+ All documents
Home > Documents > Microbial responses to solvent and alcohol stress

Microbial responses to solvent and alcohol stress

Date post: 17-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
10
© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 1 Biotechnol. J. 2008, 3 DOI 10.1002/biot.200800158 www.biotechnology-journal.com 1 Introduction The pursuit of suitable renewable alternatives to oil-based fuels has become a socio-economic pri- ority, starkly pertinent in South Africa at both the level of government and consumer because of dra- matic increases in the price of petrol at the pump and increases in the price of basic foodstuffs. The negative impacts of fossil fuel combustion on hu- man health and the local and global environments are well established [1, 2] and include the emission of sulphur oxide, nitrous oxide, carbon monoxide and carbon dioxide, potentially contributing to glo- bal warming [3, 4]. In consequence, there is significant interest in blending alcohols, particularly anhydrous ethanol and butanol, with gasoline, having the twofold ef- fect of limiting CO 2 and particulate emissions and reducing national dependency on fossil fuels. The best example of successful national transition from fossil fuel dependency to renewable fuel depend- ency in the transport sector is in Brazil where the national alcohol programme was responsible, in the 1996/97 season, for 273 million tons of har- vested (wet weight) sugar cane, leading to 13.7 mil- lion m 3 ethanol destined for transportation [5–7]. Internationally, the biological production of al- cohol for fuel supplementation has always focused on conventional fermentation using established or- ganisms (1st generation processes) such as Saccha- romyces cerevisiae [8, 9] and Zymomonas mobilis [10–12]. Such organisms have distinct advantages in terms of ethanol yields, high solvent tolerance and very well understood fermentations. However, a major drawback to these processes is that they Review Microbial responses to solvent and alcohol stress Mark Taylor 1,2 , Marla Tuffin 1 , Stephanie Burton 3 , Kirstin Eley 2 and Don Cowan 1 1 Institute for Microbial Biotechnology and Metagenomics (IMBM), University of the Western Cape, Cape Town, South Africa 2 TMO Renewables Ltd, Guildford, Surrey, UK 3 Department of Chemical Engineering, University of Cape Town, Cape Town, South Africa Increasing fuel prices and doubts over the long-term availability of oil are currently major global concerns. Such concerns have led to national policies and objectives to develop microbially pro- duced alcohols as fuel additives or substitutes. However, in South Africa this solution poses the further dilemma of sourcing a suitable fermentative carbohydrate that will not impact negatively on the availability of staple foods. The solution lies in the use of lignocellulosic materials, currently a waste product of the food and agriculture industries, which could be used in conjunction with a catabolically suitable production strain. In the pursuit of lignocellulosic biofuel production, con- ventional fermentation strains have been shown to have limited catabolic versatility. However, catabolically versatile engineered strains and novel isolates engineered with ethanologenic path- ways have subsequently been shown to exhibit limitations in solvent tolerance, hindering their full potential as economically viable production strains. A considerable volume of research has been reported on the general cellular mechanisms and physiological responses to solvent shock as well as adaptive changes responsible for solvent tolerant phenotypes in mutant progeny. Here we re- view a number of the more common cell responses to solvents with particular focus on alcohol tolerance. Keywords: Ethanol · Solvent · Tolerance · Biofuels · Fermentation Correspondence: Professor Donald Cowan, Institute for Microbial Biotech- nology and Metagenomics, University of the Western Cape, Bellville 7535, Cape Town, South Africa E-mail: [email protected] Received 1 August 2008 Revised 2 September 2008 Accepted 3 September 2008
Transcript

© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 1

Biotechnol. J. 2008, 3 DOI 10.1002/biot.200800158 www.biotechnology-journal.com

1 Introduction

The pursuit of suitable renewable alternatives tooil-based fuels has become a socio-economic pri-ority, starkly pertinent in South Africa at both thelevel of government and consumer because of dra-matic increases in the price of petrol at the pumpand increases in the price of basic foodstuffs. Thenegative impacts of fossil fuel combustion on hu-man health and the local and global environmentsare well established [1, 2] and include the emissionof sulphur oxide, nitrous oxide, carbon monoxideand carbon dioxide, potentially contributing to glo-bal warming [3, 4].

In consequence, there is significant interest inblending alcohols, particularly anhydrous ethanoland butanol, with gasoline, having the twofold ef-fect of limiting CO2 and particulate emissions andreducing national dependency on fossil fuels. Thebest example of successful national transition fromfossil fuel dependency to renewable fuel depend-ency in the transport sector is in Brazil where thenational alcohol programme was responsible, inthe 1996/97 season, for 273 million tons of har-vested (wet weight) sugar cane, leading to 13.7 mil-lion m3 ethanol destined for transportation [5–7].

Internationally, the biological production of al-cohol for fuel supplementation has always focusedon conventional fermentation using established or-ganisms (1st generation processes) such as Saccha-romyces cerevisiae [8, 9] and Zymomonas mobilis[10–12]. Such organisms have distinct advantagesin terms of ethanol yields, high solvent toleranceand very well understood fermentations. However,a major drawback to these processes is that they

Review

Microbial responses to solvent and alcohol stress

Mark Taylor1,2, Marla Tuffin1, Stephanie Burton3, Kirstin Eley2 and Don Cowan1

1 Institute for Microbial Biotechnology and Metagenomics (IMBM), University of the Western Cape, Cape Town, South Africa2 TMO Renewables Ltd, Guildford, Surrey, UK3 Department of Chemical Engineering, University of Cape Town, Cape Town, South Africa

Increasing fuel prices and doubts over the long-term availability of oil are currently major globalconcerns. Such concerns have led to national policies and objectives to develop microbially pro-duced alcohols as fuel additives or substitutes. However, in South Africa this solution poses thefurther dilemma of sourcing a suitable fermentative carbohydrate that will not impact negativelyon the availability of staple foods. The solution lies in the use of lignocellulosic materials, currentlya waste product of the food and agriculture industries, which could be used in conjunction with acatabolically suitable production strain. In the pursuit of lignocellulosic biofuel production, con-ventional fermentation strains have been shown to have limited catabolic versatility. However,catabolically versatile engineered strains and novel isolates engineered with ethanologenic path-ways have subsequently been shown to exhibit limitations in solvent tolerance, hindering their fullpotential as economically viable production strains. A considerable volume of research has beenreported on the general cellular mechanisms and physiological responses to solvent shock as wellas adaptive changes responsible for solvent tolerant phenotypes in mutant progeny. Here we re-view a number of the more common cell responses to solvents with particular focus on alcoholtolerance.

Keywords: Ethanol · Solvent · Tolerance · Biofuels · Fermentation

Correspondence: Professor Donald Cowan, Institute for Microbial Biotech-nology and Metagenomics, University of the Western Cape, Bellville 7535, Cape Town, South AfricaE-mail: [email protected]

Received 1 August 2008Revised 2 September 2008Accepted 3 September 2008

BiotechnologyJournal Biotechnol. J. 2008, 3

2 © 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

rely on readily fermentable C6 sugars (starch, su-crose and glucose) making such processes viableonly in countries with the agricultural capacitiesfor growing large quantities of carbohydrate-richcrops in excess of those required for consumption.

Dramatic increases in global populations andfood demand have motivated some researchers todevelop novel production strains with broadercatabolic properties [13–17]. The aim is that suchstrains will efficiently ferment sugars (and othercarbon fractions) derived from the hydrolysis ofhemicellulosic biomass to generate ethanol or bu-tanol and concurrently conserve food grade carbo-hydrate supplies for human consumption.

2 Fuel alcohol

The general observation can be made that many ofthe original process issues that stood in the way ofthe development of economically viable biofuel inthe latter half of the 20th century have re-emergedwith the resurgence of interest in fuel alcohol pro-duction [18–20]. Other than economically viable al-cohol yields, two principal requirements, seen bymany researchers as crucial in the development ofa process strain are: (i) catabolic versatility andhence the ability to ferment the carbohydrate frac-tion of biomass hydrolysates and (ii) the adaptationor development of process strains to grow and tol-erate high (>40 g/L) exogenous ethanol concentra-tions [21, 22].

The principal anaerobic metabolic pathwaysthat result in alcohol production have been de-scribed in some detail [9, 11, 12, 18]. One of the keyintermediary enzymes involved in ethanol synthe-sis is pyruvate decarboxylase (PDC), variants ofwhich have been described from a number ofmesophilic yeasts [23–29] and a limited number ofmesophilic bacterial species [30–34]. PDC catalysesthe Mg2+- and thiamine pyrophosphate (TPP)-de-pendent decarboxylation of pyruvate to acetalde-hyde and CO2. The subsequent conversion of ac-etaldehyde to ethanol is catalysed by a specific al-cohol dehydrogenase with the concurrent regener-ation of the cofactor NAD+, thus helping maintain aphysiological redox balance within the cell. Otheranaerobic metabolic routes to ethanol have beendescribed in both mesophilic and thermophilicspecies, principally via pyruvate catabolism catal-ysed by pyruvate-formate lyase or variants of pyru-vate dehydrogenase [35–38].

The anaerobic production of butanol forms partof the acetone, butanol and ethanol fermentation(ABE) process of anaerobic mesophiles, principal-ly characterised in Clostridium acetobutylicum and

Clostridium beijerinckii [39, 40]. Butanol productionresults from of a number of sequential reductionsof metabolic intermediates derived from acetylCoA, its production occurring concurrently withthe production of acetone, ethanol and other minorend products such as butyric acid [41].

Major advances have been made in the devel-opment of a number of catabolically versatile mu-tants in the established ethanologenic strains, Sac-charomyces cerevisiae [15, 16, 42] and Zymomonasmobilis [11, 43, 44] and in a number of other yeastand bacterial species that possess the Entner-Doudoroff metabolic pathway. This pathway in-cludes the key enzyme directing carbon flow toethanol, PDC [45, 46]. An alternative approach hasalso been the expression of the pdc gene and a suit-able alcohol dehydrogenase (adh) in metabolicallyversatile hosts such as E. coli [47–51].

For butanol production, some degree of catabol-ic versatility is inherent in the strains that haveformed the major focus in this research (C. aceto-butylicum and C. beijerinckii) [39]. However, solventtolerance has been cited as a major constraint inthe attainment of high production yields [52]. Thesensitivity of many microorganisms to high exoge-nous alcohol concentrations thus appears to be one(if not the most) important limiting factor in the de-velopment of new non-yeast biofuel productionstrategies.

Numerous cellular mechanisms have been re-ported to be involved in the responses to high ex-ogenous alcohol concentrations. Such mechanismsinclude responses at the site of solvent contact (i.e.compositional alterations of the cell membrane),enhanced capacity for membrane repair and theup-regulation or de novo expression of solvent ef-flux systems [10–12, 53–55].

3 Aspects of solvent tolerance in yeasts

Although not the primary focus of this review, in-trinsic alcohol and general stress tolerance is acharacteristic of many yeast species and has beenstudied extensively with general aim of improvingtheir applicability to many biotechnological pro-cesses [56, 57]. An understanding of the responsemechanism in these yeast strains has led to a morefocused and rational approach to similar studies inbacteria. Most notably, ethanol tolerance in S. cere-visiae has been well documented to correlate withan increased degree of fatty acid unsaturation inmembrane lipids [58, 59].

Ethanol tolerance in yeasts has also been attrib-uted to a number of other factors involved in gen-eral stress response, such as plasma membrane

© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 3

composition/change [60] and accumulation of in-tracellular ethanol and osomoprotectants [61].Tol-erance has also been shown to vary with environ-mental changes such as medium composition [62],temperature [63] and osmotic influences [64].

Specific examples of yeast response mecha-nisms include two yeast strains (S. cerevisiae andKloeckera apiculata), which were initially reportedto respond to increasing ethanol by up-regulatingthe proportion of ergosterol and unsaturated fattyacid levels in their lipid membranes as well asmaintaining phospholipid biosynthesis [60]. In an-other unsaturated fatty acid auxotrophic yeastspecies, the role of ergosterol in tolerance was su-perseded by specific membrane lipid compositionchanges particularly with respect to oleic, linoleicand linolenic acid composition [59, 65].

More recently several groups examining globalgene expression in ethanol-tolerant yeast strainshave shown that multiple synergistic responses areresponsible for the desired phenotype, i.e. im-proved alcohol tolerance. Over 250 genes have beenimplicated in tolerance [66], including a number ofATPases [67–69] and genes encoding heat shockproteins. A summary of ethanol-elicited responsesfrom a variety of yeasts is given in Table 1.

4 General solvent responses in mesophilicbacteria

The general physiological responses to solventshock have been previously reviewed and are mod-erately understood in a number of Pseudomonasspp. [70–72], E. coli [55, 71] and Oenococcus oeni[73–75]. Toluene exposure in Pseudomonas putidaelicits both outer and cytoplasmic membrane adap-tations, regarded as physical mechanisms that pre-vent solvent penetrance [55, 70]. It has been re-ported that in strain DOT-T1 exposure to solventsmediates both short-term [increased membranerigidity achieved through rapid transformation ofthe cis-9,10-methylene hexadecanoic fatty acid tothe unsaturated 9-cis-hexadecenoic acid (C16:1,9cis) fatty acid and subsequently the trans isomer]and long-term responses (proportion of cardiolipinincreases and phosphatidylethanolamine decreas-es in the phospholipid polar head groups) [70, 72].

Several well-characterised solvent responsemechanisms involve solvent exclusion systems andenergy-dependent efflux pumps of the resistance-nodulation-cell division (RND) family, which ex-port toxic organic solvents [71]. The best under-stood efflux pump responses to solvent stress arethose involved in toluene exposure in E. coli,

Pseudomonas putida, and Ps. aeruginosa [71]. In Ps.putida three efflux pumps (TtgABC, TtgDEF, andTtgGHI), coupled to a proton motive force via theTonB system, are thought to be involved [71]. Theefflux pumps AcrAB-TolC and AcrEF-TolC per-form similar roles in E. coli in imparting n-hexanetolerance [71, 76] and are regulated at the tran-scriptional level by both specific and global regula-tors.

Solvent tolerance in E. coli is strain specific anddetermined genetically. Mutants with increasedtolerance have been constructed that were defec-tive in the marR gene that encodes a repressor pro-tein for the mar operon (responsible for environ-mental stress factors) [76]. In addition, increasingthe expression of the stress-response genes (soxS,marA, and robA encoding DNA-binding proteins/transcriptional activators) was shown to increasetolerance in several strains of E. coli through regu-lation of the AcrAB-TolC system [76, 77]. Overex-pression of the manXYZ genes, which code for asugar transporter of the phosphotransferase sys-tem, have also been implicated in solvent resist-ance [78]. A number of other adaptive responseshave been observed in various species, as shown inTable 1. Surprisingly, little has been reported onethanol tolerance responses in the ethanologen,Zymomonas mobilis, apart from the general induc-tion of expression of heat shock proteins [79].

It has been suggested that high exogenous alco-hol concentrations would elicit deleterious effectson bacterial growth, viability, and metabolism dueto the stimulation of leakage within the membrane[53]. This proposal is consistent with the knownbiophysical interactions of phospholipid bilayerswith alcohols. It was also suggested that increasedethanol tolerance phenotypes could result from ac-quired adaptive and evolutionary changes in cellmembrane rigidity, a mechanistic response illus-trated by Oenococcus oeni in response to ethanoland monitored by carboxyfluorescein retention(indicative of membrane integrity) [74, 80].

The principal target for cellular adaptation topotentially toxic levels of extracellular alcohol isthe cell membrane.This was recognised as early as1976 by Ingram [81], who reported that the mem-brane fatty acid composition of E. coli K-12 alteredradically when the strain was grown in the pres-ence of ethanol and that the proportion of 18:1 fat-ty acids increases at the expense of saturated fattyacids.

More recently, proteomic analysis of this strainunder ethanol stress has shown that inosine-5’-monophosphate dehydrogenase, phosphoglucona-te dehydrogenase and glutathione reductase werestrongly regulated, suggesting a key role of redox

Biotechnol. J. 2008, 3 www.biotechnology-journal.com

BiotechnologyJournal Biotechnol. J. 2008, 3

4 © 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

balance maintenance in ethanol adaptation [73]. Inaddition, significant increases in the expression ofcell wall biosynthetic proteins were demonstrated.

Ethanol tolerant cells of Oenococcus oeni havebeen reported to contain a higher proportion of un-saturated fatty acids and a decreased proportion of

total cell lipid content, decreasing membrane fluid-ity [74, 75] (Interestingly the opposite mechanismoccurs in some Staphylococcus spp which increasemembrane fluidity in response to solvent stress).However, in a subsequent report ethanol shock(14%) was found to elicit immediate fluidisation

Table 1. Solvent response processes in microorganisms

Species Solvent Genes implicated Phenotypic response Reference

Oceanomonas baumannii Phenol Fatty acid Induced increases in fatty acid saturation [104, 105]synthase genes and decreases in the mean chain length

of the fatty acids.

Staphylococcus haemolyticus Toluene Cis-trans isomerise, Induced increase in membrane fluidity [106]various fatty acid through increases in the anteiso-fatty acids synthase genes and decreases in straight-chain fatty acids.

Staphylococcus Benzene None specified Tolerance up to 40% benzene. [107]saprophyticus M36

Helicobacter pylori Hexane Imp/ostA Increased membrane permeability and [108]organic solvent tolerance.

Vibrio parahaemolyticus Ethanol Various HSP genes, Increased thermal tolerance and ethanol [109, 110]fatty acid sythase resistivity post initial ethanol shock. Increasedgenes vaccenic acid (18:1) content but decreases

in the proportion of palmitic acid (16:0) and he ratio of saturated fatty acid to unsaturated fatty acid.

Schizosaccharomyces Ethanol tps1 Increased production of trehalose. [65, 111]pombe

Mucor rouxii Ethanol Delta(9)-, Delta(12)- Decrease in ratio of unsaturated/saturated [112]and Delta(6)- fatty acids.desaturase genes

Schizosaccharomyces Ethanol Lanosterol synthase Increased production of lanosterol and [113]pombe unsaturated fatty acids, anaerobically.

Saccharomyces cerevisiae Ethanol OLE1 Increased ATPase activity, oleic acid (C-18:1) [59, 67, var. capensis and palmitoleic acid (C-16:1) synthesis in 114]

plasma membrane.

Kluyveromyces lactis Ethanol/1- KlADH2 Decreasing the unsaturation of its membrane [115]octanol fatty acids and hence decreasing membrane

fluidity.

Saccharomyces. cerevisiae Ethanol HSP12, HSP26, HSP78, Upregulation of multiple genes involved stress [60, 99, HSP104, STRE, TPS1, response, ionic homeostasis, heat protection, 116–118]TLS1, TPS2, UGP1, trehalose synthesis and antioxidant defence. PGM2, GPD1, HOR2, Also increased ATPase activity.DAK1 etc

Lactobacillus plantarum Butanol Various HSP genes Induced expression of heat shock proteins [119, 120]and ethanol (HSP).

Bacillus psychrosaccha- Multiple Various HSP genes Induced expression of HSP. [121]rolyticus solvents

Sake yeast kyokai no. 11 Ethanol STRE Increased expression of genes under the control [61, 122]msn2p of stress response element (STRE) and msn2p.

© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 5

(attaining rigidity gradually after shock) in O. oenias monitored by fluorescence anisotropy withdiphenyl-1,3,5-hexatriene [75].

One of the few investigations to improve solventtolerance in E. coli was reported in 1998 by Aono[76], who isolated several E. coli strains with im-proved tolerance to solvents. He reported that themutants were defective in the marR gene (encod-ing a repressor protein for the mar operon, respon-sible for an environmental stress factor). The highexpression of the stress-response genes, soxS,marA, and robA, also increased organic solvent tol-erance in several strains of E. coli.

Butanol tolerance in several Clostridium spp.has been reported to correlate with the decreasedexpression of glycerol dehydrogenase as demon-strated by transposon mutational analysis and tol-erance studies in C. beijerinckii [82]. Overexpres-sion of the groESL operons in C. acetobutylicumATCC 824 has resulted in higher yields and bu-tanol-tolerant variants with increased expressionof motility and chemotaxis genes and decreasedexpression of other major stress response genes[83, 84]. These researchers have also shown thatSpo0A overexpression increases butanol toleranceas part of a general stress response that includesthe regulation of genes involved in DNA synthesis,cell division, glycolysis and butanol synthesis andvarious heat shock proteins [85].

Many physiological responses, including growth,survival and reproductive ability have been shownto respond to the presence of increased concentra-tions of alcohol. In E. coli, glycolysis was found to beparticularly tolerant to alcohol shock, probably dueto its exergonic nature, which permits glycolyticflux to proceed even in the absence of active growthand cell division [53]. Subsequently, ethanol toler-ant E. coli variants of strain KO11, capable of pro-ducing more than 60 g/L ethanol on xylose, havebeen isolated.These isolates apparently exceed thetolerance threshold of their engineered S. cerevisi-ae and Z. mobilis counterparts. Ethanol tolerantvariants were subsequently found to possess anon-functional fnr gene [86, 87].

Most recently the expression levels of manX,manY, and manZ genes (encoding a sugar trans-porter of the phosphotransferase system) weremonitored in alcohol-shocked E. coli and found tobe strongly up-regulated [78]. The gcv, betIBA andbetT genes and the compounds glycine and betainehave also been linked to increased ethanol toler-ance, the latter as protective osmolytes [86].

5 Solvent responses in thermophilic bacteria

Because of significant interest in the discovery anddevelopment of second generation biofuel process-es designed around strains possessing broad cata-bolic ranges and enhanced alcohol production ca-pacities, several research groups have sought newmetabolically diverse and adaptable isolates frommore extreme environments.There is particular fo-cus on thermophilic species from the generaClostridium [21], Thermoanaerobacter [88–90] andGeobacillus [91, 92]. Thermophilic fermentationsoffer process advantages in product removal by application of a gas stream or low-grade vacuum[52, 93].

In an early attempt to develop an ethanol toler-ant strain of C. thermocellum, it was concluded thatadaptive responses included control of the alcoholproduction rate, yield and concentration [94]. Sub-sequently, it was reported that an increase in num-bers of shortened, unsaturated and anteio-bran-ched fatty acids was found to be associated withethanol shock resulting in increased membrane“fluidity” in C. thermocellum [95]. In C. thermohy-drosulfuricum low ethanol tolerance has beenshown to be due to membrane disruption or inhibi-tion of glycolysis [96].

In 2002, it was reported that a mutant strain ofThermoanaerobacter ethanolicus (39E H8) that wastolerant of 8% ethanol, lacked the primary adh (as-sociated with ethanol consumption) and conse-quently increased the percentage of transmem-brane fatty acids (long chain C30 fatty acids) in re-sponse to the increasing levels of ethanol generat-ed from the functional secondary ADH enzyme(ethanol producing) [97]. It was suggested that thebiochemical basis for alcohol tolerance in ther-mophilic ethanologens varied from species tospecies. For example, C. thermocellum alcohol de-hydrogenase is inhibited by 1% ethanol while thereversible activity T. brockii and C. thermohydrosul-furicum homologues is stimulated by 5% solvent[38]. Further studies of tolerance in C. thermohy-drosulfuricum showed that ethanol tolerance wastemperature dependant and not an artefact ofmembrane disruption or glycolytic inhibition. In asubsequent report it was shown that tolerance toethanol in a tolerant mutant of C. thermohydrosul-furicum was linked to maintenance of a redox bal-ance and reductions in the activities of the reducedferredoxin-linked NAD reductase and NAD-linkedalcohol dehydrogenase [96, 98].

Biotechnol. J. 2008, 3 www.biotechnology-journal.com

BiotechnologyJournal Biotechnol. J. 2008, 3

6 © 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

6 Conclusions and future prospects

To advance any production scale biofuel process,solvent tolerance must be exhibited in the fermen-tative strain and the biochemical and adaptivemechanisms of tolerance must be understood. Sig-nificant advances have been made in the develop-ment of a number of strains with the ability to fer-ment the carbohydrate fraction of biomass hy-drolysate. It has long been recognised that alcoholtolerance would be as important, if not more so,than catabolic flexibility in strain developmentprogrammes. However, although many alcohol tol-erant strains of various species were developed orisolated and several reports made on the mecha-nisms of tolerance in these strains, it would be fairto say that an intimate understanding of solventtolerance in some key ethanologenic organismshas taken second place in lieu of engineering cata-bolic versatility. This is a trend that is now beingreaddressed by exploiting some of the tools of the‘-omics’ era. High coverage and high qualitygenome sequencing is becoming increasingly ac-cessible at an affordable price, allowing re-searchers to compare, clone, express and deletespecific genes more readily and with a higher de-gree of confidence. Coupling genomic data withDNA microarray technology also allows globalgene expression studies (transcriptomics) to be un-dertaken. In a recent study investigating short-term ethanol stress in S. cerevisiae, ethanol shockwas found to have a significant impact on gene ex-pression over a range of physiological areas in-cluding ionic homeostasis, heat protection and en-ergy management [99]. Similar approaches havebeen applied to help understand global butanolstress responses in C. acetobutylicum [100] and toquantify differential mRNA expression in a set of17 cellulose degrading genes of C. thermocellum[101].

Proteomics technologies have been extensivelyapplied to the study of physiological responses toalcohol stress in various microbial strains. Recentapplications of this technology have been reportedfor O. oeni [73] and C. thermocellum [102], bothshowing significant alterations in cell membraneprotein profiles.

An understanding of the physiological process-es effected by environmental change cannot onlyuncover potential gene targets for mutational stud-ies but may also increase our appreciation of theglobal physiological impact that single mutationsmay have on the expression of genes other thanthose targeted. However, it is only when genomicanalysis is combined with transcriptomic and pro-teomic data that a true understanding of global reg-

ulatory machinery is obtained, particularly at thephenotypic level which is so important in anyprocess with a biotechnological application [103].

By combining -omics data with techniques suchas genome shuffling, global transciptome machin-ery engineering and directed evolution, new toolsand integrative platforms can be developed thatwill allow facilitated engineering of industriallysignificant strains. In the context of biofuels, suchmethods would be part of a process engineeringand development strategy to extend beyond firstgeneration biofuels processes towards second gen-eration and cellulosic alcohol [103].

The authors have declared no conflict of interest.

7 References

[1] Hahn-Hagerdal, B., Galbe, M., Gorwa-Grauslund, M. F., Li-den, G. et al., Bio-ethanol – The fuel of tomorrow from theresidues of today. Trend. Biotechnol. 2006, 24, 549–556.

[2] Kampa, M., Castanas, E., Human health effects of air pol-lution. Environ. Pollut. 2008, 151, 362–367.

[3] Hughes, E., Benemann, J. R., Biological fossil CO2 mitiga-tion. Energ. Convers. Manage. 1997, 38, 467–473.

[4] Mathews, J., Seven steps to curb global warming. Energ.Policy 2007, 35, 4247–4259.

[5] Carvalho, L. C. C., Prospects for ethanol and sugar in Brazil.Int. Sugar. J. 1999, 101, 574–577

[6] Goldemberg, J., The ethanol program in Brazil. Environ.Res. Lett. 2006, 1, 1–5.

[7] Macedo, I. D., Greenhouse gas emissions and energy bal-ances in bio-ethanol production and utilization in Brazil.Biomass. Bioenerg. 1996, 14, 77–81.

[8] van Zyl,W. H., Lynd, L. R., den Haan, R., McBride, J. E., Con-solidated bioprocessing for bioethanol production usingSaccharomyces cereviside, Adv. Biochem. Eng. Biotechnol.2007, 108, 205–235.

[9] Lin, Y., Tanaka, S., Ethanol fermentation from biomass re-sources: Current state and prospects. Appl. Microbiol.Biotechnol. 2006, 69, 627–642.

[10] Gunasekaran, P., Raj, K. C., Ethanol fermentation technol-ogy – Zymomonas mobilis. Curr. Sci. 1999, 77, 56–68.

[11] Rogers, P. L., Jeon,Y. J., Lee, K. J., Lawford, H. G., Zymomonasmobilis for fuel ethanol and higher value products. Adv.Biochem. Eng. Biotechnol. 2007, 108, 263–288.

[12] Panesar, P. S., Marwaha, S. S., Kennedy, J. F., Zymomonasmobilis: an alternative ethanol producer. J. Chem. Technol.Biotechnol. 2006, 81, 623–635.

[13] Yanase, H., Nozaki, K., Okamoto, K., Ethanol productionfrom cellulosic materials by genetically engineered Zy-momonas mobilis. Biotechnol. Lett. 2005, 27, 259–263.

[14] Jeffries, T. W., Engineering yeasts for xylose metabolism.Curr. Opin. Biotechnol. 2006, 17, 320–326.

[15] Watanabe, S., Abu Saleh, A., Pack, S. P., Annaluru, N. et al.,Ethanol production from xylose by recombinant Saccha-romyces cerevisiae expressing protein-engineered NADH-preferring xylose reductase from Pichia stipitis. Microbiol-ogy 2007, 153, 3044–3054.

© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 7

[16] Watanabe, S., Saleh, A. A., Pack, S. P., Annaluru, N. et al.,Ethanol production from xylose by recombinant Saccha-romyces cerevisiae expressing protein engineered NADP+-dependent xylitol dehydrogenase. J. Biotechnol. 2007, 130,316–319.

[17] Matsushika, A., Watanabe, S., Kodaki, T., Makino, K. et al.,Bioethanol production from xylose by recombinant Sac-charomyces cerevisiae expressing xylose reductase,NADP(+)-dependent xylitol dehydrogenase, and xyluloki-nase. J. Biosci. Bioeng. 2008, 105, 296–299.

[18] Bai, F. W., Anderson, W. A., Moo-Young, M., Ethanol fer-mentation technologies from sugar and starch feedstocks.Biotechnol. Adv. 2008, 26, 89–105.

[19] Doran-Peterson, J., Cook, D. M., Brandon, S. K., Microbialconversion of sugars from plant biomass to lactic acid orethanol. Plant. J. 2008, 54, 582–592.

[20] Shanmugam, K. T., Ingram, L. O., Engineering biocatalystsfor production of commodity chemicals. J. Mol. Microbiol.Biotechnol. 2008, 15, 8–15.

[21] Dien, B. S., Cotta, M. A., Jeffries, T. W., Bacteria engineeredfor fuel ethanol production: Current status. Appl. Microbi-ol. Biotechnol. 2003, 63, 258–266.

[22] Zaldivar, J., Nielsen, J., Olsson, L., Fuel ethanol productionfrom lignocellulose:A challenge for metabolic engineeringand process integration. Appl. Microbiol. Biotechnol. 2001,56, 17–34.

[23] Franzblau, S. G., Sinclair, N. A., Induction of pyruvate de-carboxylase in Candida utilis. Mycopathologia 1983, 83, 29–33.

[24] Lockington, R. A., Borlace, G. N., Kelly, J. M., Pyruvate de-carboxylase and anaerobic survival in Aspergillus nidu-lans. Gene 1997, 191, 61–67.

[25] Lu, P., Davis, B. P., Jeffries, T. W., Cloning and characteriza-tion of two pyruvate decarboxylase genes from Pichiastipitis CBS 6054. Appl. Environ. Microbiol. 1998. 64, 94–97.

[26] Krieger, F., Spinka, M., Golbik, R., Hubner, G. et al., Pyru-vate decarboxylase from Kluyveromyces lactis. An enzymewith an extraordinary substrate activation behaviour. Eur.J. Biochem. 2002, 269, 3256–3263.

[27] Skory, C. D., Induction of Rhizopus oryzae pyruvate decar-boxylase genes. Curr. Microbiol. 2003, 47, 59–64.

[28] Flikweert, M. T., Van Der Zanden, L., Janssen, W. M.,Steensma, H. Y. et al., Pyruvate decarboxylase: an indis-pensable enzyme for growth of Saccharomyces cerevisiaeon glucose. Yeast 1996. 12, 247–257.

[29] Kellermann, E., Seeboth P. G., Hollenberg C. P., Analysis ofthe primary structure and promoter function of a pyruvatedecarboxylase gene (PDC1) from Saccharomyces cerevisi-ae. Nucleic Acids Res. 1986, 14, 8963–8977.

[30] Conway, T., Osman, Y. A., Konnan, J. I., Hoffmann, E. M. etal., Promoter and nucleotide sequences of the Zymomonasmobilis pyruvate decarboxylase. J. Bacteriol. 1987, 169, 949–954.

[31] Lowe, S. E., Zeikus, J. G., Purification and characterizationof pyruvate decarboxylase from Sarcina ventriculi. J. Gen.Microbiol. 1992, 138, 803–807.

[32] Neale, A. D., Scopes, R. K., Wettenhall, R. E., Hoogenraad,N. J., Nucleotide sequence of the pyruvate decarboxylasegene from Zymomonas mobilis. Nucleic Acids Res. 1987, 15,1753–1761.

[33] Raj, K. C.,Talarico, L.A., Ingram, L. O., Maupin-Furlow, J.A.,Cloning and characterization of the Zymobacter palmaepyruvate decarboxylase gene (pdc) and comparison to bac-

terial homologues. Appl. Environ. Microbiol. 2002, 68, 2869–2876.

[34] Chandra Raj, K., Ingram, L. O., Maupin-Furlow, J. A., Pyru-vate decarboxylase: A key enzyme for the oxidative me-tabolism of lactic acid by Acetobacter pasteurianus. Arch.Microbiol. 2001, 176, 443–451.

[35] Dawes, E. A., Foster, S. M., The formation of ethanol in Es-cherichia coli. Biochim. Biophys. Acta 1956, 22, 253–265.

[36] Lamed, R., Zeikus, J. G., Ethanol production by ther-mophilic bacteria: Relationship between fermentationproduct yields of and catabolic enzyme activities inClostridium thermocellum and Thermoanaerobium brockii.J. Bacteriol. 1980, 144, 569–578.

[37] Ben-Bassat,A., Lamed, R., Zeikus, J. G., Ethanol productionby thermophilic bacteria: metabolic control of end productformation in Thermoanaerobium brockii. J. Bacteriol. 1981,146, 192–199.

[38] Zeikus, J. G., Ben-Bassat, A., Ng, T. K., Lamed, R. J., Ther-mophilic ethanol fermentations. Basic Life Sci. 1981, 18,441–461.

[39] Awang, G. M., Jones, G.A., Ingledew,W. M.,The acetone-bu-tanol-ethanol fermentation. Crit. Rev. Microbiol. 1988, 15,33–67.

[40] Campos, E. J., Qureshi, N., Blaschek, H. P., Production ofacetone butanol ethanol from degermed corn usingClostridium beijerinckii BA101 12. Appl. Biochem. Biotech-nol. 2002, 98–100, 553–561.

[41] Ezeji, T. C., Qureshi, N., Blaschek, H. P., Bioproduction ofbutanol from biomass: From genes to bioreactors. Curr.Opin. Biotechnol. 2007, 18, 220–227.

[42] Walfridsson, M., Bao, X., Anderlund, M., Lilius, G. et al.,Ethanolic fermentation of xylose with Saccharomyces cere-visiae harboring the Thermus thermophilus xylA gene,which expresses an active xylose (glucose) isomerase.Appl. Environ. Microbiol. 1996, 62, 4648–4651.

[43] Lawford, H. G., Rousseau, J. D., Cellulosic fuel ethanol: Al-ternative fermentation process designs with wild-type and

Biotechnol. J. 2008, 3 www.biotechnology-journal.com

Don Cowan was born in New Zealand

in 1954 and underwent his undergradu-

ate and graduate training at the Univer-

sity of Waikato in Hamilton, NZ. After

several years post doctoral research

with the Thermophile Research Unit,

led by Professor Roy Daniel, he moved

to London to take up a Lectureship in

the Department of Biochemistry and

Molecular Biology at University College

London. In 2001, he moved from a Readership in the same depart-

ment to the Chair of Microbiology at the University of the Western

Cape (UWC) in Cape Town, South Africa. In 2006 he established and

is currently Director of the Institute for Microbial Biotechnology and

Metagenomics (IMBM) at UWC. The Institute, with a team of some 30

researchers, is involved in a range of studies from microbial ecology

and phylogenetic, gene discovery to enzyme structure-function and en-

gineering and applied microbiology. The Biofuels Research Program in

IMBM is one of the larger such programs in South Africa.

BiotechnologyJournal Biotechnol. J. 2008, 3

8 © 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

recombinant Zymomonas mobilis. Appl. Biochem. Biotech-nol. 2003, 105–108, 457–469.

[44] Buchholz, S. E., Eveleigh, D. E., Genetic modification of Zy-momonas mobilis. Biotechnol. Adv. 1990, 8, 547–581.

[45] Wood, B. E., Ingram, L. O., Ethanol production from cel-lobiose, amorphous cellulose, and crystalline cellulose byrecombinant Klebsiella oxytoca containing chromosomallyintegrated Zymomonas mobilis genes for ethanol produc-tion and plasmids expressing thermostable cellulasegenes from Clostridium thermocellum. Appl. Environ. Mi-crobiol. 1992, 58, 2103–2110.

[46] Zhou, S., Ingram. L., Engineering endoglucanase-secretingstrains of ethanologenic Klebsiella oxytoca P2. J. Ind. Mi-crobiol. Biotechnol. 1999, 22, 600–607.

[47] Alterthum, F., Ingram, L. O., Efficient ethanol productionfrom glucose, lactose, and xylose by recombinant Es-cherichia coli. Appl. Environ. Microbiol. 1989, 55, 1943–1948.

[48] Banerjee, M., Kinetics of ethanolic fermentation of D-xy-lose by Klebsiella pneumoniae and its mutants. Appl. Envi-ron. Microbiol. 1989, 55, 1169–1177.

[49] de Vos, W. M., Hugenholtz, J., Engineering metabolic high-ways in Lactococci and other lactic acid bacteria. TrendsBiotechnol. 2004, 22, 72–79.

[50] Golias, H., Dumsday, G. J., Stanley, G. A., Pamment, N. B.,Evaluation of a recombinant Klebsiella oxytoca strain forethanol production from cellulose by simultaneous sac-charification and fermentation: comparison with nativecellobiose-utilising yeast strains and performance in co-culture with thermotolerant yeast and Zymomonas mobilis.J. Biotechnol. 2002, 96, 155–168.

[51] Ingram, L. O., Conway,T., Clark, D. P., Sewell, G.W. et al., Ge-netic engineering of ethanol production in Escherichia coli.Appl. Environ. Microbiol. 1987, 53, 2420–2425.

[52] Ezeji, T. C., Qureshi, N., Blaschek, H. P., Butanol fermenta-tion research: Upstream and downstream manipulations.Chem. Rec. 2004, 4, 305–314.

[53] Ingram, L. O., Ethanol tolerance in bacteria. Crit. Rev.Biotechnol. 1990, 9, 305–319.

[54] Isken, S., de Bont, J. A., Bacteria tolerant to organic sol-vents. Extremophiles 1998, 2, 229–238.

[55] Sardessai, Y., Bhosle, S., Tolerance of bacteria to organicsolvents. Res. Microbiol. 2002, 153, 263–268.

[56] Santos, J., Sousa, M. J., Carcloso, H., Inacio, J. et al., Ethanoltolerance of sugar transport, and the rectification of stuckwine fermentations. Microbiology 2008, 154, 422–430.

[57] Attfield, P., Myers, D., Hazell, B., The importance of stresstolerance to baker’s yeast. Australas. Biotechnol. 1997, 7,149–154.

[58] Mishra, P., Kaur, S., Lipids as modulators of ethanol toler-ance in yeast. Appl. Microbiol. Biotechnol. 1991, 34, 697–702.

[59] You, K. M., Rosenfield, C. L., Knipple, D. C., Ethanol toler-ance in the yeast Saccharomyces cerevisiae is dependenton cellular oleic acid content. Appl. Environ. Microbiol.2003, 69, 1499–1503.

[60] Alexandre, H., Rousseaux, I., Charpentier, C., Relationshipbetween ethanol tolerance, lipid composition and plasmamembrane fluidity in Saccharomyces cerevisiae andKloeckera apiculata. FEMS Microbiol. Lett. 1994, 124, 17–22.

[61] Ogawa,Y., Nitta, A., Uchiyama, H., Imamura,T. et al.,Toler-ance mechanism of the ethanol-tolerant mutant of sakeyeast. J. Biosci. Bioeng. 2000, 90, 313–320.

[62] Kadokura, T., Nakazato, A., Takeda, M., Chikubu, S. et al.,Influence of culture liquid of fungi on alcohol tolerance ofyeast. Seibutsu-Kogaku Kais. 1996, 74, 167–170.

[63] Ciesarova, Z., Smogrovicova, D., A study of ethanol toler-ance in yeast. Chem. Listy. 1996, 90, 497–501.

[64] Ciesarova, Z., Smogrovicova, D., Domeny, Z., Enhancementof yeast ethanol tolerance by calcium and magnesium. Fo-lia Microbiol. 1996, 41, 485–488.

[65] Swan, T. M., Watson, K., Stress tolerance in a yeast lipidmutant: Membrane lipids influence tolerance to heat andethanol independently of heat shock proteins and tre-halose. Can. J. Microbiol. 1999, 45, 472–479.

[66] Hu, X. H., Wang, M. H., Tan, T., Li, J. R. et al., Genetic dis-section of ethanol tolerance in the budding yeast Saccha-romyces cerevisiae. Genetics 2007, 175, 1479–1487.

[67] Aguilera, F., Peinado, R.A., Millan, C., Ortega, J. M. et al., Re-lationship between ethanol tolerance, H+-ATPase activityand the lipid composition of the plasma membrane in dif-ferent wine yeast strains. Int. J. Food Microbiol. 2006, 110,34–42.

[68] Fujita, K., Matsuyama, A., Kobayashi, Y., Iwahashi, H., Thegenome-wide screening of yeast deletion mutants to iden-tify the genes required for tolerance to ethanol and otheralcohols. FEMS Yeast Res. 2006, 6, 744–750.

[69] Lei, J., Zhao, X., Ge, X., Bai, F., Ethanol tolerance and thevariation of plasma membrane composition of yeast flocpopulations with different size distribution. J. Biotechnol.2007, 131, 270–275.

[70] Ramos, J. L., Duque, E., Rodriguez-Herva, J. J., Godoy, P.,Haidour, A., Reyes, F., Fernandez-Barrero, A., Mechanismsfor solvent tolerance in bacteria. J. Biol. Chem. 1997, 272,3887–3890.

[71] Ramos, J. L., Duque, E., Gallegos, M. T., Godoy, P. et al.,Mechanisms of solvent tolerance in gram-negative bacte-ria. Annu. Rev. Microbiol. . 2002, 56, 743–768.

[72] Segura, A., Duque, E., Rojas, A., Godoy, P. et al., Fatty acidbiosynthesis is involved in solvent tolerance inPseudomonas putida DOT-T1E. Environ. Microbiol. 2004, 6,416–423.

[73] Silveira, M. G., Baumgartner, M., Rombouts, F. M., Abee, T.,Effect of adaptation to ethanol on cytoplasmic and mem-brane protein profiles of Oenococcus oeni. Appl. Environ.Microbiol. 70, 2748–2755.

[74] Da Silveira, M. G., Golovina, E. A., Hoekstra, F. A., Rom-bouts, F. M. et al., Membrane fluidity adjustments inethanol-stressed Oenococcus oeni cells. Appl. Environ. Mi-crobiol. 2003, 69, 5826–5832.

[75] Tourdot-Marechal, R., Gaboriau, D., Beney, L., Divies, C.,Membrane fluidity of stressed cells of Oenococcus oeni. Int.J. Food Microbiol. 2000, 55, 269–273.

[76] Aono, R., Improvement of organic solvent tolerance levelof Escherichia coli by overexpression of stress-responsivegenes. Extremophiles 1998, 2, 239–248.

[77] Asako, H., Kobayashi, K., Aono, R., Organic solvent toler-ance of Escherichia coli is independent of OmpF levels inthe membrane. Appl. Environ. Microbiol. 1999, 65, 294–296.

[78] Okochi, M., Kurimoto, M., Shimizu, K., Honda, H., Increaseof organic solvent tolerance by overexpression of manXYZin Escherichia coli. Appl. Microbiol. Biotechnol. 2007, 73,1394–1399.

[79] Michel, G. P., Starka, J., Effect of ethanol and heat stresseson the protein pattern of Zymomonas mobilis. J. Bacteriol.1986, 165, 1040–1042.

[80] Graca da Silveira, M., Vitoria San Romao, M., Loureiro-Dias, M. C., Rombouts, F. M. et al., Flow cytometric assess-ment of membrane integrity of ethanol-stressed Oenococ-cus oeni cells. Appl. Environ. Microbiol. 2002, 68, 6087–6093.

© 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 9

[81] Ingram, L. O., Adaptation of membrane lipids to alcohols.J. Bacteriol. 1976, 125, 670–678.

[82] Liyanage, H., Young, M., Kashket, E. R., Butanol toleranceof Clostridium beijerinckii NCIMB 8052 associated withdown-regulation of gldA by antisense RNA. J. Mol. Micro-biol. Biotechnol. 2000, 2, 87–93.

[83] Tomas, C. A., Beamish, J., Papoutsakis, E. T., Transcription-al analysis of butanol stress and tolerance in Clostridiumacetobutylicum. J. Bacteriol. 2004, 186, 2006–2018.

[84] Tomas, C. A., Welker, N. E., Papoutsakis, E. T., Overexpres-sion of groESL in Clostridium acetobutylicum results in in-creased solvent production and tolerance, prolonged me-tabolism, and changes in the cell’s transcriptional pro-gram. Appl. Environ. Microbiol. 2003, 69, 4951–4965.

[85] Alsaker, K. V., Spitzer, T. R., Papoutsakis, E. T., Transcrip-tional analysis of spo0A overexpression in Clostridium ace-tobutylicum and its effect on the cell’s response to butanolstress. J. Bacteriol. 2004, 186, 1959–1971.

[86] Gonzalez, R.,Tao, H., Purvis, J. E.,York, S.W. et al., Gene ar-ray-based identification of changes that contribute toethanol tolerance in ethanologenic Escherichia coli: Com-parison of KO11 (parent) to LY01 (resistant mutant).Biotechnol. Prog. 19, 612–623.

[87] Yomano, L. P.,York, S.W., Ingram, L. O., Isolation and char-acterization of ethanol-tolerant mutants of Escherichia coliKO11 for fuel ethanol production. J. Ind. Microbiol. Biotech-nol. 1998, 20, 132–138.

[88] Georgieva, T. I., Ahring, B. K., Evaluation of continuousethanol fermentation of dilute-acid corn stover hydroly-sate using thermophilic anaerobic bacterium Thermoan-aerobacter BG1L1. Appl. Microbiol. Biotechnol. 2007, 77,61–68.

[89] Georgieva,T. I., Mikkelsen, M. J.,Ahring, B. K., Ethanol pro-duction from wet-exploded wheat straw hydrolysate bythermophilic anaerobic bacterium thermoanaerobacterBG1L1 in a continuous immobilized reactor. Appl.Biochem. Biotechnol. 2008, 145, 99–110.

[90] Klinke, H. B., Thomsen, A, B., Ahring, B. K., Potential in-hibitors from wet oxidation of wheat straw and their effecton growth and ethanol production by Thermoanaerobactermathranii. Appl. Microbiol. Biotechnol. 2001, 57, 631–638.

[91] Thompson, A. H., Studholme, D. J., Green, E. M., Leak, D. J.,Heterologous expression of pyruvate decarboxylase inGeobacillus thermoglucosidasius. Biotechnol. Lett. 2008, 30,1359–1365.

[92] Taylor, M. P., Esteban, C., Leak, D. J., Development of a ver-satile shuttle vector for gene expression in Geobacillus spp.Plasmid 2008, 60, 45–52.

[93] Ezeji, T. C., Karcher, P. M., Qureshi, N., Blaschek, H. P., Im-proving performance of a gas stripping-based recoverysystem to remove butanol from Clostridium beijerinckiifermentation. Bioprocess Biosyst. Eng. 2005, 27, 207–214.

[94] Herrero, A. A., Gomez, R. F., Development of ethanol toler-ance in Clostridium thermocellum: Effect of growth tem-perature. Appl. Environ. Microbiol. 1980, 40, 571–577.

[95] Herrero,A.A., Gomez, R. F., Roberts, M. F., Ethanol-inducedchanges in the membrane lipid composition of Clostridiumthermocellum. Biochim. Biophys. Acta 1982, 693, 195–204.

[96] Lovitt, R.W., Shen, G. J., Zeikus, J. G., Ethanol production bythermophilic bacteria: biochemical basis for ethanol andhydrogen tolerance in Clostridium thermohydrosulfuricum.J. Bacteriol. 1988, 170, 2809–2815.

[97] Burdette, D. S., Jung, S. H., Shen, G. J., Hollingsworth, R. I. etal., Physiological function of alcohol dehydrogenases and

long-chain (C(30)) fatty acids in alcohol tolerance of Ther-moanaerobacter ethanolicus. Appl. Environ. Microbiol. 2002,68, 1914–1918.

[98] Lovitt, R.W., Longin, R., Zeikus, J. G., Ethanol production bythermophilic bacteria: Physiological comparison of solventeffects on parent and alcohol-tolerant strains of Clostridi-um thermohydrosulfuricum. Appl. Environ. Microbiol. 1984,48, 171–177.

[99] Alexandre, H., nsanay-Galeote, V., Dequin, S., Blondin, B.,Global gene expression during short-term ethanol stressin Saccharomyces cerevisiae. FEBS Lett. 2001, 498, 98–103.

[100] Borden, J. R., Papoutsakis, E. T., Dynamics of genomic-li-brary enrichment and identification of solvent tolerancegenes for Clostridium acetobutylicum. Appl. Environ. Micro-biol. 2007, 73, 3061–3068.

[101] Stevenson, D. M., Weimer, P. J., Expression of 17 genes inClostridium thermocellum ATCC 27405 during fermenta-tion of cellulose or cellobiose in continuous culture. Appl.Environ. Microbiol. 2005, 71, 4672–4678.

[102] Williams, T. I., Combs, J. C., Lynn, B. C., Strobel, H. J., Pro-teomic profile changes in membranes of ethanol-tolerantClostridium thermocellum. Appl. Microbiol. Biotechnol.2007, 74, 422–432.

[103] Patnaik, R., Engineering complex phenotypes in industri-al strains. Biotechnol. Prog. 2008, 24, 38–47.

[104] Brown, G. R., Sutcliffe, I. C., Bendell, D., Cummings, S. P.,The modification of the membrane of Oceanomonas bau-mannii when subjected to both osmotic and organic sol-vent stress. FEMS Microbiol. Lett. 2000, 189, 149–154.

[105] Brown, G. R., Sutcliffe, I. C., Cummings, S. P., Combined sol-vent and water activity stresses on turgor regulation andmembrane adaptation in Oceanimonas baumannii ATCC700832. Anton Van. Leeuw. 2003, 83, 275–283.

[106] Nielsen, L. E., Kadavy, D. R., Rajagopal, S., Drijber, R. et al.,Survey of extreme solvent tolerance in gram-positive coc-ci: Membrane fatty acid changes in Staphylococcushaemolyticus grown in toluene. Appl. Environ. Microbiol.2005, 71, 5171–5176.

[107] Fang,Y., Lu, Z., Lv, F., Bie, X. et al., A newly isolated organ-ic solvent tolerant Staphylococcus saprophyticus M36 pro-duced organic solvent-stable lipase. Curr. Microbiol. 2006,53, 510–515.

[108] Chiu, H. C., Lin,T. L.,Wang, J.T., Identification and charac-terization of an organic solvent tolerance gene in Heli-cobacter pylori. Helicobacter 2007, 12, 74–81.

[109] Chiang, M. L., Ho, W. L., Chou, C. C., Response of Vibrioparahaemolyticus to ethanol shock. Food Microbiol. 2006,23, 461–467.

[110] Chiang, M. L., Ho,W. L., Chou, C. C., Ethanol shock changesthe fatty acid profile and survival behaviour of Vibrio para-haemolyticus in various stress conditions. Food Microbiol.2008, 25, 359–365.

[111] Soto,T., Fernandez, J.,Vicente-Soler, J., Cansado, J. et al.,Ac-cumulation of trehalose by overexpression of tps1, codingfor trehalose-6-phosphate synthase, causes increased re-sistance to multiple stresses in the fission yeast Schizosac-charomyces pombe. Appl. Environ. Microbiol. 1999, 65,2020–2024.

[112] Laoteng, K., Jitsue, S., Dandusitapunth, Y., Cheevadha-narak, S., Ethanol-induced changes in expression profilesof cell growth, fatty acid and desaturase genes of Mucorrouxii. Fungal Genet. Biol. 2008, 45, 61–67

Biotechnol. J. 2008, 3 www.biotechnology-journal.com

BiotechnologyJournal Biotechnol. J. 2008, 3

10 © 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

[113] Koukkou, A. I., Tsoukatos, D., Drainas, C., Effect of ethanolon the sterols of the fission yeast Schizosaccharomyces-pombe. FEMS Microbiol. Lett. 1993, 111, 171–175.

[114] Choi, J. Y., Stukey, J., Hwang, S. Y., Martin, C. E., Regulatoryelements that control transcription activation and unsatu-rated fatty acid-mediated repression of the Saccharomycescerevisiae OLE1 gene. J. Biol. Chem. 1996, 271, 3581–3589.

[115] Heipieper, H. J., Isken, S., Saliola, M., Ethanol tolerance andmembrane fatty acid adaptation in adh multiple and nullmutants of Kluyveromyces lactis. Res. Microbiol. 2000, 151,777–784.

[116] Alexandre, H., Rousseaux, I., Charpentier, C., Ethanoladaptation mechanisms in Saccharomyces-cerevisiae.Biotechnol. Appl. Biochem. 1994, 20, 173–183.

[117] Alexandre, H., Berlot, J. P., Charpentier, C., Effect of ethanolon membrane fluidity of protoplasts from Saccharomycescerevisiae and Kloeckera apiculata grown with or withoutethanol, measured by fluorescence anisotropy. Biotechnol.Tech. 1994, 8, 295–300.

[118] Alexandre, H., Charpentier, C.,The plasma-membrane AT-Pase of Kloeckera apiculata purification, characterization

and effect of ethanol on activity. World J. Microbiol. Biotech-nol. 1994, 10, 704–708.

[119] Fiocco, D., Capozzi, V., Goffin, P., Hols, P. et al., Improvedadaptation to heat, cold, and solvent tolerance in Lacto-bacillus plantarum. Appl. Microbiol. Biotechnol. 2007, 77,909–915

[120] Alegria, G. E., Lopez, I., Ruiz, J. I., Saenz, J. et al., High tol-erance of wild Lactobacillus plantarum and Oenococcusoeni strains to lyophilisation and stress environmentalconditions of acid pH and ethanol. FEMS Microbiol. Lett.2004, 230, 53–61.

[121] Kang, H. J., Heo, D. H., Choi, S. W., Kim, K. N. et al., Func-tional characterization of Hsp33 protein from Bacillus psy-chrosaccharolyticus, additional function of HSP33 on re-sistance to solvent stress. Biochem. Biophys. Res. Commun.2007, 358, 743–750.

[122] Watanabe, M.,Tamura, K., Magbanua, J. P.,Takano, K. et al.,Elevated expression of genes under the control of stressresponse element (STRE) and msn2p in an ethanol-toler-ance sake yeast kyokai no. 11. J. Biosci. Bioeng. 2007, 104,163–170.


Recommended