You are on page 1of 7

Applied Catalysis B: Environmental 18 (1998) 215221

Hydrodechlorination and hydrogenation of aromatic compounds over palladium on alumina in hydrogen-saturated water
Christoph Schuth1, Martin Reinhard*
Environmental Engineering and Science, Department of Civil Engineering, Stanford University, Stanford, CA 94305-4020, USA Received 3 September 1997; accepted 9 April 1998

Abstract The catalytic transformations of 1,2-dichlorobenzene, chlorobenzene, 4-chlorobiphenyl, g-hexachlorocyclohexane (Lindane), naphthalene and phenanthrene were studied over palladium on alumina in hydrogen-saturated water (Pd/Al2O3/ H2) at room temperature and ambient pressure. The chlorinated benzenes were rapidly hydrodechlorinated and Lindane was dehydrochlorinated to benzene. Partial or complete hydrogenation was observed for biphenyl and the polycyclic aromatic hydrocarbons. The phenanthrene ring was cleaved at the 9,10-position. In general dechlorination reactions were faster than hydrogenation reactions. # 1998 Elsevier Science B.V. All rights reserved. Keywords: Hydrodechlorination; Hydrogenation; Palladium; Chlorinated compounds; PCB; PAH; Lindane; Chlorobenzene; Water treatment

1. Introduction Nobel metals are widely used in chemical synthesis and for the catalysis of hydrogenation and hydrodehalogenation reactions, typically in the gas phase or in organic liquids under high hydrogen pressure and temperature conditions [13]. These processes are potentially applicable for the removal of halogenated hydrocarbon and polycyclic aromatic hydrocarbon compounds (PAHs) from uid waste streams. Recent studies have focused on the hydrodechlorination of
*Corresponding author. Tel.: 001 650 723 0308; fax: 001 650 725 3162; e-mail: reinhard@ce.stanford.edu 1 Present address: Institute of Geology, University of Tubingen, Sigwartstrasse 10, D-72076 Tu bingen, Germany. E-mail: schueth@cive.stanford.edu.

halogenated hydrocarbon compounds in hazardous organic liquids [46] and water [79]. Schreier and Reinhard [8] have shown that under ambient conditions in water, perchloroethylene (PCE) and trichloroethylene (TCE) are rapidly transformed into ethylene, which is further hydrogenated to ethane. Hoke et al. [10] reported the dechlorination of 4chlorophenol under mild conditions in water. The catalytic dechlorination of polychlorinated biphenyl compounds was demonstrated using a Pd/Fe bimetallic catalyst in a water/methanol/acetone solution at ambient temperature [11]. Dechlorination of PCB's was also demonstrated in a diglyme suspension at 1258C using a titanium catalyst and sodium borohydrate as the reductant [12]. Zhang and Rusling [13] reported the dechlorination of PCB in soils and clays by electrolysis in a bicontinuous microemulsion of

0926-3373/98/$ see front matter # 1998 Elsevier Science B.V. All rights reserved. PII: S0926-3373(98)00037-X

216

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

didodecyldimethylammonium bromide (DDAB), dodecane, and water. Murena [4] observed the hydrodechlorination of monochlorobiphenyls on sulde treated NiMo/g-Al2O3 in hexadecane at 250 3508C. In all cases biphenyl was the nal product. Hagh and Allen [14] studied the hydrodechlorination of hexachlorobenzene to benzene with a NiMo/ g-alumina catalyst at 3258C in the gas-phase. Only some of the possible intermediates were observed and the dechlorination steps had comparable rates. Gioia et al. [15] reported the hydrodechlorination of 1,2,3trichlorobenzene over the same type of catalyst in a batch reactor in hexadecane at 2003508C under 100 bar H2. In their system all possible intermediates were observed and the dechlorination rate decreased with decreasing number of chlorine substituents. Balko et al. [5] described the hydrodechlorination of polychlorobenzenes with up to four chlorine atoms in ethanol at 308C and several atmospheres hydrogen pressure. These authors observed that the dechlorination rates increased with decreasing number of chlorine substituents. Benzene hydrogenation was not observed and benzene was the nal product in all cases. Catalytic hydrogenation of PAHs was studied in organic solvents under mild conditions [1,2]. The observed transformations included direct hydrogenation, isomerization, and disproportionation. The product distribution was found to be highly dependent on catalyst type, temperature, and solvent. Hydrogenation is believed to proceed stepwise but intermediates with partially hydrogenated rings are typically not detected. Biphenyl is reported to undergo partial hydrogenation over palladium on alumina at 1008C and 500 psi hydrogen pressure in methanol to yield cyclohexylbenzene as the main product [2]. This study reports the hydrodehalogenation and hydrogenation of selected environmental contaminants under ambient temperature and pressure in water using Pd over g-alumina as the catalyst and hydrogen gas as the reductant. Products were measured and rate laws developed for the following compounds: 1,2dichlorobenzene, chlorobenzene, 4-chlorobiphenyl, naphthalene, phenanthrene and g-hexachlorocyclohexane (Lindane). These compounds are prominent and often refractory environmental contaminants. 4chlorobiphenyl was used as a representative for the class of polychlorinated biphenyls. Naphthalene and

phenanthrene were selected as model polycyclic aromatic hydrocarbon compounds. Palladium was chosen as the catalyst metal because of its broad reactivity for dehalogenation and hydrogenation reactions [2,5]. To minimize mass balance problems due to the sorption of organic compounds onto the support material, galumina was used as a low sorbing support. 2. Materials and methods 2.1. Experimental All experiments were performed at room temperature (218C28C) and ambient pressure in deionized water in a 100 ml glass syringe (Popper, NY). The syringe was modied to allow sampling via a glasssampling port capped with a Mininert valve at the side of the syringe and placed vertically on a stirrer (Fig. 1). Initially, the water was saturated with hydrogen gas (approximately 0.8 mM). The plunger was inserted leaving no head-space by venting the excess gas through the sampling port. No further hydrogen was added during the experiment. The compound of interest was injected as a concentrated methanol solution (1020 ml methanol). The nal concentration was well below the water solubility. The solution was allowed to equilibrate for several minutes and the initial sample was taken. The approximate initial concentrations were: 3.9 mM for phenanthrene and 13.5 mM for 1,2-dichlorobenzene, 17.9 mM chlorobenzene, 10.7 mM 4-chlorobiphenyl, 6.9 mM Lindane and 15.6 mM naphthalene. The reaction was started by adding the pre-reduced catalyst (1 wt% palladium on alumina powder; Aldrich 20 570-2) as a slurry in deionized water through the sampling port. The amount of catalyst used was 0.5 g/l for the 1,2-dichlorobenzene and chlorobenzene experiments, and 1 g/l for all other experiments. Changes in water volume in the syringe due to sampling were adjusted by the movement of the glass plunger thus avoiding any head-space. Water samples (1 ml) were withdrawn with a gastight glass syringe (Hamilton, Reno, NV) and liquid/ liquid extracted with pentane or hexane. Intermediates and products were characterized by GC-MS (HP 5890 equipped with a 30 m0.32 mm (0.25 mm lm thickness) DB-5 fused silica capillary column (J and W

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

217

dCadt k1 CX

(1)

If the data could not be tted using a simple rstorder model Eq. (2) was used: dC k1 C Y dt 1 k2 C (2)

where C is the substrate concentration and k1 and k2 are constants. This type of rate law is often observed for heterogeneous reactions in batch reactors with a constant volume and accounts for reactions that shift in order from zero to one as the substrate is utilized [15,16]. The rate law is rst-order if k2 C is much smaller than 1 and zero-order if k2 C is much greater than 1. In some cases, reactions were reversible and Eq. (3) was used: dc k1 C Ceq Y dt 1 k2 C (3)

Fig. 1. Experimental setup. A 100 ml glass syringe modified to allow sampling via a sampling port at the side of the syringe. No headspace is created due to sampling.

where Ceq indicates the concentration at the reaction equilibrium. Rate constants were evaluated by tting the data to the above models using a least squares minimizing routine (Scientist1 2.01, MicroMath, Salt Lake City, Utah). The tting routine is based on a modied Powell algorithm. All data points were weighted inversely proportional to the squares of the observed values. 3. Results and discussion The tted rate constants are summarized in Table 1. Numbers indicate the reaction steps as shown in the reaction schemes below. Five reactions (R2, RA5, R6, R7, R10) followed simple rst-order kinetics. In the other cases, the data tted a variable reaction model (Eq. (2) or Eq. (3)). Hydrodechlorination occurred rapidly for all chlorinated compounds investigated and, in general, preceded hydrogenation. Polyaromatic compounds reacted faster than monoaromatic compounds and hydrogenation was incomplete in some cases. For hydrogen pseudo-rst order conditions were assumed since it was present in excess concentrations. 3.1. 1,2-Di- and monochlorobenzenes Hydrodechlorination of 1,2-dichlorobenzene was complete within approximately 13 min and yielded

Scientic, Folsom, CA)) and quantied with internal standards using a HP 5890 GC equipped with a ECD, FID or PID (HNU-Systems, Newton, MA). Standards were purchased in the highest purity available from Aldrich, Milwaukee, WI. No standards were available for 1,2,3,4-tetrahydrophenanthrene and the octahydrophenanthrene isomers. For these compounds the FIDresponse was assumed to be equal to the response of 9,10-dihydrophenanthrene. 2.2. Data analysis From the kinetic data overall rate laws were developed by tting the data to reaction models. Because the hydrogen concentration was at least 40 times greater than the substrate concentration and the catalyst was initially reduced, rates were assumed to be independent of the hydrogen concentration. The rst approximation was to t the data to a rst-order ratelaw:

218

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

Table 1 Reaction rates for k1 and k2 as determined by fitting the kinetic data Reaction R1 R2* R3* R4* R5 R6 R7 R8 R9 R10
*

k1 (min1) 0.031 (0.035) 0.63 (0.30) 0.447 (0.042) 0.652 (0.047) 0.478 (0.005) 0.0055 (0.0002) 0.0114 (0.0005) 0.294 (0.004) 0.56 (0.28) 0.011 (0.005)

k2 (l/mmol) 0.15 (0.23) 0.092 (0.023) 0.189 (0.047) 0.056 (0.006) 0.82 (0.45)

Scheme 1.

Numbers indicate reactions as shown in the schemes. Concentrations of catalyst in the aqueous phase are 1 g/l except 0.5 g/l where indicated with *.

rate of chlorobenzene (R2) was 0.63 min1. This value is in agreement with the rate obtained in separate experiments using chlorobenzene as the starting compound (R4, 0.65 min1) (data not shown). The observed rate of 1,2-dichlorobenzene dechlorination is thus slightly (approximately 25%) slower than the dechlorination of chlorobenzene. This observation is in agreement with that of Balko and coworkers [5] who determined these rates using palladium at 4 atm hydrogen pressure in methanol at 308C. 3.2. 4-Chlorobiphenyl 4-Chlorobiphenyl was hydrodechlorinated within 10 min (Fig. 3) mainly via R5. Biphenyl was slowly hydrogenated to yield the cyclohexylbenzene which was then further hydrogenated to dicyclohexyl as the nal detected product (Scheme 2) (data not shown in Fig. 3). The hydrodechlorination rate of chlorobiphenyl (R5R7) was approximately a factor of 100 faster

Fig. 2. Hydrodechlorination of 1,2-dichlorobenzene over 0.5 g/l palladium on alumina in deionized water. Data points represent duplicate experiments. Rate constants are shown in Table 1.

benzene which was stable. Chlorobenzene was detected at low concentrations (<1% of C0). Mass balances were better than 90%. A satisfactory t for the kinetic data (Fig. 2) was obtained by assuming sequential hydrodechlorination reactions (R1 and R2) in parallel with reaction R3 as indicated in Scheme 1. The data for the conversion of 1,2-dichlorobenzene (Fig. 2) t to reaction model 2. For the transformation of chlorobenzene only the rst-order process was observed. The rate for the direct (two-chlorines are removed in one-step) dechlorination process (R3, 0.45 min1) is more than an order of magnitude faster than R1, the one-chlorine reaction. The dechlorination

Fig. 3. Hydrodechlorination and hydrogenation of 4-chlorobiphenyl, initial part. 1 g/l Pd on alumina in deionized hydrogen saturated water. Rate constants are shown in Table 1.

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

219

Scheme 2.

than the hydrogenation rate of biphenyl (R6). The intermediate biphenyl is accumulated in the aqueous phase with CaC0 exceeding 0.9. A direct transformation of 4-chlorobiphenyl to cyclohexylbenzene had to be assumed (R7) to t the data. Dicyclohexyl was rst detected after approximately 7 h and complete conversion of cyclohexylbenzene to dicyclohexyl took approximately seven days. Mass balances were close to 100% for R5 and about 85% for R6. The yield for the dicyclohexyl formation was poor, however, perhaps due to increased sorption of dicyclohexyl. Unknown peaks in the FID-chromatogram were not observed and benzene or cyclohexane was not detected. 3.3. Lindane Lindane was converted to benzene with a 99% yield within 18 min (Scheme 3). The mass balance was close to 100% (Fig. 4). The reaction involves elimination of six moles of HCl per mole of substrate. Intermediates were not detected indicating that the transformation reactions are much faster than the desorption of the partially dechlorinated intermediates. The reaction rate was zero-order at high concentrations and rst-order at low concentrations.

Fig. 4. Dechlorination of Lindane. 1 g/l Pd on alumina in deionized hydrogen saturated water. Rate constants obtained by fitting are shown in Table 1.

Scheme 4.

3.4. Polycyclic aromatic hydrocarbons (PAH) The PAHs naphthalene and phenanthrene were partially hydrogenated. Hydrogenation of naphthalene stopped at 1,2,3,4-tetrahydronaphthalene (Tetralin) (Scheme 4). This is expected due to the reported ability of palladium to selectively hydrogenate olens in the presence of aromatics. Naphthalene is assumed to adsorb on the metal surface as if it were a cyclic diolen fused to an aromatic ring [17]. The reaction rate was zero-order at high concentrations and rst-order at low concentrations (Fig. 5) (Table 1). The transformation was incomplete and

Scheme 3.

220

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

Fig. 5. Hydrogenation of naphthalene. 1 g/l Pd on alumina in deionized hydrogen saturated water. Rate constants obtained by fitting are shown in Table 1.

Fig. 6. Hydrogenation of phenanthrene. 1 g/l Pd on alumina in deionized hydrogen saturated water. Lines do not represent data fit. I: phenanthrene. II: 9,10-dihydrophenanthrene. III: 1,2,3,4-tetrahydrophenanthrene. IV: 1,2,3,4,5,6,7,8-octahydrophenanthrene. V: 1,2,3,4,4a,8,9,10,10a-octahydrophenanthrene. VI: dicyclohexyl.

equilibrium was reached after about 40 min. with the back reaction from 1,2,3,4-tetrahydronaphthalene being about 50 times slower than the hydrogenation reaction. The formation of naphthalene from the tetrahydronaphthalene was veried in a separate experiment using tetrahydronaphthalene as the starting compound (data not shown). The data t to Eq. (3). The hydrogenation of phenanthrene [I] (Scheme 5) yielded 1,2,3,4,4a,9,10,10a-octahydrophenanthrene [IV] and 1,2,3,4,5,6,7,8-octahydrophenanthrene [V] as the main products with 9,10-dihydrophenanthrene [II] and 1,2,3,4-tetrahydrophenanthrene [III] being transient intermediates. After 25 min dicyclohexyl [VI] was detected (Fig. 6). Scheme 5 shows the complex conversion pattern of phenanthrene with

observed reactions (solid arrows) and hypothesized reactions (dashed arrows). After 15 min 98% of the phenanthrene was transformed; thereafter, the phenanthrene concentration remained stable. Respiking the system with a similar phenanthrene concentration showed the same conversion pattern at similar rates. The nal phenanthrene concentration in the respiked system was approximately twice the concentration before the respike indicating the existence of an equilibrium between the products and phenanthrene. Using II as the starting compound yielded all the intermediates and products described above including I, supporting the proposed pathway. Rapid formation

Scheme 5.

C. Schuth, M. Reinhard / Applied Catalysis B: Environmental 18 (1998) 215221

221

of II from I is expected since palladium is reported to rapidly reduce the 9,10-position [18]. The formation of III from II can proceed directly or indirectly via I. Durland and Atkins [19] have shown that migration of hydrogen can take place over Raney nickel under nitrogen by reacting II to I and III in a ratio of 2:1. The concentration ratio of IV and V (1:2.5) in the experiments with II as the starting compound was very similar to the ratio observed in the experiments with I as the starting compound (1:2.3). However, in this case, the concentration of III was signicantly lower. These results indicate that migration of hydrogen from the 9,10-position is also likely to be responsible for the high concentrations of V. Durland and Atkins observed isomerization of IV and V under nitrogen atmosphere over Raney nickel with V being the more stable compound. The formation of IV requires ring cleavage at the 9,10-position but the parent compound for this reaction is not known. The mass balance in the phenanthrene as well as in the 9,10-dihydrophenanthrene experiments showed a signicant mass loss (about 30%) within the rst 5 min (Fig. 6). The reason for this loss is not known. Identical experiments with phenanthrene but without hydrogen showed no conversion and no signicant initial mass loss indicating that no signicant sorption to the catalyst occurred. However, no satisfactory model t for the kinetic data could be obtained indicating the possibility of unknown pathways, intermediates and products. 4. Conclusions The feasibility of treating water contaminated with chlorinated aromatic compounds, Lindane, and PAHs has been demonstrated in DI water. Under ambient pressure and temperature conditions, chlorinated aromatics are rapidly hydrodechlorinated and biphenyl and PAHs are partially or fully hydrogenated. Lindane is dehydrodechlorinated to benzene, the nal product. Hydrodechlorinations were more rapid than the biphenyl and PAH hydrogenations. Nonchlorinated one ring aromatic compounds were stable.

These processes may transform biologically refractory pollutants to compounds that are harmless or amenable to biodegradation. The applicability of the Pd/Al2O3/H2 treatment for removing these contaminants from aqueous waste streams remains to be evaluated under actual treatment conditions. Acknowledgements Funding was provided by the US Environmental Protection Agency, Ofce of Research and Development, under agreement R-825421 and by the German Research Foundation through a fellowship to C. Schuth. The content of this study does not represent the views of these agencies. References
[1] B.C. Gates, Catalytic Chemistry, Wiley, New York, 1991. [2] P. Rylander, Catalytic Hydrogenation in Organic Synthesis, Academic Press, New York, 1979. [3] M. Freifelder, Practical Catalytic Hydrogenation, Wiley/ Interscience, New York, 1971. [4] F. Murena, Environ. Technol. 18 (1997) 317. [5] E.N. Balko, F. Von Trentini, Appl. Catal. B 2 (1993) 1. [6] F. Gioia, J. Hazard. Mater. 26 (1991) 243. [7] C. Schlimm, E. Heitz, Environ. Prog. 15(1) (1996) 38. [8] C.G. Schreier, M. Reinhard, Chemosphere 31(6) (1995) 3475. [9] S. Kovenklioglu, Z. Cao, D. Shah, R.J. Farrauto, E.N. Balko, J. AIChE 38(7) (1992) 1003. [10] J.B. Hoke, G.A. Gramiccioni, E.N. Balko, Appl. Catal. B 1 (1992) 285. [11] C. Grittini, M. Malcomson, Q. Fernando, N. Korte, Environ. Sci. Technol. 29(11) (1995) 2898. [12] Y. Liu, J. Schwartz, C.L. Cavallaro, Environ. Sci. Technol. 29(3) (1995) 836. [13] S. Zang, J.F. Rusling, Environ. Sci. Technol. 29(5) (1995) 1195. [14] B.F. Hagh, T. Allen, Chem. Eng. Sci. 45(8) (1990) 2695. [15] F. Gioia, V. Famiglietti, F. Murena, J. Hazard. Mater. 33 (1993) 63. [16] O. Levenspiel, The Chemical Reactor Omnibook, OSU Book Stores, Corvallis Oregon, 1996. [17] A.W. Weitkamp, J. Catal. 6 (1966) 431. [18] P.P. Fu, R.G. Harvey, Tetrahedron Lett. 5 (1977) 415. [19] J.R. Durland, H. Atkins, J. Am. Chem. Soc. 60 (1938) 1501.

You might also like