Ethane Dehydrogenation Using A High-Temperature Catalytic Membrane Reactor

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

joumal of

MEE%:E
ELSEVIER Journal of Membrane Science 90 ( 1994) 1 l- 19

Ethane dehydrogenation using a high-temperature catalytic


membrane reactor
E. Gobina, R. Hughes*
Department of Chemical Engineering, Universityofsalford, Salford h4.5 4 W, UK

(ReceivedAugust 27, 1993; accepted in revised form December 17, 1993)

Abstract

Experiments have been conducted for the dehydrogenation of ethane to ethylene using a high-temperature cat-
alytic membrane reactor under isothermal conditions. Typically, conversions of up to 7 times (cocurrent mode)
and 8 times (countercurrent mode) higher than the equilibrium value achievable in conventional fixed-bed reac-
tors have been attained at high sweep flow rates. The membrane used in this study was a thin layer of Pd-23 wt%
Ag on porous Vycor glass. The significant improvement in the ethane conversion is attributed to the exclusive and
continuous permeation of hydrogen through the membrane.

Key words: Catalytic dehydrogenation; Metal membranes; Membrane reactors; Gas separations

1. Introduction availability of high-temperature, porous, ce-


ramic and metallic membranes has opened up
Membrane reactors are unique in that they are new avenues to researchers and engineers for
capable of combining chemical reaction and sep- carrying out catalytic processes at elevated tem-
aration in a single-unit operation. The mem- peratures. This is because ceramic and metallic
brane, being selective to one or more of the prod- membranes in contrast to their organic counter-
uct or reactant species, can be used to improve parts, are robust, can function for prolonged pe-
the yields of thermodynamically limited reac- riods at high temperatures with great stability,
tions. Although the basic concept of membrane possess improved pore size controllability and
reactors has been known for almost three dec- have higher rates of heat transfer. This develop-
ades [ l-3 1, it is only recently that the applica- ment has renewed interest in the use of high-
tion of the concept is being recognised [ 4-6 1. temperature catalytic membrane reactors and is
Low-temperature applications ( < 1OOoC) the subject of recent reviews [ 8,9]. Table 1 pre-
were the first to benefit from this concept and an sents a selection of experimental studies, for gas-
extensive review has been carried out by Chang phase reactions, utilising porous ceramic and
and Furusaki [ 7 1. Most of the low-temperature metallic membranes, attempted by several work-
applications utilise organic membranes. The ers to enhance the conversion of equilibrium-
limited reactions.
*Correspondingauthor. Tel: 61-745-5081, Fax: 61-745-5999, One particular class of reactions of industrial
Email: R. Hughes @ Chemistry. Salford. ac.uk. importance is the dehydrogenation of lower al-

0376-7388/94/$07.00 0 1994 Elsevier Science B.V. All rights reserved


SSDZO376-7388(94)00011-M
12 E. Gobina, R. Hughes /Journal of Membrane Science 90 (1994) I I- 19

Table 1
Experimental studies of gas-phase reactions using membranes

Membrane Catalyst Chemical reaction Temperature Enhancement Ref.


material studied range ( “C) achieved

Porous A1203 MoSz in pores of membrane Decomposition of H2S 800 Two-fold increase on the 10
to form H2 and S equilibrium conversion
Porous A1203 Cr209 (20 wtl) Al,Os Dehydrogenation of 575 Conversion increased from an II
propane to propene equilibrium value of 40.1 to
58.7%
Microporous glass Pt/A1203 Dehydrogenation of 215 Conversion increased from 35% 12
cyclohexane without membrane to 80% with
membrane
Porous supported Dehydrogenation of 600-640 Conversion of ethylbenzene in 13
multrlayer A1203 ethylbenzene to styrene the presence of the membrane
were 15% higher than those in
the absence of the membrane
Porous Vycor glass Pt cat. within pores of the Dehydrogenation of 270 Conversion from an equilibrium 14
membrane (34 wt% Pt ) cyclohexane value of 26 to 56% in the
membrane
Microporous Thin dense film (20 pm) of Water gas shift reaction 400 Complete conversion of CO 15
glass/thin Pd film Pd coating the external attained in the membrane
surface of the glass reactor compared to the
equilibrium conversion of 60%
Dense Pd tube Pt/Al,O, (0.5 wt% Pt) Dehydrogenation of 200 Conversion of C6Hr2 was 99.7% 16
cyclohexane in the membrane reactor
compared with an equilibrium
conversion of only 18.7%
Non-porous Pd/Ag Gas-phase decomposition Decomposition of 500 The conversion of HI was 4%, a 17
tube hydrogen iodide twenty-fold increase from a
value of 0.2% in the absence of a
membrane

kanes. These yield valuable commodity chemi- ties, even more avenues are opened and, for ex-
cals for the production of alcohols, gasoline ample, one can couple reactions in a membrane
blends and polymers. However, these reactions reactor.
are also limited by their thermodynamic equilib- Experimental investigations that have been at-
rium and to obtain higher yields, very high tem- tempted for reactions accompanying membrane
peratures are required resulting in losses in both separation with the objective of shifting the
selectivity and overall catalyst activity. An in- equilibrium towards increased conversion in-
centive therefore exists for a process that has the clude amongst others (see Table 1) the decom-
capability of overcoming the thermodynamic position of hydrogen sulfide [ 18 1, and decom-
equilibrium while maintaining the integrity of the position of hydrogen iodide [ 191 using
catalyst and improving the product yield. Mem- microporous glass membranes. Microporous al-
brane reactor technology appears to be promis- umina membranes have been used to investigate
ing since the selective removal of hydrogen from the reaction of propane to propylene [ 201, ethane
the reaction products would inevitably tend to to ethylene [ 2 1 ] and for the dehydrogenation of
shift the equilibrium towards increased product ethylbenzene to styrene [ 131. In these investi-
yield. The integration of chemical reaction and gations, the flow of reactant and product species
separation in a membrane reactor would then re- through the membrane followed a Knudsen dif-
sult in better rate control because of shorter con- fusion mechanism and reactor performance was
tact time and reduced temperature of operation. much improved compared with a conventional
When the membrane has also catalytic proper- catalyst-packed reactor (see Table 1) . However,
E. Gobina, R. Hughes /Journal of Membrane Science 90 (I 994) I l-1 9 13

computer simulations and experiments [ 141 that the concept of overcoming thermodynamic
have established that there is a limitation on the equilibrium limitations with consequent im-
performance of microporous membrane reactors proved product yields is feasible with a careful
and it is almost impossible to make the reaction choice of membrane.
proceed beyond a certain level of conversion
without the need to recycle the unreacted feed
since some ‘slip’ of feed also passes through the
micropores of the membranes into the separa- 2. Experimental
tion side.
This limitation can be overcome by using a The experimental equipment used in this study
membrane that is selective only to product(s). is similar to that employed earlier [ 231 and con-
This principle has been demonstrated by Itoh sists of a feed gas delivery system, a high-temper-
[ 161 who succeeded in obtaining almost 100% ature membrane reactor, a gas chromatograph
conversion for the dehydrogenation of cycloh- and an on-line data acquisition system. The
exane when using a palladium membrane in- membrane is a composite consisting of a porous
stead of microporous glass membrane. A 20-fold Vycor glass tubular substrate ( 10 mm o.d., 1.1
increase was obtained for the decomposition of mm thick and pore size of 40 A) onto which a
hydrogen iodide by using a non-porous palla- thin continuous film of Pd-Ag was deposited by
dium/silver tube [ 171, far higher than that ob- the process of magnetron sputtering. The process
tained by using a porous membrane [ 19 1. This of magnetron sputtering can be used to deposit
high increase was attributed to the exclusive sep- very thin films on almost any porous substrate.
aration of hydrogen. It involves ion bombardment of the target, thus
Palladium and palladium alloys are the cur- ejecting atoms which then adhere onto the po-
rent preferred methods for removing hydrogen rous substrate. By exposing the substrate to a ra-
exclusively. However, existing commercially dio frequency (rf ) glow discharge (plasma), the
available palladium membranes are too thick adhesion strength of the film can be enhanced.
( N 53 pm) to provide economic rates of permea- Film thickness (and hence deposition rates) were
tion and palladium being an expensive metal, a estimated using additional samples of porous
great thickness also means more cost. Although Vycor glass in the deposition chamber placed ad-
the selectivity of microporous membranes is jacent to the actual substrate of interest. Direct
limited, they could be employed as substrates on measurement of the film thickness on these sam-
which a thin metal layer of a dense membrane ples was carried out using scanning electron mi-
material such as palladium may be deposited to croscopy (SEM ). The membrane thickness in
form a coherent layer having exclusive permea- this investigation was 6 ,um. Film thickness uni-
tion to hydrogen. formity was measured at various locations along
We have previously reported on the applica- and across the samples by electrical resistivity.
tion of a high-temperature membrane reactor for The variation along the membrane averaged
the selective permeation of hydrogen from an 6.7%. Film stoichiometry was determined by the
N2-Hz gaseous mixture [ 22 ] and the character- electron probe microanalysis technique
isation of Pd-metallised porous Vycor glass pro- (EPMA ) . A total of 50 analyses was carried out
duced by the process of magnetron sputtering over a randomly selected area, and an average
[ 231. In this paper, as part of an ongoing re- value obtained. There was very little difference
search into membrane reactor applications, we between the Pd: Ag content in the target (77 : 23
present results for the catalytic dehydrogenation wt%), and that on the substrate (23 + 3 wt% Ag ) .
of ethane in a high temperature catalytic mem- There was no light transmission across the mem-
brane reactor promoted by 0.5 wt% Pd/A1203 branes indicating a dense structure. Over an ex-
catalysts packed in the tube side of the reactor. tended period of use in the membrane reactor,
The significant equilibrium shift observed shows there was no physical peeling from the substrate.
14 E. Gobina, R. Hughes /Journal of Membrane Science 90 (1994) 1 I-I 9

This is indicative of the good adherence strength ing BET surface area using nitrogen adsorption,
of the film. SEM and EPMA. The catalyst was used mainly
The high-temperature membrane reactor, de- to promote the dehydrogenation reaction.
tails of which are schematically shown in Fig. 1, All experiments were performed at a total feed
consists of an outer stainless steel shell (with pressure of 128.7 kPa ( 1.27 bar) while the shell
provision for either inflow or outflow of sweep pressure was maintained at 10 1.3 kPa ( 1 bar).
gas) and an inner tube. The membrane tube is The catalytic membrane and the catalyst pellets
centralised in the reactor by a set of moulded were reduced in a flow of pure hydrogen at 2.5
graphite rings packed at each end of the reactor. cm3 (STP) /min for 30 min at a temperature of
Careful tightening of stainless steel plugs on both 400°C prior to the reaction experiments. The
ends compresses the rings and completes the seal feed gas mixture consisting of a single feed of 50
between the shell and tube. The thickness of the ~01% ethane in nitrogen was then allowed to flow
Pd-Ag film was 6 pm and the area available for at a predetermined rate over the catalyst pellets;
hydrogen permeation was 47.14 cm2. this continued until temperature stability was
Three 0.5 mm thermocouples were located achieved. Due to the endothermicity of the de-
along the axis of the catalyst bed to determine hydrogenation reaction, a small temperature
the temperature profile inside the reactor. A sim- drop in the catalyst bed was observed. This was
ilar size thermocouple was located at the mid- compensated by increasing the heat input to
point of the shell side. Steady-state temperature maintain the desired temperature in the bed.
was attained when the variation in the thermo- Temperature variation along the reactor was then
couple readings were within 22°C of the de- less than 3 “C. Membrane performance was as-
sired temperature. sessed by comparing the shift observed from the
The catalyst for the dehydrogenation reaction equilibrium conversion of ethane. This equilib-
was located inside the tube and consisted of 2.55 rium conversion was determined by closing the
g of 0.5 wt% Pd/A1203 cylindrical pellets (3.35 shell-side inlet, so that shell-side sweep flow was
mm o.d.x3.63 mm long) which were carefully
absent and continuously monitoring the effluent
and uniformly packed in the bed. The catalysts
from the catalyst bed via an air-activated auto-
were characterised by various techniques includ-
matic sampler. This was continued until con-
stant concentrations were obtained as indicated
by identical values of the peak chromatograph
areas. The ethane conversion calculated under
these circumstances was the equilibrium conver-
sion and was then used for comparison with con-
versions obtained when the shell side was opened
and the sweep gas consisting of pure nitrogen was
allowed to flow. Both the permeate and retentate
streams were sampled. The shell-side sweep gas
flow was then varied and flow directions re-
versed for either cocurrent or countercurrent op-
erations while maintaining the temperature and
feed mole fraction of ethane constant.
Reaction experiments were conducted in the
reactor using a single feed consisting of a 50 ~01%
mixture of ethane in nitrogen, the nitrogen serv-
ing as an internal standard for the analysis of gas
composition. All experiments were carried out at
Fig. 1. Essential details of the catalytic membrane reactor. 660 K under isothermal conditions.
E. Gobina, R. Hughes /Journal of Membrane Science 90 (I 994) 1 I-l 9 15

3. Results and discussion in the square roots of the upstream and down-
stream hydrogen partial pressures. This is con-
3.1. Permeability of membrane under non- sistent with Sieve&s law [ 25 1, and indicates that
reactive conditions permeation through the bulk of the metal is the
rate-limiting step. A closer inspection of Fig. 2
The permeability of the membrane to hydro- also reveals that there is a temperature effect on
gen was studied using a 60% Ha: 40% Nz gas the permeability constant. This was further in-
mixture, varying the driving force and the tem- vestigated by maintaining a constant pressure
perature in order to obtain quantitative infor- difference while varying the temperature. The
mation regarding the permeability of the mem- results correspond to the Arrhenius law and the
brane. If the permeate rate of hydrogen gas permeability constant (Q ) of hydrogen gas
through the Pd : Ag membrane [ QH ( cm3 ) ] is as- through the film can then be expressed as follows:
sumed to obey the half-power pressure law [ 241, Q,,=7.174x 10m5exp( -6.380/RT) (2)
then:
where R is the ideal constant (kJ/mol K) and T
QH =Qopt” -pi”) (A/d,) (1) the absolute temperature (K). The activation
where Q0 is the permeability constant of hydro- energy of 6.38 kJ/mol obtained in this investi-
gen gas through the membrane (cm3 cm/cm2 s gation compares well with reported values of 5.73
atm”‘) ; A is the membrane area available for [26], 6.60 [27], and 5.86 kJ/mol [28]. In these
flow (cm2); and dt is the film thickness of the previous investigations, similar Pd-Ag alloy
membrane (cm). A typical plot of Qu versus membranes were employed.
(P ;I2 -pii ) at various operating temperatures
is given in Fig. 2. The permeation rate of hydro- 3.2. Ethane dehydrogenation experiments
gen gas is directly proportional to the difference
The equilibrium conversion of ethane to eth-
ylene was determined twice (before and after ex-
periments in the membrane reactor). An aver-
age value of 2.57% was obtained and used for
1.6 .
comparison with the membrane reactor conver-
sions. This value is simply the maximum con-
version achievable in the absence of hydrogen
separation and may be compared with a value of
0.87% estimated from thermodynamic data for
the same temperature. A similar three-fold in-
crease in the measured equilibrium constant for
this reaction compared to the thermodynamic
value was obtained by Champagnie et al. [ 2 I 1.
Operation of the catalytic membrane reactor was
achieved by varying the flows of the feed and/or
sweep gas at constant temperature using a fixed
weight of catalyst. The overall conversion of
ethane to ethylene, allowing for volume changes
was obtained using the equation,

X=( FA;rOFA)x 100% (3)


Fig. 2. Effect of pressure and temperature on the permeation
rate of hydrogen. where X is the overall ethane conversion to eth-
16 E. Gobina, R. Hughes /Journal ofMembrane Science 90 (1994) 11-I 9

ylene; FAOand FA are the respective flow rates of stream. The mass transfer flux, N, to the inner
ethane in the feed and product in the tube side. surface of the membrane is given by
No ethane or ethylene was detected in the shell
side exit confirming that the membrane is selec- N= (~1 -pz)Sh D,,ld (4)
tive only to hydrogen. where p1 and p2 are the molar densities of hydro-
gen on the reaction and permeate side, respec-
tive ( mol/cm3), DABis the diffusivity of hydro-
3.3. Conversion vs. time factor gen in the reaction mixture (cm2/s), d is the
reactor diameter (cm) and Sh is the Sherwood
The effect of time factor ( W/F,,) defined as number (k,d/D) . The permeate-side concentra-
the weight of catalyst in the bed, W (g), divided tion of hydrogen was very small due to the pres-
by the molar flow rate of ethane, FAO(mol/s), ence of the sweep gas. Under the conditions of
on the conversion of ethane to ethylene is pre- the present study, the flow on the reaction (tube)
sented in Fig. 3. The conversion of ethane de- side was laminar and the Sherwood number is
pends on the mean residence time of ethane in given by [ 29 ] :
the reactor. This means that conversion in-
creases with the time factor as expected. In the
catalytic membrane reactor, at long contact
times, the conversion of ethane to ethylene be- where Re and SC are the Reynolds and Schmidt
comes higher than the equilibrium value because numbers, h and h are the fluid viscosities at av-
product hydrogen is preferentially and exclu- erage bulk temperature and the wall, respec-
sively diffused through the membrane. This se- tively, and L the length of reactor across which
lective and continuous removal of hydrogen from permeation of hydrogen occurred. Using this
the reaction zone, results in a shift in the equilib- correlation for Sh and the flux equation (4)) the
rium towards increased ethane conversion. At hydrogen mass transfer flux from the bulk gas to
very short contact times, equilibrium shift is pre- the inner membrane surface was always greater
vented because the mean residence time of ethane than the permeation flux and therefore mass
in the reactor is too short. transfer in the tube side was not limiting.
Under conditions of varying feed flow rate at A comparison of the efficiency of the modes of
constant catalyst weight, the external mass trans- operation of the catalytic membrane reactor is
fer rate changes with the velocity of the gas shown in Fig. 4. Both countercurrent and cocur-
rent modes of operation were examined. In the
countercurrent mode, the flow of the sweep gas
was in the opposite direction to that of the feed
while in the cocurrent mode both were in the
same direction. Experimental results show that
operation of the membrane reactor in the coun-
tercurrent mode is more desirable compared to
that in the cocurrent mode. This is because higher
conversions can be obtained for the same sweep
flow in the countercurrent mode than for the co-
current mode. This conclusion was also arrived
JA I
0.5 1.5 2.5 3.5 at by computer simulation results, in a mem-
TIblE PKTOH
,!4,FA<>i x 105 vi-at Z,WOl brane reactor hy Itoh [ 16 ] for the dehydrogena-
tion of I-butene and shown experimentally for
Fig. 3. Influence of time factor ( W/F,,) on the conversion
of ethane in the membrane reactor. Temperature, 600 K; cat- the dehydrogenation of cyclohexane [ 281. In
alyst weight, 2.55 g; flow rate of sweep gas, 300 cm3 (STP)/ these studies Pd-Ag membranes similar to that
min. used in the present study were employed.
E. Gobina, R. Hughes /Journal of Membrane Science 90 (1994) 11-I 9 17

tinuously from the reaction zone. Since these Pd-


Ag membranes are selective only to hydrogen, an
increase in the sweep ratio ensures a large hydro-
gen partial pressure difference between the shell
side and tube side. The partial pressure differ-
ence across the membrane is the driving force for
hydrogen permeation which results in an equi-
librium shift towards higher ethane conversion.
At a sweep ratio of 5, for example, the ethane
conversion increases from the equilibrium value
of 2.5 (without any flow of sweep gas) to 17.66%.
This is a 7-fold increase, and is attributed to the
selective and continuous removal of hydrogen
from the reaction zone. This result compares with
a 6-fold increase reported by Champagnie et al.
[ 2 I] who employed a Pt-impregnated porous al-
Fig. 4. Effect of sweep gas flow rate on ethane conversion.
(-- ) Equilibrium; ( 0 ) cocurrent flow; ( A ) countercurrent umina membrane having a pore size of 40 A, for
flow. Time factor ( W/F,,) = 1.144 x 10’ g cat s/g mol; tem- ethane dehydrogenation. By using a dense metal-
perature, 660 K, space time, 8.4 s. lic membrane in this study, we have thus elimi-
nated the problem of back-permeation of reac-
tant from the permeate side to the feed side at
high conversions and re-equilibration [ 301 often
encountered in porous membrane reactor
systems.

3.5. Performance of the membrane reactor

Table 2 shows the relationship between mem-


brane performance and rate of reaction at high
performance, i.e., at a conversion of N 18%. The
rate of Hz permeation is 6.7 times the produc-
tion rate. This suggests that the membrane has
Fig. 5. Membrane reactor conversion as a function of the sufficient capacity to permeate the hydrogen and
sweep gas/feed gas flow ratio. Temperature, 660 K; cocur-
that the flow through the membrane is not rate
rent flow.
limiting. Although the hydrogen removal rate is
far in excess of the production rate, Fig. 3 sug-
3.4. Effect of sweepjlow/feedjlow
gests that at sufficiently long contact times the
conversion attains a virtually constant value. The
Fig. 5 shows the conversion of ethane to eth- reason for this is not clear; one possibility is that
ylene at a temperature of 660 K as a function of
the sweep gas ratio; defined as the ratio of the Table 2
flow rate of sweep gas in the shell side to the feed Estimation of catalytic membrane reactor performance
inlet flow rate. The feed-side inlet consisted of a
Temperature Ethane H2 production H2 permeation rate
50 ~01% ethane in nitrogen mixture. Increasing (K) conversion rate [usingpermeability
the sweep ratio increases the ethane conversion. W) (gmof/s) Eqs. (2) and (3)]
This is due to the fact that in the catalytic mem- (gmol/s)
brane reactor, conversion is increased by the se- 18 4.0x 10-h 26.7X IO-6
660
lective and exclusive removal of hydrogen con-
18 E. Gobina. R. Hughes /Journal of Membrane Science 90 (I 994) I I- 19

readsorption of ethylene and hydrogen on the Johnson Matthey Research Centre, Sonning
surface of the catalyst pellets occurs under these Common to whom we express our thanks.
conditions and this is related to the relative high
surface area of the catalyst employed
(-200 m2/g).
The dehydrogenation of 1 mol of ethane yields 6. References
equimolar amounts of ethylene and hydrogen.
The selectivity of ethylene calculated when no
sweep gas was present was 50%. However, when [ 1]A.S. Michaels, New separation technique for the CPI,
a sweep gas flow was used this selectivity was Chem. Eng. Prog., 64( 12) (1968) 31.
2lD.I.C. Wang, A.E. Humphrey and E. Arthur, Enzyme
close lOO%, indicating that almost all of the hy- detergents, already a multimillion dollar industry, are
drogen in the gas phase was removed by the but one concrete result from biochemical engineering.
membrane. Chem. Eng., 86(27) (1969) 108.
3 ] W.C. Pfefferle, Process for dehydrogenation, US Pat.
3,290,406 (1966).
[ 41 J.F. Roth, Future catalysis for the production of chemi-
4. Conclusions cals, in J.W. Ward (Ed.), Catalysis 1987, Elsevier, Am-
sterdam, 1988.
It has been shown experimentally that a much [ 5 ] V.M. Gryaznov, Hydrogen permeable palladium mem-
higher level of ethane conversion can be attained brane catalysts, an aid to the efficient production of ul-
trapure chemicals and pharmaceuticals. Platinum Met.
even in a single-stage catalytic membrane reac-
Rev., 30(2) (1986) 68.
tor. The significant characteristic advantages are 6 ] V.M. Gryaznov., VS. Smimov and M.G. Slinko, in J.W.
that it is possible to carry out the reaction at a Hightowes (Ed.), Proc. 5th Int. Congr. Catal., Vol. 2,
lower reaction temperature and shorter contact North Holland, Amsterdam, 1973, p. 1139.
time. This results in a shorter reactor length and 17lH.N. Chang and S. Furusaki, Membrane Bioreactors:
Present and Prospects, Adv. Biochem. Eng. Biotechnol.,
lower energy requirement for eventual separa-
44 (1991) 27.
tion of the product(s) and reactant (s ) from each 8]J. Shu, B.P.A. Grandjean, A. Van Neste and S. Kalia-
other and from the reaction mixture. Such a re- quine, Catalytic palladium-based membrane reactors: a
actor could be applied to study other dehydro- review, Can. J. Chem. Eng., 69 ( 199 1) 1036.
genation reactions of industrial importance such [ 9 ] J.N. Armour, Catalysis with permselective inorganic
membranes, Appl. Catal., 49 ( 1989) 1.
as the conversion of ethylbenzene to styrene, bu-
[ lO]T. Kameyama, K. Fukuda, M. Fujishige, H. Yokohawa
tene to butadiene and methanol to formalde- and M. Dokiya, Production of hydrogen from hydrogen
hyde. Because the metallic surface is in itself cat- sulfide by means of selective diffusion membrane, Adv.
alytic to a host of hydrogenation reactions, there Hydrogen Energy Prog., 2 ( 1981) 569.
exists a possibility of coupling the endothermic [ 1llJ.G.A. Bitter, Br. Pat. Appl. 8629135 (1986).
dehydrogenation with an exothermic hydrogen- [ 1210. Shinji, M. Misono and Y. Yoneda, The dehydroge-
nation of cyclohexane by the use of porous-glass reactor,
ation reaction. This would assist transport of both Bull. Chem. Sot. Jpn., 55 (1982) 2760.
hydrogen and heat in the system. Work in this [ 13 ] J.C.S. Wu, T.E. Gerdes, J.L. Pszczolkowski, R.R. Bhave
area is continuing. and P.K.T. Liu, Dehydrogenation ethylbenzene to sty-
rene using commercial ceramic membranes as reactors,
Sep. Sci. Technol., 25 (1990) 1489.
[ 141Y.M. Sun and S.J. Khang, Catalytic membranes for si-
5. Acknowledgements multaneous chemical reaction and separation applied to
dehydrogenation reaction. Ind. Eng. Chem. Res., 27
We thank the SERC for continued financial (1988) 1136.
support in this research. We are also grateful to [ 15 ] E. Kikuchi, S. Uemiya, N. Sato, H. Inoue, H. Ando and
T. Matsuda, Membrane reactor using microporous glass-
Dr. Dermot Monaghan for his assistance in car-
supported thin film of palladium. Application to the
rying out the magnetron depositions in the sub- water gas shift reaction, Chem. Lett., 3 ( 1989) 489.
strate. Characterisation of the catalyst pellets and [ 16]N. Itoh, A membrane reactor using palladium. AIChE
palladium film stoichiometry was undertaken by J., 33 (1987) 1576.
E. Gobina, R. Hughes /Journal oJMembrane Science 90 (1994) 1 l-l 9 19

[ 171 J. Yeheskel, D. Leger and P. Courvoisier, Thermal de- [ 24]G. Bohmholdt and E. Wicke, Diffusion von H und D in
composition of hydroiodic acid and hydrogen separa- Pd und Pd-Legierungen. Z. Physik. Chem. N.F., 56
tion, Adv. Hydrogen Energy, 2 (1979) 569. (1967) 133.
[ 18]T. Kameyama, M. Dokiya, M. Fujishige, H. Yokohawa [25]A. Sieverts and W. Kumbhaar, Solubility of gases in
and K. Fukuda, Possibility for effective production of metals and alloys, Ber. Dtsch. Chem. Ges., 43 ( 1910)
hydrogen from hydrogen sulfide by means of a porous 893.
glass membrane, Ind. Eng. Chem. Fundam., 20 ( 198 1) [ 26]H. Yoshida, S. Konishi and Y. Naruse, Effects of impur-
97. ities on hydrogen permeability through palladium alloy,
[ 19]N. Itoh, Y. Shindo, K. Obata, T. Hakuta and H. Yoshi- J. Less-Common Met., 89 (1983) 429.
tome, Simulation of a reaction accompanied by separa- [27]F.J. Ackerman and G.J. Koskinas, Permeation of hy-
tion, Int. Chem. Eng., 25 ( 1985) 138. drogen and deuterium through palladium-silver alloys,
[20]Z.D. Ziaka, R.G. Minet and T.T. Tsotsis, Propane de- J. Chem. Eng. Data, 17 ( 1972) 5 1.
hydrogenation in a packed-bed membrane reactor, [ 281 J. Chabot, J. Lecomte, C. Grumet and J. Sannier, Fuel
AIChE J. 39 (1993) 526. clean-up system. Poisoning of palladium-silver mem-
[Zl]A.M. Champagnie, T.T. Tsostsis, R.G. Minet and I.A. branes by gaseous impurities, Fusion Technol., 14
Webster, A high temperature membrane reactor for (1988) 614.
ethane dehydrogenation, Chem. Eng. Sci., 45 ( 1990) [ 29lC.O. Bennett and J. Myers, Momentum, Heat and Mass
2423. Transfer, McGraw-Hill, New York, 1982, p. 557.
[22]E. Gobina and R. Hughes, Selective high temperature [ 30lK.C. Cannon and J.J. Hacskaylo, Evaluation of palla-
membranes for hydrogen separation, I993 IChemE Re- dium-impregnation on the performance of Vycor glass
search Event, Institution of Chemical Engineers, Rugby, catalytic membrane reactor, J. Membrane Sci., 65
UK, 1993, p. 522. (1992) 259.
[ 231 E. Gobina and R. Hughes, High temperature selective
membranes for hydrogen separation, Dev. Chem. Eng.
Min. Process., accepted for publication.

You might also like