis”; de Boer, J. H., et ai., Eds.; Elsevier: Amsterdam, 1960; p 66.
78, 987. North, A. M. “The Kinetics of Free Radical Polymerization”; Pergamon Press:
Boudart, M.; Loffier, D. G.; Gottifredi, J. C. I n t . J. Chem. Kinet. 1985, 77, Oxford, 1966.
1119. Ozaki, A.; Aika, K. “Catalysis, Science and Technology”; Anderson, J. R.,
Christlansen, J. A&. Catal. Re/. Subj. 1953, 5 , 311. Boudart, M., Eds.; Springer-Veriag: Heidelberg, 1981; Voi. 1, p 87.
Cimino. A.; Boudart, M.; Taylor, H. S. J. Phys. Chem. 1954, 58, 796. Prigogine, I.; Outer, P.; Herbo, CI. J. Phys. Colloid Chem. 1948, 52, 321.
De Donder, Th. “L’AftlnitB”; Gauthier-Viiiars: Paris, 1927; p 43. Satterfieid, C. N.; Roberts, G. W. AIChE J. 1988, 14, 159.
Emmett, P. H. “New Approaches to the Study of Catalysis”; Phi Lambda Sinfelt, J. H.; Hurwitz, H.; Shuiman, R. A. J. Phys. Chem. 1960, 64, 892.
Upsilon: University Park, PA, 1962; Chapter 5. Temkin, M. I . I n t . Chem. Eng. 1971, I f , 709.
Eyring, H. J. Chem. Phys. W35, 3, 107. Temkin, M. I.; Pyzhev, V. Acts Physicochim. URSS 1940, 12, 217.
Garfinkle, M. J. Chem. Phys. 1983, 79, 2779. Ti”, B. Proc. I n t . Congr. Catal., 8th, 1984 1985 1, 7.
Glassman, I . “Combustion”; Academic Press: New York, 1977. Vannice, A. “Catalysis, Science and Technology”; Anderson, J. R., Boudart,
Gcddard, W. A., 111 Science 1985, 227, 917. M., Eds.; Springer-Veriag: Heidelberg, 1982; Voi. 3, p 139.
Hampton, R. F., Ed. J. M y s . Chem. Ref. Data 1973, 2 , 267. Van Tiggekn, A. “Oxidations et Combustions”; Technip: Paris, 1968; Voi. 11,
Happel, J. Catal. Rev. 1972, 6 . 221. Chapter 12.
Herbo, CI. Bull. SOC. Chim. Be@. 1942, 51, 44. Wagner, C. Adv. Catal. Re/. Subj. 1970, 21, 323.
Hjortkjaer, Jes ”Rhodium Complex Catalyzed Reactions”; Polyteknisk Vorlag:
Lyngby, Denmark, 1982; Voi. 2.
Horiuti, J. J. Res. Inst. Catal., Hokkaido Unlv. 1953, 2 , 87. Received f o r review June 21, 1985
Horiutl, J. Ann. N.Y. Acad. Sci. 1973, 273, 5. Accepted October 17, 1985
Horiuti. J.; Nakamura, T. Adv. Catal. Re/. Sub]. 17, 1.
Johnston, H. S. “Gas Phase Reaction Kinetics of Neutral Oxygen Species”;
U.S. Government Printing Office: Washington, D.C., 1968; NSRDS-NBS This paper was presented at the National Meeting of the American
20. p 12. Chemical Society, Miami, April 1985.

GENERAL ARTICLES

Cubic Chain-of-Rotators Equation of State


Hwayong Kim,+ Ho-Mu Lln, and Kwang-Chu Chao’
School of Chemical Engineering, Purdue lJnivers& West Lafayette, Indiana 47907

A cubic equation of the perturbation type is developed to express pressure as being made up of contributions due
to repulsive, rotational, and attractive forces. Use of the equation requires p,, T,, and w of a substance to be
known. Calculated pvT, vapor pressure, and enthalpy are compared with data and with the Soave equation and
the Peng-Robinson equation for a variety of substances over wide ranges of temperature and pressure. The
equation is extended to mixtures by using van der Waals one-fluid mixing rules for the equation parameters.
Gas-liquid equillbria of fluid mixtures are calculated for low-pressure symmetric mixtures as well as for highpressure
asymmetric mixtures of a heavy solvent with a light gas such as hydrogen, methane, carbon dioxide, and nitrogen.
Calculated pvT of mixtures is illustrated with two binary systems for gas and liquid states%p to the critical point.

Introduction equation of state to include Carnahan and Starling’s re-


Equations of state are useful for the calculation of fluid pulsive pressure and, additionally, to account for the ro-
thermodynamic properties. This usefulness has prompted tational motion of polyatomic molecules in terms of
a continual development of new equations. The pertur- equivalent translational degrees of freedom. Chien et al.
bation type of approach is noteworthy for being productive (1983) obtained an expression for the rotational pressure
of some very useful equations of which the cubic equations from Boublik’s (1975) equation for hard dumbbell mole-
have received much attention. Equations such as Red- cules. The chain-of-rotators (COR) equation of state that
lich-Kwong (1949), Soave (1972), and Peng-Robinson Chien et al. developed using the Carnahan and Starling
(1976) are in wide use due to their simplicity and generality repulsive pressure expression and their new rotational
combined with reasonable accuracy. AU of these equations pressure expression appears to be accurate, but complex.
contain van der Waals excluded-volume expression for the In this work we develop a cubic equation as a simplified
repulsive pressure RT/(u - b). The excluded-volume form of the COR equation of Chien et al. in order to
concept is valid for a dilute gas but breaks down at high provide the computational ease of cubic equations while
densities such as those of liquids (Vera and Prausnitz, retaining the structure of the COR equation. The repulsive
1972; Gubbins, 1973; Abbott, 1979; Henderson, 1979). and rotational pressure contributions are simulated with
In place of van der Waals form, Carnahan and Starling simpler functions. The attractive pressure expression is
(1969, 1972) obtained an expression for the repulsive also simplified. The equation is then fitted to the vapor
pressure based on molecular dynamics calculations. Do- pressure and saturated liquid density at subcritical tem-
nohue and Prausnitz (1978) developed a perturbation peratures and the pu isotherms at supercritical tempera-
tures.
The CCOR equation requires the three constants T,,p,,
‘Department of Chemical Engineering, University of Delaware, and o for the substance to be described. These constants
Newark. DE. are known for a large number of substances, in contrast
0196-4313/86/1025-0075$01.50/0 0 1986 American Chemical Society

Vous aimerez peut-être aussi