Synthesis of α,β-unsaturated esters via a chemo-enzymatic chain elongation approach by combining carboxylic acid reduction and Wittig reaction

  1. ,
  2. ,
  3. ,
  4. ,
  5. and
National Engineering Laboratory for Industrial Enzymes and Tianjin Engineering Center for Biocatalytic Technology, Tianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences, Tianjin 300308, P. R. China
  1. Corresponding author email
Guest Editor: N. Turner
Beilstein J. Org. Chem. 2015, 11, 2245–2251. https://doi.org/10.3762/bjoc.11.243
Received 20 Aug 2015, Accepted 28 Oct 2015, Published 19 Nov 2015
Full Research Paper
cc by logo

Abstract

α,β-Unsaturated esters are versatile building blocks for organic synthesis and of significant importance for industrial applications. A great variety of synthetic methods have been developed, and quite a number of them use aldehydes as precursors. Herein we report a chemo-enzymatic chain elongation approach to access α,β-unsaturated esters by combining an enzymatic carboxylic acid reduction and Wittig reaction. Recently, we have found that Mycobacterium sp. was able to reduce phenylacetic acid (1a) to 2-phenyl-1-ethanol (1c) and two sequences in the Mycobacterium sp. genome had high identity with the carboxylic acid reductase (CAR) gene from Nocardia iowensis. These two putative CAR genes were cloned, overexpressed in E. coli and one of two proteins could reduce 1a. The recombinant CAR was purified and characterized. The enzyme exhibited high activity toward a variety of aromatic and aliphatic carboxylic acids, including ibuprofen. The Mycobacterium CAR catalyzed carboxylic acid reduction to give aldehydes, followed by a Wittig reaction to afford the products α,β-unsaturated esters with extension of two carbon atoms, demonstrating a new chemo-enzymatic method for the synthesis of these important compounds.

Introduction

α,β-Unsaturated esters are versatile building blocks for organic synthesis and of significant importance for industrial applications [1-13]. A great variety of synthetic methods have been developed to access α,β-unsaturated esters [14-24]. One popular approach is the Wittig reaction which produces α,β-unsaturated esters with two more carbon atoms [25]. While fatty acids are abundant from natural resources, aromatic carboxylic acids could be prepared by the degradation of lignin, an unused and abundant component of biomass, although the effective methods for the degradation of lignin need to be developed. These carboxylic acids could be reduced to their corresponding aldehydes, which can be used as starting materials for Wittig reaction. The combination of carboxylic acid reduction and Wittig reaction would offer a new approach for the production of bio-based α,β-unsaturated esters. However, the conventional chemical methods for the carboxylic acid reduction require strong reducing reagents such as metal hydrides, posting operational danger and low selectivity. In addition, the conversion of COOH into CHO requires particular hydride reagents, to avoid the further reduction to primary alcohols [26]. On the contrary, enzymatic reduction of carboxylic acids proceeds under mild reaction conditions with high selectivity and tolerance of other functional groups [27,28]. Currently, only a few members of this interesting type of enzymes (CAR, E.C.1.2.1.30) have been biochemically characterized (Table 1). They possess similar consensus sequence characteristics and reaction mechanism (post-translational phosphopantetheinylation, ATP, Mg2+, and NADPH as cofactors) [27-30]. As such, we initiated the search for new carboxylic acid reductases and the exploration of their potential as biocatalysts for the efficient bioreduction of carboxylic acids. Herein we report a new CAR from Mycobacterium sp. (Mycobacterium CAR) and its application in a chemo-enzymatic chain elongation method for the preparation of α,β-unsaturated esters by combining an enzymatic carboxylic acid reduction and Wittig reaction.

Table 1: Identified CARs (EC 1.2.1.30).

Identified CAR Accession number Origin References
NiCARa,b AAR91681.1 Nocardia iowensis [30,31]
MsCAR WP_011855500.1 Mycobacterium sp. JLS [32]
SgCAR WP_012382217.1 Streptomyces griseus subsp. griseus NBRC 13350 [32]
MmCARa WP_012393886.1 Mycobacterium marinum M [33]
SrCARa WP_013138593.1 Segniliparus rotundus DSM 44985 [28]
Mycobacterium CARa WP_019510583.1 Mycobacterium sp. This work

aThe protein was biochemically characterized and determined as a monomeric protein. bThe protein was also purified from natural strain.

Results and Discussion

Nineteen actinomycete strains available in our laboratory were screened for the carboxylic acid reductase activity using phenylacetic acid (1a) as the substrate by GC analysis of the products. Among them, Mycobacterium sp. was found to catalyse the reduction of 1a to 2-phenyl-1-ethanol (1c) with low conversion (Supporting Information File 1, Figure S1). The genome of Mycobacterium sp. strain was sequenced, and two gene sequences having 58% and 47% protein sequence identity with the NiCAR [30] (accession number AAR91681.1) were found by local BLAST search (tblastn) (Supporting Information File 1, Figure S2). These two genes were also found to have identical sequences with those (Gene ID 17912504 and Gene ID 17917114), respectively, in the genomic sequence of Mycobacterium neoaurum VKM Ac-1815D (accession no. CP006936) [34] by NCBI BLAST search (http://www.ncbi.nlm.nih.gov/BLAST/).

Since PPTase from Mycobacterium sp. has not been identified, a known Nocardia PPTase (accession number ABI83656.1) was selected for the post-translational phosphopantetheinylation in the current study [27,28]. The gene (Gene ID 17912504) was cloned into pET30b(+) and expressed in E. coli. The recombinant enzyme (Mycobacterium CAR) showed carboxylic acid reductase activity toward 1a. The other gene (Gene ID 17917114) was cloned into pET28a(+) and expressed in E. coli. The gene was well expressed, but less fraction of soluble protein was obtained. The recombinant enzyme showed no carboxylic acid reductase activity toward 1a when it was tested under same conditions as the gene 17912504 after having been treated with the PPTase enzyme.

The His-tagged Mycobacterium CAR and His-PPTase were produced as soluble protein as described in Supporting Information File 1 and purified in one chromatographic step each using a HisTrapTM FF crude column (Supporting Information File 1, Figure S3). The molecular mass of His-CAR was estimated to be about 133 kDa by gel filtration chromatography. Since its theoretical value is 125 kDa, this enzyme is a monomeric protein. This is consistent with NiCAR [31] and SrCAR [28]. The carboxylic acid reductase activity was hardly detected when apo-CAR was used in the reaction system (1a as substrate) or the reaction system did not contain Mg2+ (Supporting Information File 1, Table S2), indicating that post-translational phosphopantetheinylation and Mg2+ were necessary for this enzymatic reduction, similar to the observations for the reaction with the carboxylic acid reductases from Nocardia [29,35] and Segniliparus [28]. The optimal pH and temperature for the enzymatic reduction of 1a with Mycobacterium CAR were pH 9 and 25 °C, respectively. The apparent Km and catalytic efficiencies (kcat/Km) of Mycobacterium CAR toward benzoic acid (2a) (Supporting Information File 1, Table S1) were 1.75 ± 0.16 mM and 0.93 mM−1·s−1, respectively. They were lower than those for NiCAR [31], SrCAR [28] and MnCAR [33].

In order to explore the application potential of Mycobacterium CAR, its substrate specificity was examined with the purified enzyme. The results in Table 2 showed that a series of aromatic and aliphatic carboxylic acids were reduced to their corresponding aldehydes. For the aliphatic acids, the reduction of nonanoic acid (3a) and lauric acid (4a) resulted in higher yields than those of hexanoic acid (5a) and pentadecanoic acid (6a), indicating that Mycobacterium CAR prefers the mid-chain fatty acids over the short-chain and long-chain ones. Mycobacterium CAR also showed activity toward different carboxylic acids with aromatic ring at the end carbon atom. For example, 1a (75% yield), 2a (94% yield), 3-phenylpropanoic acid (7a, 48% yield), and 4-phenylbutyric acid (8a) (64% yield) were reduced to the corresponding aldehydes. In contrast to ortho- and meta-methylbenzoic acid (9a and 10a), ortho-hydroxylbenzoic acid (11a) was obviously the poorer substrate than meta-counterpart (12a), suggesting that in this case the activity of Mycobacterium CAR might be strongly influenced by the electronic properties other than the steric factor of the ortho-group on the benzene ring. This enzyme was less active toward 2-methylhexanoic acid (13a) and 2-phenylpropionic acid (14a) than 5a and 1a, respectively. This may be due to the sterically hindered effect of the α-methyl substitution, which is consistent with the results for Pyrococcus furiosus (whole-cell) [36] and NiCAR [31]. However, no significant difference was noted between 3-phenylbutyric acid (15a) and 7a. Mycobacterium CAR showed good chemoselectivity for the reduction of some carboxylic acids containing C=C or C=O double bonds, such as linoleic acid (16a), cinnamic acid (17a), ferulic acid (18a), and 3-benzoylpropionic acid (19a). These acids were reduced, with C=C or C=O double bonds remaining unaffected. However, this enzyme was less active toward 17a than its saturated counterpart (7a). Interestingly, Mycobacterium CAR could effectively reduce ibuprofen (20a) more effectively than NiCAR and SrCAR, but it was not enantioselective. Rac-13a, rac-14a and rac-15a (5 mM) were almost completely transformed by Mycobacterium CAR (100 µg), suggesting it had no or low enantioselectivity toward these compounds.

Table 2: Substrate specificity of Mycobacterium CARa.

[Graphic 1]
Substrate Analytic yield (%) of aldehyde
[Graphic 2] 1a 75
[Graphic 3] 2a 94
[Graphic 4] 3a (n = 6) 100
4a (n = 9) 100
5a (n = 3) 68
6a (n = 12) 3
[Graphic 5] 7a 48
[Graphic 6] 8a 64
[Graphic 7] 9a (ortho) 57
10a (meta) 62
[Graphic 8] 11a (ortho) 7
12a (meta) 41
[Graphic 9] 13a 62; 100b
[Graphic 10] 14a (R/S) 47; 100b
[Graphic 11] 15a 44; 95b
[Graphic 12] 16ac 14
[Graphic 13] 17a 36
[Graphic 14] 18ad 40
[Graphic 15] 19a 68
[Graphic 16] 20a (R/S) 98
[Graphic 17] 21a 76
[Graphic 18] 22ac, d 18

aReaction conditions: Tris-HCl buffer (1 mL, 100 mM, pH 9) contained NADP+ (0.9 mM), GDH (1 U), glucose (60 mM), MgCl2 (10 mM), ATP (15 mM), substrate concentration (10 mM) and enzyme mixture (holo-CAR, 50 ug), 16 h, 25 °C, 200 rpm. bSame as a, but 5 mM of substrate and 100 μg of Mycobacterium CAR were used. cSame as a, but the reaction was performed in sodium phosphate buffer (100 mM, pH 7.5). dSilylation was performed before the GC analysis.

The Mycobacterium CAR-catalysed carboxylic acid reduction was combined with Wittig reaction to establish a new chemo-enzymatic approach to the synthesis of α,β-unsaturated esters. As described in the Experimental section, the holo-CAR enzyme mixture was prepared and mixed with NADP+, GDH, glucose, ATP and carboxylic acid in Tris-HCl buffer. The reaction mixture was incubated at 25 °C for 16 h and extracted with ethyl acetate. The organic extract was concentrated to about 20 mL, and mixed with ethyl (triphenylphosphoranylidene)acetate and Na2CO3. After 24 h at room temperature, the organic solvent was removed under reduced pressure and the residue was purified by silica gel column chromatography to give the product (α,β-unsaturated ester). As shown in Table 3, α,β-unsaturated esters were obtained in moderate to high yields with trans-isomer as the major product. The corresponding α,β-unsaturated esters from aromatic carboxylic acids (1a, 2a, 10a, 12a and 17a) had higher yields than those from aliphatic ones (3a, 4a and 5a), and this might be due to lower yields of aliphatic aldehydes and higher loss in the product separation.

Table 3: Synthesis of α,β-unsaturated esters via enzymatic reduction and Wittig reactiona.

[Graphic 19]
Substrate (a) Analytic yield (%) of aldehydes (b)b Isolated yield (%) of α,β-unsaturated esters (d) Ratio E/Zc
1a 72 60 68:32
2a 100 70 92:8
3a 65 41 87:13
4a 70 46 89:11
5a 67 38 96:4
10a 100 81 92:8
12a 79 59 90:10
17a 78 65 94:6

aThe carboxylic acid was first reduced to aldehyde, after being extracted with ethyl acetate, ethyl (triphenylphosphoranylidene)acetate was added for the Wittig reaction. bDetermined by GC analysis of the reaction mixture. cDetermined by GC analysis of isolated products.

Conclusion

A new CAR from Mycobacterium sp. was successfully cloned, overexpressed and identified. It exhibited a broad substrate spectrum and was active toward both aliphatic and aromatic carboxylic acids, including ibuprofen. Other functional groups such as keto groups and C=C double bonds remained unaffected. Mycobacterium CAR catalysed carboxylic acid reduction to give aldehydes, followed by a Wittig reaction to afford α,β-unsaturated esters with extension of two carbon atoms. This study demonstrates a new chemo-enzymatic chain elongation method for the synthesis of these important compounds from bio-based fatty and aromatic acids of natural resources. However, the enzymatic reduction of carboxylic acids requires CoA, ATP and NADPH, and this still presents challenge for its application at large scale, which may be overcome by using the whole cell catalyst of the engineered enzyme production strain with efficient amount of CoA, ATP and NADPH or effective regeneration systems of them.

Experimental

Cloning of Mycobacterium CAR gene

Mycobacterium sp. chromosomal DNA (gDNA) was extracted and purified using a TIANamp Bacteria DNA Kit. The Mycobacterium CAR gene (Gene ID 17912504) was amplified by PCR using Mycobacterium gDNA as template and primers containing the restriction sites NdeI and XhoI, respectively, CAR-F 5′-CATGCATATGTTCGCCGAAAATCTTGATGACCAG-3′ and CAR-R 5′-CATCTCGA GCAGCAGGCCGAGCAATTGCAGGT-3′. The PCR fragment was purified and then ligated with cloning vector pJET1.2/blunt, which was confirmed by DNA sequencing. A CAR DNA fragment was acquired from the vector pJET1.2-CAR by digesting at the restriction sites NdeI and XhoI, and then ligated by T4 DNA ligase into pET30b(+) at the same restriction sites to generate the expression vector pET30b(+)-CAR. The confirmed recombinant vector was transformed into E. coli BL21(DE3).

Expression and purification of Mycobacterium CAR and Nocardia PPTase

A culture of E. coli BL21 (DE3) cells harboring pET30b(+)-CAR or pET32a(+)-PPTase was grown overnight in LB-ampicillin (100 µg/mL) medium (5 mL) at 37 °C, and then inoculated into 1 L of LB-ampicillin (100 µg/mL) medium. The resulting culture was incubated continually at 200 rpm in a rotary shaker at 37 °C until cells reached mid-log growth (OD600 of 0.5–1.0), which was followed by the addition of 0.5 mM IPTG and further incubation for 12 h at 25 °C. Cells were harvested by centrifugation at 12 000g for 10 min at 4 °C, and disrupted by high pressure homogenizer after re-suspension in binding buffer (20 mM sodium phosphate buffer, 0.5 M NaCl, 20 mM imidazole, pH 7.4). His-CAR or His-PPTase protein in the supernatant fraction was collected from the crude cell lysate by centrifugation at 12 000g for 20 min. Protein purification was performed on a HisTrapTM FF crude column (GE Healthcare, Piscataway, USA), and the protein was desorbed with an elution buffer (20 mM sodium phosphate, 0.5 M NaCl, 0.5 M imidazole, pH 7.4). The purified proteins His-CAR or His-PPTase were dialyzed in a sodium phosphate buffer (50 mM, pH 7.5) and then stored at −20 °C for further use.

Standard reduction procedure

The His-CAR (1.3 mg) was incubated with His-PPTase (256 µg) in the presence of CoA (1 mM) as a cofactor for 1 h at 28 °C in a final volume of 520 µL of sodium phosphate buffer (100 mM, pH 7.5) containing 10 mM of MgCl2. The resulting enzyme mixture (holo-CAR, 50 or 100 µg) was mixed with NADP+ (0.9 mM), GDH (1 U, one unit corresponds to the amout of enzyme which could convert 1 µmol NADP+ to NADPH per minute using D-glucose as the substrate), glucose (60 mM), MgCl2 (10 mM), carboxylic acid (5 or 10 mM, from 1 M stock solution in DMSO), and ATP (15 mM) in Tris-HCl buffer (100 mM, pH 9) with a final volume of 1 mL. The reaction mixture was incubated at 200 rpm in a rotary shaker at 25 °C for 16 h, and extracted with 1 mL of ethyl acetate after the pH was adjusted to 2–3 with 1 M HCl solution. The organic extracts were dried over anhydrous sodium sulfate and analysed by gas chromatography (GC) to determine the amount of substrate (a) and products (aldehyde b) in the mixture. All experiments were conducted in triplicate.

Substrate specificity

The reduction of a series of carboxylic acids was carried out by following the standard reduction procedure. The yields were determined by GC analysis.

Experimental procedures for the synthesis of compounds 1d, 2d, 3d, 4d, 5d, 10d, 12d and 17d

A typical procedure was as follows using ethyl 4-phenylbut-2-enoate (1d) as the example. The enzyme mixture (holo-CAR, 0.5 mg/mL) was prepared as described above, and was mixed with NADP+ (0.45 mM), GDH (1 U), glucose (60 mM), ATP (13 mM) and phenylacetic acid (1a, 10 mM) in Tris-HCl buffer (total volume 25 mL, 100 mM, pH 9). The reaction mixture was incubated at 100 rpm in a rotary shaker at 25 °C for 16 h, and extracted 3 times with 25 mL of ethyl acetate. The organic extract was concentrated to about 20 mL under reduced pressure, and then ethyl (triphenylphosphoranylidene)acetate (100 mM) and Na2CO3 (about 0.5 g) were added, and the reaction mixture was stirred for 24 h at room temperature. The organic solvent was removed under reduced pressure and the residue was purified by silica gel column chromatography to give the product, 1d (30.1 mg, 60%) was obtained.

Supporting Information

Supporting Information File 1: Materials, bacterial screening, analytical procedures, NMR data and spectra of 1d, 2d, 3d, 4d, 5d, 10d, 12d and 17d.
Format: PDF Size: 2.0 MB Download

Acknowledgements

This work was financially supported by National Basic Research Program of China (973 Program, No. 2011CB710800) and the Chinese Academy of Sciences (KSZD-EW-Z-015).

References

  1. Shibata, K.; Chatani, N. Org. Lett. 2014, 16, 5148–5151. doi:10.1021/ol502500c
    Return to citation in text: [1]
  2. Cheng, J.; Huang, Z.; Chi, Y. R. Angew. Chem., Int. Ed. 2013, 52, 8592–8596. doi:10.1002/anie.201303247
    Return to citation in text: [1]
  3. Wang, P.; Ling, L.; Liao, S.-H.; Zhu, J.-B.; Wang, S. R.; Li, Y.-X.; Tang, Y. Chem. – Eur. J. 2013, 19, 6766–6773. doi:10.1002/chem.201204182
    Return to citation in text: [1]
  4. Hashimoto, T.; Maruoka, K. Chem. Rev. 2015, 115, 5366–5412. doi:10.1021/cr5007182
    Return to citation in text: [1]
  5. Zhu, Y.; Wang, Q.; Cornwall, R. G.; Shi, Y. Chem. Rev. 2014, 114, 8199–8256. doi:10.1021/cr500064w
    Return to citation in text: [1]
  6. He, J.; Ling, J.; Chiu, P. Chem. Rev. 2014, 114, 8037–8128. doi:10.1021/cr400709j
    Return to citation in text: [1]
  7. Yasukawa, T.; Suzuki, A.; Miyamura, H.; Nishino, K.; Kobayashi, S. J. Am. Chem. Soc. 2015, 137, 6616–6623. doi:10.1021/jacs.5b02213
    Return to citation in text: [1]
  8. Weber, F.; Brueckner, R. Eur. J. Org. Chem. 2015, 2428–2449. doi:10.1002/ejoc.201403622
    Return to citation in text: [1]
  9. Peacock, L. R.; Chapman, R. S. L.; Sedgwick, A. C.; Mahon, M. F.; Amans, D.; Bull, S. D. Org. Lett. 2015, 17, 994–997. doi:10.1021/acs.orglett.5b00103
    Return to citation in text: [1]
  10. Niu, Z.; Chen, J.; Chen, Z.; Ma, M.; Song, C.; Ma, Y. J. Org. Chem. 2015, 80, 602–608. doi:10.1021/jo5021135
    Return to citation in text: [1]
  11. Parveen, S.; Hussain, S.; Qin, X.; Hao, X.; Zhu, S.; Rui, M.; Zhang, S.; Fu, F.; Ma, B.; Yu, Q.; Zhu, C. J. Org. Chem. 2014, 79, 4963–4972. doi:10.1021/jo500338c
    Return to citation in text: [1]
  12. Hatano, M.; Horibe, T.; Ishihara, K. Angew. Chem., Int. Ed. 2013, 52, 4549–4553. doi:10.1002/anie.201300938
    Return to citation in text: [1]
  13. Časar, Z.; Steinbücher, M.; Košmrlj, J. J. Org. Chem. 2010, 75, 6681–6684. doi:10.1021/jo101050z
    Return to citation in text: [1]
  14. Oliveira, M. E. R.; da Silva Filho, E. C.; Filho, J. M.; Ferreira, S. S.; Oliveira, A. C.; Campos, A. F. Chem. Eng. J. 2015, 263, 257–267. doi:10.1016/j.cej.2014.11.016
    Return to citation in text: [1]
  15. Nakagiri, T.; Murai, M.; Takai, K. Org. Lett. 2015, 17, 3346–3349. doi:10.1021/acs.orglett.5b01583
    Return to citation in text: [1]
  16. Liu, J.; Liu, Q.; Franke, R.; Jackstell, R.; Beller, M. J. Am. Chem. Soc. 2015, 137, 8556–8563. doi:10.1021/jacs.5b04052
    Return to citation in text: [1]
  17. El-Batta, A.; Jiang, C.; Zhao, W.; Anness, R.; Cooksy, A. L.; Bergdahl, M. J. Org. Chem. 2007, 72, 5244–5259. doi:10.1021/jo070665k
    Return to citation in text: [1]
  18. Chen, Y.; Romaire, J. P.; Newhouse, T. R. J. Am. Chem. Soc. 2015, 137, 5875–5878. doi:10.1021/jacs.5b02243
    Return to citation in text: [1]
  19. Kona, J. R.; King'ondu, C. K.; Howell, A. R.; Suib, S. L. ChemCatChem 2014, 6, 749–752. doi:10.1002/cctc.201300942
    Return to citation in text: [1]
  20. Shearouse, W. C.; Korte, C. M.; Mack, J. Green Chem. 2011, 13, 598–601. doi:10.1039/c0gc00671h
    Return to citation in text: [1]
  21. Zeitler, K. Org. Lett. 2006, 8, 637–640. doi:10.1021/ol052826h
    Return to citation in text: [1]
  22. Kantam, M. L.; Kumar, K. B. S.; Balasubramanyam, V.; Venkanna, G. T.; Figueras, F. J. Mol. Catal. A: Chem. 2010, 321, 10–14. doi:10.1016/j.molcata.2010.01.012
    Return to citation in text: [1]
  23. List, B.; Doehring, A.; Hechavarria Fonseca, M. T.; Job, A.; Rios Torres, R. Tetrahedron 2006, 62, 476–482. doi:10.1016/j.tet.2005.09.081
    Return to citation in text: [1]
  24. Barma, D. K.; Kundu, A.; Bandyopadhyay, A.; Kundu, A.; Sangras, B.; Briot, A.; Mioskowski, C.; Falck, J. R. Tetrahedron Lett. 2004, 45, 5917–5920. doi:10.1016/j.tetlet.2004.05.113
    Return to citation in text: [1]
  25. Maryanoff, B. E.; Reitz, A. B. Chem. Rev. 1989, 89, 863–927. doi:10.1021/cr00094a007
    Return to citation in text: [1]
  26. Seyden-Penne, J. Reductions by the Alumino-and Borohydrides in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 1997.
    Return to citation in text: [1]
  27. Napora-Wijata, K.; Strohmeier, G. A.; Winkler, M. Biotechnol. J. 2014, 9, 822–843. doi:10.1002/biot.201400012
    Return to citation in text: [1] [2] [3]
  28. Duan, Y.; Yao, P.; Chen, X.; Liu, X.; Zhang, R.; Feng, J.; Wu, Q.; Zhu, D. J. Mol. Catal. B: Enzym. 2015, 115, 1–7. doi:10.1016/j.molcatb.2015.01.014
    Return to citation in text: [1] [2] [3] [4] [5] [6] [7]
  29. Venkitasubramanian, P.; Daniels, L.; Rosazza, J. P. N. J. Biol. Chem. 2007, 282, 478–485. doi:10.1074/jbc.M607980200
    Return to citation in text: [1] [2]
  30. He, A. M.; Li, T.; Daniels, L.; Fotheringham, I.; Rosazza, J. P. N. Appl. Environ. Microbiol. 2004, 70, 1874–1881. doi:10.1128/Aem.70.3.1874-1881.2004
    Return to citation in text: [1] [2] [3]
  31. Li, T.; Rosazza, J. P. J. Bacteriol. 1997, 179, 3482–3487.
    Return to citation in text: [1] [2] [3] [4]
  32. Behrouzian, B.; McDaniel, R.; Zhang, X.; Clark, L. Engineered Biosynthesis of Fatty Alcohols. WO Patent WO2,010,135,624, Nov 25, 2010.
    Return to citation in text: [1] [2]
  33. Akhtar, M. K.; Turner, N. J.; Jones, P. R. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 87–92. doi:10.1073/pnas.1216516110
    Return to citation in text: [1] [2]
  34. Shtratnikova, V. Y.; Bragin, E. Y.; Dovbnya, D. V.; Pekov, Y. A.; Schelkunov, M. I.; Strizhov, N.; Ivashina, T. V.; Ashapkin, V. V.; Donova, M. V. Genome Announce. 2014, 2, No. 1e01177-13. doi:10.1128/genomeA.01177-13
    Return to citation in text: [1]
  35. Li, T.; Rosazza, J. P. J. Biol. Chem. 1998, 273, 34230–34233. doi:10.1074/jbc.273.51.34230
    Return to citation in text: [1]
  36. van den Ban, E. C. D.; Willemen, H. M.; Wassink, H.; Laane, C.; Haaker, H. Enzyme Microb. Technol. 1999, 25, 251–257. doi:10.1016/s0141-0229(99)00036-8
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities