You are on page 1of 14

Biotechnology DOI 10.1002/biot.201000301 Biotechnol. J.

2011, 6, 286–299
Journal

Review

Stress-related challenges in pentose fermentation to ethanol


by the yeast Saccharomyces cerevisiae

João R. M. Almeida1*,**, David Runquist1*, Violeta Sànchez i Nogué1*, Gunnar Lidén2*,


Marie F. Gorwa-Grauslund1*
1 Applied Microbiology, Lund University, Lund, Sweden
2 Chemical Engineering, Lund University, Lund, Sweden

Conversion of agricultural residues, energy crops and forest residues into bioethanol requires hy-
Received 6 October 2010
drolysis of the biomass and fermentation of the released sugars. During the hydrolysis of the hemi- Revised 17 December 2010
cellulose fraction, substantial amounts of pentose sugars, in particular xylose, are released. Fer- Accepted 20 December 2010
mentation of these pentose sugars to ethanol by engineered Saccharomyces cerevisiae under in-
dustrial process conditions is the subject of this review. First, fermentation challenges originating
from the main steps of ethanol production from lignocellulosic feedstocks are discussed, followed
by genetic modifications that have been implemented in S. cerevisiae to obtain xylose and arabi-
nose fermenting capacity per se. Finally, the fermentation of a real lignocellulosic medium is dis-
cussed in terms of inhibitory effects of furaldehydes, phenolics and weak acids and the presence
of contaminating microbiota.

Keywords: Contamination · Ethanol · Inhibitors · Pentose fermentation · Saccharomyces cerevisiae

1 Introduction cluding both natural pentose fermenting yeasts


such as Pichia stipitis, Candida shehatae or Pachy-
The hemicellulose fractions of many non-food bio- solen tannophilus [6] or metabolically engineered
mass resources that can be used for sustainable Saccharomyces cerevisiae (recently reviewed by [7,
biofuel production – including major agricultural 8]). This review discusses stress challenges in pen-
residues such as corn stover, wheat straw and sug- tose fermentation of hydrolysates using recombi-
ar cane bagasse – are pentose-rich (Table 1). Con- nant S. cerevisiae.
version of pentoses after hydrolysis can be
achieved using a variety of either natural ferment-
ing or engineered microorganisms such as engi- 2 Challenges in lignocellulosic fermentation
neered Escherichia coli [1], Zymomonas mobilis [2],
thermophilic bacteria such as Thermoanaerobac- The biological conversion route of lignocellulose to
terium ethanolicus [3] or Th. saccharolyticum [4], ethanol consists of a pretreatment step, an enzy-
fungi such as Rhizopus oryzae [5] and yeasts in- matic hydrolysis step, a fermentation step and a fi-
nal recovery of the formed ethanol [9]. A summary
of the challenges faced by the yeast S. cerevisiae
Correspondence: Professor Gunnar Lidén, Chemical Engineering, Lund during the production of ethanol from lignocellu-
University, PO Box 124, 221 00 Lund, Sweden losic feedstocks is shown in Fig 1. The objective of
E-mail: gunnar.liden@chemeng.lth.se the pretreatment – normally a high temperature
Fax: +46-46-149156

Abbreviations: HMF, 5-hydroxymethyl-2-furaldehyde; LAB, lactic acid bacte-


ria; PPP, pentose phosphate pathway; SSF, simultaneous saccharification ** The authors contributed equally to this work.
and fermentation; SSL, spent sulphite liquor; XDH, xylitol dehydrogenase; ** Current address: EMBRAPA Agroenergy, PqEB, Brasilia, 70770-901 DF,
XI, xylose isomerase; XK, xylulokinase; XR, xylose reductase Brazil

286 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

Table 1. Reported composition (g/100 g dry matter) of some pentose-rich lignocellulosic feedstocks

Material Glucan Mannan Galactan Xylan Arabinan Lignin Acetyl Reference


Corn stover (American) 36.1 1.8 2.5 21.4 3.5 17.2 3.2 [108]
Corn stover (Italian) 36.8 0.3 2.9 22.2 5.5 21.2 1.7 [108]
Sugar cane bagasse 43.3 NR NR 24.3 2.0 22.8 2.0 [109]
Wheat straw 41.2 NR NR 26.1a) – 19.1 4.2 [110]
32.6 0.8 20.1 3.3 26.5 2.0 [111]
Barley straw 36.8 NR 2.2 17.2 5.3 14.3 NR [112]
Rice straw 34.2 NR NR 24.5 NR 11.9 NR [113]
36.6 NR NR 16.1 NR 14.9 NR [114]
Switch grass 34.2 0.5 1.5 23.3 2.0 19.9 2.4 [115]
Salix 41.5 3.0 2.1 15.0 1.8 25.2 NR [116]
a) Including arabinan.
NR, not reported.

treatment at 180–210°C for some minutes – is to fa- cesses and slow down growth [11]. Acetyl groups
cilitate the enzymatic hydrolysis by making the cel- cleaved off from the hemicelluloses will be found in
lulose more accessible for enzymatic degradation. the hydrolysate as acetate/acetic acid. The acetyl
Acid catalysts (e.g. sulphuric acid, SO2 or carboxylic content is normally higher in pentose-rich materi-
acids) improve the hydrolysis of hemicellulose, als than in materials with a low pentose content
whereas base catalysts (NaOH, ammonia, lime), or- (such as softwoods). In particular glucuronoxylans
ganic solvents or oxygen facilitate partial solubi- in hardwoods are acetylated to a higher degree
lization of lignin [10]. The pH value must be ad- than the galactoglucomannans typical of soft-
justed after pretreatment, and this will lead to the woods.The stress caused by acetic acid is thus more
formation of salt, which can cause osmotic stress important in xylose-rich hydrolysates. Furaldehy-
for the yeast. The osmotic stress will trigger glyc- des, primarily 2-furaldehyde from pentoses and
erol formation, affect membrane transport pro- 5-hydroxymethyl-2-furaldehyde (HMF) from hex-

Figure 1. Challenges faced by the yeast S. cerevisiae during the production of ethanol from lignocellulosic feedstocks.

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 287


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

oses, are formed by dehydration of the liberated ceed 5% by weight, and inhibition by ethanol is
sugar monomers at high temperatures and acidic therefore minor compared to that of many other
conditions, and a fraction of the furaldehydes is compounds.
also likely to be further degraded to other organic As described above, almost each process step
acids – e.g. levulinic acid and formic acid (as re- gives rise to environmental stress factors that may
viewed in [12]). The lignin fraction of the biomass impact the ethanol yield or productivity in the fer-
gives rise to different phenolics after pretreat- mentation. In addition, the presence of highly con-
ment, e.g. phenylpropanoid derivatives, such as taminated feedstock and non-sterile conditions
p-coumaric and ferulic acid, or phenolics without (see e.g. [16]) introduce an additional challenge of
the propanoid side chain [13].The hydrolysis of the competition with contaminating microorganisms
pretreated material may take place together with for the available sugars. Finally, the metabolic engi-
the fermentation, i.e. the so-called simultaneous neering of the yeast needed to convey pentose-util-
saccharification and fermentation (SSF) [14]. Fur- ising capacity adds genetic factors. The combina-
thermore, the liquid fraction after pretreatment tion of the environmental and genetic factors result
may either be separated before the hydrolysis, or in the physiological changes shown in Figure 1.
treated together with the solid fraction. If the enzy-
matic hydrolysis takes place as a separate step, a
temperature of 40–50°C is used, in agreement with 3 Metabolic engineering for efficient
the optimum temperature for the currently avail- fermentation and anaerobic growth
able cellulases. However, a lower temperature must on pentose sugars
be used in the SSF to stay within the allowable tem-
perature for the yeast. An improved yeast thermo- 3.1 Xylose utilization
tolerance would thus be beneficial from a process
point of view. Utilization of the pentose sugar xylose has been
Also the end-product of the fermentation, enabled in S. cerevisiae by expression of the het-
ethanol, is in itself inhibitory at high levels (see [15] erologous enzymes xylose reductase (XR) and xyl-
for a recent review), but in its present stage of de- itol dehydrogenase (XDH) or xylose isomerase (XI)
velopment, ethanol concentrations in lignocellu- [7, 17] (Fig. 2). Both non-targeted and targeted
lose-based processes using pentoses seldom ex- metabolic engineering approaches have then been

Figure 2. Xylose and arabinose pathways expressed in


recombinant S. cerevisiae. The utilization of cofactors and ATP
in the central carbon metabolism is depicted schematically.

288 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

followed to further improve xylose utilization, as however, unsuccessful in identifying alterations at


reviewed below. a single control point, but rather suggested an ad-
ditive effect of several small mutations in the
3.1.1 Improvement of xylose utilization by non-target genome [24]. The strain RWB218 was not charac-
approaches terized at the transcriptome level, however in-
Random mutagenesis and evolutionary engineer- creased transport affinity was detected in kinetic
ing has been successfully applied to increase xy- studies, which possibly resulted from increased
lose utilization in a number of recombinant S. cere- expression of high-affinity native transporters
visiae strains. The strains H2490-4 [18], TMB3400 [21].
[19] and C1 [20] were constructed based on the
oxido-reductive XR/XDH pathway (Fig. 2) and were 3.1.2 Improvement of xylose utilization by rational
selected during prolonged chemostat cultivation pathway design
following chemical mutagenesis (TMB3400 and C1) Rational design of the xylose utilization pathway in
(Fig. 3A). The strain RWB218 was constructed us- S. cerevisiae is based on identifying, altering and ul-
ing the XI pathway (Fig. 2) and was selected in timately moving metabolic control points. A major
chemostat followed by sequential batch cultivation controlling step in xylose utilization is the expres-
(Fig. 3A) [21].All strains had significantly improved sion level of the heterologous enzymes XR/XDH
ethanol productivity and growth rate on xylose, and XI. Specifically, the low enzyme activity and
compared to their parental strains. Enzymatic substrate affinity of XR and XI requires high intra-
analysis revealed increased activity of the heterol- cellular xylose concentration to push the rate of the
ogous enzymes XR and XDH in the strains C1 and reaction, which leads to decreased ethanol produc-
TMB3400, whereas the activity of several enzymes tivity as substrate is consumed [25]. Overexpres-
in the pentose phosphate pathway (PPP) (Fig. 2) sion of XR somewhat alleviated this problem and
was increased in the strain H2490-4 [18, 22]. These increased the rate of xylose consumption substan-
observations were further corroborated by prote- tially [26]. Similarly, very high expression level of
ome analysis, which identified higher expression of XI encoding gene from episomal plasmids was re-
XR, XDH and the non-oxidative PPP in TMB3400 quired for anaerobic growth on xylose by the strain
compared to the parental strain [23]. Transcrip- RWB202-AFX (Fig. 3A) [27].
tome analyses of the strains TMB3400 and C1 were,

A
Notation Strain Relevant genotype Additional step(s) Reference
in Fig. 3
1 C1 XYL1, XYL2, XKS1 Chemical mutagenesis [20]
2 TMB 3415 XYL1 (K270R), XYL2, XKS1, PPP – [46]
3 RWB202-AFX XI multicopy – [27]
4 TMB3421 XYL1 (N272D P275Q), XYL2, XKS1, PPP – [121]
5 TMB 3420 XYL1 (N272D P275Q), XYL2, XKS1, PPP Sequential batch selection [121]
6 RWB 217 XI multicopy, XKS1, PPP – [36]
7 RWB 218 XI multicopy, XKS1, PPP Batch and chemostat selection [21]

B
0,6

0,5 7
6
re (g/gDW×h)

0,4
5
0,3

0,2 4
1 23
0,1 Figure 3. (A) Engineered S. cerevisiae
strains capable of anaerobic growth on
0 xylose. (B) Relationship between the
0 0,05 0,1 0,15
ethanol productivity (re) and the anaerobic
Specific growth rate (h-1) growth rate on xylose.

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 289


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

Additional control points in xylose metabolism 3.2 Arabinose utilization


are the endogenous phosphorylating enzyme xylu-
lokinase (XK), and the PPP, respectively (Fig. 2). Arabinose utilization has been investigated to less
Since XK uses ATP as a cofactor, the expression lev- extent than xylose utilization due to the lower
el of this enzyme needs to be carefully optimized abundance of arabinose in lignocellulosic biomass
[28–30]. Overexpression of XK improved the con- (Table 1). Arabinose can be utilized by S. cerevisiae
version of xylulose [31]. However, without concur- via the introduction of either an isomerization
rent overexpression of XR, xylose utilization was de- pathway consisting of L-arabinose isomerase (AI),
creased [28, 30]. The non-oxidative PPP follows af- L-ribulokinase (RK) and L-ribulose-5-P 4-epimer-
ter XK in the conversion of xylose to glycolytic in- ase (RE) or a reduction/oxidation-based pathway
termediates (Fig. 2). Compared to glycolysis, the PPP consisting of XR, L-arabinitol 4-dehydrogenase
is kinetically constrained since the PPP reactions (LAD), L-xylulose reductase (LXR) and XDH (Fig.
are less thermodynamically favourable and operate 2) [7]. Although arabinose utilization is consider-
close to the equilibrium [32]. Fermentation of xylu- ably less efficient than xylose, it shares several con-
lose to ethanol in native S. cerevisiae is thus orders trol points with xylose utilization since both het-
of magnitude slower than that of glucose [31, 33]. erologous pathways end with the formation of xy-
When all genes of the non-oxidative PPP were over- lulose (Fig. 2). Arabinose utilization thus benefits
expressed, the xylulose utilization increased, but xy- from high expression of the genes from the initial
lose conversion was not improved unless XR and heterogonous arabinose-catabolizing enzymes
XDH activities were raised [34, 35]. A considerably [47], the endogenous XK, the non-oxidative PPP as
higher ethanol productivity was also observed when well as high affinity native transporters [48].
high XI level was combined with high expression of In practice, the fermentation of the arabinose
XK and non-oxidative PPP encoding genes in fraction in lignocellulose hydrolysates cannot be
strains RWB 217 and RWB 218 (Fig. 3A) [27, 36]. disconnected from the fermentation of xylose,
In S. cerevisiae, xylose is transported via the which raises a set of new challenges. For instance,
hexose transporters, however with a much lower XR displays a similar affinity for xylose and arabi-
affinity, which limits its utilization in the presence nose [49, 50]. Thus, in S. cerevisiae strains combin-
of glucose [37]. On xylose medium however, it was ing the oxidation/reduction pathway for xylose and
initially shown that xylose transport did not control the isomerization pathway for arabinose-utiliza-
the flux of the pathway in a recombinant S. cere- tion, almost all consumed arabinose was converted
visiae strain [25]. However, when XR was over- to arabitol (arabinitol) by XR (Fig. 2) [51–53]. Ara-
expressed the control coefficient of transport in- bitol formation can be avoided by using the XI-
creased [38], and in a strain where the PPP was based pathway for xylose utilization, which does
overexpressed, increased transport improved xy- not include XR. Nevertheless this approach proved
lose utilization at least when the substrate concen- difficult since an elaborate selection procedure was
tration was below 4 g/L [39]. required to enable arabinose utilization without
For S. cerevisiae strains employing the oxido-re- lowering the capacity of the same strain for xylose
ductive xylose pathway, an additional control point utilization [54, 55]. Arabitol formation can also be
is the imbalanced co-factor preference of the two reduced by employing the oxidation/reduction
recombinant enzymes XR and XDH. Since XR pathways for both xylose- and arabinose-utiliza-
prefers NADPH and XDH is strictly NAD+ depend- tion (Fig. 2), so that arabitol can be further metab-
ent, the cell creates an excess of NAPD+ and a de- olized to xylitol and ethanol. An oxidation/reduc-
ficiency in NADH which leads to the production of tion pathway for arabinose metabolism has thus
xylitol as a byproduct (Fig. 2) and lower ethanol been described [56, 57] and its performance was re-
yield [25, 40]. The cofactor specificity of XR and/or cently improved using a suitable combination of
XDH has thus been modified by protein engineer- enzymes [58].
ing for balanced cofactor usage of the two enzymes
[41–44]. In this way the ethanol productivity, and
consequently the anaerobic growth rate, was dras- 4 Pentose fermentation in hydrolysates
tically improved for strains TMB 3415, TMB 3420
and TMB 3421 utilizing mutant XRs (Fig. 3) [45, 46]. In reported studies in which pentose-fermenting
strains of S. cerevisiae have been evaluated in un-
detoxified pentose-rich lignocellulosic hydrolysa-
tes (Table 2), the maximum ethanol concentrations
obtained did not exceed 45 g/L and in many cases,
the xylose conversion was not complete even after

290 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

Table 2. Some studies in which pentose fermenting S. cerevisiae has been used for the conversion of lignocellulosic hydrolysates

Strain Raw material Pretreatment and Ethanol concentration and Reference


fermentation conditions xylose conversion
424A(LNH-ST) Sweet sorghum AFEX pretreatment (2:1 ammonia to Highest ethanol concentration [117]
(XR/XDH/XK) bagasse biomass loading, 140°C, 30 min) reached 42.3 g/L, with a xylose
followed by enzymatic hydrolysis. conversion of 55.6% (after about 75 h).
Approximate liquid concentrations: Very little xylitol formation.
glucose 65 g/L, xylose 33 g/L.
Initial yeast concentration 0.3 g/L

424A(LNH-ST) Poplar Comparison of many different Highest ethanol concentration reached [84]
(XR/XDH/XK) pretreatments including AFEX, after 50 h: 40 g/L (lime pretreatment)
oxidative alkali, SO2 catalysed steam with about 90% conversion xylose;
pretreatment, followed by enzymatic SO2 steam explosion yielded 26 g/L
hydrolysis. Initial liquid concentra- ethanol, about 90% xylose conversion.
tions (AFEX): glucose 62.3 g/L,
xylose 16.2 g/L, acetic acid 3.5 g/L.
Initial liquid concentrations (steam
pretreatment SO2) glucose 33.2 g/L,
xylose 25.8 g/L, acetic acid 6.2 g/L.
Initial yeast concentration about 5 g/L

TMB3400 Corn stover SSF Ethanol concentration of 36.8 g/L [78]


(XR/XDH/XK + Steam pretreatment at 200°C, 5 min, reached at 11% WIS using fed-batch
mutagenesis and SO2 catalysed. Liquid composition: addition. Residual xylose concentration
selection) glucose 7.2, xylose 36, acetic acid, in liquid about 10 g/L after 96 h
2.2 g/L, HMF 0.2 g/L g, furfural 1.5 g/L.
Substantial amounts of xylan remaining
in the solid material (6.6–9.2%), which
was degraded during SSF.
Initial yeast concentration: 5 g/L.

TMB3400 Wheat straw SSF A final ethanol concentration of 38 g/L [118]


(XR/XDH/XK) Steam pretreatment, dilute H2SO4 and a xylose conversion about 40%
catalyst, 210°C, 2.5 min. Liquid in fed-batch up to 9% WIS.
composition after pretreatment: Corresponding for fed-batch at 7%:
glucose 10.9 g/L, xylose 51.2 g/L, ethanol concentration of 34 g/L and
acetic acid 5 g/L, HMF 0.6 g/L, a xylose conversion of 74%.
furfural, 1.7 g/L. Xylan content in
fibre about 3.5%.
Fed-batch fermentation at a total solids
loading of 9%WIS. Initial yeast
concentration 4 g/L

TMB3400 Sugar cane SSF


(XR/XDH/XK) bagasse Steam pretreatment at 190°C, 5 min, Ethanol concentration 26 g/L. [119]
SO2 catalysed. Liquid composition Conversion of xylose about 40% at
after pretreatment: glucose 4.6 g/L, 7.5%WIS (Ethanol concentration
xylose 39.5 g/L, arabinose 2.7 g/L, 20 g/L, and xylose conversion 88%
furfural 0.9 g/L, acetic acid 4.2 g/L. at 5%WIS).
Yeast concentration 4 g/L.

TMB3400 Sugar cane Steam pretreatment at 190°C, 5 min, Only hemicellulose part fermented. [109]
(XR/XDH/XK) bagasse SO2 catalysed. Liquid concentrations: Highest xylose conversion for liquid
15 g/L xylose, 4 g/L glucose. from pretreatment 65% after 20 h.
Initial yeast concentration 5 g/L

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 291


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

Table 2. Continued

Strain Raw material Pretreatment and Ethanol concentration and Reference


fermentation conditions xylose conversion
MA-R4 Eucalyptus Primarily mechanical pretreatment Ethanol concentration 37.2 g/L. [120]
(XR/XDH/XK (ball milling) followed by enzymatic Conversion of 90% of the xylose
introduced into hydrolysis at high cellulase loadings. after 48 h
flocculating industrial Liquid composition: 61.1 g/L glucose,
strain IR-2) 13.0 g/L xylose, 1.2 g/L galactose,
1.0 g/L mannose, 0.6 g/L arabinose,
10 g/L yeast extract added (rich medium).
Low inhibitor concentration
Relatively high yeast density –
final value 14.3 g dw/L

RWB2182 Wheat straw Steam explosion pretreated followed Ethanol concentration: 38.1 g/L. [17]
(XI) by enzymatic hydrolysis, batch About 90% of xylose converted
Strain improved fermentation. Liquid composition: after 55 h. Almost stoichiometric
by evolution 47.8 g/L glucose, 20.7 g/L xylose, yield based on consumed sugars
3 g/L acetic acid.
Initial yeast concentration: 1.5 g/L.

RWB218 Corn stover Steam explosion pretreated Ethanol concentration about 18 g/L. [17]
(XI) (190°C, 5 min), followed by enzymatic About 90% of xylose converted
hydrolysis. Liquid composition: after 48 h.
40 g/L glucose, 9 g/L xylose,
4 g/L acetic acid. Fed-batch + batch
fermentation.
SSF, simultaneous saccharification and fermentation; WIS, water insoluble solids; AFEX, ammonia fibre expansion.

a long period of time. The specific xylose consump- 4.1.1 Metabolically engineered microorganisms
tion rates obtained in hydrolysates were also clear- for aldehyde reduction
ly lower than those on synthetic media. A fundamental observation is that as long as cer-
tain aldehydes – e.g. furfural, HMF, vanillin, vera-
4.1 Effect of aldehyde inhibitors on xylose traldehyde – are present in the medium, yeasts are
fermentation not able to grow and show reduced metabolic ac-
tivity [61, 69–71]. However, at permissive levels, the
The toxicity of lignocellulosic hydrolysates has yeast can usually slowly convert the aldehydes to
been correlated with the presence and concentra- less toxic compounds and growth and metabolic ac-
tion of various aldehydes, such as furfural [59–61]. tivity are restored [61, 72–74]. S. cerevisiae, for in-
Among 60 identified phenolic compounds in ligno- stance, can reduce furfural and HMF to their corre-
cellulosic hydrolysates, the phenylaldehydes were sponding alcohols; 2-furan methanol (FM or fur-
found to be more toxic than the phenolic acids and furyl alcohol) and furan 2,5-dimethanol (FDM or
alcohols [62, 63] underlining the importance of the HMF alcohol). A few alcohol dehydrogenases, i.e.
aldehydes as inhibitors. Adh6, Adh7 and a mutated Adh1 enzyme, as well as
Aldehydes can interact with cellular structures, the P. stipitis XR (Ps-XR) and the YKL071w gene
generate reactive oxygen species and inhibit en- product have been identified as catalyzing furalde-
zymes in the central carbon metabolism [64–67]. hyde reduction, and their positive effect upon over-
This results in reduction of fermentation rate, ces- expression of the corresponding gene has been
sation of growth, extension of lag-phase of yeast, confirmed [68, 75]. Overexpression of the tran-
and reduction of ethanol productivity and/or yield scription factor encoding gene YAP1, known to be
[12, 63]. Understanding and improving the mecha- involved in oxidative stress response, was also
nisms by which microorganisms respond to alde- shown to give an increased tolerance to hy-
hydes is clearly important for the development of drolysate and specific model inhibitors – and an in-
strains with increased tolerance towards lignocel- creased reduction capacity of furaldehydes [76].
lulosic hydrolysates [12, 68].

292 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

More recently an isolate of the gram positive oxidation of xylitol so that more carbon became
bacteria Cupriavidus basilensis (HMF14) could aer- available for biomass production.
obically completely metabolize HMF and furfural
[77], which may open new possibilities for the mi- 4.2 Pentose fermentation in the presence
crobial detoxification of hydrolysates. of weak acids

4.1.2 Interactions between xylose and furaldehyde Ethanol fermentation from lignocellulosic feed-
pathways stocks is also challenged by the presence of weak
S. cerevisiae keeps its redox balance tightly regu- acids in the medium, at concentrations that can be
lated, essentially by regulating the activity of gly- inhibitory to S. cerevisiae metabolism. Acetic acid
colytic and fermentative enzymes. The anaerobic originates from the deacetylation of the hemicellu-
glucose catabolism (when not considering draining lose fraction, whereas formic and levulinic acids
the pathway of anabolic precursors), is balanced in are formed during the breakdown of HMF and fur-
the glycolysis and fermentation pathways with re- fural [83]. The innate levels of acetic acid are high-
spect to generation/consumption of the co-factor ly dependent on the raw material. Reported acetyl
pairs NADH/NAD+ (and NADPH/NADP+). Howev- contents in some pentose rich materials are in the
er, during the fermentation of lignocellulosic hy- range 1.7–4.2 g/100 g DM (Table 1). The concentra-
drolysates there are additional compounds which tions of weak acids obtained in the fermentation
will be reduced and/or oxidized: (i) the conversion broth will vary even more widely depending on
of xylose by the XR/XDH pathway from P. stipitis is both the method of pretreatment as well as the
based on the use of NADPH and NAD+ (Fig. 2); solids loading used in the pretreatment. It can be
(ii) the detoxification of aldehydes is based on re- estimated that the concentration of acetic acid may
duction reactions using NADH and NADPH. Con- easily be above 10 g/L after a typical steam pre-
sequently, both xylose metabolism and detoxifica- treatment of poplar [84].
tion reactions will affect the redox (or more cor- Weak acids are also produced during the fer-
rectly the co-factor) balance of the cell. mentation stage, since acetic acid is a minor by-
Ethanolic fermentation with S. cerevisiae strains product of yeast fermentation. However, acetic and
overexpressing the XR/XDH pathway in defined lactic acids are mainly generated from the contam-
mineral medium containing xylose often gives sub- ination by lactic acid bacteria (LAB) and other bac-
stantial yields of xylitol as a result of the co-factor teria (see also Section 4.3). For example, acetic and
imbalance between the NADPH-preferring XR and lactic acid levels above 12 and 16 g/L, respectively,
the NAD+-dependent XDH (cf. Fig. 2). However, have been reported in a pilot-scale study using
fermentation of xylose in lignocellulosic hydroly- corn fibre as substrate (LAB) [16, 85].
sates by engineered S. cerevisiae strains showed
considerably lower xylitol yields than during the 4.2.1 Effect of weak acids on yeast metabolism
fermentation of xylose in defined mineral medium Both acetic and lactic acid are known to increase
[78, 79]. Most likely this is due to aldehydes and the lag phase and reduce the specific growth rate of
other compounds in the lignocellulosic hydrolysate S. cerevisiae, thereby affecting ethanol productivity
which act as redox sinks [80, 81]. For instance, an (see e.g. [86–89]. Acetic acid is more inhibitory than
NADH-dependent reduction of furfural regener- lactic acid, with a reported minimal inhibitory con-
ates NAD+ necessary for xylitol oxidation by XDH. centration for growth of 100 mM (6 g/L), as com-
Thereby glycerol and/or xylitol yields may de- pared to 278 mM (25 g/L) for lactic acid in non-
crease [81]. buffered defined medium [86].
The simultaneous conversion of aldehyde com- The effects of weak acids are strongly pH de-
pounds and xylose was investigated in an XR/XDH pendent. At pH value below the pKa-value of the
strain overexpressing either the NADH-dependent acid, the undissociated form of weak acids predom-
ADH1-S110P-Y295C encoding gene or the NADPH- inates. It is this form which enters the cell and once
dependent Adh6 encoding gene [82]. NADPH us- inside it dissociates (due to the higher intracellular
age during HMF reduction slightly reduced xylitol pH), thereby releasing protons. In order to main-
formation, whereas the usage of NADH in HMF re- tain a proper intracellular pH, protons must be
duction considerably decreased both xylitol and transported out of the cells with the help of the
glycerol production and increased biomass yield plasma membrane ATPase and at the expense of
[82].A metabolic flux analysis indicated that the re- ATP (Fig. 2). Whereas low levels of acids activate
duction of HMF replaced the regeneration of NAD+ the glycolytic rate by stimulating ATP production
in the glycerol pathway and provided NAD+ for the [89], higher levels become inhibitory due to the
acidification of the cytosol after depletion of the

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 293


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

available ATP. Also, as less ATP becomes available of lignocellulosic hydrolysates are preferably run
for biomass formation, ethanol productivity de- at a low pH to limit bacterial contaminations (Sec-
creases. In addition, high levels of weak acids may tion 4.3).
result in the intracellular accumulation of anions Recently, S. cerevisiae was metabolically engi-
(Fig. 2) [90], which has been shown to be inhibito- neered allowing acetate to replace dihydroxyace-
ry to several glycolytic enzymes [91]. ATP is also tone phosphate (DHAP) as a redox sink during
needed for the active efflux of the accumulating anaerobic growth in a glycerol mutant strain. This
dissociated acids, which has been suggested to oc- was achieved by introducing a bacterial NAD+-de-
cur via Pdr12p [92]. pendent acetylating acetaldehyde dehydrogenase,
by which acetyl-CoA may be reduced to acetalde-
4.2.2 Xylose fermentation in the presence hyde (and CoA). The engineered strain was able to
of weak acids grow anaerobically and acetate was reduced to
The effect of acetic acid on xylose fermentation has ethanol in the process [97]. Such approach may pave
been investigated in strains carrying either the XR- the way to other strategies enabling the conversion
XDH or the XI pathway [84, 93–96]. A high level of of the inhibitory acetate during fermentation.
acetic acid with a low pH caused dramatic reduc-
tion of xylose consumption rate for an XR-XDH 4.3 The contamination issue
strain, whereas a change in pH alone showed little
effects on the xylose consumption [84]. Growth and Large scale plants for the production of ethanol
ethanol production from xylose were also impaired from lignocellulosic biomass are not yet available.
in a recombinant XI-expressing strain at low pH However, one can anticipate that microbial compe-
(3.5) and in the presence of 3 g/L acetic acid [93], tition will become an important problem, due to the
whereas the glucose consumption rate was only presence of a large amount of pentose sugars that
slightly affected under similar conditions. are still not, or only poorly, utilized by S. cerevisiae.
The lower rate of xylose consumption, typically
three to sixfold lower than for glucose in engi- 4.3.1 Microbial contaminants in lignocellulosic
neered S. cerevisiae [93, 94] affects the maximum hydrolysates
specific ATP production rate, which explains why Systematic and acute bacterial infections are re-
weak acids have a more dramatic effect on yeast currently observed in large scale facilities during
growth on xylose than on glucose [93–95]. Indeed the ethanol production from hexose sugars, with
xylose fermentation could be restored in a recom- sugar cane in Brazil [98] and corn in United States
binant XI-strain at pH 3.5 and in the presence of of America [99]. However, little is known on the
3 g/L acetic acid by applying a continuous low feed- contaminant microbiota in lignocellulosic ethanol
ing of glucose to the medium [93]. pilot plants. The most studied cases of contamina-
Acetic acid decreases ethanol productivity but tion are in SSL, which is the waste stream of cellu-
its effect on ethanol yield is less clear. In some re- lose pulp that has been obtained by treating the
ports, the ethanol yield was shown to increase as wood mechanically and chemically. The SSL has a
the metabolism was redirected from anabolic (bio- high sugar content which can be used as feedstock
mass) to dissimilatory (ethanol) pathway in the for various microorganisms.
presence of acetic acid [93, 94]. In contrast, a lower Both yeast and bacterial contaminants have
ethanol yield from glucose and xylose was also re- been isolated and identified in SSL plants, includ-
ported in the presence of a more complex medium ing the xylose-utilising yeast Pichia membranaefa-
such as spent sulphite liquor (SSL) [96]. However, ciens, the acetic acid bacteria Acetobacter tropicalis
it is unclear whether the reported yield relates to and A. syzygii, the homofermentative bacteria Lac-
the consumed or the initial sugar levels, since the tobacillus plantarum and L. pentosus as well as the
tested XR and XDH-engineered strains displayed heterofermentative L. buchneri [100] (C. Larsson
low xylitol and unchanged glycerol yields in acetic and E. Albers, personal communication; V. Sànchez
acid rich-hydrolysate media [84, 96]. et al., unpublished). Lactobacillus contaminants
have also been identified in a pilot-plant where
4.2.3 Engineering approaches to overcome the weak corn fibres were used to produce ethanol [16]. Dur-
acid inhibition ing episodes of contamination in sugar cane
One straightforward strategy to overcome or limit ethanol plants, non-S. cerevisiae strains were also
the inhibition by acids is to run fermentations at a found as main contaminants [101].
pH value sufficiently high to keep the undissociat-
ed acids at low levels. In practice however, it may be
difficult to implement this strategy as fermentation

294 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

4.3.2 Effects of microbial contamination on yeast 5 Concluding remarks


fermentation
LAB compete with yeast for sugar utilization. Ho- Both targeted and non-targeted approaches have
mofermentative bacteria are able to consume hex- been applied to generate efficient pentose-fer-
ose and pentose sugars to produce lactic acid, menting S. cerevisiae strains. The strength of the
whereas heterofermentative bacteria produce non-targeted approach is that it requires less a pri-
acetic acid in addition to lactic acid [102]. Finally ori knowledge of the pathway and is generally
acetic acid bacteria produce acetic acid from straight forward and efficient. The targeted ap-
ethanol [103]. An immediate consequence of the proach on the other hand may require tedious
contamination is that less sugar and essential identification of specific traits that enhance metab-
growth factors become available for S. cerevisiae olism, however once identified these traits can be
and the ethanol yield decreases (see e.g. [104]). In directly transferred between strains. Both ap-
addition, the production of considerable amount of proaches have been successful in the generation of
acids decreases the pH which inhibits yeast fer- S. cerevisiae strain that can grow anaerobically on
mentation by reducing biomass formation and xylose, a trait which is shown to be linearly corre-
ethanol yield. lated to ethanol productivity (Fig. 3B). Still the xy-
lose consumption rate remains three to sixfolds
4.3.3 Strategies to control a contamination lower than the glucose consumption rate. Generat-
In the sugar cane ethanol industry, acid wash treat- ing tolerant S. cerevisiae with even higher xylose
ment is used as a common strategy for reducing consumption rate would of course benefit ethanol
acute episodes of contamination [105]. Pulses of productivity, but it would also enable the cells to
sulphuric acid are made, leading to a temporary pH better cope with the ATP-demanding acid stress
drop on the system. The bacterial population is and outcompete any pentose-utilizing LAB-con-
then reduced due to its high sensitivity for low pH taminant in the process.
[106]. A central issue in converting pentoses from hy-
Various pasteurization procedures and the ad- drolysates is also the presence of aldehydes – not
dition of biocides, such as penicillin, virginiamycin least the furaldehydes. Recent identification of en-
and nisin, have also been tested in a series of batch zymes with aldehyde reduction activity has enabled
and continuous fermentation runs in a corn fibre targeted approaches for improving the in situ re-
based pilot plant. All these methods reduce tem- duction capacity of S. cerevisiae strains. Similarly ge-
porarily the level of contamination but none has a netic engineering approaches are emerging to cope
permanent effect [16]. Also, the usage of biocides to with the problem of weak acid inhibition. In both
control contamination is not economically feasible cases, however, it is important to understand how
and not advisable from an environmental point of these modifications impact pentose metabolism.
view.
Recent studies demonstrated that the addition
of NaCl and ethanol on softwood hydrolysate could The authors wish to thank the Swedish Energy
inhibit the growth of LAB whereas baker’s yeast Agency (Energimyndigheten) for financial support.
was still able to grow (C. Larsson and E.Albers, per-
sonal communication). The authors have declared no conflict of interest.
The use of hops extract to control bacterial con-
tamination in lignocellulosic fermentation plants is
also discussed. Hop that gives a bitter taste and acts 6 References
as antiseptic agent, is well-known in beer produc- [1] Ingram, L. O., Conway, T., Clark, D. P., Sewell, G. W., Preston,
tion. Hop cones have a high content of α-acids, for J. F., Genetic engineering of ethanol production in Es-
which Gram-positive bacteria are sensitive to [106, cherichia coli. Appl. Environ. Microbiol. 1987, 53, 2420–2425.
107]. [2] Zhang, M., Eddy, C., Deanda, K., Finkestein, M., Picataggio,
In the near future, the development of S. cere- S., Metabolic engineering of a pentose metabolism pathway
in ethanologenic Zymomonas mobilis. Science 1995, 267,
visiae strains with xylose consumption rates that
240–243.
are sufficiently high to compete with LAB for the [3] Lacis, L. S., Lawford, H. G., Thermoanaerobacter ethanolicus
utilization of all sugars present in the lignocellu- growth and product yield from elevated levels of xylose or
losic hydrolysates, may provide another solution to glucose in continuous cultures. Appl. Environ. Microbiol.
the bacterial contamination issue. 1991, 57, 579–585.
[4] Shaw, A. J., Podkaminer, K. K., Desai, S. G., Bardsley, J. S., et
al., Metabolic engineering of a thermophilic bacterium to

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 295


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

produce ethanol at high yield. Proc. Natl. Acad. Sci. U. S. A. [21] Kuyper, M., Toirkens, M. J., Diderich, J. A., Winkler, A. A., et
2008, 105, 13769–13774. al., Evolutionary engineering of mixed-sugar utilization by
[5] Skory, C. D., Freer, S. N., Bothast, R. J., Screening for ethanol- a xylose-fermenting Saccharomyces cerevisiae strain.
producing filamentous fungi. Biotechnol. Lett. 1997, 19, FEMS Yeast Res. 2005, 5, 925–934.
203–206. [22] Wahlbom, C. F., Otero, R. R. C., van Zyl, W. H., Hahn-
[6] Toivola, A., Yarrow, D., Van den Bosch, E., Van Dijken, J. P., Hägerdal, B., Jönsson, L. J., Molecular analysis of a Saccha-
Scheffers, W. A., Alcoholic fermentation of D-xylose by romyces cerevisiae mutant with improved ability to utilize
yeasts. Appl. Environ. Microbiol. 1984, 47, 1221–1223. xylose shows enhanced expression of proteins involved in
[7] Hahn-Hägerdal, B., Karhumaa, K., Jeppsson, M., Gorwa- transport, initial xylose metabolism, and the pentose phos-
Grauslund, M. F., Metabolic engineering for pentose uti- phate pathway. Appl. Environ. Microbiol. 2003, 69, 740–746.
lization in Saccharomyces cerevisiae. Adv. Biochem. Eng./ [23] Karhumaa, K., Pahlman, A. K., Hahn-Hägerdal, B., Levan-
Biotechnol. 2007, 108, 147–177. der, F., Gorwa-Grauslund, M. F., Proteome analysis of the
[8] Van Vleet, J. H., Jeffries, T. W., Yeast metabolic engineering xylose-fermenting mutant yeast strain TMB 3400. Yeast
for hemicellulosic ethanol production. Curr. Opin. Biotech- 2009, 26, 371–382.
nol. 2009, 20, 300–306. [24] Bengtsson, O., Jeppsson, M., Sonderegger, M., Parachin, N.
[9] Hahn-Hägerdal, B., Galbe, M., Gorwa-Grauslund, M. F., S., et al., Identification of common traits in improved xy-
Lidén, G., Zacchi, G., Bio-ethanol – the fuel of tomorrow lose-growing Saccharomyces cerevisiae for inverse meta-
from the residues of today. Trends Biotechnol. 2006, 24, bolic engineering. Yeast 2008, 25, 835–847.
549–556. [25] Kötter, P., Ciriacy, M., Xylose fermentation by Saccha-
[10] Alvira, P., Tomás-Pejó, E., Ballesteros, M., Negro, M. J., Pre- romyces cerevisiae. Appl. Microbiol. Biotechnol. 1993, 38,
treatment technologies for an efficient bioethanol produc- 776–783.
tion process based on enzymatic hydrolysis: A review. [26] Jeppsson, M., Träff, K., Johansson, B., Hahn-Hägerdal, B.,
Bioresour. Technol. 2010, 101, 4851–4861. Gorwa-Grauslund, M. F., Effect of enhanced xylose reduc-
[11] Modig, T., Granath, K., Adler, L., Lidén, G., Anaerobic glyc- tase activity on xylose consumption and product distribu-
erol production by Saccharomyces cerevisiae strains under tion in xylose-fermenting recombinant Saccharomyces
hyperosmotic stress. Appl. Microbiol. Biotechnol. 2007, 75, cerevisiae. FEMS Yeast Res. 2003, 3, 167–175.
289–296. [27] Kuyper, M., Winkler, A. A., van Dijken, J. P., Pronk, J. T., Min-
[12] Almeida, J. R. M., Modig, T., Petersson, A., Hahn-Hägerdal, imal metabolic engineering of Saccharomyces cerevisiae for
B. et al., Increased tolerance and conversion of inhibitors in efficient anaerobic xylose fermentation: A proof of princi-
lignocellulosic hydrolysates by Saccharomyces cerevisiae. ple. FEMS Yeast Res. 2004, 4, 655–664.
J. Chem. Technol. Biotechnol. 2007, 82, 340–349. [28] Jin, Y. S., Ni, H. Y., Laplaza, J. M., Jeffries, T. W., Optimal
[13] Martín, C., Klinke, H. B., Marcet, M., García, L., et al., Study growth and ethanol production from xylose by recombi-
of the phenolic compounds formed during pretreatment of nant Saccharomyces cerevisiae require moderate D-xylu-
sugarcane bagasse by wet oxidation and steam explosion. lokinase activity. Appl. Environ. Microbiol. 2003, 69, 495–
Holzforschung 2007, 61, 483–487. 503.
[14] Olofsson, K., Bertilsson, M., Lidén, G., A short review on [29] Teusink, B.,Walsh, M. C., van Dam, K.,Westerhoff, H.V.,The
SSF – an interesting process option for ethanol production danger of metabolic pathways with turbo design. Trends
from lignocellulosic feedstocks. Biotechnol.Biofuels 2008, 1, Biochem. Sci. 1998, 23, 162–169.
7. [30] Johansson, B., Christensson, C., Hobley, T., Hahn-Hägerdal,
[15] Ding, J., Huang, X., Zhang, L., Zhao, N., et al., Tolerance and B., Xylulokinase overexpression in two strains of Saccha-
stress response to ethanol in the yeast Saccharomyces romyces cerevisiae also expressing xylose reductase and
cerevisiae. Appl. Microbiol. Biotechnol. 2009, 85, 253–263. xylitol dehydrogenase and its effect on fermentation of xy-
[16] Schell, D. J., Dowe, N., Ibsen, K. N., Riley, C. J., et al., Conta- lose and lignocellulosic hydrolysate. Appl. Environ. Micro-
minant occurrence, identification and control in a pilot- biol. 2001, 67, 4249–4255.
scale corn fiber to ethanol conversion process. Bioresour. [31] Eliasson, A., Boles, E., Johansson, B., Österberg, M., et al.,
Technol. 2007, 98, 2942–2948. Xylulose fermentation by mutant and wild-type strains of
[17] van Maris, A. J., Winkler, A. A., Kuyper, M., de Laat, W. T., et Zygosaccharomyces and Saccharomyces cerevisiae. Appl.
al., Development of efficient xylose fermentation in Sac- Microbiol. Biotechnol. 2000, 53, 376–382.
charomyces cerevisiae: Xylose isomerase as a key compo- [32] Vaseghi, S., Baumeister, A., Rizzi, M., Reuss, M., In vivo dy-
nent. Adv. Biochem. Eng./Biotechnol. 2007, 108, 179–204. namics of the pentose phosphate pathway in Saccha-
[18] Pitkänen, J. P., Rintala, E., Aristidou, A., Ruohonen, L., Pent- romyces cerevisiae. Metab. Eng. 1999, 1, 128–140.
tilä, M., Xylose chemostat isolates of Saccharomyces cere- [33] Wang, P. Y., Schneider, H., Growth of yeasts on D-xylulose.
visiae show altered metabolite and enzyme levels com- Can. J. Microbiol. 1980, 26, 1165–1168.
pared with xylose, glucose, and ethanol metabolism of the [34] Johansson, B., Hahn-Hägerdal, B., The non-oxidative pen-
original strain. Appl. Microbiol. Biotechnol. 2005, 67, 827– tose phosphate pathway controls the fermentation rate of
837. xylulose but not of xylose in Saccharomyces cerevisiae
[19] Wahlbom, C. F., van Zyl, W. H., Jönsson, L. J., Hahn-Häger- TMB3001. FEMS Yeast Res. 2002, 2, 277–282.
dal, B., Otero, R. R., Generation of the improved recombi- [35] Karhumaa, K., Hahn-Hägerdal, B., Gorwa-Grauslund, M. F.,
nant xylose-utilizing Saccharomyces cerevisiae TMB 3400 Investigation of limiting metabolic steps in the utilization
by random mutagenesis and physiological comparison of xylose by recombinant Saccharomyces cerevisiae using
with Pichia stipitis CBS 6054. FEMS Yeast Res. 2003, 3, metabolic engineering. Yeast 2005, 22, 359–368.
319–326. [36] Kuyper, M., Hartog, M. M., Toirkens, M. J., Almering, M. J., et
[20] Sonderegger, M., Sauer, U., Evolutionary engineering of al., Metabolic engineering of a xylose-isomerase-express-
Saccharomyces cerevisiae for anaerobic growth on xylose. ing Saccharomyces cerevisiae strain for rapid anaerobic xy-
Appl. Environ. Microbiol. 2003, 69, 1990–1998. lose fermentation. FEMS Yeast Res. 2005, 5, 399–409.

296 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

[37] Leandro, M. J., Fonseca, C., Goncalves, P., Hexose and pen-
tose transport in ascomycetous yeasts: An overview. FEMS
Yeast Res. 2009, 9, 511–525. Gunnar Lidén received his Ph.D. in
[38] Gárdónyi, M., Jeppsson, M., Lidén, G., Gorwa-Grauslund, Chemical Reaction Engineering from
M. F., Hahn-Hägerdal, B., Control of xylose consumption by Chalmers University, Göteborg, Swe-
xylose transport in recombinant Saccharomyces cerevisiae. den in 1993, where he remained until
Biotechnol. Bioeng. 2003, 82, 818–824.
1999. At that time he moved to Lund
[39] Runquist, D., Fonseca, C., Rådström, P., Spencer-Martins, I.,
Hahn-Hägerdal, B., Expression of the Gxf1 transporter university to take on his current posi-
from Candida intermedia improves fermentation perform- tion as professor in chemical engineer-
ance in recombinant xylose-utilizing Saccharomyces cere- ing. His main research focus concerns
visiae. Appl. Microbiol. Biotechnol. 2009, 82, 123–130. biotechnology for the production of
[40] Jeppsson, M., Johansson, B., Hahn-Hägerdal, B., Gorwa- biofuels.In particular, he worked on im-
Grauslund, M. F., Reduced oxidative pentose phosphate
proving fermentation technology for efficient lignocellulose conver-
pathway flux in recombinant xylose-utilizing Saccha-
romyces cerevisiae strains improves the ethanol yield from sion. His research group is currently involved in two large European
xylose. Appl. Environ. Microbiol. 2002, 68, 1604–1609. research projects within FP7 (NEMO and BIOLYFE) concerning these
[41] Matsushika, A., Watanabe, S., Kodaki, T., Makino, K., et al., issues.
Expression of protein engineered NADP+-dependent xyli-
tol dehydrogenase increases ethanol production from xy-
lose in recombinant Saccharomyces cerevisiae. Appl. Micro-
biol. Biotechnol. 2008, 81, 243–255.
[42] Watanabe, S., Kodaki, T., Makino, K., Complete reversal of Marie Gorwa-Grauslund is professor of
coenzyme specificity of xylitol dehydrogenase and increase Applied Microbiology, Lund University
of thermostability by the introduction of structural zinc.
since 2009. After a Ph.D. in Molecular
J. Biol. Chem. 2005, 280, 10340–10349.
[43] Jeppsson, M., Bengtsson, O., Franke, K., Lee, H., et al., The and Cellular Genetics in 1996 at INSA
expression of a Pichia stipitis xylose reductase mutant with Toulouse, France and a post-doctoral
higher K(M) for NADPH increases ethanol production fellowship at Lesaffre (France) on the
from xylose in recombinant Saccharomyces cerevisiae. development of stress-tolerant yeast,
Biotechnol. Bioeng. 2006, 93, 665–673.
she joined the Division of Applied Mi-
[44] Watanabe, S., Abu Saleh, A., Pack, S. P., Annaluru, N., et al.,
crobiology in 1999. Her current re-
Ethanol production from xylose by recombinant Saccha-
romyces cerevisiae expressing protein-engineered NADH- search projects involve the identifica-
preferring xylose reductase from Pichia stipitis. Microbiol- tion, development and engineering of biocatalysts for the production
ogy 2007, 153, 3044–3054. of fuels and bulk chemicals and for stereo-selective whole cell bio-
[45] Bengtsson, O., Hahn-Hägerdal, B., Gorwa-Grauslund, M. F., transformations.
Xylose reductase from Pichia stipitis with altered coen-
zyme preference improves ethanolic xylose fermentation
by recombinant Saccharomyces cerevisiae. Biotechnol. Bio-
fuels 2009, 2, 9. strain using evolutionary engineering. Biotechnol. Biofuels
[46] Runquist, D., Hahn-Hägerdal, B., Bettiga, M., Increased ex- 2010, 3, 13.
pression of the oxidative pentose phosphate pathway and [53] Bettiga, M., Hahn-Hägerdal, B., Gorwa-Grauslund, M. F.,
gluconeogenesis in anaerobically growing xylose-utilizing Comparing the xylose reductase/xylitol dehydrogenase
Saccharomyces cerevisiae. Microb. Cell Fact. 2009, 8, 49. and xylose isomerase pathways in arabinose and xylose
[47] Wiedemann, B., Boles, E., Codon-optimized bacterial genes fermenting Saccharomyces cerevisiae strains. Biotechnol.
improve L-arabinose fermentation in recombinant Saccha- Biofuels 2008, 1, 16.
romyces cerevisiae. Appl. Environ. Microbiol. 2008, 74, [54] Wisselink, H. W., Toirkens, M. J., del Rosario Franco Berriel,
2043–2050. M.,Winkler,A.A., et al., Engineering of Saccharomyces cere-
[48] Becker, J., Boles, E., A modified Saccharomyces cerevisiae visiae for efficient anaerobic alcoholic fermentation of
strain that consumes L-arabinose and produces ethanol. L-arabinose. Appl. Environ. Microbiol. 2007, 73, 4881–4891.
Appl. Environ. Microbiol. 2003, 69, 4144–4150. [55] Wisselink, H. W., Toirkens, M. J., Wu, Q., Pronk, J. T., van
[49] Verduyn, C., Vankleef, R., Frank, J., Schreuder, H., et al., Maris, A. J., Novel evolutionary engineering approach for
Properties of the NAD(P)H-dependent xylose reductase accelerated utilization of glucose, xylose, and arabinose
from the xylose-fermenting yeast Pichia stipitis. Biochem. J. mixtures by engineered Saccharomyces cerevisiae strains.
1985, 226, 669–677. Appl. Environ. Microbiol. 2009, 75, 907–914.
[50] Rizzi, M., Erlemann, P., Buithanh, N. A., Dellweg, H., Xylose [56] Richard, P., Verho, R., Putkonen, M., Londesborough, J.,
fermentation by yeasts. 4.Purification and kinetic-studies Penttilä, M., Production of ethanol from L-arabinose by
of xylose reductase from Pichia stipitis. Appl. Microbiol. Saccharomyces cerevisiae containing a fungal L-arabinose
Biotechnol. 1988, 29, 148–154. pathway. FEMS Yeast Res. 2003, 3, 185–189.
[51] Karhumaa, K., Wiedemann, B., Hahn-Hägerdal, B., Boles, [57] Verho, R., Putkonen, M., Londesborough, J., Penttilä, M.,
E., Gorwa-Grauslund, M. F., Co-utilization of L-arabinose Richard, P., A novel NADH-linked L-xylulose reductase in
and D-xylose by laboratory and industrial Saccharomyces the L-arabinose catabolic pathway of yeast. J. Biolo. Chem.
cerevisiae strains. Microb. Cell Fact. 2006, 5, 18. 2004, 279, 14746–14751.
[52] Garcia Sanchez, R., Karhumaa, K., Fonseca, C., Sànchez [58] Bettiga, M., Bengtsson, O., Hahn-Hägerdal, B., Gorwa-
Nogué, V., et al., Improved xylose and arabinose utilisation Grauslund, M. F., Arabinose and xylose fermentation by re-
by an industrial recombinant Saccharomyces cerevisiaee

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 297


Biotechnology Biotechnol. J. 2011, 6, 286–299
Journal

combinant Saccharomyces cerevisiae expressing a fungal [75] Heer, D., Heine, D., Sauer, U., Resistance of Saccharomyces
pentose utilization pathway. Microb. Cell Fact. 2009, 8, 40. cerevisiae to high concentrations of furfural is based on
[59] Zaldivar, J., Martinez, A., Ingram, L. O., Effect of selected NADPH-dependent reduction by at least two oxireduc-
aldehydes on the growth and fermentation of ethanolo- tases. Appl. Environ. Microbiol. 2009, 75, 7631–7638.
genic Escherichia coli. Biotechnol. Bioeng. 1999, 65, 24–33. [76] Alriksson, B., Horváth, I. S., Jönsson, L. J., Overexpression of
[60] Heer, D., Sauer, U., Identification of furfural as a key toxin Saccharomyces cerevisiae transcription factor and mul-
in lignocellulosic hydrolysates and evolution of a tolerant tidrug resistance genes conveys enhanced resistance to lig-
yeast strain. Microb. Biotechnol. 2008, 1, 497–506. nocellulose-derived fermentation inhibitors. Process
[61] Taherzadeh, M. J., Gustafsson, L., Niklasson, C., Lidén, G., Biochem. 2010, 45, 264–271.
Physiological effects of 5-hydroxymethylfurfural on Sac- [77] Koopman, F., Wierckx, N., de Winde, J. H., Ruijssenaars, H.
charomyces cerevisiae. Appl. Microbiol. Biotechnol. 2000, 53, J., Identification and characterization of the furfural and 5-
701–708. (hydroxymethyl)furfural degradation pathways of Cupri-
[62] Larsson, S., Quintana-Sainz, A., Reimann, A., Nilvebrant, N. avidus basilensis HMF14. Proc. Natl. Acad. Sci. USA. 2010,
O., Jönsson, L. J., Influence of lignocellulose-derived aro- 107, 4919–4924.
matic compounds on oxygen-limited growth and ethanolic [78] Öhgren, K., Bengtsson, O., Gorwa-Grauslund, M. F., Galbe,
fermentation by Saccharomyces cerevisiae. Appl. Biochem. M., et al., Simultaneous saccharification and co-fermenta-
Biotechnol. 2000, 84–86, 617–632. tion of glucose and xylose in steam-pretreated corn stover
[63] Klinke, H. B., Thomsen, A. B., Ahring, B. K., Inhibition of at high fiber content with Saccharomyces cerevisiae
ethanol-producing yeast and bacteria by degradation prod- TMB3400. J. Biotechnol. 2006, 126, 488–498.
ucts produced during pre-treatment of biomass. Appl. Mi- [79] Karhumaa, K., Fromanger, R., Hahn-Hägerdal, B., Gorwa-
crobiol. Biotechnol. 2004, 66, 10–26. Grauslund, M. F., High activity of xylose reductase and xyl-
[64] Modig, T., Liden, G., Taherzadeh, M. J., Inhibition effects of itol dehydrogenase improves xylose fermentation by re-
furfural on alcohol dehydrogenase, aldehyde dehydroge- combinant Saccharomyces cerevisiae. Appl. Microbiol.
nase and pyruvate dehydrogenase. Biochem. J. 2002, 363, Biotechnol. 2007, 73, 1039–1046.
769–776. [80] Bruinenberg, P. M., Debot, P. H. M., van Dijken, J. P., Schef-
[65] Allen, S. A., Clark, W., McCaffery, J. M., Cai, Z., et al., Furfur- fers, W. A., The role of redox balances in the anaerobic fer-
al induces reactive oxygen species accumulation and cellu- mentation of xylose by yeasts. Eur. J. Appl. Microbiol.
lar damage in Saccharomyces cerevisiae. Biotechnol. Biofu- Biotechnol. 1983, 18, 287–292.
els 2010, 3, 2. [81] Wahlbom, C. F., Hahn-Hägerdal, B., Furfural, 5-hydrox-
[66] Banerjee, N., Bhatnagar, R., Viswanathan, L., Inhibition of ymethyl furfural, and acetoin act as external electron ac-
glycolysis by furfural in Saccharomyces cerevisiae. Appl. Mi- ceptors during anaerobic fermentation of xylose in recom-
crobiol. Biotechnol. 1981, 11, 226–228. binant Saccharomyces cerevisiae. Biotechnol. Bioeng. 2002,
[67] Heipieper, H. J., Weber, F. J., Sikkema, J., Keweloh, H., De 78, 172–178.
Bont, J. A. M., Mechanisms of resistance of whole cells to [82] Almeida, J. R. M., Bertilsson, M., Hahn-Hägerdal, B., Lidén,
toxic organic solvents. Trends Biotechnol. 1994, 12, 409–415. G., Gorwa-Grauslund, M. F., Carbon fluxes of xylose-con-
[68] Almeida, J. R. M., Bertilsson, M., Gorwa-Grauslund, M. F., suming Saccharomyces cerevisiae strains are affected dif-
Gorsich, S., Lidén, G., Metabolic effects of furaldehydes and ferently by NADH and NADPH usage in HMF reduction.
impacts on biotechnological processes. Appl. Microbiol. Appl. Microbiol. Biotechnol. 2009, 84, 751–761.
Biotechnol. 2009, 82, 625–638. [83] Palmqvist, E., Hahn-Hägerdal, B., Fermentation of ligno-
[69] Delgenes, J. P., Moletta, R., Navarro, J. M., Effects of ligno- cellulosic hydrolysates. II: Inhibitors and mechanisms of
cellulose degradation products on ethanol fermentations of inhibition. Bioresour. Technol. 2000, 74, 25–33.
glucose and xylose by Saccharomyces cerevisiae, Zy- [84] Lu, Y., Warner, R., Sedlak, M., Ho, N., Mosier, N. S., Compar-
momonas mobilis, Pichia stipitis, and Candida shehatae. En- ison of glucose/xylose cofermentation of poplar hy-
zyme Microb. Technol. 1996, 19, 220–225. drolysates processed by different pretreatment technolo-
[70] Endo, A., Nakamura, T., Ando, A., Tokuyasu, K., Shima, J., gies. Biotechnol. Prog. 2009, 25, 349–356.
Genome-wide screening of the genes required for toler- [85] Schell, D. J., Riley, C. J., Dowe, N., Farmer, J., et al., A
ance to vanillin, which is a potential inhibitor of bioethanol bioethanol process development unit: Initial operating ex-
fermentation, in Saccharomyces cerevisiae. Biotechnol. Bio- periences and results with a corn fiber feedstock. Biore-
fuels 2008, 1, 3. sour. Technol. 2004, 91, 179–188.
[71] Larroy, C., Fernández, M. R., González, E., Parés, X., Biosca, [86] Narendranath, N. V., Thomas, K. C., Ingledew, W. M., Effects
J. A., Characterization of the Saccharomyces cerevisiae of acetic acid and lactic acid on the growth of Saccha-
YMR318C (ADH6) gene product as a broad specificity romyces cerevisiae in a minimal medium. J. Ind. Microbiol.
NADPH-dependent alcohol dehydrogenase: Relevance in Biotechnol. 2001, 26, 171–177.
aldehyde reduction. Biochem. J. 2002, 361, 163–172. [87] Graves,T., Narendranath, N.V., Dawson, K., Power, R., Effect
[72] Liu, Z. L., Slininger, P. J., Dien, B. S., Berhow, M. A., et al., of pH and lactic or acetic acid on ethanol productivity by
Adaptive response of yeasts to furfural and 5-hydrox- Saccharomyces cerevisiae in corn mash. J. Ind. Microbiol.
ymethylfurfural and new chemical evidence for HMF con- Biotechnol. 2006, 33, 469–474.
version to 2,5-bis-hydroxymethylfuran. J. Ind. Microbiol. [88] Thomas, K. C., Hynes, S. H., Ingledew, W. M., Influence of
Biotechnol. 2004, 31, 345–352. medium buffering capacity on inhibition of Saccharomyces
[73] Villa, G. P., Bartroli, R., López, R., Guerra, M., et al., Micro- cerevisiae growth by acetic and lactic acids. Appl. Environ.
bial transformation of furfural to furfuryl alcohol by Sac- Microbiol. 2002, 68, 1616–1623.
charomyces cerevisiae. Acta Biotechnol. 1992, 12, 509–512. [89] Pampulha, M. E., Loureiro-Dias, M. C., Energetics of the ef-
[74] Boopathy, R., Bokang, H., Daniels, L., Biotransformation of fect of acetic acid on growth of Saccharomyces cerevisiae.
furfural and 5-hydroxymethyl furfural by enteric bacteria. FEMS Microbiol. Lett. 2000, 184, 69–72.
J. Ind. Microbiol. 1993, 11, 147–150.

298 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


Biotechnol. J. 2011, 6, 286–299 www.biotechnology-journal.com

[90] Russell, J., Another explanation for the toxicity of fermen- in tolerance of Saccharomyces cerevisiae to Hop Iso-alfa-
tation acids at low pH: Anion accumulation versus uncou- acids. Appl. Environ. Microbiol. 2010, 76, 318–328.
pling. J. Appl. Bacteriol. 1992, 73, 363–370. [108] Öhgren, K., Bura, R., Saddler, J., Zacchi, G., Effect of hemi-
[91] Pampulha, M. E., Loureiro-Dias, M. C., Activity of glycolytic cellulose and lignin removal on enzymatic hydrolysis of
enzymes of Saccharomyces cerevisiae in the presence of steam pretreated corn stover. Bioresour. Technol. 2007, 98,
acetic acid. Appl. Microbiol. Biotechnol. 1990, 34, 375–380. 2503–2510.
[92] Piper, P., Mahe, Y., Thompson, S., Pandjaitan, R., et al., The [109] Carrasco, C., Baudel, H. M., Sendelius, J., Modig, T., et al.,
pdr12 ABC transporter is required for the development of SO2-catalyzed steam pretreatment and fermentation of en-
weak organic acid resistance in yeast. EMBO J. 1998, 17, zymatically hydrolyzed sugarcane bagasse. Enzyme Microb.
4257–4265. Technol. 2010, 46, 64–73.
[93] Bellissimi, E., van Dijken, J. P., Pronk, J.T., van Maris, A. J. A., [110] Remond, C., Aubry, N., Cronier, D., Noel, S., et al., Combina-
Effects of acetic acid on the kinetics of xylose fermentation tion of ammonia and xylanase pretreatments: Impact on
by an engineered, xylose-isomerase-based Saccharomyces enzymatic xylan and cellulose recovery from wheat straw.
cerevisiae strain. FEMS Yeast Res. 2009, 9, 358–364. Bioresour. Technol. 2010, 101, 6712–6717.
[94] Casey, E., Sedlak, M., Ho, N. W. Y., Mosier, N. S., Effect of [111] Linde, M., Jakobsson, E. L., Galbe, M., Zacchi, G., Steam in-
acetic acid and pH on the cofermentation of glucose and retreatment of dilute H2SO4-impregnated wheat straw and
xylose to ethanol by a genetically engineered strain of Sac- SSF with low yeast and enzyme loadings for bioethanol
charomyces cerevisiae. FEMS Yeast Res. 2010, 10, 385–393. production. Biomass Bioenergy 2008, 32, 326– 332.
[95] Helle, S., Cameron, D., Lam, J., White, B., Duff, S., Effect of [112] Linde, M., Galbe, M., Zacchi, G., Simultaneous saccharifica-
inhibitory compounds found in biomass hydrolysates on tion and fermentation of steam-pretreated barley straw at
growth and xylose fermentation by a genetically engi- low enzyme loadings and low yeast concentration. Enzyme
neered strain of S. cerevisiae. Enzyme Microb.Technol. 2003, Microb. Technol. 2007, 40, 1100–1107.
33, 786–792. [113] Wiselolgel, A., Biomass feedstock resources and composi-
[96] Helle, S. S., Murray,A., Lam, J., Cameron, D. R., Duff, S. J., Xy- tion. in: Wyman, C. (Ed.), Handbook on Bioethanol Produc-
lose fermentation by genetically modified Saccharomyces tion and Utilization, Taylor and Francis Washington, DC
cerevisiae 259ST in spent sulfite liquor. Bioresour. Technol- 1996, pp. 108–118.
nol. 2004, 92, 163–171. [114] Hsu, T. C., Guo, G. L., Chen, W. H., Hwang, W. S., Effect of di-
[97] Medina, V. G., Almering, M. J. H., van Maris, A. J. A., Pronk, J. lute acid pretreatment of rice straw on structural proper-
T., Elimination of glycerol production in anaerobic cultures ties and enzymatic hydrolysis. Bioresour.Technol. 2010, 101,
of a Saccharomyces cerevisiae strain engineered to use 4907–4913.
acetic acid as an electron acceptor. Appl. Environ. Microbi- [115] Faga, B. A., Wilkins, M. R., Banat, I. M., Ethanol production
ol. 2010, 76, 190–195. through simultaneous saccharification and fermentation of
[98] de Oliva-Neto, P., Yokoya, F., Evaluation of bacterial con- switchgrass using Saccharomyces cerevisiae D(5)A and
tamination in a fed-batch alcoholic fermentation process. thermotolerant Kluyveromyces marxianus IMB strains.
World J. Microbiol. Biotechnol. 1994, 10, 697–699. Bioresour. Technolnol. 2010, 101, 2273–2279.
[99] Skinner, K.A., Leathers,T. D., Bacterial contaminants of fuel [116] Sassner, P., Galbe, M., Zacchi, G., Steam pretreatment of
ethanol production. J. Ind. Microbiol. Biotechnol. 2004, 31, Salix with and without SO2 impregnation for production of
401–408. bioethanol. Appl. Biochem. Biotechnol. 2005, 121–124, 1101–
[100] Lindén,T., Peetre, J., Hahn-Hägerdal, B., Isolation and char- 1117.
acterization of acetic acid-tolerant galactose-fermenting [117] Li, B. Z., Balan, V., Yuan, Y. J., Dale, B. E., Process optimiza-
strains of Saccharomyces cerevisiae from a spent sulfite tion to convert forage and sweet sorghum bagasse to
liquor fermentation plant. Appl. Environ. Microbiol. 1992, ethanol based on ammonia fiber expansion (AFEX) pre-
58, 1661–1669. treatment. Bioresour. Technol. 2010, 101, 1285–1292.
[101] Basílio, A. C. M., de Araújo, P. R. L., de Morais, J. O. F., da Sil- [118] Olofsson, K., Rudolf, A., Lidén, G., Designing simultaneous
va Filho, E. A., et al., Detection and identification of wild saccharification and fermentation for improved xylose
yeast contaminants of the industrial fuel ethanol fermen- conversion by a recombinant strain of Saccharomyces cere-
tation process. Curr. Microbiol. 2008, 56, 322–326. visiae. J. Biotechnol. 2008, 134, 112–120.
[102] andler, O., Carbohydrate metabolism in lactic acid bacteria. [119] Rudolf,A., Baudel, H., Zacchi, G., Hahn-Hägerdal, B., Lidén,
Antonie Van Leeuwenhoek 1983, 49, 209–224. G., Simultaneous saccharification and fermentation of
[103] Rao, M. R., Stokes, J. L., Utilization of ethanol by acetic acid steam-pretreated bagasse using Saccharomyces cerevisiae
bacteria. J. Bacteriol. 1953, 66, 634–638. TMB3400 and Pichia stipitis CBS6054. Biotechnol. Bioeng.
[104] Narendranath, N. V., Hynes, S. H., Thomas, K. C., Ingledew, 2008, 99, 783–790.
W. M., Effects of lactobacilli on yeast-catalyzed ethanol fer- [120] Matsushika, A., Inoue, H., Murakami, K., Takimura, O.,
mentations. Appl. Environ. Microbiol. 1997, 63, 4158– 4163. Sawayama, S., Bioethanol production performance of five
[105] de Souza Liberal, A. T., Basílio, A. C. M., do Monte Resende, recombinant strains of laboratory and industrial xylose-
A., Brasileiro, B. T. V., et al., Identification of Dekkera brux- fermenting Saccharomyces cerevisiae. Bioresour. Technol-
ellensis as a major contaminant yeast in continuous fuel nol. 2009, 100, 2392–2398.
ethanol fermentation. J. Appl. Microbiol. 2007, 102, 538–547. [121] Runquist, D., Hahn-Hägerdal, B., Bettiga, M., Increased
[106] Sakamoto, K., Konings, W. N., Beer spoilage bacteria and ethanol productivity in xylose-utilizing Saccharomyces
hop resistance. Int. J. Food Microbiol. 2003, 89, 105–124. cerevisiae via a randomly mutagenized xylose reductase.
[107] Hazelwood, L. A., Walsh, M. C., Pronk, J. T., Daran, J. M., In- Appl. Environ. Microbiol. 2010, 76, 7796–7802.
volvement of vacuolar sequestration and active transport

© 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 299

You might also like