Engineered Saccharomyces cerevisiaecapable of simultaneous ... · PDF fileEngineered...

14
Engineered Saccharomyces cerevisiae capable of simultaneous cellobiose and xylose fermentation Suk-Jin Ha a,b,1 , Jonathan M. Galazka c,1 , Soo Rin Kim a,b , Jin-Ho Choi a,b , Xiaomin Yang d , Jin-Ho Seo e , N. Louise Glass f , Jamie H. D. Cate c,g,2 , and Yong-Su Jin a,b,2 a Department of Food Science and Human Nutrition, University of Illinois, Urbana, IL 61801; b Institute for Genomic Biology, University of Illinois, Urbana, IL 61801; c Department of Molecular and Cell Biology, University of California, Berkeley, CA 94720; d BP Biofuels Business Unit, Berkeley, CA 97420; e Department of Agricultural Biotechnology, Seoul National University, Seoul 152-742, Korea; f Department of Plant and Microbial Biology, University of California, Berkeley, CA 94720; and g Physical Biosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720 Edited by Lonnie O'Neal Ingram, University of Florida, Gainesville, FL, and approved December 1, 2010 (received for review July 22, 2010) The use of plant biomass for biofuel production will require effi- cient utilization of the sugars in lignocellulose, primarily glucose and xylose. However, strains of Saccharomyces cerevisiae presently used in bioethanol production ferment glucose but not xylose. Yeasts engineered to ferment xylose do so slowly, and cannot utilize xylose until glucose is completely consumed. To overcome these bottlenecks, we engineered yeasts to coferment mixtures of xylose and cellobiose. In these yeast strains, hydrolysis of cellobiose takes place inside yeast cells through the action of an intracellular β-glucosidase following import by a high-affinity cellodextrin transporter. Intracellular hydrolysis of cellobiose mini- mizes glucose repression of xylose fermentation allowing cocon- sumption of cellobiose and xylose. The resulting yeast strains, cofermented cellobiose and xylose simultaneously and exhibited improved ethanol yield when compared to fermentation with either cellobiose or xylose as sole carbon sources. We also observed improved yields and productivities from cofermentation experi- ments performed with simulated cellulosic hydrolyzates, suggest- ing this is a promising cofermentation strategy for cellulosic biofuel production. The successful integration of cellobiose and xylose fer- mentation pathways in yeast is a critical step towards enabling economic biofuel production. biofuels cellodextrin transporter cofermentation intracellular β-glucosidase T here is a long history of microbial fermentation of glucose derived from cornstarch or sugarcane to fuels and chemicals. Modifying this process to ferment sugars produced from the depolymerization of plant biomass is a central paradigm for cellulosic biofuel production (1). While hydrolyzates from corn- starch and sugarcane contain only hexoses, hydrolyzates from cel- lulosic biomass contain various hexoses and pentoses. Hexoses are produced following the hydrolysis of cellulose by fungal cellulases, which primarily generate cellobiose that can be further hydrolyzed to glucose by β-glucosidases. Pentoses are produced by acid treatment of hemicellulose, which releases xylose and arabinose. While the composition is variable, plant biomass hydrolyzates consists of 70% cellodextrins and glucose, and 30% of xylose (2). Successful conversion of cellulosic biomass into biofuel will therefore require organisms capable of efficient utilization of xylose as well as cellodextrins and glucose (3, 4). The traditional glucose fermenting yeast, Saccharomyces cerevisiae, cannot ferment xylose because it does not possess a functional xylose assimilation pathway. However, as wild-type S. cerevisiae can metabolize xylulose using the pentose phosphate pathway, the introduction of genes (XYL1 and XYL2) encoding xylose reductase (XR) and xylitol dehydrogenase (XDH) from Pichia stipitis (taxonomic classification has been changed to Schef- fersomyces stipitis) facilitated xylose assimilation in this yeast (5, 6). In addition, either overexpression of XKS1 coding for endogenous xylulokinase (7, 8), or introduction of XYL3 coding for P. stipitis xylulokinase (9), significantly improved the rate and yield of xylose fermentation in S. cerevisiae. Despite these endea- vors, reported ethanol yields and productivities from xylose fermentation by engineered strains were much inferior to those of xylose fermentation by P. stipitis, as well as those of glucose fermentation by S. cerevisiae (10, 11). Moreover, both natural and engineered microorganisms showed reduced ethanol toler- ance during xylose fermentation as compared to glucose fermen- tation (12). Combined with the lower fermentation rate, the reduced ethanol tolerance during xylose fermentation poses a significant problem in the fermentation of sugar mixtures con- taining high concentrations of glucose (70100 gL) and xylose (4060 gL) present in cellulosic hydrolyzates. Because micro- organisms utilize glucose preferentially, at the time of glucose de- pletion (when cells begin to use xylose) the ethanol concentration is already high enough (3545 gL of ethanol) to further reduce the xylose fermentation rate. As a result, sequential utilization of xylose after glucose depletion because of so called glucose repressionis a significant barrier to successful utilization of mixed sugars in cellulosic hydrolyzates. To bypass the problems caused by glucose repression in a mixed sugar fermentation, we developed a unique strategy to co- ferment cellobiose, a dimer of glucose, with xylose simultaneously (Fig. 1). Wild-type S. cerevisiae cannot assimilate cellobiose be- cause it lacks both a cellobiose transporter and a β-glucosidase capable of hydrolyzing cellobiose into glucose. Therefore, we introduced a newly discovered cellodextrin transporter and intra- cellular β-glucosidase from the cellulolytic fungi, Neurospora crassa (13), into S. cerevisiae strains engineered to ferment xylose. A different approach enabling cofermentation of cellobiose and xylose through a display of β-glucosidase on the surface of xylose- fermenting S. cerevisiae has been reported (14). Although both approaches target cofermentation of cellobiose and xylose, the two differ in the location of cellobiose hydrolysis and its effect on xylose transport. The surface display of β-glucosidase gener- ates glucose extracellularly. In contrast, in the present design, cel- lobiose is hydrolyzed intracellularly following transport. Because xylose uptake by S. cerevisiae is facilitated by hexose transporters which are severely inhibited by glucose (1517), the extracellular glucose generated by surface display of β-glucosidase can inhibit xylose uptake (14) whereas the intracellular glucose from intra- cellular hydrolysis of cellobiose will not. Perhaps due to this fundamental distinction, our strategy achieves faster xylose and Author contributions: S.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. designed research; S.-J.H., S.R.K., J.-H.C., and Y.-S.J. performed research; J.-H.S. and N.L.G. contributed new reagents/analytic tools; S.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. analyzed data; and S.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1 S.-J.H. and J.M.G. contributed equally to this work. 2 To whom correspondence may be addressed. E-mail: [email protected] or [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1010456108/-/DCSupplemental. 504509 PNAS January 11, 2011 vol. 108 no. 2 www.pnas.org/cgi/doi/10.1073/pnas.1010456108

Transcript of Engineered Saccharomyces cerevisiaecapable of simultaneous ... · PDF fileEngineered...

Engineered Saccharomyces cerevisiae capable ofsimultaneous cellobiose and xylose fermentationSuk-Jin Haa,b,1, Jonathan M. Galazkac,1, Soo Rin Kima,b, Jin-Ho Choia,b, Xiaomin Yangd, Jin-Ho Seoe,N. Louise Glassf, Jamie H. D. Catec,g,2, and Yong-Su Jina,b,2

aDepartment of Food Science and Human Nutrition, University of Illinois, Urbana, IL 61801; bInstitute for Genomic Biology, University of Illinois, Urbana,IL 61801; cDepartment of Molecular and Cell Biology, University of California, Berkeley, CA 94720; dBP Biofuels Business Unit, Berkeley, CA 97420;eDepartment of Agricultural Biotechnology, Seoul National University, Seoul 152-742, Korea; fDepartment of Plant and Microbial Biology, University ofCalifornia, Berkeley, CA 94720; and gPhysical Biosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720

Edited by Lonnie O'Neal Ingram, University of Florida, Gainesville, FL, and approved December 1, 2010 (received for review July 22, 2010)

The use of plant biomass for biofuel production will require effi-cient utilization of the sugars in lignocellulose, primarily glucoseand xylose. However, strains of Saccharomyces cerevisiae presentlyused in bioethanol production ferment glucose but not xylose.Yeasts engineered to ferment xylose do so slowly, and cannotutilize xylose until glucose is completely consumed. To overcomethese bottlenecks, we engineered yeasts to coferment mixturesof xylose and cellobiose. In these yeast strains, hydrolysis ofcellobiose takes place inside yeast cells through the action of anintracellular β-glucosidase following import by a high-affinitycellodextrin transporter. Intracellular hydrolysis of cellobiose mini-mizes glucose repression of xylose fermentation allowing cocon-sumption of cellobiose and xylose. The resulting yeast strains,cofermented cellobiose and xylose simultaneously and exhibitedimproved ethanol yield when compared to fermentation witheither cellobiose or xylose as sole carbon sources.We also observedimproved yields and productivities from cofermentation experi-ments performed with simulated cellulosic hydrolyzates, suggest-ing this is a promising cofermentation strategy for cellulosic biofuelproduction. The successful integration of cellobiose and xylose fer-mentation pathways in yeast is a critical step towards enablingeconomic biofuel production.

biofuels ∣ cellodextrin transporter ∣ cofermentation ∣intracellular β-glucosidase

There is a long history of microbial fermentation of glucosederived from cornstarch or sugarcane to fuels and chemicals.

Modifying this process to ferment sugars produced from thedepolymerization of plant biomass is a central paradigm forcellulosic biofuel production (1). While hydrolyzates from corn-starch and sugarcane contain only hexoses, hydrolyzates from cel-lulosic biomass contain various hexoses and pentoses. Hexosesare produced following the hydrolysis of cellulose by fungalcellulases, which primarily generate cellobiose that can be furtherhydrolyzed to glucose by β-glucosidases. Pentoses are producedby acid treatment of hemicellulose, which releases xylose andarabinose. While the composition is variable, plant biomasshydrolyzates consists of ∼70% cellodextrins and glucose, and∼30% of xylose (2). Successful conversion of cellulosic biomassinto biofuel will therefore require organisms capable of efficientutilization of xylose as well as cellodextrins and glucose (3, 4).

The traditional glucose fermenting yeast, Saccharomycescerevisiae, cannot ferment xylose because it does not possess afunctional xylose assimilation pathway. However, as wild-typeS. cerevisiae can metabolize xylulose using the pentose phosphatepathway, the introduction of genes (XYL1 and XYL2) encodingxylose reductase (XR) and xylitol dehydrogenase (XDH) fromPichia stipitis (taxonomic classification has been changed to Schef-fersomyces stipitis) facilitated xylose assimilation in this yeast(5, 6). In addition, either overexpression of XKS1 coding forendogenous xylulokinase (7, 8), or introduction of XYL3 codingfor P. stipitis xylulokinase (9), significantly improved the rate and

yield of xylose fermentation in S. cerevisiae. Despite these endea-vors, reported ethanol yields and productivities from xylosefermentation by engineered strains were much inferior to thoseof xylose fermentation by P. stipitis, as well as those of glucosefermentation by S. cerevisiae (10, 11). Moreover, both naturaland engineered microorganisms showed reduced ethanol toler-ance during xylose fermentation as compared to glucose fermen-tation (12). Combined with the lower fermentation rate, thereduced ethanol tolerance during xylose fermentation poses asignificant problem in the fermentation of sugar mixtures con-taining high concentrations of glucose (∼70–100 g∕L) and xylose(∼40–60 g∕L) present in cellulosic hydrolyzates. Because micro-organisms utilize glucose preferentially, at the time of glucose de-pletion (when cells begin to use xylose) the ethanol concentrationis already high enough (∼35–45 g∕L of ethanol) to further reducethe xylose fermentation rate. As a result, sequential utilizationof xylose after glucose depletion because of so called “glucoserepression” is a significant barrier to successful utilization ofmixed sugars in cellulosic hydrolyzates.

To bypass the problems caused by glucose repression in amixed sugar fermentation, we developed a unique strategy to co-ferment cellobiose, a dimer of glucose, with xylose simultaneously(Fig. 1). Wild-type S. cerevisiae cannot assimilate cellobiose be-cause it lacks both a cellobiose transporter and a β-glucosidasecapable of hydrolyzing cellobiose into glucose. Therefore, weintroduced a newly discovered cellodextrin transporter and intra-cellular β-glucosidase from the cellulolytic fungi, Neurosporacrassa (13), into S. cerevisiae strains engineered to ferment xylose.A different approach enabling cofermentation of cellobiose andxylose through a display of β-glucosidase on the surface of xylose-fermenting S. cerevisiae has been reported (14). Although bothapproaches target cofermentation of cellobiose and xylose, thetwo differ in the location of cellobiose hydrolysis and its effecton xylose transport. The surface display of β-glucosidase gener-ates glucose extracellularly. In contrast, in the present design, cel-lobiose is hydrolyzed intracellularly following transport. Becausexylose uptake by S. cerevisiae is facilitated by hexose transporterswhich are severely inhibited by glucose (15–17), the extracellularglucose generated by surface display of β-glucosidase can inhibitxylose uptake (14) whereas the intracellular glucose from intra-cellular hydrolysis of cellobiose will not. Perhaps due to thisfundamental distinction, our strategy achieves faster xylose and

Author contributions: S.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. designed research; S.-J.H.,S.R.K., J.-H.C., and Y.-S.J. performed research; J.-H.S. and N.L.G. contributed newreagents/analytic tools; S.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. analyzed data; andS.-J.H., J.M.G., X.Y., J.H.D.C., and Y.-S.J. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1S.-J.H. and J.M.G. contributed equally to this work.2To whom correspondence may be addressed. E-mail: [email protected] or [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1010456108/-/DCSupplemental.

504–509 ∣ PNAS ∣ January 11, 2011 ∣ vol. 108 ∣ no. 2 www.pnas.org/cgi/doi/10.1073/pnas.1010456108

cellobiose coconsumption rates, and may prove optimal for bio-fuel production from cellulosic feedstocks.

ResultsConstruction of Cellobiose-Fermenting S. cerevisiae Through Introduc-tion of Cellodextrin Transporter and β-glucosidase from N. crassa.Three N. crassa cellodextrin transporters (cdt-1, cdt-2, andNCU00809) were introduced into a S. cerevisiae strain expressingan intracellular β-glucosidase (gh1-1). All three transformantswere able to grow and produce ethanol with cellobiose as the solecarbon source (Fig. S1), but the three transformants exhibited dif-ferent cellobiose fermentation rates (cdt-1>cdt-2>NCU00809).The fastest cellulose-fermenting transformant (D801-130), ex-pressing both cdt-1 and gh1-1, consumed 40 g∕L of cellobiosewithin 24 h, producing 16.8 g∕L of ethanol. The volumetric pro-ductivity of cellobiose fermentation (PEthanol∕Cellobiose ¼ 0.7 g∕L·h)was lower than that of glucose fermentation (PEthanol∕Glucose ¼1.2 g∕L·h), and ethanol yield from cellobiose (YEthanol∕Cellobiose ¼0.42 g∕g or 0.26 mol∕molC) was slightly lower than ethanol yieldfrom glucose (YEthanol∕Glucose ¼ 0.43 g∕g or 0.28 mol∕molC) un-der the same culture conditions. However, the observed cellobioseconsumption rate and ethanol yield by the D801-130 were animprovement over S. cerevisiae strains engineered to fermentcellobiose through surface display of β-glucosidase (14, 18). Theseresults suggest that simultaneous expression of cdt-1 and gh1-1 inS. cerevisiae can result in an efficient cellobiose-fermenting yeast.

Construction of an Engineered S. cerevisiae Capable of Fermenting Xy-lose both Rapidly and Efficiently Through Rational and CombinatorialStrategies. Attempts to engineer xylose fermentation into S. cer-evisiae have suffered from low yields due to xylitol accumulation,and low productivities due to slow xylose fermentation rates(10, 11). In order to improve yield and productivity of xylosefermentation, we undertook both rational and combinatorialapproaches. First, we drastically reduced xylitol accumulationthrough reconstituting an efficient xylose metabolic pathway.XR prefers NADPH to NADH while XDH solely uses NADþ

as cofactor (19, 20), resulting in a redox imbalance that is thoughtto result in xylitol accumulation. While this imbalance has beencorrected by replacing wild-type XR with various mutant XRsexhibiting higher preferences to NADH (21, 22), these efforts re-sulted in much slower xylose assimilation rates despite reducingthe amounts of xylitol produced, as the mutant XRs had reducedspecific activities as compared to wild-type XR.

In contrast to previous studies that relied on either wild-typeXR or a mutant XR, we expressed both wild-type XR and a mu-

tant XR (R276H) in S. cerevisiae, along with XDH and XK toconstruct a functional xylose metabolic pathway in S. cerevisiae.The XR mutant (R276H) had been reported to exhibit muchhigher preference for NADH whereas wild-type XR showed two-fold higher preference for NADPH (22). The resulting xylose-fermenting S. cerevisiae strain (DA24) expressing both wild-typeXR and mutant XR (R276H) exhibited comparable XR activitieswith both NADPH and NADH in an in vitro XR activity assay.Unlike other engineered strains showing higher XR activitieswith preferred cofactors, strain DA24 showed similar XR activ-ities regardless of cofactors (Table 1).

Second, we further improved the xylose fermentation rateof DA24 using an evolutionary engineering approach (28). TheDA24 strain rapidly consumed xylose and produced ethanol withconsistent ethanol yields (YEthanol∕Xylose ¼ 0.31–0.32 g∕g or0.21–0.22 mol∕molC) in both shaker-flask and bioreactor fer-mentation experiments (Fig. S2). The DA24 strain also producednegligible amounts of xylitol. Still, the DA24 strain consumedxylose slower than the naturally existing xylose-fermenting yeast,P. stipitis. Using evolutionary engineering, one strain (DA24-16)isolated after repeated subcultures of the DA24 on xylose con-taining medium showed much faster xylose fermentation rateswhen compared to the parental strain under various cultureconditions (Table S1). The specific xylose uptake rates, ethanolyields, and ethanol production rates of the DA24 and DA24-16strains were much higher than engineered S. cerevisiae strains re-ported previously (Table S2). Interestingly, the DA24-16 strainconsumed xylose almost as fast as P. stipitis, the fastest xylose-fermenting yeast known. However, the ethanol yield fromDA24-16 was slightly lower than that from P. stipitis (Fig. 2).

Cofermentation of Cellobiose and Xylose Lead to Increased EthanolYield and Productivity. To further improve the newly engineeredxylose-fermenting strains, we introduced genes coding for acellodextrin transporter and a β-glucosidase (cdt-1 and gh1-1)into the efficient xylose-fermenting strain, DA24-16, to constructa strain capable of consuming cellobiose and xylose simulta-neously. We hypothesized that glucose repression of xylose utili-zation may be alleviated in this strain, due to the intracellularhydrolysis of cellobiose. One resulting strain with cdt-1 integratedinto the genome of DA24-16 and expressing gh1-1 from a multi-copy plasmid, strain DA24-16-BT3, coconsumed cellobiose andxylose and produced ethanol with yields of 0.38–0.39 g∕g in allconditions (Table S3). We also investigated potential synergisticeffects of cofermentation by culturing DA2416-BT3 under threedifferent conditions: 40 g∕L of cellobiose, 40 g∕L of xylose, and

Cellulosic biomass

Pretreatment Hemicellulose

Xylose

Cellulose

Cellobiose

Glucose

β-glucosidase(gh1-1)

Xylose

Xylitol

Xylulose

PPP

Glycolysis EthanolS. cerevisiaeDA24-16BT3

XYL1 and mXYL1

XYL2

XKS1

Cellodextrin transporter(cdt-1)

Time

[Glu

cose

&X

ylos

e]

[Eth

anol

]

Time

[Cel

lobi

ose

&X

ylos

e]

[Eth

anol

]

A B

C

Fig. 1. Strategy for simultaneous cofermentation ofcellobiose and xylose without glucose repression.(A) A strain improvement strategy to engineer yeastcapable of fermenting two nonmetabolizable sugars(cellobiose and xylose). The cellodextrin assimilationpathway consists of a cellodextrin transporter (cdt-1)and an intracellular β-glucosidase (gh1-1) from thefilamentous fungus N. crassa. The modified xylosemetabolic pathway utilizes xylose reductase isoen-zymes (wild-type XR and amutant XRR276H), xylitol de-hydrogenase (XYL2), and xylulokinase (XKS1) fromthe xylose-fermenting yeast P. stipitis. (B) Schematicfermentation profile of a sugar mixture containingglucose and xylose by the engineered S. cerevisiae.Glucose fermentation represses xylose fermentationcompletely so that xylose fermentation begins onlyafter glucose depletion (analogous fermentation re-sult shown in Fig. 5A). (C) Schematic fermentationprofile of a sugar mixture containing cellobiose andxylose by the engineered S. cerevisiae. Cellobioseand xylose are simultaneously utilized, as neithercarbon source represses consumption of the other(analogous fermentation result shown in Fig. 5B).

Ha et al. PNAS ∣ January 11, 2011 ∣ vol. 108 ∣ no. 2 ∣ 505

APP

LIED

BIOLO

GICAL

SCIENCE

S

40 g∕L of both sugars (total 80 g∕L of sugars). Interestingly,DA24-16BT3 was able to coconsume 80 g∕L of a cellobiose/xylose mixture within the same period that was required to con-sume 40 g∕L of cellobiose or 40 g∕L xylose separately (Fig. 3).Moreover, DA24-16BT3 produced ethanol with a higher yield(0.39 g∕g or 0.24 mol∕molC) from a mixture of cellobiose andxylose as compared to ethanol yields (0.31–0.33 g∕g or0.21–0.22 mol∕molC) from single sugar fermentations (cello-biose or xylose) (SI Text). Ethanol productivity also dramaticallyincreased from 0.27 g∕L·h to 0.65 g∕L·h during cofermentation.These results suggest that cofermentation of cellobiose andxylose can enhance overall ethanol yield and productivity.

Fermentation of Simulated Cellulosic Hydrolyzates by the EngineeredS. cerevisiae Capable of Cofermenting Cellobiose and Xylose.We alsoperformed fermentation experiments comparing our engineered

S. cerevisiae strain (DA24-16BT3) to P. stipitis, a yeast capable ofcofermenting cellobiose and xylose efficiently. Using a simulatedhydrolyzate (10 g∕L of glucose, 80 g∕L of cellobiose, and 40 g∕Lof xylose) based on the composition of energycane (SI Text), the

Table 1. Comparison of xylose reductase activity of DA24 with those of other engineered S. cerevisiae strainsexpressing wild-type or mutant XYL1

StrainsSpecific xylose reductase

activity (U∕mg)Ratio of XR activity

with NADH and NADPH ReferenceNADH NADPH

S. cerevisiae with wild-type XYL1 0.07 ± 0.03 0.09 ± 0.02 0.8 Amore et al. (23)0.10 ± 0.02 0.12 ± 0.03 0.8 Takuma et al. (24)

- 2.7 ± 2.5 - Walfridsson et al. (25)2.3 ± 0.3 3.4 ± 0.5 0.68 Tantirungkij et al. (5)

- 0.411 - Ho et al. (8)- 0.43 ± 0.03 - Eliasson et al. (26)- 1.1 ± 0.08 - Wahlbom et al. (27)

S. cerevisiae with mutant XYL1 9.3 ± 0.1 0.4 ± 0.0 23.1 Watanabe et al. (22)DA24 0.32 ± 0.15 0.29 ± 0.10 1.1 This study

Fig. 2. Comparisons of xylose consumption and ethanol production be-tween DA24-16 (A) and P. stipitis (B). Symbols: xylose (green square), ethanol(orange diamond), and cell growth (open circle).

Fig. 3. Synergistic effects from cofermentation of cellobiose and xylose bystrain DA24-16BT3. Symbols: cellobiose (red triangle up), xylose (greensquare), ethanol (orange diamond), and cell growth (open circle). (A) 40 g∕Lof cellobiose, (B) a mixture of 40 g∕L of cellobiose and 40 g∕L of xylose, and(C) 40 g∕L of xylose.

506 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1010456108 Ha et al.

DA24-16BT3 consumed glucose first and coconsumed cellobiose,and xylose rapidly. A total of 130 g∕L of sugars were consumedwithin 60 h even though small inoculums were used (OD600 ¼ 1.3,or ∼0.4 g dried cell/L). In contrast, P. stipitis could not finishfermenting the sugar mixture within the same period underidentical culture conditions (Fig. 4). DA24-16BT3 produced48 g∕L of ethanol within 60 h (YEthanol∕Sugars ¼ 0.37 g∕g or0.23 mol∕molC, and PEthanol∕Sugars ¼ 0.79 g∕L·h). Fermentationexperiments using a bioreactor with the same or higher inoculumsyielded consistent results (Table S4). We also performed fermen-tation experiments of sugar mixtures containing extreme concen-trations of glucose and xylose (100 g∕L of glucose and 60 g∕L ofxylose) and cellobiose and xylose (100 g∕L of cellobiose and60 g∕L xylose). As illustrated in Fig. 1B, xylose fermentation isseverely delayed after glucose consumption when the mixtureof glucose and xylose was used. However, cofermentation of cel-lobiose and xylose showed a slower ethanol production initially,but resulted in higher ethanol production at the end (Fig. 5).

DiscussionAlthough we obtained promising results from our cofermentationexperiments, further improvements to these yeast strains canlikely be made. For example, we observed transient accumulationof cellodextrins in the medium during cellobiose fermentation(Figs. S3 and S4). It is likely that the accumulated cellodextrinswere generated by the transglycosylation activity (29) of β-gluco-sidase (gh1-1), and secreted by the cellodextrin transporter(cdt-1), which likely facilitates the transport of cellodextrins inboth directions (intracellular ↔ extracellular). This transient cel-lodextrin accumulation may not reduce product yields becausethe accumulated cellodextrins are consumed by our engineeredyeast after cellobiose depletion. However, the accumulatedcellodextrins might decrease productivity because the transportrates of cellotriose and cellotetraose might be slower than thatof cellobiose. We also observed that small amounts of glucosewere constantly detected in the medium during cofermentation.Because even low amounts of glucose accumulation can repressxylose fermentation, glucose accumulation must be controlled atthe lowest possible levels. We hypothesize that the relative ex-pression level of the cellodextrin transporter and β-glucosidaseis likely to affect glucose accumulation, because more glucoseis accumulated in the medium when cdt-1 is introduced on a mul-ticopy plasmid than when cdt-1 is integrated into the genome. Forexample, the strain (DA24-16-BT) with both cdt-1 and gh1-1 on amulticopy plasmid showed relatively slower xylose utilizationthan DA24-16-BT3, potentially because of glucose repression(Fig. S5). Future optimization of cellodextrin transporter andβ-glucosidase expression levels, or the identification of β-glucosi-

dases with reduced transglycosylastion activities, may eliminatethe accumulation of glucose and cellodextrins during cofermen-tation.

Here, we demonstrated a unique strategy to allow the cofer-menation of hexose and pentose sugars by S. cerevisiae. By com-bining an efficient xylose utilization pathway with a cellobiosetransport system, we bypassed problems caused by glucoserepression. As a result, the engineered yeast cofermented twononmetabolizable sugars in cellulosic hydrolyzates, synergisti-cally, into ethanol. The unique cofermentation method advanceslignocellulosic technologies on both the saccharification andfermentation fronts. Most traditional fungal cellulase cocktailsrequire extra β-glucosidase enzyme to be added to fully convertcellobiose into glucose for yeast fermentation. The cellobiose/xylose cofermentation yeast makes it possible to use cellulasecocktails with limited β-glucosidase activities, lowering the en-zyme usage and cost associated with the cellulose saccharificationprocess. On the other hand, the synergy between the cellobioseand xylose cofermentation significantly increases ethanol produc-tivity, improving the fermentation economics. Furthermore, thepresence of small amounts of glucose from the pretreatment andhydrolysis of lignocellulosic materials should not affect the capa-city of the engineered yeast to convert the hexose and pentosesugar mixtures into ethanol.

In this study, we measured the capacity of our engineeredstrain to ferment various mixtures of sugars meant to mimichydrolyzates from plant biomass. However, we predict that thecapacity of our strain to coferment cellodextrins and xylose willmake it particularly useful during the Simultaneous Saccharifica-tion and CoFermentation (SSCF) of pretreated plant biomass(30). During SSCF, hemicellulose would first be hydrolyzed byacid pretreatment, resulting in xylose and cellulose. Then, fungalcellulases and the yeast strain developed here would be added,allowing the coconversion of xylose and cellobiose released bythe cellulases, into ethanol. Because little extracelluar glucoseis produced in this scheme, repression of xylose utilization willbe minimized, and cofermentation will proceed rapidly and syner-gistically.

Although the S. cerevisiae strain used in this study was a labora-tory strain, the fermentation performance of the engineeredstrain was very impressive as compared to published results(31). We envision that key fermentation parameters (yield andproductivity) can be improved further if we employ industrialstrains as a platform. Finally, applications of our cofermentationstrategy would not be limited to ethanol production. Because it isa foundational technology, the strategy presented here can becombined with any other product diversification technologiesto produce commodity chemicals and advanced biofuels.

Fig. 4. Cofermentation of glucose, cellobiose, and xylose by strain DA24-16BT3 and P. stipitis. (A) DA24-16BT3 fermentation profile of a sugar mixture contain-ing 80 g/L of cellobiose, 40 g∕L of xylose, and 10 g∕L of glucose. (B) P. stipitis fermentation of a sugar mixture containing 80 g∕L of cellobiose, 40 g∕L of xylose,and 10 g∕L of glucose. Symbols: cellobiose (red triangle up), xylose (green square), ethanol (orange diamond), OD (open circle), and glucose (blue triangledown).

Ha et al. PNAS ∣ January 11, 2011 ∣ vol. 108 ∣ no. 2 ∣ 507

APP

LIED

BIOLO

GICAL

SCIENCE

S

Materials and MethodsStrains and Plasmid Constructions. S. cerevisiae D452-2 (MATalpha, leu2, his3,ura3, and can1) was used for engineering of xylose and cellobiose metabo-lism in yeast. Escherichia coli DH5 (F–recA1 endA1 hsdR17 [rK– mK+] supE44thi-1 gyrA relA1) (Invitrogen) was used for gene cloning and manipulation.P. stipitis CBS 6054 (32) was obtained from Thomas Jeffries at University ofWisconsin-Madison. Strains and plasmids used in this work are described inTable S5. The primers used for confirming the transformation of expressioncassettes containing cdt-1, cdt-2, and gh1-1 are listed in Table S6.

Medium and Culture Conditions. E. coli was grown in Luria-Bertani medium;50 μg∕mL of ampicillin was added to the medium when required. Yeaststrains were routinely cultivated at 30 °C in YP medium (10 g∕L yeast extract,20 g∕L Bacto peptone) with 20 g∕L glucose. To select transformants using anamino acid auxotrophic marker, yeast synthetic complete (YSC) medium wasused, which contained 6.7 g∕L yeast nitrogen base plus 20 g∕L glucose,20 g∕L agar, and CSM-Leu-Trp-Ura (Bio 101) which supplied appropriate nu-cleotides and amino acids.

Fermentation Experiments. Yeast cultures were grown in YP medium contain-ing 20 g∕L of glucose or 20 g∕L of cellobiose to prepare inoculums for xyloseor cellobiose fermentation experiments, respectively. Cells at midexponentialphase from YP media containing 20 g∕L of glucose or cellobiose were har-vested and inoculated after washing twice by sterilized water. Flask fermen-tation experiments were performed using 50 mL of YP medium containingappropriate amounts of sugars in 250 mL flask at 30 °C with initial OD600 of∼1 or 10 under oxygen limited conditions. All of the flask fermentation ex-periments were repeated independently. The variations between indepen-dent fermentations were less than 5%. Fermentation profiles shown infigures are from one representative fermentation. Bioreactor fermentationswere performed in 400 mL of YP medium containing appropriate amounts ofsugars using Sixfors Bioreactors (Appropriate Technical Resources, Inc) at30 °C with an agitation speed of 200 rpm under oxygen limited conditions.Initial cell densities were adjusted to OD600 ¼ ∼1 or 10.

Yeast Transformation. Transformation of expression cassettes for construct-ing xylose and cellobiose metabolic pathways was performed using theyeast EZ-Transformation kit (BIO 101). Transformants were selected onYSC medium containing 20 g∕L glucose. Amino acids and nucleotides wereadded as necessary. For the construction of cellobiose consuming recombi-nant S. cerevisiae, transformations of cdt-1 and gh1-1 were selected onYSC medium containing 20 g∕L cellobiose. Introduction of expression cas-settes into yeast was confirmed by colony PCR with specific primers(Table S6).

Analytical Methods. Cell growth was monitored by optical density (OD) at600 nm using UV-visible Spectrophotometer (Biomate 5). Glucose, xylose,xylitol, glycerol, acetate, and ethanol concentrations were determined byhigh performance liquid chromatography (HPLC, Agilent Technologies1200 Series) equipped with a refractive index detector using a Rezex ROA-Organic Acid Hþ (8%) column (Phenomenex Inc.). The column was elutedwith 0.005 N of H2SO4 at a flow rate of 0.6 mL∕min at 50 °C. The analysisof cellodextrin in fermentation samples was performed using high perfor-mance anion exchange chromatography (HPAEC) analysis. HPAEC analysiswas performed with an analytical column for carbohydrate detection (Car-boPac PA100, Dionex Co.) and an electrochemical detector (ED40, DionexCo.). Filtered samples were eluted with a linear gradient from 100% bufferA (100 mM NaOH in water) to 60% buffer B (500 mM of sodium acetate inbuffer A) over 70 min. The flow rate of the mobile phase was maintainedat 1.0 mL∕min.

ACKNOWLEDGMENTS. The authors thank Dr. Huimin Zhao for sharing hispreliminary results, Dr. Isaac Cann for HPAEC analysis, and William Beesonfor helpful discussions. We also thank Dr. Thomas Jeffries for generously pro-viding P. stipitis CBS 6054. This work was supported by funding from EnergyBiosciences Institute to J.H.D.C. and Y.-S.J. Strain DA26-16 will be deposited inthe American Type Culture Collection.

1. Wyman CE (2007) What is (and is not) vital to advancing cellulosic ethanol. TrendsBiotechnol 25:153–157.

2. Carroll A, Somerville C (2009) Cellulosic biofuels. Annu Rev Plant Biol 60:165–182.3. Stephanopoulos G (2007) Challenges in engineering microbes for biofuels production.

Science 315:801–804.4. Ragauskas AJ, et al. (2006) The path forward for biofuels and biomaterials. Science

311:484–489.5. Tantirungkij M, Nakashima N, Seki T, Yoshida T (1993) Construction of xylose-

assimilating Saccharomyces cerevisiae. J Ferment Bioeng 75:83–88.6. Kotter P, Ciriacy M (1993) Xylose fermentation by Saccharomyces cerevisiae. Appl Mi-

crobiol Biotechnol 38:776–783.7. Toivari MH, Aristidou A, Ruohonen L, Penttila M (2001) Conversion of xylose to

ethanol by recombinant Saccharomyces cerevisiae: importance of xylulokinase(XKS1) and oxygen availability. Metab Eng 3:236–249.

8. Ho NWY, Chen Z, Brainard AP (1998) Genetically engineered Saccharomyces yeastcapable of effective cofermentation of glucose and xylose. Appl Environ Microbiol64:1852–1859.

9. Jin YS, Ni H, Laplaza JM, Jeffries TW (2003) Optimal growth and ethanol productionfrom xylose by recombinant Saccharomyces cerevisiae require moderate D-xyluloki-nase activity. Appl Environ Microbiol 69:495–503.

10. Jeffries TW, Jin YS (2004) Metabolic engineering for improved fermentation ofpentoses by yeasts. Appl Microbiol Biotechnol 63:495–509.

11. Hahn-Hagerdal B, Karhumaa K, Jeppsson M, Gorwa-Grauslund MF (2007) Metabolicengineering for pentose utilization in Saccharomyces cerevisiae. Adv Biochem EngBiotechnol 108:147–177.

12. Jeffries TW, Jin YS (2000) Ethanol and thermotolerance in the bioconversion of xyloseby yeasts. Adv Appl Microbiol 47:221–268.

13. Galazka JM, et al. (2010) Cellodextrin transport in yeast for improved biofuel produc-tion. Science 330:84–86.

14. Nakamura N, et al. (2008) Effective xylose/cellobiose co-fermentation and ethanolproduction by xylose-assimilating S. cerevisiae via expression of β-glucosidase on itscell surface. Enzyme Microb Tech 43:233–236.

15. LeeWJ, KimMD, Ryu YW, Bisson LF, Seo JH (2002) Kinetic studies on glucose and xylosetransport in Saccharomyces cerevisiae. Appl Microbiol Biotechnol 60:186–191.

Fig. 5. Fermentation profiles by DA24-16BT3 during the fermentation of sugar mixtures containing extreme concentrations of glucose and xylose, (A) 100 g∕Lof glucose and 60 g∕L of xylose and cellobiose and xylose, (B) 100 g∕L of cellobiose and 60 g∕L xylose. Actual fermentation profiles are similar to the patterns ofsugar utilization and ethanol production illustrated in Fig. 1. Symbols: cellobiose (red triangle up), xylose (green square), ethanol (orange diamond), OD (opencircle), and glucose (blue triangle down).

508 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1010456108 Ha et al.

16. Jojima T, Omumasaba CA, Inui M, Yukawa H (2010) Sugar transporters in efficientutilization of mixed sugar substrates: current knowledge and outlook. Appl MicrobiolBiotechnol 85:471–480.

17. Runquist D, Hahn-Hèagerdal B, Rêadstrèom P (2010) Comparison of heterologous xy-lose transporters in recombinant Saccharomyces cerevisiae. Biotechnol Biofuels 3:5.

18. Kotaka A, et al. (2008) Direct ethanol production from barley β-glucan by sake yeastdisplaying Aspergillus oryzae β-glucosidase and endoglucanase. J Biosci Bioeng105:622–627.

19. Verduyn C, et al. (1985) Properties of the NAD(P)H-dependent xylose reductase fromthe xylose-fermenting yeast Pichia stipitis. Biochem J 226:669–677.

20. Rizzi M, Harwart K, Bui-Thanh N-A, Dellweg H (1989) A kinetic study of the NADþ-xylitol-dehydrogenase from the yeast Pichia stipitis. J Ferment Bioeng 67:25–30.

21. Bengtsson O, Hahn-Hagerdal B, Gorwa-Grauslund MF (2009) Xylose reductase fromPichia stipitis with altered coenzyme preference improves ethanolic xylose fermenta-tion by recombinant Saccharomyces cerevisiae. Biotechnol Biofuels 2:9.

22. Watanabe S, et al. (2007) Ethanol production from xylose by recombinant Saccharo-myces cerevisiae expressing protein-engineered NADH-preferring xylose reductasefrom Pichia stipitis. Microbiology 153:3044–3054.

23. Amore R, Kötter P, Küster C, Ciriacy M, Hollenberg CP (1991) Cloning and expression inSaccharomyces cerevisiae of the NAD(P)H-dependent xylose reductase-encoding gene(XYL1) from the xylose-assimilating yeast Pichia stipitis. Gene 109:89–97.

24. Takuma S, et al. (1991) Isolation of xylose reductase gene of Pichia stipitis and itsexpression in Saccharomyces cerevisiae. Appl Biochem Biotechnol 28–29:327–340.

25. Walfridsson M, Anderlund M, Bao X, Hahn-Hägerdal B (1997) Expression of differentlevels of enzymes from the Pichia stipitis XYL1 and XYL2 genes in Saccharomyces cer-evisiae and its effects on product formation during xylose utilization. Appl MicrobiolBiotechnol 48:218–224.

26. Eliasson A, Christensson C, Wahlbom CF, Hahn-Hägerdal B (2000) Anaerobic xylose fer-mentation by recombinant Saccharomyces cerevisiae carrying XYL1, XYL2, and XKS1in mineral medium chemostat cultures. Appl Environ Microbiol 66:3381–3386.

27. Wahlbom CF, Van Zyl WH, Jonsson LJ, Hahn-Hägerdal B, Cordero Otero RR (2003)Generation of the improved recombinant xylose-utilizing Saccharomyces cerevisiaeTMB 3400 by random mutagenesis and physiological comparison with Pichia stipitisCBS 6054. FEMS Yeast Res 3:319–326.

28. Sauer U (2001) Evolutionary engineering of industrially important microbial pheno-types. Adv Biochem Eng Biot 73:129–169.

29. Christakopoulos P, Bhat MK, Kekos D, Macris BJ (1994) Enzymatic synthesis of trisac-charides and alkyl beta-D-glucosides by the transglycosylation reaction of beta-glucosidase from Fusarium oxysporum. Int J Biol Macromol 16:331–334.

30. Doran-Peterson J, et al. (2009) Simultaneous saccharification and fermentation andpartial saccharification and co-fermentation of lignocellulosic biomass for ethanolproduction. Springer Protocols 581:263–280.

31. Hahn-Hagerdal B, Karhumaa K, Fonseca C, Spencer-Martins I, Gorwa-Grauslund MF(2007) Towards industrial pentose-fermenting yeast strains. Appl Microbiol Biotechnol74:937–953.

32. Jeffries TW, et al. (2007) Genome sequence of the lignocellulose-bioconverting andxylose-fermenting yeast Pichia stipitis. Nat Biotechnol 25:319–326.

Ha et al. PNAS ∣ January 11, 2011 ∣ vol. 108 ∣ no. 2 ∣ 509

APP

LIED

BIOLO

GICAL

SCIENCE

S

Supporting InformationHa et al. 10.1073/pnas.1010456108SI TextSI Analysis. Improved ethanol yield of cofermentation when comparedto single sugar fermentations (cellobiose or xylose). First, the massbalances between substrate (xylose, cellobiose, and a mixture ofxylose and cellobiose) and fermentation products (cell mass andethanol) explain the improved yield of cofermentation. In bothcases of xylose (Fig. 3A) and cellobiose (Fig. 3C) fermentations,about 4.8 g∕L of cells (OD ∼ 16) were produced after consuming40 g∕L of sugars. In other words, the yields of biomass(YBiomass∕xylose) from either xylose or cellobiose were about0.12 g cell∕g sugar. In cofermentation (Fig. 3B), the final cell den-sity was only 6.2 g∕L (OD ∼ 22) even though twice amounts ofsugars (total 80 g∕L of sugars) were consumed. Therefore, theyield of biomass from the cofermentation was only 0.08 g∕g su-gars. These data explain how cofermentation results in higherethanol yield than single sugar fermentation. Second, xylose fer-mentation by engineered S. cerevisiae requires oxygen-limitedconditions for efficient ethanol production (1, 2). As shown inFig. 3A, ethanol production from xylose begins only when the celldensity is high enough to cause oxygen-limitation (after 12 h).However, in the case of cofermentation, yeast cells grew fasterand oxygen-limitation started earlier than for single sugarfermentation conditions. As such, consumption of sugars duringcofermentation can be less oxidative (or more fermentative) thansingle sugar fermentation, which resulted in more ethanol pro-duction. In addition, the slow release of glucose from intracellularhydrolysis of glucose may exert partial glucose repression, whichbrings about more fermentative sugar metabolism resulting im-proved ethanol production while xylose transport is not limited.

Prediction of sugar concentrations in cellulosic hydrolyzates. Thecomposition of different lignocellulosic plants varies broadly.For instance, the US Department of Energy biomass databaselists the composition of more than 150 biomass samples (http://www1.eere.energy.gov/biomass/m/feedstock_databases.html).The cellulose to hemicellulose ratios of these samples are between1.4 and 19, and the average is 2.3. Energy crops typically havehigher hemicellulose content than woody biomass. The averagecellulose to hemicellulose ratios of sugarcane bagasse, cornstover, and sorghum are 2.0, 1.85, and 2.14, respectively.We there-fore used a glucan/xylan ratio of 2 in our simulated sugar experi-ment design. The engineered yeast will likely be used inconjunction with traditional cellulase cocktails that are deficient

in β-glucosidase activities for the biofuels production. The bio-mass hydrolysis process may result in small amounts of glucosein the lignocellulosic hydrolysates as 6–30% glucan-to-glucoseconversions with incomplete cellulase cocktails were reported(3). Considering all the above factors, a sugar combination of80 g∕L cellobiose, 10 g∕L glucose, and 40 g∕L xylose was chosenin the simulated sugar experiments.

SI Discussion. Advantages of intracellular hydrolysis of cellobiose overextracellular hydrolysis. Our approach holds several advantagesover the cell surface display strategy employed by Nakamuraet al. (4). First, intracellular hydrolysis of cellobiose via the cel-lodextrin transport system (5) can reduce the glucose transportload of hexose/pentose sugar transporters. The transport ofthe extracellular glucose, generated from extracellular hydrolysisby a displayed enzyme on cell surface, may compete with xylosefor cross-membrane transportation because glucose inhibitsxylose transport competitively (2, 3, 5). In addition, the systempresented in here exploits the higher affinity that cellodextrintransporter have for cellobiose (KM ≈ 3–4 μM) as compared toβ-glucosidases [KM ≈ 100–1;000 μM (5)] and the S. cerevisiaehexose transporters’ apparent affinity for glucose [KM ≈ 1;000–10;000 μM (5)]. The surface display of a β-glucosidase relieson the extracellular hydrolysis of cellobiose to glucose by alow-affinity beta-glucosidase followed by transport via low-affinity hexose transporters, and will be compromised at bothsteps. These inefficiencies will become particularly importantduring simultaneous saccharification and fermentation, when so-luble sugars much be kept at a concentration that does not inhibitcellulases [19–410 μM (5)]. Second, expression levels of β-gluco-sidase on the cell surface needs careful optimization under givenconditions in order to prevent excessive hydrolysis of cellobiose,as noted by Nakamura et al. (4). Excessive hydrolysis of cellobiosewould result in glucose accumulation at high concentrations,which would impede cofermentation of xylose. In this sense,the cellodextrin transport system is more amenable for construct-ing cofermenting strain under various conditions. Third, the sta-bility of intracellular β-glucosidase will be higher than a displayedβ-glucosidase because the intracellular enzyme can be protectedfrom harsh external environments. Intracellular expression couldprovide a significant benefit in fermentation of lignocellulosichydrolyzates, which contain uncharacterized toxic or poisoningcompounds to enzymes.

1. Jin YS, Laplaza JM, Jeffries TW (2004) Saccharomyces cerevisiae engineered for xylosemetabolism exhibits a respiratory response. Appl Environ Microbiol 70:6816–6825.

2. Souto-Maior AM, Runquist D, Hahn-Hägerdal B (2009) Crabtree-negative charac-teristics of recombinant xylose-utilizing Saccharomyces cerevisiae. J Biotechnol143:119–123.

3. Medve J, Karlsson J, Lee D, Tjerneld F (1998) Hydrolysis of microcrystalline cellulose bycellobiohydrolase I and endoglucanase II from Trichoderma reesei: adsorption, sugarproduction pattern, and synergism of the enzymes. Biotechnol Bioeng 59:621–634.

4. Nakamura N, et al. (2008) Effective xylose/cellobiose cofermentation and ethanol pro-duction by xylose-assimilating S. cerevisiae via expression of beta-glucosidase on itscell surface. Enz Microb Technol. 43:233–236.

5. Galazka JM, et al. (2010) Cellodextrin transport in yeast for improved biofuel produc-tion. Science 330:84–86.

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 1 of 8

Fig. S1. Comparison of cellobiose utilization by yeast strains expressing one of three cellobiose transporters and a β-glucosidase (gh1-1). Figure shows utliza-tion by strains expressing (A) cdt-1, (B) NCU00809, and (C) cdt-2. Symbols: cellobiose (red triangle), ethanol (orange diamond), and OD (open circle).

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 2 of 8

Fig. S2. Xylose fermentation by strain DA24 under various conditions. (A) 40 g/L of xylose in a shaker flask, (B) 80 g/L of xylose in a flask, (C) 80 g/L of xylose in abioreactor. Symbols: xylose (green square), ethanol (orange diamond), and OD (open circle).

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 3 of 8

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 4 of 8

Fig. S3. HPLC chromatograms from each time point suggesting cellotriose and cellotetraose accumulation during cofermentation of cellobiose and xylose bystrain DA24-16BT3.

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 5 of 8

Fig. S4. HPAEC analysis showing cellodextrin accumulation in the fermentation medium after 22 h fermentation by strain DA24-16BT3 during cofermentationof cellobiose and xylose. (G1: glucose; G2: cellobiose; G3: cellotriose; G4: cellotetraose; and G5:cellopentaose).

Fig. S5. Comparison of sugar utilization patterns by transformants with an integrated copy of cdt-1 (A), and cdt-1 on a multicopy plasmid (B) during co-fermentation of 40 g/L of cellobiose and 40 g/L of xylose mixture. Symbols: cellobiose (red triangle), xylose (green square), ethanol (orange diamond), and OD(open circle).

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 6 of 8

Table S1. Comparison of fermentation parameters by DA24 and DA24-16 under two different sugar conditions

Carbon source StrainsProduced

ethanol (g∕L)Sugar consumption

rate (g∕L·h) Yield (g∕g)Productivity

(g∕L·h)

Xylose (80 g∕L) DA24 24.2 1.16 0.34 0.40DA24-16 27.9 1.32 0.35 0.47

Glucose (70 g∕L) and xylose (40 g∕L) DA24 34.8 1.45 0.39 0.74DA24-16 45.1 1.78 0.42 0.96

Table S2. Comparison of fermentation parameters of DA24 and DA24-16 with other engineered S. cerevisiae strains

Strain Relevant genotype/phenotypeSpecific xylose consumption

rate (g xylose∕g cell · h)Ethanol yield

(g∕g)Xylitol yield

(g∕g) Reference

DA24 XYL1, a mutant XYL1, XYL2, and XKS1 0.53 0.32 0.08 This studyDA24-16 Evolved isolate from DA24 0.71 0.35 0.04 This studyH1693 XYL1 and XYL2 0.09 0.04 0.47 (10)H1691 XYL1, XYL2, and XKS1 0.2 0.12 0.41 (10)TMB3399 XYL1, XYL2, and XKS1 NA 0.05 0.59 (11)TMB3400 Chemical mutant of TMB3399 NA 0.18 0.25 (11)C1 Evolved isolate from TMB3001 0.56 0.24 0.32 (12)H2674 (control) XYL1, XYL2, and XKS1 0.07 0.14 0.56 (13)H2673 (GPD1) XYL1, XYL2, XKS1, and GPD1 overexpression 0.06 0.17 0.49 (13)H2723 (Δzwf1) XYL1, XYL2, XKS1, and Δzwf1 0.05 0.18 0.29 (13)H2684 (GPD1Δzwf1) XYL1, XYL2, XKS1, GPD1 overexpression, and

Δzwf10.06 0.31 0.35 (13)

RWB202-AFX XI, evolved isolate 0.21 0.42 0.02 (14)RWB217 XI, XK, ΔGRE3, and overexpressed pentose

phosphate pathway (PPP) enzymesNA 0.43 0.003 (15)

RWB218 XI, XK, ΔGRE3, overexpressed PPP, andselected for enhanced glucose uptake

NA 0.41 0.001 (16)

H2490-4 XYL1, XYL2, XKS1, and adaptation 0.58 0.14 0.82 (17)TMB3001 XYL1, XYL2, and XKS1 0.15 0.31 0.29 (18)TMB3255 XYL1, XYL2, XKS1, and Δzwf1 0.02 0.41 0.05 (18)TMB3008 XYL1, XYL2, XKS1, and Δgnd1 0.08 0.38 0.13 (18)TMB3250 XYL1, XYL2, XKS1 0.1 0.3 0.3 (18)TMB3251 XYL1, XYL2, XKS1, and attenuated PGI 0.07 0.34 0.21 (18)TMB3256 XYL1, XYL2, XKS1, and ZWF1 0.06 0.36 0.13 (19)TMB3037 XYL1, XYL2, XKS1, and YRP13-ZWF1 0.11 0.34 0.19 (19)TMB3260 XYL1, XYL2, XKS1, and high XR activity 0.25 0.3 0.13 (20)TMB3253 control XYL1, XYL2, XKS1, strain for TMB3254 0.16 0.28 0.34 (19)TMB3254 XYL1, XYL2, XKS1, overproducing

transhydrogenase0.16 0.28 0.3 (19)

Toivari MH, Aristidou A, Ruohonen L, Penttilä M (2001) Conversion of xylose to ethanol by recombinant Saccharomyces cerevisiae: Importance of xylulokinase (XKS1) and oxygenavailability. Metab Eng 3:236–249.Wahlbom CF, Van Zyl WH, Jönsson LJ, Hahn-Hägerdal B, Cordero Otero RR (2003) Generation of the improved recombinant xylose-utilizing Saccharomyces cerevisiae TMB 3400 byrandom mutagenesis and physiological comparison with Pichia stipitis CBS 6054. FEMS Yeast Res 3:319–326.Sonderegger M, Sauer U (2003) Evolutionary engineering of Saccharomyces cerevisiae for anaerobic growth on xylose. Appl Environ Microbiol 69:1990–1998.Verho R, Londesborough J, Penttilä M, Richard P (2003) Engineering redox cofactor regeneration for improved pentose fermentation in Saccharomyces cerevisiae. Appl EnvironMicrobiol 69:5892–5897.Kuyper M, Winkler AA, Van Dijken JP, Pronk JT (2004) Minimal metabolic engineering of Saccharomyces cerevisiae for efficient anaerobic xylose fermentation: A proof of principle.FEMS Yeast Res 4:655–664.Kuyper M, et al. (2005) Metabolic engineering of a xylose-isomerase-expressing Saccharomyces cerevisiae strain for rapid anaerobic xylose fermentation. FEMS Yeast Res 5:399–409.Kuyper M, et al. (2005) Evolutionary engineering of mixed-sugar utilization by a xylose-fermenting Saccharomyces cerevisiae strain. FEMS Yeast Res 5:925–934.Pitkänen J-P, Rintala E, Aristidou A, Ruohonen L, Penttilä M (2005) Xylose chemostat isolates of Saccharomyces cerevisiae show altered metabolite and enzyme levels compared withxylose, glucose, and ethanol metabolism of the original strain. Appl Microbiol Biotechnol 67:827–837.JeppssonM, Johansson B, Hahn-Hägerdal B, Gorwa-GrauslundMF (2002) Reduced oxidative pentose phosphate pathway flux in recombinant xylose-utilizing Saccharomyces cerevisiaestrains improves the ethanol yield from xylose. Appl Environ Microbiol 68:1604–1609.Jeppsson M, Johansson B, Jensen PR, Hahn-Hägerdal B, Gorwa-Grauslund MF (2003) The level of glucose-6-phosphate dehydrogenase activity strongly influences xylose fermentationand inhibitor sensitivity in recombinant Saccharomyces cerevisiae strains. Yeast 20:1263–1272.

JeppssonM, Träff K, Johansson B, Hahn-Hägerdal B, Gorwa-GrauslundMF (2003) Effect of enhanced xylose reductase activity on xylose consumption and product distribution in xylose-fermenting recombinant Saccharomyces cerevisiae. FEMS Yeast Res 3:167–175.

Table S3. Cofermentation of cellobiose and xylose using mixtures containing various concentrations of cellobioseand xylose by DA24-16BT3 strain

Cellobiose/xylose (g∕L) Produced ethanol (g∕L) Sugar consumption rate (g∕L·h) Yield (g∕g) Productivity (g∕L·h)

20∕20 14.8 1.12 0.37 0.4130∕30 23.3 1.33 0.37 0.5040∕40 30.5 1.67 0.39 0.65

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 7 of 8

Table S4. Cofermentation of glucose, cellobiose, and xylose (10 g∕L, 80 g∕L, and 40 g∕L, respectively) using bioreactor byDA24-16BT3 strain using different inoculums

Target initial OD (A600) Produced ethanol (g∕L) Sugar consumption rate (g∕L·h) Ethanol yield (g∕g) Productivity (g∕L·h)

∼1 (1.2) 47.9 1.93 0.37 0.71∼10 (10.2) 48.1 2.18 0.37 0.82

Table S5. Strains and plasmids used in this study

Strain or plasmid Description Reference

StrainsD452-2 MATa, leu2, his3,ura3,can1 Hosaka et al. (6)D801-130 D452-2 expressing β-glucosidase (gh1-1) and cdt-1 (NCU00801) In this studyD809-130 D452-2 expressing β-glucosidase (gh1-1) and NCU00809 In this studyD8114-130 D452-2 expressing β-glucosidase (gh1-1) and cdt-2 (NCU08114) In this studyDA24 D452-2 expressing XYL1, mXYL1, XYL2, and XKS1 (Isogenic of D452-2 except for

leu2 ∷ TDH3P -XYL1-TDH3T ,ura3 ∷ URA3-PGKP-mXYL1-PGKT -PGKP -XYL2-PGKT ,Ty3 ∷ neo-TDHP -XKS1-TDHT )

In this study

DA24-16 Evolved strain of DA24 in xylose containing media In this studyDA24-16BT3 DA24-16 expressing β-glucosidase (gh1-1) in a multicopy plasmid and cdt-1

(NCU00801) though single-copy integrationIn this study

DA24-16BT-M DA24-16 expressing β-glucosidase (gh1-1) and cdt-1 (NCU00801) in multicopyplasmids

In this study

P. stipitis CBS 6054 NRRL Y-11545 ¼ ATCC58785 ¼ IFO10063 Jeffries et al. (7)PlasmidspRS425 LEU2, a multicopy plasmid Christianson et al. (8)pRS426 URA3, a multicopy plasmid Christianson et al. (8)pRS403 HIS3, anintegrative plasmid Sikorski et al. (9)pRS405 URA3, an integrative plasmid Sikorski et al. (9)pRS425-β-glucosidase β-glucosidase (gh1-1) under the control of PGK promoter in pRS425 Galazka et al. (5)pRS426-cdt-1 cdt-1 under the control of PGK promoter in pRS426 Galazka et al. (5)pRS426-cdt-2 cdt-2 under the control of PGK promoter in pRS426 Galazka et al. (5)pRS426-NCU00809 NCU00809 under the control of PGK promoter in pRS426 Galazka et al. (5)pRS403-cdt-1 cdt-1 under the control of PGK promoter in pRS403 In this study

Hosaka K, Nikawa J, Kodaki T, Yamashita S (1992) A dominant mutation that alters the regulation of INO1 expression in Saccharomyces cerevisiae. J Biochem 111:352–358.Jeffries TW, et al. (2007) Genome sequence of the lignocellulose-bioconverting and xylose-fermenting yeast Pichia stipitis. Nat Biotechnol 25:319–326.Christianson TW, Sikorski RS, Dante M, Shero JH, Hieter P (1992) Multifunctional yeast high-copy-number shuttle vectors. Gene 110:119–122.Sikorski RS, Hieter P (1989) A system of shuttle vectors and yeast host strains designed for efficient manipulation of DNA in Saccharomyces cerevisiae. Genetics 122:19–27.

Table S6. Synthetic oligonucleotides used in this study

Name Sequences

NCU00801-F ATGGATCCAAAAATGTCGTCTCACGGCTCCNCU00801-R ATGAATTCCTACAAATCTTCTTCAGAAATCAATTTTTGTTCAGCAACGATAGCTTCGGACNCU08114-F ATACTAGTAAAAATGGGCATCTTCAACAAGAAGCNCU08114-R GCATATCGATCTACAAATCTTCTTCAGAAATCAATTTTTGTTCAGCAACAGACTTGCCCTCATGNCU00130-F GCATACTAGTAAAAATGTCTCTTCCTAAGGATTTCCTCTNCU00130-R ATACTGCAGTTAATGATGATGATGATGATGGTCCTTCTTGATCAAAGAGTCA AAG

Ha et al. www.pnas.org/cgi/doi/10.1073/pnas.1010456108 8 of 8