Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

More efficient redox biocatalysis by utilising 1,4-butanediol as a ‘smart cosubstrate’

Selin Kara a, Dominik Spickermann b, Joerg H. Schrittwieser a, Christian Leggewie b, Willem J. H. van Berkel c, Isabel W. C. E. Arends a and Frank Hollmann *a
aDepartment of Biotechnology, Delft University of Technology, Julianalaan 136, 2628BL Delft, The Netherlands. E-mail: f.hollmann@tudelft.nl; Fax: (+31) 152 781415; Tel: (+31) 152 781957
bevocatal GmbH, Merowingerplatz 1A, 40225 Düsseldorf, Germany
cLaboratory of Biochemistry, Wageningen University, Dreijenlaan 3, 6703 HA Wageningen, The Netherlands

Received 12th November 2012 , Accepted 19th December 2012

First published on 20th December 2012


Abstract

1,4-Butanediol is shown to be an efficient cosubstrate to promote NAD(P)H-dependent redox biocatalysis. The thermodynamically and kinetically inert lactone coproduct makes the regeneration reaction irreversible. Thereby not only the molar surplus of cosubstrate is dramatically reduced but also faster reaction rates are obtained.


Enzymes are amongst the catalysts of choice if selectivity is desired. This is particularly true for oxidoreductases catalysing preparatively important reactions ranging from reduction of C[double bond, length as m-dash]O, C[double bond, length as m-dash]C and other functional groups to specific oxyfunctionalisation reactions such as hydroxylation, epoxidation or Baeyer–Villiger reactions.1 Most of these reactions depend on the supply with reducing equivalents, delivered to the enzymes through the nicotinamide cofactors (NAD(P)H). For economic and practical reasons, NAD(P)H has to be applied in catalytic amounts combined with a suitable in situ regeneration system.1,2 After more than 2 decades of intensive research3 this ‘cofactor challenge’ is generally considered to be solved. Amongst a variety of different regeneration systems, ADH-mediated oxidation of simple alcohols such as ethanol or isopropanol represents one of the most common NAD(P)H regeneration systems (Scheme 1). This approach is most elegant if the regenerating ADH is also the production enzyme mediating an (enantio)selective reduction reaction.4 This ‘substrate coupled’ approach represents a biocatalytic version of the well-known Meerwein–Ponndorf–Verley (MPV) reduction.5
Comparison of the ‘classical’ biocatalytic MPV-reduction e.g. using isopropanol and the proposed ‘smart cosubstrate’ approach using 1,4-butanediol. The lactone coproduct renders the regeneration reaction irreversible. The lower part shows the equilibrium conversions as calculated from the law of mass action (lines) and the waste generated (bars).
Scheme 1 Comparison of the ‘classical’ biocatalytic MPV-reduction e.g. using isopropanol and the proposed ‘smart cosubstrate’ approach using 1,4-butanediol. The lactone coproduct renders the regeneration reaction irreversible. The lower part shows the equilibrium conversions as calculated from the law of mass action (lines) and the waste generated (bars).

However, as is common amongst all MPV reductions, the reversibility and poor thermodynamic driving force of the reaction necessitates (unless elaborate coproduct removal is applied) significant molar surpluses of the cosubstrate. From an environmental point of view this is not desirable as the significant waste generated (Scheme 1) has to be dealt with.

Inspired by recent work of Lavandera et al.6 we hypothesised that thermodynamically stable coproducts may represent a facile way to shift the equilibrium of ADH-catalysed MPV reductions and thereby reduce the molar surplus of the cosubstrate used. In that respect α,ω-diols such as 1,4-butanediol (1,4-BD) appeared most promising. First, 1,4-BD can be oxidised twice thereby doubling the yield of reducing equivalents liberated from the cosubstrate and, secondly, the resulting γ-butyrolactone (GBL) represents a thermodynamically stable and kinetically inert coproduct (Scheme 1). Hence, the waste generated for >95% conversion would be reduced approximately 40-fold (see ESI for further details). Overall, a ‘smart cosubstrate’ approach solving the above-mentioned limitations was envisaged.

In a first set of experiments we screened a range of commercially available ADHs for activity towards 1,4-BD. Out of the 12 ADHs evaluated 8 showed significant activity (Fig. 1). Both NAD- and NADP-dependent ADHs showed significant activity towards 1,4-BD (Fig. 1). Amongst them, the well-known ADH from horse liver (HLADH) showed the highest activity and therefore was chosen as a biocatalyst for further investigation. Interestingly, HLADH has been known for decades to accept a broad range of 1,4-diols yielding enantiopure lactones but a ‘smart cosubstrate’ application of this reaction has not been proposed yet.7


Screening of ADHs for the oxidation of 1,4-BD. Reaction conditions: c(1,4-BD) = 0–4200 mM, c(NAD+) = 0.5 mM/c(NADP+) = 0.4 mM, c(ADH) = 0.12–0.36 g L−1, buffer: Tris-HCl (50 mM, pH 7.0), T = 30 °C.
Fig. 1 Screening of ADHs for the oxidation of 1,4-BD. Reaction conditions: c(1,4-BD) = 0–4200 mM, c(NAD+) = 0.5 mM/c(NADP+) = 0.4 mM, c(ADH) = 0.12–0.36 g L−1, buffer: Tris-HCl (50 mM, pH 7.0), T = 30 °C.

The kinetic parameters of 1,4-BD (Fig. S1) as well as EtOH and iPrOH were determined (Table S1) showing that HLADH exhibits a reasonable apparent KM value of 23 mM towards 1,4-BD (together with a mild substrate inhibition, Ki,app = 1.3 M).

Next we compared the performance of 1,4-BD to isopropanol (iPrOH) as a sacrificial electron donor in the HLADH-driven reduction of cinnamaldehyde (Fig. 2). HLADH has been reported as a suitable enzyme for cinnamyl alcohol/cinnamaldehyde substrate/product coupling among the other enzyme preparations.8


Equilibrium conversion values of HLADH-catalysed reduction of cinnamaldehyde in the presence of 0.5 equiv. of 1,4-BD, 0.5–5 equiv. of iPrOH. Conditions: c(cinnamaldehyde) = 5 mM, c(NAD+) = 0.1 mM, c(HLADH) = 1 g L−1, buffer: Tris-HCl (50 mM, pH 7.0), T = 30 °C, time = 72 h.
Fig. 2 Equilibrium conversion values of HLADH-catalysed reduction of cinnamaldehyde in the presence of 0.5 equiv. of 1,4-BD, 0.5–5 equiv. of iPrOH. Conditions: c(cinnamaldehyde) = 5 mM, c(NAD+) = 0.1 mM, c(HLADH) = 1 g L−1, buffer: Tris-HCl (50 mM, pH 7.0), T = 30 °C, time = 72 h.

Even when a 5-fold molar excess of iPrOH was applied, initial rate and maximal conversion fell back significantly behind the results obtained with 0.5 equiv. of 1,4-BD. Similar observations were also made using ethanol as a cosubstrate and/or using further substrates.

It is worth mentioning that using 1,4-BD as a cosubstrate always gave GBL in the expected 1[thin space (1/6-em)]:[thin space (1/6-em)]2 molar ratio to the product formed; the intermediate hydroxyaldehyde or its corresponding hemiacetal was not observed.

Unfortunately, cinnamaldehyde proved to be a poor model substrate due to significant product inhibition. Already in the presence of 0.2 equiv. of alcohol (approx. corresponding to 20% conversion), the initial reduction rate decreased by more than 75% (Fig. S2). Nevertheless, 1,4-BD enabled significantly higher conversions than iPrOH or EtOH, which we attribute to the higher thermodynamic driving force exerted by the irreversible regeneration half-reaction.

To circumvent inhibition issues, we drew our attention to the reduction of α-arylpropionaldehydes (Profen aldehydes) enabling full conversion of the racemic starting material in a reductive dynamic kinetic resolution approach (RDKR).8 Obviously, product inhibition in this case is less pronounced. We therefore evaluated the ‘smart cosubstrate’ approach to promote the synthesis of enantiopure Profen alcohols. Gratifyingly, we found indeed that only 0.5 equiv. of 1,4-BD as a ‘smart cosubstrate’ was necessary to achieve complete conversion of the racemic starting material (Table 1). With comparable amounts of EtOH or iPrOH, conversions of 24 and 14% were observed, respectively. These numbers are significantly lower than the expected equilibrium conversion assuming an equilibrium constant of 1. Most probably, thermodynamic reasons account for this as an MPV reduction between aldehydes and secondary alcohols should be overall thermodynamically uphill.9 This probably also is the reason for the huge molar excess of cosubstrate utilised in previous studies (Table 1).8

Table 1 HLADH-catalysed RDKR of 2-phenyl-1-propanal using various cosubstrates

Cosubstrate MRa [mol mol−1] Conversion [%] ee (S) [%]
Conditions: c(2-phenyl-1-propanal) = 5 mM, c(cosubstrate) = 2.5 mM, c(NAD+) = 0.1 mM, c(HLADH) = 0.1 g L−1, buffer: Tris-HCl (50 mM, pH 7.5, 1% v/v MeCN), T = 30 °C, reaction time: 24 h.a MR: molar ratio of cosubstrate to substrate.b Values taken from ref. 8.c c(HLADH) = 1 g L−1.
i PrOH 0.5 14 >99
EtOH 0.5 24 >99
EtOHb 1000 99.4 66
1,4-BDc 0.5 99 56
1,4-BD 0.5 98 95


Interestingly, the optical purity of the final product decreased with increasing enzyme concentration (Table 1). In the presence of 10 times more biocatalyst the optical purity of the product dropped from 95% ee to 56% ee. We attribute this to the comparably slow in situ racemisation rate of the starting material under the given reaction conditions combined with an imperfect enantiodiscrimination of HLADH.10 Hence, at high HLADH concentrations, the enzymatic reduction activity outperforms the racemisation rate leading to decreased optical purity of the product.

Overall, the ‘smart cosubstrate’ approach appears to be a promising alternative to the established methods of substrate coupled biocatalytic reduction of alcohols. Further exploration of the scope is currently ongoing in our laboratories.

Encouraged by these promising results we also became interested in whether the ‘smart cosubstrate’ method might be generally applicable to promote NADH-dependent redox reactions such as (1) an enoate reductase-catalysed reduction of conjugated C[double bond, length as m-dash]C-double bonds and (2) a monooxygenase-catalysed oxyfunctionalisation reaction. It is worth mentioning that these reactions are thermodynamically favoured and hence irreversible. As a model reaction for the reduction of conjugated C[double bond, length as m-dash]C-double bonds we chose the enantioselective reduction of ketoisophorone using the enoate reductase from Thermus scotoductus SA-01 (TsER, Fig. 3).11



          TsER-catalysed reduction of ketoisophorone to (R)-levodione using 1,4-BD (●) or EtOH (■) as a cosubstrate. Reaction conditions: c(ketoisophorone) = 10 mM, c(cosubstrate) = 5 mM, c(NAD+) = 1 mM, c(TsER) = 0.25 g L−1, c(HLADH) = 1.0 g L−1, buffer: MOPS (50 mM, pH 7.0, 5 mM CaCl2, 1% v/v MeCN), T = 30 °C.
Fig. 3 TsER-catalysed reduction of ketoisophorone to (R)-levodione using 1,4-BD (●) or EtOH (■) as a cosubstrate. Reaction conditions: c(ketoisophorone) = 10 mM, c(cosubstrate) = 5 mM, c(NAD+) = 1 mM, c(TsER) = 0.25 g L−1, c(HLADH) = 1.0 g L−1, buffer: MOPS (50 mM, pH 7.0, 5 mM CaCl2, 1% v/v MeCN), T = 30 °C.

Compared to EtOH as a cosubstrate 1,4-BD gave almost 3 times higher initial rates thus comparing well with the initial rates obtained for the biocatalytic MPV reduction. The product (R)-levodione was produced in high optical purity (ee > 95%) in both cases but slowly racemised over time;12 after 24 h the ee-value had dropped to 86% underlining the necessity for fast reaction kinetics as achieved with the ‘smart cosubstrate’ approach. Hence, we conclude that the ‘smart cosubstrate’ approach is advantageous in terms of rates and amount of waste product even for a thermodynamically favourable reaction such as the alcohol-oxidation promoted reduction of conjugated C[double bond, length as m-dash]C-bonds.13

The second model reaction chosen was the regioselective hydroxylation of 3-hydroxybenzoate yielding 2,5-dihydroxybenzoate as catalysed by 3-hydroxybenzoate-6-hydroxylase (3HB6H) from Rhodococcus jostii RHA1 (Fig. 4).14


Hydroxylation of 3-hydroxybenzoate to 2,5-dihydroxybenzoate using iPrOH (▲), EtOH (■) or 1,4-BD (●). Reaction conditions: c(3-hydroxybenzoate) = 5 mM, c(cosubstrate) = 2.5 mM, c(NAD+) = 1 mM, c(3HB6H) = 0.25 g L−1, c(HLADH) = 1 g L−1, buffer: Tris-SO4 (20 mM, pH 8.0), T = 30 °C.
Fig. 4 Hydroxylation of 3-hydroxybenzoate to 2,5-dihydroxybenzoate using iPrOH (▲), EtOH (■) or 1,4-BD (●). Reaction conditions: c(3-hydroxybenzoate) = 5 mM, c(cosubstrate) = 2.5 mM, c(NAD+) = 1 mM, c(3HB6H) = 0.25 g L−1, c(HLADH) = 1 g L−1, buffer: Tris-SO4 (20 mM, pH 8.0), T = 30 °C.

Again, 1,4-BD proved to be the most efficient cosubstrate enabling initial product formation rates 7.5 and 24.5 times higher than with ethanol and isopropanol, respectively. As a result, full conversion of the starting material was achieved in the time frame of the experiment only using 1,4-BD. Interestingly, conversions obtained using EtOH were somewhat higher than the expected maximal 50%. Possibly, an endogenous E. coli aldehyde dehydrogenase present in the commercial enzyme preparation may further oxidise the acetaldehyde coproduct thereby providing more NADH equivalents.15

Encouraged by these results, we further scaled up the reaction to the semi-preparative scale (7 g L−1 with full conversion into 2,5-dihydroxybenzoate within 8 h, Fig. S3). It should be mentioned here that the catalytic performance of the nicotinamide cofactor under the reaction conditions chosen here still is far from economic feasibility. However, we are convinced that after further optimization and upscaling, high total turnover numbers for the nicotinamide cofactors can be achieved.

Conclusions

The applicability of the ‘smart cosubstrate’ approach was demonstrated on a range of NAD(P)H-dependent redox reactions ranging from enantioselective reduction of carbonyl- or C[double bond, length as m-dash]C-bonds to specific aromatic hydroxylation. 1,4-BD can be oxidised twice forming a thermodynamically and kinetically stable lactone coproduct. Especially compared to the ‘traditional’ substrate-coupled regeneration approach, the smart cosubstrate concept excels by a significant reduction in cosubstrate needed to achieve quantitative conversion, thereby significantly reducing waste product formation. Furthermore, the coproduct (γ-butyrolactone) serves as an activated precursor for the synthesis of biodegradable polyesters.16

Of course a broad range of NAD(P)H regeneration systems is known to promote enoate reductase- and monooxygenase-reactions (Table 2).1,2,17–22 Being intrinsically favourable reactions, the equilibrium issue here, if existing at all, is less pronounced. Nevertheless, apart from the electrochemical methods and hydrogenations, the ‘smart cosubstrate’ approach ranges amongst the least waste-intensive methods shown in Table 2.

Table 2 Comparison of the waste generated by various NAD(P)H regeneration systems
Cosubstrate Coproduct Catalyst(s) Wastea [g mol−1 product] Ref.
Hase: hydrogenase; Rh:[Cp*Rh(bpy)(H2O)]2+; FDH: formate dehydrogenase; ADH: alcohol dehydrogenase; AldDH: aldehyde dehydrogenase; PDH: phosphite dehydrogenase; GDH: glucose dehydrogenase.a Theoretical value according to the reaction equation.b Depending on the way the electrical current was generated.c Double oxidation of the cosubstrate.d A phosphite buffer is transformed into a phosphate buffer.e This study.
Cathode Hase 0b 17
Rh
H2 Hase 0 18
HCO2H CO2 FDH 44 19
Rh
iProp Acetone ADH 58 4
EtOH Acetaldehyde ADH 44
EtOH Acetic acid ADH/AldDH 30c 20
H3PO3 H3PO4 PDH 98d 21
Rh
Glucose Gluconic acid GDH 196 22
1,4-BD GBL HLADH 43c


We believe that the ‘smart cosubstrate’ approach is a robust and versatile concept to promote a broad range of NAD(P)H-dependent reactions. Ongoing research in our laboratories focuses on broadening the ‘smart cosubstrate’ scope (e.g. yielding enantiopure lactone products) and optimisation and application of this approach.

Experimental

HLADH, isoform E, recombinantly expressed in Escherichia coli is commercially available from evocatal GmbH (Düsseldorf, Germany). TsER was produced according to a literature procedure11d by recombinant expression in E. coli followed by heat-purification. 3HB6H was produced by recombinant expression in E. coli following a literature procedure.14 All chemicals were purchased from Sigma Aldrich (Zwijndrecht, The Netherlands) in the highest purity available and used as received. A detailed description of the experimental procedures as well as the analytical protocols can be obtained from the ESI.

The authors thank Deutsche Bundesstiftung Umwelt (DBU) for financial support of the project (AZ 13261). Serena Gargiulo, Remco van Oosten and Maarten Gorseling are acknowledged for fruitful discussions and technical support.

Notes and references

  1. (a) K. Faber, Biotransformations in Organic Chemistry, Springer-Verlag, Berlin, Heidelberg, 2011 Search PubMed; (b) K. Drauz, H. Groeger and O. May, Enzyme Catalysis in Organic Synthesis, Wiley-VCH, Weinheim, 2012 Search PubMed; (c) F. Hollmann, I. W. C. E. Arends and D. Holtmann, Green Chem., 2011, 13, 2285–2313 RSC; (d) F. Hollmann, I. W. C. E. Arends, K. Buehler, A. Schallmey and B. Bühler, Green Chem., 2011, 13, 226–265 RSC; (e) U. Kragl, M. Mueller, M. Wolberg, T. Schubert and W. Hummel, in Technology Transfer in Biotechnology, Springer-Verlag, Berlin Heidelberg, 2005, vol. 92, p. 261 Search PubMed; (f) S. Wenda, S. Illner, A. Mell and U. Kragl, Green Chem., 2011, 13, 3007–3047 RSC.
  2. (a) A. Weckbecker, H. Groger and W. Hummel, Adv. Biochem. Eng./Biotechnol., 2010, 120, 195–242 CrossRef CAS; (b) W. Kroutil, H. Mang, K. Edegger and K. Faber, Curr. Opin. Chem. Biol., 2004, 8, 120–126 CrossRef CAS; (c) F. Hollmann, I. W. C. E. Arends and K. Buehler, ChemCatChem, 2010, 2, 762–782 CrossRef CAS; (d) R. Wichmann and D. Vasic-Racki, in Technology Transfer in Biotechnology, Springer-Verlag, Berlin, Heidelberg, 2005, p. 225 Search PubMed.
  3. H. Chenault and G. Whitesides, Appl. Biochem. Biotechnol., 1987, 14, 147–197 CrossRef CAS.
  4. (a) M. Wolberg, M. V. Filho, S. Bode, P. Geilenkirchen, R. Feldmann, A. Liese, W. Hummel and M. Muller, Bioprocess Biosyst. Eng., 2008, 31, 183–191 CrossRef CAS; (b) M. Villela Filho, T. Stillger, M. Muller, A. Liese and C. Wandrey, Angew. Chem., Int. Ed., 2003, 42, 2993–2996 CrossRef CAS; (c) T. Daussmann, T. C. Rosen and P. Dünkelmann, Eng. Life Sci., 2006, 6, 125–129 CrossRef CAS; (d) K. Baer, M. Krausser, E. Burda, W. Hummel, A. Berkessel and H. Groger, Angew. Chem., Int. Ed., 2009, 48, 9355–9358 CrossRef CAS; (e) T. Schubert, W. Hummel and M. Mueller, Angew. Chem., Int. Ed., 2002, 41, 634–635 CrossRef CAS; (f) M. Wolberg, W. Hummel, C. Wandrey and M. Mueller, Angew. Chem., Int. Ed., 2000, 39, 4306–4308 CrossRef CAS; (g) W. Stampfer, K. Edegger, B. Kosjek, K. Faber and W. Kroutil, Adv. Synth. Catal., 2004, 346, 57–62 CrossRef CAS; (h) D. de Gonzalo, I. Lavandera, K. Faber and W. Kroutil, Org. Lett., 2007, 9, 2163–2166 CrossRef; (i) M. Amidjojo and D. Weuster-Botz, Tetrahedron: Asymmetry, 2005, 16, 899–901 CrossRef CAS; (j) K. Schroer, E. Tacha, S. Bringer-Meyer, W. Hummel, T. Daussmann, R. Pfaller and S. Lutz, J. Biotechnol., 2007, 131, 95–96 CrossRef; (k) A. Jakoblinnert, R. Mladenov, A. Paul, F. Sibilla, U. Schwaneberg, M. B. Ansorge-Schumacher and P. D. de Maria, Chem. Commun., 2011, 47, 12230–12232 RSC; (l) L. J. Wang, C. X. Li, Y. Ni, J. Zhang, X. Liu and J. H. Xu, Bioresour. Technol., 2011, 102, 7023–7028 CrossRef CAS; (m) J. Liang, E. Mundorff, R. Voladri, S. Jenne, L. Gilson, A. Conway, A. Krebber, J. Wong, G. Huisman, S. Truesdell and L. James, Org. Process Res. Dev., 2010, 14, 188–192 CrossRef CAS; (n) J. Liang, J. Lalonde, B. Borup, V. Mitchell, E. Mundorff, N. Trinh, D. A. Kochrekar, R. N. Cherat and G. G. Pai, Org. Process Res. Dev., 2010, 14, 193–198 CrossRef CAS.
  5. (a) C. F. Degraauw, J. A. Peters, H. van Bekkum and J. Huskens, Synthesis, 1994, 1007–1017 CrossRef CAS; (b) H. Meerwein and R. Schmidt, Justus Liebigs Ann. Chem., 1925, 444, 221–238 CrossRef CAS; (c) W. Ponndorf, Angew. Chem., 1926, 39, 138–143 CrossRef CAS; (d) M. U. Raja, R. Ramesh and K. H. Ahn, Tetrahedron Lett., 2009, 50, 7014–7017 CrossRef CAS.
  6. I. Lavandera, A. Kern, V. Resch, B. Ferreira-Silva, A. Glieder, W. M. Fabian, S. de Wildeman and W. Kroutil, Org. Lett., 2008, 10, 2155–2158 CrossRef CAS.
  7. (a) A. J. Irwin and J. B. Jones, J. Am. Chem. Soc., 1977, 99, 556–561 CrossRef CAS; (b) A. J. Irwin and J. B. Jones, J. Am. Chem. Soc., 1977, 99, 1625–1630 CrossRef CAS; (c) A. J. Irwin, K. P. Lok, K. W.-C. Huang and J. B. Jones, J. Chem. Soc., Perkin Trans. 1, 1978, 1636–1642 RSC; (d) K. P. Lok, I. J. Jakovac and J. B. Jones, J. Am. Chem. Soc., 1985, 107, 2521–2526 CrossRef CAS; (e) I. J. Jakovac, H. B. Goodbrand, K. P. Lok and J. B. Jones, J. Am. Chem. Soc., 1982, 104, 4659–4665 CrossRef CAS; (f) A. J. Bridges, P. S. Raman, G. S. Y. Ng and J. B. Jones, J. Am. Chem. Soc., 1984, 106, 1461–1467 CrossRef CAS; (g) S. Gargiulo, I. W. C. E. Arends and F. Hollmann, ChemCatChem, 2011, 3, 338–342 CrossRef CAS; (h) F. Boratyński, G. Kiełbowicz and C. Wawrzeńczyk, J. Mol. Catal. B: Enzym., 2010, 65, 30–36 CrossRef; (i) G. Hilt, B. Lewall, G. Montero, J. H. P. Utley and E. Steckhan, Liebigs Ann./Rec, 1997, 2289–2296 CAS; (j) I. Schröder, E. Steckhan and A. Liese, J. Electroanal. Chem., 2003, 541, 109–115 CrossRef.
  8. (a) P. Galletti, E. Emer, G. Gucciardo, A. Quintavalla, M. Pori and D. Giacomini, Org. Biomol. Chem., 2010, 8, 4117–4123 RSC; (b) D. Giacomini, P. Galletti, A. Quintavalla, G. Gucciardo and F. Paradisi, Chem. Commun., 2007, 4038–4040 RSC; (c) T. Kawamoto, A. Aoki, K. Sonomoto and A. Tanaka, J. Ferment. Bioeng., 1989, 67, 361–362 CrossRef CAS.
  9. M. L. Mavrovouniotis, Biotechnol. Bioeng., 1990, 36, 1070–1082 CrossRef CAS.
  10. P. Könst, H. Merkens, S. Kara, S. Kochius, A. Vogel, R. Zuhse, D. Holtmann, I. W. Arends and F. Hollmann, Angew. Chem., Int. Ed., 2012, 51, 9914–9917 CrossRef.
  11. (a) S. Gargiulo, D. J. Opperman, U. Hanefeld, I. W. Arends and F. Hollmann, Chem. Commun., 2012, 48, 6630–6632 RSC; (b) J. Bernard, E. van Heerden, I. W. C. E. Arends, D. J. Opperman and F. Hollmann, ChemCatChem, 2012, 4, 196–199 CrossRef CAS; (c) D. J. Opperman, B. T. Sewell, D. Litthauer, M. N. Isupov, J. A. Littlechild and E. van Heerden, Biochem. Biophys. Res. Commun., 2010, 393, 426–431 CrossRef CAS; (d) D. J. Opperman, L. A. Piater and E. van Heerden, J. Bacteriol., 2008, 190, 3076–3082 CrossRef CAS.
  12. A. Fryszkowska, H. Toogood, M. Sakuma, J. M. Gardiner, G. M. Stephens and N. S. Scrutton, Adv. Synth. Catal., 2009, 351, 2976–2990 CrossRef CAS.
  13. K. Tauber, M. Hall, W. Kroutil, W. M. F. Fabian, K. Faber and S. M. Glueck, Biotechnol. Bioeng., 2011, 108, 1462–1467 CrossRef CAS.
  14. S. Montersino and W. J. van Berkel, Biochim. Biophys. Acta, 2012, 1824, 433–442 CrossRef CAS.
  15. A. Chang, M. Scheer, A. Grote, I. Schomburg and D. Schomburg, Nucleic Acids Res., 2009, 37, 588–592 CrossRef.
  16. T. Moore, R. Adhikari and P. Gunatillake, Biomaterials, 2005, 26, 3771–3782 CrossRef CAS.
  17. (a) F. Hollmann, A. Schmid and E. Steckhan, Angew. Chem., Int. Ed., 2001, 40, 169–171 CrossRef CAS; (b) K. Delecouls-Servat, A. Bergel and R. Basseguy, Bioprocess Biosyst. Eng., 2004, 26, 205–215 CrossRef CAS; (c) F. Hildebrand and S. Lütz, Tetrahedron: Asymmetry, 2007, 18, 1187–1193 CrossRef CAS; (d) R. Ruppert, S. Herrmann and E. Steckhan, Tetrahedron Lett., 1987, 28, 6583–6586 CrossRef CAS; (e) J. Cantet, A. Bergel and M. Comtat, Bioelectrochem. Bioenerg., 1992, 27, 475–486 CrossRef CAS.
  18. (a) J. Ratzka, L. Lauterbach, O. Lenz and M. B. Ansorge-Schumacher, Biocatal. Biotransform., 2011, 29, 246–252 CAS; (b) R. Mertens, L. Greiner, E. C. D. van den Ban, H. Haaker and A. Liese, J. Mol. Catal. B: Enzym., 2003, 24–25, 39–52 CrossRef CAS; (c) Y. Ni, P.-L. Hagedoorn, J.-H. Xu, I. W. C. E. Arends and F. Hollmann, Chem. Commun., 2012, 48, 12056–12058 RSC.
  19. (a) K. Hofstetter, J. Lutz, I. Lang, B. Witholt and A. Schmid, Angew. Chem., Int. Ed., 2004, 43, 2163–2166 CrossRef CAS; (b) C. N. Jensen, J. Cartwright, J. Ward, S. Hart, J. P. Turkenburg, S. T. Ali, M. J. Allen and G. Grogan, ChemBioChem, 2012, 13, 872–878 CrossRef CAS; (c) K. Kuehnel, S. C. Maurer, Y. Galeyeva, W. Frey, S. Laschat and V. B. Urlacher, Adv. Synth. Catal., 2007, 349, 1451–1461 CrossRef CAS; (d) K. Durchschein, S. Wallner, P. Macheroux, W. Schwab, T. Winkler, W. Kreis and K. Faber, Eur. J. Org. Chem., 2012, 2012, 4963–4968 CrossRef CAS; (e) M. Hall, C. Stückler, B. Hauer, R. Stürmer, T. Friedrich, M. Breuer, W. Kroutil and K. Faber, Eur. J. Org. Chem., 2008, 1511–1516 CrossRef CAS; (f) M. Hall, C. Stückler, H. Ehammer, E. Pointner, G. Oberdorfer, K. Gruber, B. Hauer, R. Stürmer, W. Kroutil, P. Macheroux and K. Faber, Adv. Synth. Catal., 2008, 350, 411–418 CrossRef CAS; (g) M. Hall, C. Stueckler, W. Kroutil, P. Macheroux and K. Faber, Angew. Chem., Int. Ed., 2007, 46, 3934–3937 CrossRef CAS; (h) V. Köhler, Y. M. Wilson, M. Durrenberger, D. Ghislieri, E. Churakova, T. Quinto, L. Knorr, D. Haussinger, F. Hollmann, N. J. Turner and T. R. Ward, Nat. Chem., 2013 DOI:10.1038/nchem.1498; (i) D. Westerhausen, S. Herrmann, W. Hummel and E. Steckhan, Angew. Chem., Int. Ed., 1992, 31, 1529–1531 CrossRef; (j) R. Ruppert, S. Herrmann and E. Steckhan, J. Chem. Soc., Chem. Commun., 1988, 1150–1151 RSC.
  20. S. Broussy, R. W. Cheloha and D. B. Berkowitz, Org. Lett., 2009, 11, 305–308 CrossRef CAS.
  21. (a) B. Kosjek, F. J. Fleitz, P. G. Dormer, J. T. Kuethe and P. N. Devine, Tetrahedron: Asymmetry, 2008, 19, 1403–1406 CrossRef CAS; (b) C. Rodriguez, G. de Gonzalo and V. Gotor, J. Mol. Catal. B: Enzym., 2011, 74, 138–143 CrossRef; (c) D. E. Torres Pazmiño, A. Riebel, J. D. Lange, F. Rudroff, M. D. Mihovilovic and M. W. Fraaije, ChemBioChem, 2009, 10, 2595–2598 CrossRef; (d) D. E. Torres Pazmiño, R. Snajdrova, B.-J. Baas, M. Ghobrial, M. D. Mihovilovic and M. W. Fraaije, Angew. Chem., Int. Ed., 2008, 47, 2307–2310 CrossRef.
  22. (a) J. D. Zhang, A. T. Li, H. L. Yu, T. Imanaka and J. H. Xu, J. Ind. Microbiol. Biotechnol., 2011, 38, 633–641 CrossRef CAS; (b) Y. Lu and L. H. Mei, J. Ind. Microbiol. Biotechnol., 2007, 34, 247–253 CrossRef CAS; (c) H. J. Park, J. Jung, H. Choi, K. N. Uhm and H. K. Kim, J. Microbiol. Biotechnol., 2010, 20, 1300–1306 CrossRef CAS; (d) J. F. Chaparro-Riggers, T. A. Rogers, E. Vazquez-Figueroa, K. M. Polizzi and A. S. Bommarius, Adv. Synth. Catal., 2007, 349, 1521–1531 CrossRef CAS; (e) H. S. Toogood, J. M. Gardiner and N. S. Scrutton, ChemCatChem, 2010, 2, 892–914 CrossRef CAS; (f) D. J. Bougioukou, A. Z. Walton and J. D. Stewart, Chem. Commun., 2010, 46, 8558–8560 RSC.

Footnote

Electronic supplementary information (ESI) available: Detailed description of the experimental and analytical procedures and further experimental data. See DOI: 10.1039/c2gc36797a

This journal is © The Royal Society of Chemistry 2013