7
APPLIED AND ENVIRONMENTAL MICROBIOLOGY, Feb. 1989, P. 323-329 0099-2240/89/020323-07$02.00/0 Copyright C) 1989, American Society for Microbiology Coenzyme A Transferase from Clostridium acetobutylicum ATCC 824 and Its Role in the Uptake of Acids DENNIS P. WIESENBORN,1 FREDERICK B. RUDOLPH,2 AND ELEFTHERIOS T. PAPOUTSAKISlt* Department of Chemical Engineering' and Department of Biochemistry,2 Rice University, Houston, Texas 77251-1892 Received 12 October 1988/Accepted 25 November 1988 Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence of 0.5 M ammonium sulfate and 20% (vol/vol) glycerol, whereas activity was rapidly lost in the absence of these stabilizers. The kinetic binding mechanism was Ping Pong Bi Bi, and the Km values at pH 7.5 and 30°C for acetate, propionate, and butyrate were, respectively, 1,200, 1,000, and 660 mM, while the Km value for acetoacetyl-CoA ranged from about 7 to 56 ,uM, depending on the acid substrate. The Km values for butyrate and acetate were high relative to the intracellular concentrations of these species; consequently, in vivo enzyme activity is expected to be sensitive to changes in those concentrations. In addition to the carboxylic acids listed above, this CoA transferase was able to convert valerate, isobutyrate, and crotonate; however, the conversion of formate, n-caproate, and isovalerate was not detected. The acetate and butyrate conversion reactions in vitro were inhibited by physiological levels of acetone and butanol, and this may be another factor in the in vivo regulation of enzyme activity. The optimum pH of acetate conversion was broad, with at least 80% of maximal activity from pH 5.9 to greater than 7.8. The purified enzyme was a heterotetramer with subunit molecular weights of about 23,000 and 25,000. Coenzyme A (CoA) transferases activate carboxylic acids to the respective CoA thioesters at the expense of the CoA thioester of another species of carboxylic acid. The reaction is easily reversed. The presence of CoA transferase in cell extracts was first demonstrated by Stadtman for Clostridium kluyveri (26). Different types of CoA transferase have been subsequently shown to be present in several other species of bacteria. These enzymes have been prepared in very pure form from Escherichia coli (24), Peptostreptococcus elsdenii (28), and Clostridium sp. strain SB4 (3). Some physical and kinetic properties of these enzymes have been character- ized. All studied CoA transferases show relatively broad substrate specificities, but each enzyme has characteristic preferred substrates. Among clostridia that are used to carry out the acetone- butanol fermentation, a CoA transferase activity that trans- fers CoA from acetoacetyl-CoA to butyrate or to acetate has been found in crude extracts (1, 10, 30). However, the direct transfer of CoA from acetyl-CoA to butyrate has not been detected (1). The role of CoA transferase in butanol-forming clostridia is fundamentally different from that in other bac- teria for which it has been characterized. Generally, in this latter group of bacteria the enzyme is involved in the uptake of substrates for energy and structural use (3, 24, 28). In butanol-forming clostridia, however, the CoA transferase acts mainly to detoxify the medium by removing the acetate and butyrate excreted earlier in the fermentation (10, 11, 17). This provides acetoacetate for decarboxylation to acetone and provides butyryl-CoA for subsequent conversion to butanol during solventogenesis. Consequently, CoA trans- ferase plays a key role in this economically important fermentation. Andersch et al. (1) found in batch solvent fermentations with Clostridium acetobutylicum DSM 1732 that CoA trans- ferase activity measured with butyrate or acetate increases * Corresponding author. t Present address: Department of Chemical Engineering, North- western University, 2145 Sheridan Road, Evanston, IL 60208. gradually, reaching peak activity during solventogenesis. Hartmanis et al. (10) determined the relative activity with various carboxylic acids, using crude extracts of C. aceto- butylicum ATCC 824. They reported that a wide variety of carboxylic acids, including formate, straight-chain acids up to heptanoate, and several branched-chain and unsaturated carboxylic acids, were used as substrates. More recently, Yan et al. (30) studied the change in levels of CoA transferase in batch fermentations of Clostridium beijerinckii NRRL B592 and NRRL B593. The transferase activity was not measured separately, but was measured in combination with acetoacetyl-CoA hydrolase as an indica- tion of total acetoacetate-forming (acetoacetyl-CoA-utiliz- ing) activity. This combined activity showed only a two- to threefold increase at the onset of solventogenesis, in con- trast to more dramatic increases in the levels of other enzymes also involved in solventogenesis. Those workers suggested that expression of acetoacetate-forming activity might be an early event in the switch from acidogenesis to solventogenesis. A more complete understanding of CoA transferase in butanol-forming clostridia is required to better understand the regulation of acid uptake and solvent formation. A better understanding of the regulatory features of CoA transferase and other key metabolic enzymes will help guide the im- provement of this fermentation. This paper reports the purification to homogeneity and the physical and kinetic characterization of CoA transferase in C. acetobutylicum. MATERIALS AND METHODS Organism and growth conditions. C. acetobutylicum ATCC 824 was grown in a 14-liter fermentor (Microferm; New Brunswick Scientific Co., Inc., Edison, N.J.) sparged with nitrogen and controlled at 37°C and pH 4.5. Mainte- nance of the organism, including storage, transfer, and heat shocking, has been described previously (20). The medium contained the following (per liter of distilled water): KH2PO4, 0.75 g; K2HPO4, 0.75 g; MgSO4 H20, 0.4 g; MnSO4. H20, 0.01 g; FeSO4 7H20, 0.01 g; NaC1, 1.0 g; 323 Vol. 55, No. 2 on March 10, 2021 by guest http://aem.asm.org/ Downloaded from

Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

APPLIED AND ENVIRONMENTAL MICROBIOLOGY, Feb. 1989, P. 323-3290099-2240/89/020323-07$02.00/0Copyright C) 1989, American Society for Microbiology

Coenzyme A Transferase from Clostridium acetobutylicumATCC 824 and Its Role in the Uptake of Acids

DENNIS P. WIESENBORN,1 FREDERICK B. RUDOLPH,2 AND ELEFTHERIOS T. PAPOUTSAKISlt*

Department of Chemical Engineering' and Department ofBiochemistry,2 Rice University, Houston, Texas 77251-1892

Received 12 October 1988/Accepted 25 November 1988

Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold tohomogeneity. This enzyme was stable in the presence of 0.5 M ammonium sulfate and 20% (vol/vol) glycerol,whereas activity was rapidly lost in the absence of these stabilizers. The kinetic binding mechanism was PingPong Bi Bi, and the Km values at pH 7.5 and 30°C for acetate, propionate, and butyrate were, respectively,1,200, 1,000, and 660 mM, while the Km value for acetoacetyl-CoA ranged from about 7 to 56 ,uM, dependingon the acid substrate. The Km values for butyrate and acetate were high relative to the intracellularconcentrations of these species; consequently, in vivo enzyme activity is expected to be sensitive to changes inthose concentrations. In addition to the carboxylic acids listed above, this CoA transferase was able to convertvalerate, isobutyrate, and crotonate; however, the conversion of formate, n-caproate, and isovalerate was notdetected. The acetate and butyrate conversion reactions in vitro were inhibited by physiological levels ofacetone and butanol, and this may be another factor in the in vivo regulation of enzyme activity. The optimumpH of acetate conversion was broad, with at least 80% of maximal activity from pH 5.9 to greater than 7.8. Thepurified enzyme was a heterotetramer with subunit molecular weights of about 23,000 and 25,000.

Coenzyme A (CoA) transferases activate carboxylic acidsto the respective CoA thioesters at the expense of the CoAthioester of another species of carboxylic acid. The reactionis easily reversed. The presence of CoA transferase in cellextracts was first demonstrated by Stadtman for Clostridiumkluyveri (26). Different types of CoA transferase have beensubsequently shown to be present in several other species ofbacteria. These enzymes have been prepared in very pureform from Escherichia coli (24), Peptostreptococcus elsdenii(28), and Clostridium sp. strain SB4 (3). Some physical andkinetic properties of these enzymes have been character-ized. All studied CoA transferases show relatively broadsubstrate specificities, but each enzyme has characteristicpreferred substrates.Among clostridia that are used to carry out the acetone-

butanol fermentation, a CoA transferase activity that trans-fers CoA from acetoacetyl-CoA to butyrate or to acetate hasbeen found in crude extracts (1, 10, 30). However, the directtransfer of CoA from acetyl-CoA to butyrate has not beendetected (1). The role of CoA transferase in butanol-formingclostridia is fundamentally different from that in other bac-teria for which it has been characterized. Generally, in thislatter group of bacteria the enzyme is involved in the uptakeof substrates for energy and structural use (3, 24, 28). Inbutanol-forming clostridia, however, the CoA transferaseacts mainly to detoxify the medium by removing the acetateand butyrate excreted earlier in the fermentation (10, 11, 17).This provides acetoacetate for decarboxylation to acetoneand provides butyryl-CoA for subsequent conversion tobutanol during solventogenesis. Consequently, CoA trans-ferase plays a key role in this economically importantfermentation.Andersch et al. (1) found in batch solvent fermentations

with Clostridium acetobutylicum DSM 1732 that CoA trans-ferase activity measured with butyrate or acetate increases

* Corresponding author.t Present address: Department of Chemical Engineering, North-

western University, 2145 Sheridan Road, Evanston, IL 60208.

gradually, reaching peak activity during solventogenesis.Hartmanis et al. (10) determined the relative activity withvarious carboxylic acids, using crude extracts of C. aceto-butylicum ATCC 824. They reported that a wide variety ofcarboxylic acids, including formate, straight-chain acids upto heptanoate, and several branched-chain and unsaturatedcarboxylic acids, were used as substrates.More recently, Yan et al. (30) studied the change in levels

of CoA transferase in batch fermentations of Clostridiumbeijerinckii NRRL B592 and NRRL B593. The transferaseactivity was not measured separately, but was measured incombination with acetoacetyl-CoA hydrolase as an indica-tion of total acetoacetate-forming (acetoacetyl-CoA-utiliz-ing) activity. This combined activity showed only a two- tothreefold increase at the onset of solventogenesis, in con-trast to more dramatic increases in the levels of otherenzymes also involved in solventogenesis. Those workerssuggested that expression of acetoacetate-forming activitymight be an early event in the switch from acidogenesis tosolventogenesis.A more complete understanding of CoA transferase in

butanol-forming clostridia is required to better understandthe regulation of acid uptake and solvent formation. A betterunderstanding of the regulatory features of CoA transferaseand other key metabolic enzymes will help guide the im-provement of this fermentation. This paper reports thepurification to homogeneity and the physical and kineticcharacterization of CoA transferase in C. acetobutylicum.

MATERIALS AND METHODS

Organism and growth conditions. C. acetobutylicumATCC 824 was grown in a 14-liter fermentor (Microferm;New Brunswick Scientific Co., Inc., Edison, N.J.) spargedwith nitrogen and controlled at 37°C and pH 4.5. Mainte-nance of the organism, including storage, transfer, and heatshocking, has been described previously (20). The mediumcontained the following (per liter of distilled water):KH2PO4, 0.75 g; K2HPO4, 0.75 g; MgSO4 H20, 0.4 g;MnSO4. H20, 0.01 g; FeSO4 7H20, 0.01 g; NaC1, 1.0 g;

323

Vol. 55, No. 2

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 2: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

324 WIESENBORN ET AL.

TABLE 1. Purification of CoA transferase fromC. acetobutylicum

Procedure Protein Activity Sp act cPatrifin YieldProcedure (mg) (U) (U/Mg) (ftold (%

Crude extract 3,710 1,320 0.36 1 100Ammonium sulfate frac- 855 720 0.84 2.3 55

tionationOctyl Sepharose columns 55.4 322 5.8 16.1 24Phenyl Sepharose column 33.7 229 6.8 18.9 17Gel filtration 9.8 171 17.4 48.3 13PolyPROPYL A column 3.4 99 29.1 80.8 7.5

asparagine, 2.0 g; yeast extract, 5.0 g; (NH4)2SO4, 2.0 g;antifoam C, 0.2 ml; and glucose, 50 g. Cells were harvestedby centrifugation approximately 3 h after the fermentationreached the stationary growth phase and then were storedunder nitrogen at -20°C.CoA transferase assay. CoA transferase was routinely

assayed at room temperature in the butyrate conversiondirection by monitoring the decrease in A310 as an indicationof the disappearance of the enolate form of acetoacetyl-CoA(24). The assay mixture contained, in a final volume of 1.0ml, 100 mM Tris hydrochloride (pH 7.5), 100 mM potassiumbutyrate (pH 7.5), 20 mM MgCl2, 0.10 mM acetoacetyl-CoA,5% (vol/vol) glycerol, and less than 0.05 U of CoA transfer-ase. Addition of acetoacetyl-CoA initiated the reaction. Thehydrolysis of acetoacetyl-CoA was measured by omittingpotassium butyrate and was subtracted from the slope ob-tained when butyrate was included in the cuvette. Thishydrolysis was less than 10% of the uncorrected slope whencrude extracts were used. One unit of activity is defined asthe amount of enzyme which converts 1 ,umol of acetoacetyl-CoA per min under these conditions. The extinction coeffi-cient is 8.0 mM-1 cm-' (27).

Other assays. Protein was assayed by the method ofBradford (4) with bovine serum albumin as the standard.The concentration of acetoacetyl-CoA used in the stan-

dard assay and in kinetic and substrate specificity studieswas determined by incubating the acetoacetyl-CoA withneutral hydroxylamine for 1 min to release the reduced formof CoA (CoASH) (7). CoASH was then measured by addingDTNB [5,5'-dithiobis(2-nitrobenzoic acid)] and measuringthe total change in A412. The extinction coefficient is 13.6mM-1 cm-' (7). The calculated concentration was correctedfor the presence of CoASH in the stock solution of ace-toacetyl-CoA by adding DTNB to stock solution that had notbeen hydrolyzed with neutral hydroxylamine and then mea-suring the initial change in A412.

Purification of CoA transferase. All of the steps describedbelow were carried out at room temperature, and anaerobicconditions were not used; however, between steps theenzyme preparation was stored under nitrogen at 4°C. BufferA contained 25 mM MOPS (4-morpholinepropanesulfonicacid) (pH 7.0), 0.5 M (NH4)2SO4, and 20% (vol/vol) glycerol.Buffer B contained 25 mM MOPS (pH 7.0) and 15% (vol/vol)glycerol. Buffer C was buffer B with 0.75 M (NH4)2SO4. Theresults of a typical preparation are presented in Table 1.

(i) Preparation of cell extract. Frozen cells (95 g) were

thawed in two volumes of buffer A and then passed once

through a French pressure cell (model FA-073; SLM Instru-ments, Inc., Urbana, Ill.) at 15,000 lb/in2. Unlysed cells wereremoved by centrifugation, resuspended in buffer A to a totalvolume of 40 ml, and passed through the pressure cell a

second time at 15,000 lb/in2. The disrupted-cell suspensions

were combined and centrifuged for 45 min at 30,000 x g toremove cell debris.

(ii) Ammonium sulfate precipitation. Recrystallized ammo-nium sulfate was added to a final concentration of 1.8 M.After 30 min of stirring, the suspension was centrifuged at20,000 x g for 15 min and the pellet was discarded. Ammo-nium sulfate was then added to give the supernatant a finalconcentration of 2.2 M. After being stirred for 30 min, thesuspension was centrifuged at 20,000 x g for 15 min and thesupernatant was discarded. Pellets from this second precip-itation were resuspended in buffer A to a final volume ofabout 40 ml.

(iii) Octyl Sepharose CL-4B chromatographies. The resus-pended ammonium sulfate fraction was diluted with buffer Bcontaining 2.0 M (NH4)2SO4 to a final concentration of 0.8 M(NH4)2SO4. This enzyme preparation was then applied di-rectly to an Octyl Sepharose CL-4B (Pharmacia Fine Chem-icals, Piscataway, N.J.) column (2.5 by 6.3 cm) equilibratedwith 1.0 M (NH4)2SO4 in buffer B. The column was thenwashed with equilibration buffer. A flow rate of about 80ml/h was maintained at all times. The CoA transferase doesnot bind to the column at this level of ammonium sulfate, buta large amount of contaminating protein is removed. Frac-tions containing high activity were combined to give a totalvolume of about 80 ml. Crystalline ammonium sulfate wasadded to a final concentration of 1.5 M in the combinedfractions. This preparation was then applied directly to asecond Octyl Sepharose CL-4B column (2.5 by 14 cm)equilibrated with 1.5 M (NH4)2SO4 in buffer B at 80 ml/h.The column was washed with equilibration buffer until theA280 was less than 0.8, and then activity was eluted by alinear gradient from 1.5 to 1.0 M (NH4)2SO4 in a total volumeof 240 ml. Washing and elution were performed at 160 ml/h.

(iv) Phenyl Sepharose CL-4B chromatography. The com-bined sample from the previous step, totalling 90 ml, wasapplied directly to a Phenyl Sepharose CL-4B (Pharmacia)column (2.5 by 3.8 cm) equilibrated with buffer C at 110 ml/h.The column was washed with buffer C until the A280 was lessthan 0.4, and then activity was eluted by a linear gradientfrom 100% buffer C to 75% (vol/vol) buffer C plus 25%(vol/vol) ethylene glycol, in a total volume of 240 ml.Washing and elution were performed at 160 ml/h. Fractionswith high activity, totalling 190 ml, were concentrated toabout 40 ml with type PM10 ultrafiltration membranes in astirred cell (Amicon, Danvers, Mass.) and then concentratedfurther to 2.8 ml with filtering cones (Amicon type CF25ultrafiltration membrane).

(v) Sephacryl S-300 chromatography. The concentratedsample was applied to a Sephacryl S-300 (Pharmacia) col-umn (1.6 by 90 cm) equilibrated with buffer A and eluted at15 ml/h. Fractions with high activity, totalling about 10 ml,were concentrated to about 3 ml with filtering cones (Amicontype CF25 ultrafiltration membranes), and then crystallineammonium sulfate was added to a final concentration of 1.5M.

(vi) High-pressure liquid chromatography with hydropho-bic-interaction column. A polyPROPYL A (Custom LC,Inc., Houston, Tex.) column (4.6 by 200 mm) was used witha high-pressure liquid chromatography system (LaboratoryControl Data, Riviera Beach, Fla.). The column was equili-brated with 1.1 M (NH4)2SO4 in buffer B. After applicationof the sample, the column was washed with 15 ml ofequilibration buffer, and then activity was eluted with alinear gradient from 1.1 to 0.9 M (NH4)2SO4 in buffer B in atotal volume of 15 ml. The flow rate was 1.0 mlUmin through-out this step. Fractions with high activity were combined

APPL. ENVIRON. MICROBIOL.

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 3: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

CoA TRANSFERASE IN CLOSTRIDIUM ACETOBUTYLICUM 325

into three groups corresponding to the beginning, middle,and end of the activity peak, and then glycerol was added toa final concentration of 20% (vol/vol).

Determination of native and subunit molecular weights. Thenative molecular weight was determined by measuring elu-tion volume on a calibrated gel filtration column (2). Thecolumn (1.6 by 90 cm) contained Sephacryl S-300 equili-brated with buffer A. Samples of 2.0 ml were applied andthen eluted at 15 ml/h with collection of 5-min fractions. Thefollowing standards were used to calibrate the column (mo-lecular weights in parentheses): carbonic anhydrase(29,000), serum albumin (66,000), alcohol dehydrogenase(150,000), ,-amylase (200,000), apoferritin (443,000), andthyroglobulin (669,000).The subunit molecular weight was determined by gel

electrophoresis on a 12% polyacrylamide gel in the presenceof sodium dodecyl sulfate (18).

Kinetic studies. The following kinetic studies were con-ducted at 30°C and were initiated by addition of acetoacetyl-CoA. With the exception of studies of optimum pH andrelative rate of simultaneous conversion with acetate andbutyrate, the standard assay for CoA transferase was usedwith modifications as noted below. All measurements weredone at least in duplicate.The conversion of butyrate was studied with potassium

butyrate concentrations ranging from 33 to 300 mM andacetoacetyl-CoA concentrations ranging from 3.9 to 100 ,uM.The conversion of propionate was studied with potassiumpropionate concentrations ranging from 60 to 300 mM andacetoacetyl-CoA concentrations ranging from 3.9 to 35 ,uM.The conversion of acetate was studied with potassium ace-tate concentrations ranging from 60 to 300 mM and ace-toacetyl-CoA concentrations ranging from 2.8 to 25 ,uM.The substrate specificity was studied at 100 mM concen-

trations of each carboxylic acid tested. Stock solutions of thefree acids were first adjusted to pH 7.5 by the addition ofKOH.The effect of metabolites was studied by using 300 mM

potassium butyrate or 300 mM potassium acetate. Theaddition of NAD, ADP, and ATP, and to a lesser extentCoASH, acetyl-CoA, and butyryl-CoA, caused a markedreduction of the extinction coefficient of acetoacetyl-CoA,probably by chelating the magnesium ions which are addedto the assay. The extinction coefficient is sensitive to theconcentration of magnesium ions and was measured in eachcase for use in calculating the enzyme activity.The optimum pH of acetate conversion was studied by

adding the coupling enzyme phosphotransacetylase to con-vert acetyl-CoA to acetyl phosphate (24). The rate of break-age of acyl-CoA bonds was determined by monitoring thedecrease in A232. The cuvette contained, in a final volume of1.0 ml, 15 mM MES (morpholineethanesulfonic acid) buffer,15 mM MOPS buffer, 100 mM potassium phosphate, 5%(vol/vol) glycerol, 0.1 mM acetoacetyl-CoA, 3 U of phos-photransacetylase, 100 mM potassium acetate, and sufficientCoA transferase to cause a maximal rate of absorbancechange of about 0.3 min-'. Phosphotransacetylase was notlimiting over the range of pH 5.4 to 7.8. The hydrolysis ofacetoacetyl-CoA was measured by omitting acetate from thecuvette and was less than 3% of the rate of absorbancechange when acetate was present.The simultaneous use of acetate and butyrate as alterna-

tive substrates was studied by measuring the rate of conver-sion of acetoacetyl-CoA (and thus the total rate of formationof acetyl-CoA and butyryl-CoA) by monitoring the decreasein A310. Then, under identical assay conditions, only the rate

of acetyl-CoA formation was measured, by monitoring thedecrease in A232 (24). The rate of butyryl-CoA formation wascalculated by subtracting the rate of acetyl-CoA formationfrom the rate of acetoacetyl-CoA conversion. Each assaycontained, in a final volume of 1.0 ml, 100 mM Tris hydro-chloride (pH 7.5), 5% (vol/vol) glycerol, 20 mM MgCl2, 25mM potassium arsenate (pH 7.5), 0.1 mM acetoacetyl-CoA,potassium acetate ranging in concentration from 0 to 200 mM(pH 7.5), potassium butyrate ranging in concentration from 0to 200 mM (pH 7.5), 3 U of phosphotransacetylase, andabout 0.02 U of pure CoA transferase. The phosphotrans-acetylase, which was purified from C. kluyveri (Sigma Chem-ical Co., St. Louis, Mo.), hydrolyzes acetyl-CoA but notbutyryl-CoA in the presence of arsenate.

Source of chemicals. The dye-binding protein assay kit waspurchased from Bio-Rad Laboratories, Richmond, Calif. Allother chemicals were purchased from Sigma Chemical Co.,except where noted.

RESULTS

Purification of CoA transferase. The results of the purifi-cation of CoA transferase are presented in Table 1. Theprocedures for each step are described in Materials andMethods. The high concentrations of ammonium sulfate andglycerol added to buffers were necessary to stabilize activityin both crude and pure preparations of the enzyme. Omittingeither salt or glycerol or both led to rapid loss of activity.EDTA and dithiothreitol apparently are not needed to pro-tect enzyme activity. Activity was somewhat cold labile, butfull activity was restored by allowing the enzyme prepara-tion to stand at room temperature for at least 1 h. Pure andpartially pure enzyme preparations stored for 1 week at 4°Cunder nitrogen showed less than 5% loss in activity. Cellextracts could be stored at room temperature for up to 4 dayswithout detectable loss in activity.The high concentrations of salt and glycerol required to

stabilize activity prevented CoA transferase and other pro-teins from binding to ion-exchange and dye-ligand affinitycolumns. Consequently, the purification procedure reportedhere relies mainly on various hydrophobic-interaction chro-matographies. No dialysis is required at any step. Thispurification clearly illustrates the potential of new chroma-tography techniques for purification of proteins with unusualstability requirements.

Fractions from the high-pressure liquid chromatographyhydrophobic-interaction column were electrophoresed on apolyacrylamide gel in the presence of sodium dodecyl sulfate(Fig. 1). Fractions corresponding to the beginning, middle,and end of the activity peak all showed two bands ofessentially equal intensities that corresponded to molecularweights of 23,000 and 25,000. The elution volume corre-sponded to a native molecular weight of about 93,000. Thus,it is likely that this CoA transferase is a heterotetramer oftwo a and two I subunits.

Kinetic studies. The butyrate conversion data as analyzedby double-reciprocal plots were best fitted by a family oflines that are approximately parallel at moderate concentra-tions of both substrates but that become hyperbolic concave-up at high concentrations of the varied substrate (Fig. 2). Atthe highest concentration of the fixed substrate, the fittedline intersects the other lines to the right of the reciprocal-velocity axis. This behavior, known as double competitivesubstrate inhibition, is consistent only with a Ping Pong Bi Bikinetic binding mechanism (5, 6, 23). The inhibitory effect ofhigh concentrations of one substrate is relieved by increasing

VOL. 55, 1989

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 4: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

326 WIESENBORN ET AL.

24-" jP

.24-

FIG. 1. Sodium dodecyl sulfate-polyacrylamide electrophoresisgel showing CoA transferase and standards. Proteins were stainedwith Coomassie blue. The left lane contained molecular weightstandards (molecular weight is given below in parentheses and ismarked beside the pertinent band in thousands) bovine serumalbumin (66,000), ovalbumin (45,000), glyceraldehyde-3-phosphatedehydrogenase (36,000), carbonic anhydrase (29,000), trypsinogen(24,000), trypsin inhibitor (20,100), and a-lactalbumin (14,200). Theright three lanes contained purified CoA transferase from threedifferent fractions within the activity peak from the last step ofpurification.

the level of the other substrate. Values of Km. and inhibitionconstants (K,) for both butyrate and acetoacetyl-CoA werecalculated from intercept and slope replots by using theprocedure described by Segel (23). Those values are pre-sented in Table 2.On double-reciprocal plots for acetate-conversion and

propionate conversion (not shown), the data were best fittedby a family of parallel lines and did not exhibit significantsubstrate inhibition at the levels of substrates used. Kmvalues for acetoacetyl-CoA, acetate, and propionate werecalculated from intercept and slope replots (Table 2). TheKm, values for the carboxylic acids were remarkably high.

Relative-activity measurements for the acetate conversionreaction in the range of pH 5.4 to 7.8 indicated that relativeactivity was greater than 90% from at least pH 6.4 to 7.8. Therelative activity was 82% at pH 5.9 and 56% at pH 5.4.

Substrate specificity. Table 3 shows the relative activity ofCoA transferase with a 100 mM concentration of the potas-sium salt of various carboxylic acids. The relative activitywas highest with acetate, propionate, and butyrate, while theactivity with valerate, isobutyrate, and crotonate was signif-icant but low. The activity with the other carboxylic acidstested was below the level of detection.

Studies of acetate and butyrate as alternative substrates.Measurements of the relative rate of conversion of acetateand butyrate when both acids were present confirmed the

0

cm

C"

a)e-4000

tD

Lo_L

o-.

.,

620CD

u

c- o

o0 ..L . . L

0.000 0.005 0.010 0.015 0.020 0.025 0.030 0.0351/ lbutyratel (1/mM)

X0 .s

00~~~~~~~~~~~

u

~0

00

CL

cu .00,

G2

0 50 too 150 200 2501/ lAcAcCoA] ( 1/mM)

FIG. 2. Double-reciprocal plots of butyrate conversion data. (A)Butyrate was varied with the following fixed concentrations ofacetoacetyl-CoA: O, 3.9 ,uM; O, 5.0 ,uM; x, 7.0 ,uM; +, 11.7 ILM;A, 35 ,uM; anid O, 100 ,uM. (B) Acetoacetyl-CoA was varied with thefollowing fixed concentrations of butyrate: O, 33.3 ,uM; x, 42.9mM; +, 60 mM; A, 100 mM; and O, 300 mM. The dashed linescorrespond to 100 ,uM acetoacetyl-CoA (A) and 300 mM butyrate(B).

general trend predicted for a Ping Pong Bi Bi mechanismwhen two alternative substrates are present (22). Addition ofacetate depressed the rate of conversion of butyrate, andaddition of butyrate depressed the rate of conversion ofacetate. When equimolar acetate and butyrate were presenttogether, the rate of conversion of butyrate relative toacetate was 27% with each substrate at 100 mM and 32% at200 mM. In comparison, when the two acids were presentseparately, the rate of conversion of butyrate relative toacetate was 26% with each substrate at 100 mM and 25% at200 mM. When 100 mM acetate and 200 mM butyrate were

A

APPL. ENVIRON. MICROBIOL.

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 5: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

CoA TRANSFERASE IN CLOSTRIDIUM ACETOBUTYLICUM 327

TABLE 2. Kinetic constants of conversion of acetate,propionate, and butyrate by CoA transferasea

Km (mM) Ki (mM) RelativeSubstrate Vmax

Acid AACoA Acid AACoA (%)

Acetate 1,200 0.021 ND ND 100Propionate 1,000 0.007 ND ND 32Butyrate 660 0.056 410 0.21 13

a Abbreviations: AACoA, acetoacetyl-CoA; ND, not determined.

present together, the rate of conversion of butyrate was 71%of that of acetate.

Effects of metabolites, cofactors, and inorganic salts on CoAtransferase activity. Table 4 summarizes the in vitro effects ofvarious metabolites and cofactors on activity with acetateand butyrate. Butanol and acetone at 200 mM resulted insignificant inhibition; the highest physiological levels ofthese products are 200 and 100 mM, respectively (17, 19).Significant inhibition by 200 mM 1-propanol and 2-butanone,which are not found in this fermentation, suggests that theCoA transferase is inhibited by solvents in general. Inhibi-tion by acetoacetate, acetyl-CoA, and butyryl-CoA was tobe expected, because those metabolites are products ofacetate and butyrate conversion. Inhibition by pyruvate wascomparable to that of acetoacetate and may be a result of thesomewhat similar structure of these two compounds. Unfor-tunately, there are little or no data on in vivo levels of theseintermediate metabolites and cofactors in butanol-producingclostridia. The cofactors ADP, ATP, and NAD did not causesignificant inhibition, even at the relatively high levelstested. Acetate conversion was more sensitive to inhibitionby CoASH, acetyl-CoA, and butyryl-CoA than was butyrateconversion, but generally the two reactions showed similardegrees of inhibition by the other metabolites and cofactorstested.

In addition to the results presented in Table 4, the in vitroeffects of KCl, NaCl, LiCl, and (NH4)2SO4 on butyrateactivation were tested. These inorganic salts caused lessthan 10% inhibition at 100 mM. Furthermore, there was nodifference in activity when sodium butyrate was used inplace of potassium butyrate.

DISCUSSION

The CoA transferase purified in this study showed onlysuperficial similarities to other bacterial CoA transferaseswhich have been purified and characterized. The size wassimilar, consisting of two subunits with molecular weights of

TABLE 3. Substrate specificity of CoA transferase with respectto various carboxylic acids at 100 mM

Carboxylic acid Relativerate (%)

Formate .............. ................................ <1Acetate ............. ................................. 100Propionate ................ .............................. 48n-Butyrate ................ .............................. 26n-Valerate .............................................. 2.7n-Caproate .............................................. <1Isobutyrate ................. ............................. 2.5Isovalerate .............................................. <1Isocaproate .............................................. <1(DL)-3-Methyl-n-valerate............................................... <1Crotonate ............... ............................... 3.9

TABLE 4. In vitro effects of metabolites and cofactorson CoA transferase

Metabolite or Assay concn Relative activity (%) with:cofactor (mM) Acetate Butyrate

Control 100 100Butanol 200 63 53Ethanol 200 95 98Acetone 200 70 71I-Propanola 200 81 782-Butanonea 200 48 50Acetoacetate 20 61 59Pyruvate 20 61 64ATP 5 92 99ADP 5 99 100NAD 5 91 89CoASH 0.9 75 94Acetyl-CoA 1.0 73 80Butyryl-CoA 1.0 62 77

a Not a metabolite of this fermentation.

about 23,000 and 25,000, compared with 23,000 and 26,000 inE. coli (24) and 23,000 and 25,000 in Clostridium sp. strainSB4 (3). Also, as with all other CoA transferases studied, thekinetic binding mechanism was Ping Pong Bi Bi (15, 23, 28).The metabolic role of CoA transferase in C. acetobutylicum,however, was fundamentally different from that of otherCoA transferases which have been characterized, and thiswas reflected in the much higher Km values for the carbox-ylic acid and different specificities. With the C. acetobutyli-cum CoA transferase, the Km values for the preferredcarboxylic acid substrates butyrate, propionate, and acetatewere very high, ranging from 660 to 1,200 mM. In contrast,for CoA transferase in Clostridium sp. strain SB4, theapparent Km values for eight preferred carboxylic acidsubstrates, including butyrate, propionate, and acetate,ranged from 0.8 to 29 mM (3). Similarly, with E. coli CoAtransferase, a preliminary Km value for acetate of about 20mM was reported (25). All of these Km values were deter-mined with acetoacetyl-CoA as the acyl-CoA substrate. Inaddition, other CoA transferases which have been purifiedare not highly unstable in the absence of salt and glycerol,unlike this CoA transferase.A major objective in purifying this CoA transferase was to

carry out in vitro experiments to investigate the effect onenzyme activity of substrate concentration, pH, and variousmetabolites and cofactors to assess what factors are impor-tant in vivo for the metabolic regulation of this enzyme. Theintracellular concentrations of acetate, butyrate, butanol,ethanol, acetone, and some cofactors (14, 21; C. L. Meyer,Ph.D. thesis, Rice University, Houston, Tex., 1987) and thepH (8, 12, 13) are known to vary over a significant rangeunder batch fermentation conditions.The intracellular concentrations of acetate and butyrate do

not exceed about 300 and 700 mM, respectively, and aregenerally much lower than the apparent Km values of 1,200and 660 mM, respectively (8, 12, 14). At such subsaturatinglevels, CoA transferase activity in vivo should be verysensitive to changes in the intracellular acetate and butyrateconcentrations. The high Km values suggest that the enzymehas evolved in such a way that the uptake of acetate andbutyrate from the medium is a graded response to theprogressive toxic effects of increasing levels of these carbox-ylic acids.The intracellular concentration of acetoacetyl-CoA in bu-

tanol-forming clostridia has not been reported. Thus, it is not

VOL. 55, 1989

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 6: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

328 WIESENBORN ET AL.

possible to use the Km values reported here to predict theeffect of fluctuations in levels of that intermediate on in vivoactivity.Of the metabolites and cofactors tested for their effects on

enzyme activity, butanol and acetone are of particularinterest, because they caused significant inhibition at phys-iological levels. Butanol and acetone accumulate in themedium during the solventogenic phase of the fermentation,up to maximum concentrations of about 200 and 100 mM,respectively (19). These nonpolar solvents passively diffusethrough the cell membrane so that the internal concentrationis slightly higher than the external concentration. Acetone isa direct product of the CoA transferase-acetoacetate decar-boxylase pathway, and butanol is an indirect product; there-fore, inhibition by these two solvents may act as a feedbackcontrol to slow the accumulation of toxic levels of butanol. Itis interesting that high levels of ethanol, which is a relativelyminor fermentation product, had little effect on enzyme

activity.Of the remaining metabolites that were tested for their

effects on acetate and butyrate conversion, either theyshowed no inhibition or there are not yet sufficient data ontheir concentrations in vivo to determine whether the inhi-bition is physiologically important.The relative flux in vivo of acetate versus butyrate through

the CoA transferase pathway has not yet been determined,because both species may be simultaneously produced ortaken up by two other pathways (21), namely, the acetatekinase-phosphotransacetylase and butyrate kinase-phospho-transbutyrylase pathways, respectively. During solventoge-nesis in batch fermentations, the level of butyrate oftendrops much more than that of acetate, suggesting that someregulatory mechanism may cause CoA transferase to con-vert butyrate in preference to acetate (1, 10, 21). The in vitrostudies reported here show that this CoA transferase con-verts acetate at a higher rate than butyrate when bothsubstrates are present together at equal levels; however, therelative rate of conversion in vivo should be sensitive to therelative concentrations of acetate and butyrate and to thelevels of CoASH, acetyl-CoA, and butyryl-CoA, which tendto inhibit acetate conversion more than butyrate conversion.The intracellular level of butyrate is often higher than that ofacetate in vivo (8, 12, 14). Nevertheless, the in vitro resultssuggest that butyrate is probably not usually converted byCoA transferase in preference to acetate. Therefore, thegreater decrease in butyrate concentration described aboveprobably does not result solely from the CoA transferasereaction; at least one of the other two pathways mentionedabove is probably also involved.The intracellular pH in C. acetobutylicum has been shown

to decrease from 7.0 to as low as 5.5 in a batch fermentation(13), and this CoA transferase apparently has fairly highactivity throughout this range. The decrease in activity at thelower end of this range was not expected, because lowinternal pH is generally the result of high levels of acetateand butyrate. It seems unlikely that this characteristic of theenzyme is beneficial to the metabolism of the organism, andit may be of little consequence.

Substrate specificity was studied, because of the potentialapplication of C. acetobutylicum in novel fermentation. Forexample, Jewell et al. (16) showed that propionate andvalerate, but not isobutyrate, could be taken up by C.acetobutylicum and converted to the respective alcohol.Although this CoA transferase has broad substrate speci-ficity, like other CoA transferases described in the literature,it should be noted that carboxylic acids may also be con-

verted to the respective acyl-CoA by the butyrate kinase-phosphotransbutyrylase pathway (29). Both of those en-zymes also have broad specificities (9, 29a).The properties of the CoA transferase described here are

different from those reported by Hartmanis et al., whostudied this enzyme in crude extracts of C. acetobutylicumATCC 824. Hartmanis et al. reported significantly higherrelative activity with propionate, butyrate, n-valerate, isobu-tyrate, and crotonate, and they also reported significantrelative activity with formate, n-caproate, isovalerate, andisocaproate (10). Those workers used the same assay condi-tions with respect to pH and concentrations of substratesand magnesium ions as we did. The difference is not theresult of their use of crude extracts; we measured in a crudeextract the relative activity with acetate, propionate, andbutyrate and observed the same ratio of activity of about4:2:1, respectively, that was measured by using pure en-zyme. Also, at least in the cases of acetate and butyrate, thepresence of ammonium sulfate and glycerol in the assay didnot affect the relative rate. Hartmanis et al. reported that theenzyme is activated by potassium salts of carboxylic acidsand shows no activity with the sodium salts (10), whereas wefound no difference in activity between equimolar sodiumbutyrate and potassium butyrate.The results presented by Hartmanis et al., as well as those

presented by Andersch et al. for C. acetobutylicum DSM1732 (1) and by Yan et al. for C. beijerinckii (30), are basedon work with crude extracts that did not contain ammoniumsulfate or glycerol to stabilize activity. We found little or noactivity in such extracts of C. acetobutylicum ATCC 824,although some activity was recovered if ammonium sulfateand glycerol were added after the crude extract was pre-pared. Even though Andersch et al. reported specific activ-ities with acetate of up to 1.47 U/mg (1), which are compa-rable to the highest activities we have observed, wecalculated from sample spectrophotometric data included intheir paper that their specific activities are more probably onthe order of about 0.01 U/mg. Furthermore, if only lowlevels of CoA transferase are present, measurement is com-plicated, because thiolase together with enzymatic and non-enzymatic hydrolysis of acetoacetyl-CoA create a relativelyhigh background that cannot be accurately accounted for.Consequently, in the case of crude extracts of butanol-forming clostridia that have not been stabilized with salt andglycerol, the CoA transferase assay may give artifactualresults.Evidence that this has been the case can be seen when

studying the profiles of CoA transferase and acetoacetatedecarboxylase in solvent-producing batch fermentations.Andersch et al. (1) and Yan et al. (30) both found that theincrease in CoA transferase activity was relatively smalland/or gradual, in contrast to the abrupt and significantinduction of the acetoacetate decarboxylase activity. Inmany butanol-forming clostridia, CoA transferase and ace-toacetate decarboxylase form a two-enzyme pathway thatmight be expected to be coordinately induced. When CoAtransferase activity was stabilized with salt and glycerol, theinduction of CoA transferase paralleled that of acetoacetatedecarboxylase (unpublished data; M. H. W. Husemann,personal communication).

ACKNOWLEDGMENTS

We are grateful to Michael Husemann and Jeff Cary for technicalassistance and to Marie Monroe and Amy Winter for typing.

This work was supported by a graduate fellowship to D.P.W. from

APPL. ENVIRON. MICROBIOL.

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from

Page 7: Coenzyme Transferase Clostridium ATCC · Coenzyme A (CoA) transferase from Clostridium acetobutylicum ATCC 824 was purified 81-fold to homogeneity. This enzyme was stable in the presence

CoA TRANSFERASE IN CLOSTRIDIUM ACETOBUTYLICUM 329

the National Science Foundation and by National Science Founda-tion grants CPT-8415815 and CBT-8518959.

LITERATURE CITED1. Andersch, W., H. Bahl, and G. Gottschalk. 1983. Level of

enzymes involved in acetate, butyrate, acetone and butanolformation by Clostridium acetobutylicumn. Eur. J. Appl. Micro-biol. Biotechnol. 18:327-322.

2. Andrews, P. 1964. Estimation of the molecular weight of pro-teins by Sephadex gel-filtration. Biochem. J. 91:222-233.

3. Barker, H. A., I.-M. Jeng, N. Neff, J. M. Robertson, F. K. Tam,and S. Hosaka. 1978. Butyryl-CoA:acetoacetate CoA-trans-ferase from a lysine-fermenting Clostridium. J. Biol. Chem.253:1219-1225.

4. Bradford, M. M. 1976. A rapid and sensitive method for thequantitation of microgram quantities of protein utilizing theprinciple of protein-dye binding. Anal. Biochem. 72:248-254.

5. Cleland, W. W. 1970. Steady state kinetics, p. 1-65. In P. D.Boyer (ed.), The enzymes, 3rd ed., vol. 2. Academic Press,Inc., New York.

6. Fromm, H. J. 1975. Initial rate enzyme kinetics, p. 41-81. In A.Kleinzeller, G. F. Springer, and H. G. Wittmann (ed.), Molec-ular biology, biochemistry and biophysics, vol. 22. Springer-Verlag KG, Berlin.

7. Gilbert, H. F., B. J. Lennox, C. D. Mossman, and W. C. Carle.1981. The relation of acyl transfer to the overall reaction ofthiolase I from porcine heart. J. Biol. Chem. 256:7371-7377.

8. Gottwald, M., and G. Gottschalk. 1985. The internal pH ofClostridium acetobutylicum and its effect on the shift from acidto solvent formation. Arch. Microbiol. 143:42-46.

9. Hartmanis, M. G. N. 1987. Butyrate kinase from Clostridiumacetobutylicum. J. Biol. Chem. 262:617-621.

10. Hartmanis, M. G. N., T. Klason, and S. Gatenbeck. 1984.Update and activation of acetate and butyrate in Clostridiumacetobutylicum. Appl. Microbiol. Biotechnol. 20:66-71.

11. Herrero, A. A., R. F. Gomez, B. Snedecor, C. J. Tolman, andM. F. Roberts. 1985. Growth inhibition of Clostridium thermo-cellum by carboxylic acids: a mechanism based on uncouplingby weak acids. Appl. Microbiol. Biotechnol. 22:53-62.

12. Huang, L., C. W. Forsberg, and L. N. Gibbins. 1986. Influenceof external pH and fermentation products on Clostridium ace-tobutylicum intracellular pH and cellular distribution of fermen-tation products. Appl. Environ. Microbiol. 51:1230-1234.

13. Huang, L., L. N. Gibbins, and C. W. Forsberg. 1985. Trans-membrane pH gradient and membrane potential in Clostridiumacetobutylicum during growth under acetogenic and solvento-genic conditions. Appl. Environ. Microbiol. 50:1043-1047.

14. Husemann, M. H. W., and E. T. Papoutsakis. 1988. Solvento-genesis in Clostridium acetobutylicum fermentations related tocarboxylic acid and proton concentration. Biotechnol. Bioeng.32:843-852.

15. Jencks, W. P. 1973. Coenzyme A transferases, p. 483-496. InP. D. Boyer (ed.), The enzymes, 3rd ed., vol. 9. AcademicPress, Inc., New York.

16. Jewell, J. B., J. B. Coutinho, and A. M. Kropinski. 1986.Bioconversion of propionic, valeric, and 4-hydroxybutyric acidsinto the corresponding alcohols by Clostridium acetobutylicumNRRL 527. Curr. Microbiol. 13:215-219.

17. Jones, D. T., and D. R. Woods. 1986. Acetone-butanol fermen-tation revisited. Microbiol. Rev. 50:484-524.

18. Laemmli, U. K. 1970. Cleavage of structural proteins during theassembly of the head of bacteriophage T4. Nature (London)227:680-685.

19. Marlatt, J. A., and R. Datta. 1986. Acetone-butanol fermenta-tion process development and economic evaluation. Biotechnol.Prog. 2:23-28.

20. McLaughlin, J. K., C. L. Meyer, and E. T. Papoutsakis. 1985.Gas chromatography and gateway sensors for on-line stateestimation of complex fermentations (butanol-acetone fermen-tations). Biotechnol. Bioeng. 27:1246-1257.

21. Reardon, K. F., T.-H. Scheper, and J. E. Bailey. 1987. Metabolicpathway rates and culture fluorescence in batch fermentationsof Clostridium acetobutylicum. Biotechnol. Prog. 3:153-167.

22. Rudolph, F. B., and H. J. Fromm. 1970. Use of isotopecompetition and alternative substrates for studying the kineticmechanism of enzyme action. I. Experiments with hexokinaseand alcohol dehydrogenase. Biochemistry 9:4660-4665.

23. Segel, I. H. 1975. Enzyme kinetics, p. 818-831. John Wiley &Sons, Inc., New York.

24. Sramek, S. J., and F. E. Frerman. 1975. Purification andproperties of Escherichia coli coenzyme A-transferase. Arch.Biochem. Biophys. 171:14-26.

25. Sramek, S. J., and F. E. Frerman. 1975. Escherichia colicoenzyme A-transferase: kinetics, catalytic pathway and struc-ture. Arch. Biochem. Biophys. 171:27-34.

26. Stadtman, E. R. 1953. The coenzyme A transphorase system inClostridium kluyveri. J. Biol. Chem. 203:501-512.

27. Stern, J. R. 1956. Optical properties of acetoacetyl-S-coenzymeA and its metal chelates. J. Biol. Chem. 221:33-44.

28. Tung, K. K., and W. A. Wood. 1975. Purification, new assay,and properties of coenzyme A transferase from Peptostrep-toc occus elsdenii. J. Bacteriol. 124:1462-1474.

29. Valentine, R. C., and R. S. Wolfe. 1960. Purification and role ofphosphotransbutyrylase. J. Biol. Chem. 235:1948-1952.

29a.Wiesenborn, D. P., F. B. Rudolph, and E. T. Papoutsakis. 1989.Phosphotransbutyrylase from Clostridium acetobutylicumATCC 824 and its role in acidogenesis. Appl. Environ. Micro-biol. 55:317-322

30. Yan, R.-T., C.-X. Zhu, C. Golemboski, and J.-S. Chen. 1988.Expression of solvent-forming enzymes and onset of solventproduction in batch cultures of Clostridium beijerinckii ("Clos-tridium butylicum"). Appl. Environ. Microbiol. 54:642-648.

VOL. 55, 1989

on March 10, 2021 by guest

http://aem.asm

.org/D

ownloaded from