CHAPTER 2
Biocatalysis – A Greener Alternative in Synthetic Chemistry
Mylan Labs Ltd., India
*E-mail: Madhuresh.sethi@mylan.in
Biocatalysis can be defined as the synthesis of chemicals with the assistance of biological materials like enzymes and whole cells.1 The scope of biocatalysis includes fermentation, biotransformations using whole cells/enzymes, cloning and expression of enzymes, directed evolution of enzymes, substrate specificity, and stability. Biocatalysis, still described as a field of emerging science,2 is picking up substantial momentum nowadays. The most evident reason for its growth can be attributed to the increasing awareness of the depletion of the environment’s natural resources and the dire need to preserve them. This awareness has driven industries to realize their serious contributions to environmental pollution, and have made them adapt slowly to make their upcoming and existing processes greener and cleaner wherever possible.3 The snail-paced green revolution can only be catalyzed by implementing more stringent regulations from relevant authorities.
1It is but a paradox when one takes the example of the pharmaceutical industry. Its primary aim is to continue providing quality medicines to mankind to improve their health standards through extensive research. However, it is surprising how much hazardous waste it generates while producing the same healthcare products.4 These hazardous wastes make a major contribution to the forever increasing statistics of environmental pollution in the world. Even though the growth curve of biocatalysis is not a sharp one when it comes to its application in pharmaceuticals, the future seems to hold strong potential.
Biocatalysis gives an alternative route to the chemical pathway, in a more environment-friendly way. It is the cleanest and greenest method for chemical syntheses, and also offers a wider scope for selective production.5–10 In this area, with the advancement of pharmaceutical sciences, the focus has shifted towards production of stereospecific chiral molecules rather than their racemic mixtures. Herein lies the role of the biocatalysts, which are highly stereospecific and stereoselective and can be readily recycled. With the recent development of tools and technology in protein engineering and molecular biology, biocatalysts can be tailor-designed according to the requirements of different industrial processes.
In pharmaceutical, agricultural as well as biotechnological industries, chiral compounds are in high demand for the preparation of bulk drug substances and fine chemicals.11 This demand can be met with greater ease by chemoenzymatic synthesis than by employing classic chemical routes. That ease is because of the use of highly stereospecific enzymes. Furthermore, there is a considerable increase in the commercial availability of enzymes nowadays. This has encouraged synthetic organic chemists to effortlessly venture into the biocatalytic field and explore this option in manufacturing industries.2 Bioinformatics and cloning through gene banks have made the easy availability of different ranges of enzymes possible.12
‘Only those who will risk going too far can possibly find out how far one can go.’ T.S. Eliot
So let us take that extra step to adopt green chemistry in our industries to help mankind and the surrounding environment to travel far ahead in time, harmoniously walking hand in hand. Let us hope that our industries get a ‘green’ makeover much sooner rather than later. Incorporating eco-efficient strategies in pharmaceutical/chemical industries to re-define/re-visit some of the existing manufacturing methodologies is the need of the hour. The commitment towards a common goal, i.e. to keep our environment pollution-free, has to be advocated for, and the focus should not be diluted from this.
Enzymes are exquisitely selective catalysts, which target a single substrate from even a wide range of similar compounds. The specificity of enzymes is more due to the substrate rate of reaction rather than due to substrate binding affinity. Specificity of enzymes rests in the three‐dimensional structure of the active site of the enzyme, which is complementary to the transition reaction state. In some cases, a good substrate induces an active conformation that is not available to a poor substrate. Enzymes also contain another site called a ‘proofreading’ site that can increase selectivity. Specificity is also evident in the way that enzymes control the decomposition of unstable intermediates, restricting their conformation so that the reaction is channeled down one pathway, to yield a single product. Enzyme selectivity is often not absolute; optimization can be achieved by the evolution of new enzymes. Enzymes have a broad field of applications on the industrial scale.13 However, this chapter primarily focuses on the pharmaceutical industry, and Table 2.1 provides selected examples of chemical segments where enzymes have found application.
Table 2.1 Overview (not comprehensive) of the application of enzymes in various industries
Company | Company | Company |
---|---|---|
Food | Amylases | Hydrolysis of starch |
Proteases | Processing of cheese and meat | |
Pectinases | Clarification of juices | |
Lipases | Modification of fats | |
Glucose isomerases | Production of fructose | |
Feed | Hemicellulases | Digestibility of feed |
Cellulases | Increased nutritional value | |
Phytases | Improved phosphate uptake | |
Pharmaceutical | Penicillin acylases | Production of penicillin derivativesProduction of optically pure compounds through kinetic resolutions, dynamic kinetic resolution, desymmetrization of prochiral substrates, etc. |
Hydrolases: Lipases, proteases and amino-acylase | ||
Oxidoreductases | ||
Transaminases | ||
Lyases | ||
Textile | Amylases | Starch removal; de-sizing |
Cellulases | Denim stone washing, de-pilling | |
Pectinases | Treatment of flax and other fibers | |
Proteases | Degumming of silk, detergents | |
Laccases | Denim bleaching | |
Catalases | Removal of residual hydrogen peroxide | |
Pulp and paper | Hemicellulases | Improved bleachability |
Cellulases | Paper manufacture | |
Lipases | Removal of pitch components | |
Cellulases/hemicellulase | De-inking of recycled fibers | |
Laccase | Bleaching, fiber treatments |
Industrial implementation of biocatalysis needs a push, as it is yet to reach its peak. In this section some of the drivers and existing tools that might help to give the required impetus to industrialists – to make biocatalysis a large scale reality in the pharmaceutical industry – are discussed:
From the above-mentioned points, it can be easily inferred that biocatalysis/biotransformations, if incorporated in pharmaceutical process development, can help these industries in meeting their goals. A continuous process development is seen in making a compound/product, with a goal to achieve the following:
It is very often seen that organizations are reluctant to invest in expensive instrumentation/set-ups that a bioprocess might demand. However, that does not imply that processes like biocatalysis/biotransformation should be abandoned. It is important to streamline the bio-chemical methods within a chemical set-up, the best possible way after considering all the inputs and outputs. It is rightly said, ‘Great works are performed not by strength but by perseverance.’
Some of the important aspects to be considered for streamlining are summarized below:
Chemical processes in any pharmaceutical industry follow a pattern for the journey from development stage to process optimization stage. It might not be easy to define the scope of chemical development, encompassing the complete spectrum of work between research and production, but below is a sincere effort to concisely outline the path:18–72
While there are advantages, there are certainly some disadvantages as well. It is equally important to discuss the cons along with the pros. Highlighting the disadvantages is not to discourage people, but to help them to find suitable solutions for every problem that might occur in adapting this route:
Determining the mass balance of an immobilized biocatalyst is essential to understand the contamination related issues associated with the leaching of the immobilization support.78 Mass of a fresh biocatalyst should be the sum of recovered (dried) biocatalyst and mass of the leached support as well as protein (quantified).
Challenges will come in every aspect of life, but it is overcoming them that becomes the main challenge:
‘Never cut a tree down in the wintertime. Never make a negative decision in the low time. Never make your most important decisions when you are in your worst moods. Wait. Be patient. The storm will pass. The spring will come.’ – Robert H. Schuller
Thanks to the constant advancement of science over time, the future prospect of biocatalysis seems to be bright. But to realize this it is important to adapt the latest technologies that have come up. Awareness and understanding of the new developments are extremely important in any scientific field. The resistance to adapting newer technologies by industries arises from their lack of knowledge in the field. Thus, it is imperative to keep adapting and evolving, especially in the pharmaceutical industry. Some key-aspects for implementing biocatalysis at practical scale are:
A considerable amount of work dealing with biocatalysis is currently going on at the industrial scale in the pharmaceutical industry worldwide.88 Some relevant examples of pharma products which are already on the market, containing intermediates produced by biocatalysis, are summarized in Table 2.2.
Table 2.2 Selected examples of pharmaceutical products, the synthesis of which involved biocatalytic steps
Company | Company | Company | |
---|---|---|---|
1 | Abacavir89 | Beta lactamase | Stereoselective hydrolysis |
2 | Atazanavir sulfate90 | Leucine dehydrogenase, formate dehydrogenase | Transamination reaction, stereoselective |
3 | Atorvastatin calcium91 | Aldolases | Aldol condensation |
4 | Betametasone92 | Bacillus sphaericus var. fusiformis (ATCC 7055) | Ene formation in A-ring |
5 | Cefoxitin93 | Citrus acetyl esterase | Deacylation |
6 | Cortisone94 | Rhizopus arrhizus Fischer (ATCC 1145) | Hydroxylation |
7 | Didanosine95 | Acinetobacter lwoffi (ATCC 9036) | Amine to ketone |
8 | Diflucortolone valerate96 | Bacillus lentus MB 284 | Dehydrogenation |
9 | Diltiazem97 | Lipase or protease | Resolution |
10 | Dorzolamide98 | Neurospora crassa US 5580764 | Reduction, stereoselective |
11 | Drospirenone99 | Botryodiplodia malorum | Hydroxylation regioselective |
12 | Esomeprazole100 | Penicillium frequentans and Rhodobacterium capsulatum DSM 938 | Selective sulfoxidation |
13 | Zopiclone101 | Candida antarctica | Resolution |
14 | Ezetimibe102 | Aspergillus terreus ATCC 24839 | Resolution – reduction |
15 | Flumetasone103 | Curvularia lunata | Hydroxylation |
16 | Flunisolide104 | Cunnighamella blakesleeana ATCC 8688b | Hydroxylation |
Corynebacterium simplex ATCC 4964 | Dehydrogenation | ||
Streptomyces roseochromogenes | Hydroxylation | ||
17 | Gestodene105 | Penicillium raistrickii ATCC 10490 | Hydroxylation |
18 | Hydrocortisone106,107 | Curvularia lunata | Hydroxylation |
Rhizopus arrhizus Fischer ATCC 1145 or Rhizopus nigricans | Hydroxylation | ||
19 | Idarubicin108 | Streptomyces peucetius corneus, S. caeruleus, S. coeruleorubidas | Selective ether linkage formation |
20 | Indinavir sulfate109 | Dioxygenase | Hydroxylation |
21 | Lamivudine110 | 5′-Nucleotidase from Crotalus atrox, bacterial alkaline phosphatase | Dephosphorylation |
22 | Levodopa111 | Enzymatic | Hydroxylation |
Takadiastase | Selective deamidation to form amine | ||
23 | Methylprednisolone112 | Septomyxa affinis ATCC 6737 | Dehydrogenation |
24 | Pacilitaxel113 | Microbial | Selective reduction |
Lipase PS 30 | Resolution | ||
25 | Pravastatin114 | Mucor hiemalis or Phaseolus coccineus, Rhizoctonia solani, Nocardia | Hydroxylation |
26 | Predisolone115 | Corynebacterium hoagie ATCC 7005 | Dehydrogenation |
27 | Rosiglitasone116 | Rhodotorula rubra | Selective ene reduction |
28 | Travoprast117 | Rhizomucor meihei lipase, Chirazyme L9 | Resolution |
Candida antarctica lipase |
Several industries with activities in different pharmaceutical and chemical sectors are active in biocatalysis (see also Table 2.1). As relevant examples, BASF is among the leaders in biocatalysis as a major producer of vitamin B-2, the amino acid lysine, and several optically active amines. Apart from BASF, a precursor of aspartame, N-(benzyloxycarbonyl)-L-aspartyl-L-phenylalanine methyl ester (Z-APM), has been synthesized from N-(benzyloxycarbonyl)-L-aspartic acid (Z-L-ASP) and L-phenylalanine methyl ester (L-PM) with thermolysin. This is synthesized on a multi-thousand ton/year scale by DSM/Tosoh joint venture (Holland Sweetener Company, Geleen, The Netherlands). Likewise, L-carnitine is produced by Lonza (Basel, Switzerland) on an industrial scale using an enzyme from Agrobacterium HK1349, in a whole cell biotransformation process. The approach actually consists of a dehydrogenation of γ-butyrobetaine to 4-(trimethylamino)butenoic acid followed by selective addition of a water molecule by L-carnitine lyase. In a different area, Degussa-Huels (now Evonik) has used whole-cell biocatalysis systems to produce L-tert-leucine. The synthesis is conducted through combination of leucine dehydrogenase and formate dehydrogenase at industrial ton scale.
A remarkable strategy (as discussed above) is the implementation of protein engineering and directed evolution methods to design improved enzymes for industrial requirements. For instance, it is worth discussing the synthesis of intermediates involved in the initial stages of the manufacture of atorvastatin at industrial scale. Thus, for ethyl (S)-4-chloro-3-hydroxybutanoate, directed evolution was needed to improve factors like its coenzyme specificity, activity, as well as stability of the enzyme. Moreover, when using tert-butyl 6-cyano-3,5-dihydroxyhexanoate, protein engineering was conducted to improve activity and stability to 30% organic substrate (the hydroxyketone substrate and a tert-butyl acetoacetate impurity, both liquids) during the synthesis of this intermediate for atorvastatin.118,119 In the same field, cyano-3-hydroxybutyrate has been adapted through directed evolution of halohydrin dehalogenases to improve activity, stability and substrate tolerance of a catalyst.120 Apart from atorvastatin, directed evolution has also been used for substrates like (S)-1-(2,6-dichloro-3-fluorophenyl)ethanol,121 or for methyl (S,E)-2-(3-{3-[2-(7-chloroquinolin-2-yl)vinyl]phenyl}-3-hydroxypropyl)benzoate as intermediate for the synthesis of montelukast. Likewise for the synthesis of an intermediate for boceprevir ((1R,2S,5S)-6,6-dimethyl-3-azabicyclo[3.1.0]hexane-2-carboxylic acid), a fungal amine oxidase was improved for its expression in Escherichia coli, together with an increase in the activity by improving substrate and product tolerance.122 The activity of another amine oxidase was improved for an intermediate in telepravir synthesis ((1S,3aR,6aS)-octahydrocyclopenta[c]pyrrole-1-carboxylic acid).122 Likewise, for the synthesis of intermediates of ezetimibe ((4S)-3[(5S)-5-(4-fluorophenyl)-5-hydroxy-pentanoyl]-4-phenyl-1,3-oxazolidin-2-one), protein engineering was undertaken to design variants with improved activity and solvent stability.123 Another example is the improvement of activities for enzyme variants for the production of an intermediate for the synthesis of atazanavir (tert-butyl (2S,3R)-4-chloro-3-hydroxy-1-phenylbutan-2-yl-carbamate).124
Apart from asymmetric synthesis of building blocks for the pharmaceutical industry, another important area in which biocatalysis has found applications is in polysaccharide modification (e.g. dextrans, starch, etc.). In the case of dextrans, their industrial application can be attributed to its nonionic nature and good stability under normal operating conditions. Commercial applications may be found in pharmaceutical, food and textile industries, as well as a chromatographic media. Large-scale dextran production involves batch-wise culture of Leuconostoc mesenteroides NRRL B-512 in the presence of sucrose. Low-molecular-weight dextrans (40 000–70 000 Da) – also called clinical dextran (dextran 40 and dextran 70) – find use in the pharmaceutical industry as a raw material in medicine, a blood plasma extender, blood flow improver, ophthalmic solutions, as well as to preserve human organs during surgeries. Dextran 40 and 70 are used for prophylactic treatment in deep venous thrombosis and post-operative fatal pulmonary emboli. A solution of a complex containing 5% iron and 20% dextran is suitable for intramuscular and intravenous injection for treating iron deficiency anemia. Notably, however, the dextran linkage proportion (α-1,6; α-1,3; α-1,4; α-1,4; or α-1,2) undergoes changes with the change in the microorganism type.
Apart from dextrans, starch modification has found use with enzymes. The hydrolytic products derived from starch, like cyclodextrins and sugar syrups, are commonly used as excipients. The hydrolysis of starch is conducted with enzymes until the final desired dextrose equivalent (DE), or a certain carbohydrate profile, is reached.
Alpha-amylases and beta-amylases are the enzymes used for the hydrolysis of starch. Herein, bacterial or fungal enzymes (alpha-amylases) hydrolyze α-1,4 linkages in both amylose and amylopectin, to produce dextrose and maltose, and enzymes of barley and yeast (beta-amylases) act on the non-reducing ends of starch molecules and hence produce maltose in the beta form from the starch polymers. These enzymes are used to produce high-maltose syrups. Glucoamylases are fungal enzymes which hydrolyze maltose to produce glucose (dextrose). These enzymes catalyze hydrolysis of α-1,3, α-1,6 and β-l,6 linkages. Furthermore, debranching enzymes like pullulanase and isoamylase catalyze the hydrolysis of the 1,6 linkages without effect on the 1 and 4 linkages. These enzymes find use in the production of extremely high maltose syrups.125–128 Likewise, immobilized glucose isomerase enzyme from sources like Streptomyces, Bacillus, Actinoplanes and Arthrobacter species are used for the further conversion of high-fructose syrup. These enzymes require divalent cations (Co, Mg or Mn); inhibiting cations are Cu, Ni, Ag, Hg, Ca and Zn.129–131
As an outcome, maltodextrins (MDs), having a dextrose equivalent (DE) < 20, are formed by enzymatic and/or acid hydrolysis of starch. They consist of α-(1,4) linked D-glucose oligomers and/or polymers. The enzyme-catalyzed reaction of maltodextrin can be described as follows: Firstly, the starch slurry (30–40% dry solids) is pasted at a temperature of 80–90 °C, following its treatment with a ‘heat-stable’ bacterial alpha-amylase for liquefaction. Alpha-amylases from B. licheniformis or B. stearothermophilus can withstand temperatures up to 90–105 °C for at least 30 min, when stabilized with calcium ions; 30 min provides sufficient process time to facilitate the splitting of the 1,4 bonds and form maltose and limit dextrins. The fragmentation reaction continues until there is abundance of maltohexoses and maltoheptoses and the liquor has a DE of 12–15. Since acid-catalyzed hydrolysis of sweeteners to 55 DE or above creates products of reversion – such as gentiobiose, isomaltose and trehalose –, which give unacceptable flavors to the syrup, these syrups are usually made by acid–enzyme processes. The physical properties of a syrup depend heavily on its carbohydrate profile. The carbohydrate profile in turn is determined by the type of conversion and the nature of the enzyme treatment. Functional properties depend (obviously) on the degree of conversion. In this sense, browning, sweetness, fermentability, flavor enhancement, flavor transfer, freezing point depression and hygroscopicity increase with increasing degree of conversion, whereas bodying, cohesiveness, foam stabilization, prevention of sucrose crystallization, prevention of coarse ice crystal formation during freezing, and viscosity decrease with decreasing degree of conversion.132,133
In analogous lines, cyclodextrin-glycosyltransferases (cGTase) – acting on starch – produce cyclodextrin (CD). Different strains of bacilli, and some other bacteria, produce CGTases naturally.134 Among the α-, β- and γ-cyclodextrins formed, the β-form is predominant. Cyclodextrins can be isolated from the enzymatic reaction mixture in two ways. The ‘solvent process,’ involves addition of suitable organic substance to form an insoluble complex with the cyclodextrins, and the ‘non-solvent process’ is a chromatographic separation technique.
We have implemented enzymatic syntheses in some of our work. Our approach to developing and implementing large scale production of chemicals, intermediates and active pharmaceutical ingredients is achieved by principally following different approaches with various degrees of complexity:
To develop a particular chemoenzymatic process, the following general steps are followed:
Furthermore, to set up a practical bioconversion process, screening and optimization work is required at the following different levels:
To perform bio-reactions, the biocatalyst can be applied under various process conditions:
Finally, with respect to the choice of a reactor for biocatalytic reactions, no special equipment is needed for biocatalysis in many cases, and ‘ordinary’ stirred tanks, used in large-scale chemical synthesis with temperature and pH control, are sufficient, thus fitting a biocatalytic process in existing facilities & equipment.
Valganciclovir, an antiviral drug, is used to treat cytomegalovirus infections.135–141 It is a valyl ester of ganciclovir (a prodrug) which provides better bioavailability of ganciclovir when administered orally to cytomegalovirus infected patients. This prodrug for ganciclovir (valganciclovir) is found in the form of mixture of two diastereomers. According to the FDA specifications, the diastereomeric ratio of valganciclovir should be maintained in the range 55 : 45 to 45 : 55. After its administration, these diastereomers are rapidly converted into ganciclovir by hepatic and intestinal esterase. The biocatalytic step involved the prochiral hydrolysis catalyzed by hydrolases (Figure 2.1).
Figure 2.1 Enzymatic scheme representing the enzymatic synthesis of valganciclovir intermediate.
To this end, various hydrolases were screened towards the mono-hydrolysis of the di-valine ester of ganciclovir (e.g. lipases, CAL-A, CAL-B, proteases, esterases, etc.). The starting material was sparingly soluble in water, and hence solvents were also screened (dioxane, dimethylformamide, methyl alcohol, methylene dichloride, acetone, tetrahydrofuran, acetonitrile, ethyl alcohol, toluene, isopropyl alcohol, 2-methoxyethanol, diglyme, tert-amyl alcohol, 2-Me-THF, 2-BuOH, 1-BuOH, 1-propanol). The reaction was found to occur in solvent–water mixtures, and the enzyme Protex 6L (provided by Genencor) displayed positive results in a water–dimethylformamide mixture at pH 6.0–8.0 and temperature of 45–50 °C. Once these conditions were set, a scale-up reaction was carried out in a 500 mL round-bottom flask with an overhead stirrer and pH stat connection. The reaction mixture consisted of 20 g of the starting material in 180 mL of DMF and 120 mL of 0.1 M phosphate buffer with pH 8.0. The pH was maintained using 0.5 M NaOH. After completion of reaction, the product was isolated, giving a yield of 25% and a purity >95%. The above process suggested that Protex 6L was a versatile biocatalyst that was able to hydrolyze the starting material with a good yield and purity of >95%.
The chemoenzymatic synthesis of chiral intermediate [(S)-N-ethyl-N-methylcarbamic acid 3-(1-hydroxyethyl)phenyl ester] – a precursor of rivastigmine – was developed using alcohol dehydrogenase from Baker's yeast (Figure 2.2).
Figure 2.2 Enzymatic synthesis of rivastigmine intermediate.
To that end, N-ethyl-N-methylcarbamoyl acetophenone – synthesized by condensation of N-ethyl-N-methylcarbamoyl chloride with 3-hydroxyacetophenone – was screened for the stereoselective reduction by means of alcohol dehydrogenases. An in-process purification of crude alcohol dehydrogenase from Baker's yeast144 has been adopted and used in an enzymatic step. Various references are available on the purification of alcohol dehydrogenase and stereoselective reduction catalyzed by alcohol dehydrogenases.145–159 Apart from Baker's yeast, a number of alcohol dehydrogenases from various sources like Thermoanaerobium brockii (in recombinant form in E. coli), Parvibaculum lavamentivorans, or Deinococcus radiodurans (in recombinant form in E. coli) were screened towards enantioselective reduction of ketone to alcohol.
Type 2 diabetes has posed a severe threat to humankind worldwide. Insulin resistance is one of the primary cause as seen in type II diabetes. Even though the exact cause of the insulin resistance remains unknown, it can be partially attributed to obesity and associated lipotoxicity. N-Nonyl and N-hydroxyethyl (miglitol) of nojirimycin are drugs used in the treatment of diabetes type II symptoms and certain other metabolic disorders like Gaucher’s disease. Treatment of type-2 diabetes with miglitol (along with other drugs) is mainly prescribed for people whose diabetes cannot be controlled by dietary controls. Its mechanism of action involves retarding the breakdown and adsorption of table sugar/oligosaccharides in the small intestine, thus blood sugar levels following meals are lowered.160
Azasugars – also known as iminosugars – are mimetics of native sugars whereby the ring oxygen is replaced by a nitrogen atom. Due to their structural similarity to native carbohydrates, azasugars are able to act as glycosidase inhibitors and therefore have enormous therapeutic potential in the treatment of a variety of diseases including viral infection, bacterial infection, lysosomal storage disorders, cancer and diabetes. In particular 1-deoxyazasugars – such as 1-deoxynojirimycin and 1-deoxygalactonojirimycin – have been reported to inhibit various glycosidases in a reversible or competitive manner due to their structural resemblances to the sugar moieties of natural substrates.161,162 For the synthesis of deoxynojirimycin, aminosorbitol (1-amino-1-deoxyglucitol) is a key intermediate.
As depicted in Figure 2.3, anhydrous dextrose, benzylamine and Raney nickel in methanol were heated to 50 °C to form MGL-IA, simultaneously hydrogenated under 10–12 kg of hydrogen gas pressure to give MGL-IB. After completion of the reaction, water was added to the reaction mixture which was then filtered, palladium carbon was added to the filtrate, and hydrogenated at 50 °C, under 10–12 kg of hydrogen gas pressure at 50 °C to give aminosorbitol. The first step for either the reductive amination or the reductive alkylation is the reversible formation of the imine intermediate and water. The imine is then hydrogenated to the amine. Herein, a possible by-product formation results from the addition of the product amine to the imine, resulting in a dimer-like product with loss of ammonia (Figure 2.4).
Figure 2.3 Step I reactions to obtain aminosorbitol (MGL-I), intermediate in the synthesis of deoxynojirimycin.
Figure 2.4 Dimer structure formed as by-product in the formation of MGL-I.
Filtration is a key parameter in this process, as heavy metals like nickel and palladium are used together. To remove these metals they have to be filtered through a Celite® bed. Extreme precaution has to be taken while filtering both nickel and palladium, as both could catch fire. If the filtration is not properly conducted, then a heavy metal residue may be found in the filtrate. Subsequently, the reaction carried out between aminosorbitol and methyl formate in methanol (at reflux) gives MGL-II (Figure 2.5). The solid is filtered off and washed with methanol. The amount of dimer by-product – formed in the previous stage – is reduced during this filtration, presumably due to higher dimer solubility in methanol.
Figure 2.5 Step II, showing the formation of MGL-II by formylation of aminosorbitol.
Later on, the pH of the solution of N-formyl aminosorbitol in water was adjusted to 4–6 using orthophosphoric acid and then oxidation was performed with Gluconobacter oxydans DSM2003 whole cells with oxygen gas purging, resulting in N-formyl aminosorbose. Subsequently, reaction with sodium hydroxide results in 1,2-dehydro-2-(methylhydroxy)piperidine-3,4,5-triol, which further reacts with sodium borohydride followed by neutralization with hydrochloric acid, rendering crude deoxynojirimycin (Figure 2.6). Final purification proceeds by means of Indion 225 H resin, ammonia solution, and crystallization in 2-methoxyethanol/water to yield pure deoxynojirimycin.
Figure 2.6 Step III, showing the formation of deoxynojirimycin hydrochloride (MGL-III) from formyl aminosorbitol.
The reaction mixture of deoxynojirimycin contains several organic and inorganic impurities. Among the organic ones, the dimer impurity of amino-sorbitol (Figure 2.4) and the formylated dimer impurity of amino-sorbitol may be pinpointed. With regard to inorganic impurities, sodium chloride, sodium borohydride, boric acid, residual palladium, residual nickel and iron may be found. As the solubility of deoxynojirimycin is very high in water, it is very tedious to remove these water soluble inorganic salts. For their removal from deoxynojirimycin, resin purification was done, using an Indion 225 H resin packed in a proper column. Organic impurities such as dimer impurity of aminosorbitol and formylated dimer impurity of aminosorbitol do not bind to resin and hence are removed from deoxynojirimycin. The complete binding of deoxynojirimycin from the reaction mixture is assured by assay of elute. The resin is then washed with water until elute is free of inorganic impurities. Deoxynojirimycin is then eluted form the resin using ammonia solution. Elute is subsequently distilled and crystallized to get pure deoxynojirimycin base with a purity greater than 96%. Inorganic impurities and heavy metals were also removed during this process of resin purification. The isolated free base is then converted into hydrochloride salt in quantitative yield and purity greater than 99.9%.
As described above, separation of the inorganic impurities is tedious work while preparing these compounds of high purity, especially in highly regulated markets like Japan, where the total impurity level requirements may not exceed more than 0.1%, and other inorganic impurities cannot exceed 20 ppm. Preparing compounds for such a market represents a challenge for the synthetic community, especially at the purification steps. Apart from the above-described resin method, several other methods have been tried for purification of deoxynojirimycin, such as membrane dialysis, carboxymethyl-Sepharose chromatography, preparative thin-layer chromatography,163 as well as using hydrochloric acid.164 Of all of them, purification using resin seems to be the most viable. The better resin for the preparation of deoxynojirimycin was Lewatit l SP 112 H, which is a strong acid exchange resin. However, the availability of this resin prompted a search for another one that could be reusable, easily available and cheap. Of all the methods used for the preparation of deoxynojirimycin and their salts,164–176 the biocatalytic one involving the preparation of N-formyl-aminosorbitol, conversion into N-formyl aminosorbose through biotransformation, and finally cyclization and reduction to yield deoxynojirimycin seems to be the best from a commercial point of view.
For this project, apart from the successful biocatalyst employed (see above), various routes and enzymes were assessed previously. In the following, some selected examples are discussed. In these cases, a number of tests were run with locally available Baker's yeast, as well as with yeasts from Saccharomyces cerevisiae Type I obtained from Sigma Aldrich. Thus, locally available Baker's yeast was grown on D-sorbitol media containing 6% D-sorbitol, 2.4% yeast extract, and 4.8% KH2PO4 in Milli-Q water for 48 h. The cells obtained were then centrifuged, washed with 0.02 M MgSO4, and stored below 4 °C for future use. Wet cells were then inoculated in autoclaved reaction media containing 2% 1-amino-1-deoxy-D-glucitol solution containing 4% of yeast extract, 20% of D-sorbitol and 2% potassium dihydrogen phosphate in Milli-Q water, pH 5.0 (adjusted using 2.0 M HCl). The reaction media was stirred for 72 h, centrifuged and separated by column chromatography using Amberlyst-15 H+ resin. The acid wash was distilled off and the residue obtained was precipitated with ethanol to obtain white crystals of 1-deoxynorjorimycin. In another line, Saccharomyces cerevisiae Type I from Sigma Aldrich was inoculated in a reaction buffer pH 7.0 containing 1% 1-amino-1-deoxy-D-glucitol, 1% NAD+, 5% D-glucose in Milli-Q water, 100 mM potassium phosphate. The reaction media was stirred for 72 h, centrifuged and separated by column chromatography using Amberlyst-15 H+ resin. The acid wash was distilled off and the residue obtained was precipitated with ethanol to obtain white crystals of 1-deoxynorjorimycin.
Various lessons arose during the process development:
Absorbance study of bacterial growth is mandatory. It should be carried out both for the shake flask method and at a mini-fermenter scale.
Significant variability was observed in parameters in large-scale versus lab-scale batches. Example 1: In lab scale batches, OD absorbance (600 nm) after inoculation was found to be 0.03–0.05 while in large-scale fermenter batches the OD absorbance (610 nm) after inoculation was found to be 0.16.
Example 2: The shake flask method gave rise to maximum OD of 1.2–1.8 in 36–40 h with a viability of 89% whereas in large-scale fermenter batches enormous growth was observed within 6 h with an absorbance of 2.4 and viability of 75%.
Before inoculating into a fermenter, a performance test should be carried out at lab scale level and the cells harvested should be reactive with the substrate. Optimization of various reaction parameters was carried and statistically their significance was also determined in the finalized process.
During the biotransformation process, i.e. during oxidation of N-formyl using Gluconobacter oxydans DSM2003 whole cells, three main unknown impurities peaks (with defined retention times) were observed in a HPLC chromatogram while reaction monitoring. Since higher levels of impurities affect the yield of the process, efforts were carried out to study the factors that can reduce the formation of process impurities. Thus, a study of the effect of pH during the stage III reaction of preparation of deoxynojirimycin base was conducted. Experiments involved reaction of N-formyl aminosorbitol in the presence of water, oxygen and Gluconobacter oxydans DSM2003 whole cells, followed by addition of sodium hydroxide and sodium borohydride to give deoxynojirimycin. Post work-up and subsequent crystallization with 2-methoxyethanol yielded deoxynojirimycin base. In this experiment the pH of the reaction was varied to study its effect during the reaction. A regression analysis was also performed using Minitab to obtain clarity about the role of the pH during the reaction. It was observed that at the extreme pH limits, e.g. at pH 2.0, reaction did not occur, whereas at pH 8.0 reaction did not reach completion. The pH range 4–6 showed a certain effect on yield and purity. Interestingly, it was observed that the pH had a negative correlation with one of the impurities (number 3). Thus, when a null hypothesis p-test was carried out, no significant effect of pH was to be found on product purity, impurity 1 and impurity 2, but a significant influence was seen in minimizing impurity 3. Furthermore, large-scale batches were statistically analyzed to achieve better understanding of the influence of the list of parameters on the output obtained. The parameters pH, RPM, and oxygen cylinders consumed along the course of the reaction were studied during stage III of the reaction described above. Their effect on the output and reaction completion time was studied. It was seen that only RPM showed a statistically significant effect on the reaction completion time while the rest of the factors did not contribute to any significant effect on the output or reaction completion time.
Notably, in pharmaceutical industries – where the regulations and guidelines laid are for public safety – the raw materials used to produce a particular chemical precursor, or formulations, should also undergo scrutiny (e.g. maltodextrin, dextrans, etc.), especially when raw materials used are themselves synthesized via biocatalytic routes. Microbial synthesis is very critical, and the slightest change in strains or species of organisms can alter the desired product in various ways. Moreover, when raw materials (excipients) are of animal origin, and solvents are used for extraction, then it is mandatory that the following parameters are accounted for:
To summarize the thoughts reported in this chapter, biocatalysis implementation in the pharmaceutical industry is crucial because chemical and biological knowledge needs to be integrated together. It is also important to be aware of, and be able to ultimately remove, all the impurities arising from microorganisms or enzymes (biological impurities). Along the entire supply chain from the developers to the field force, everyone has to be transparent. Disclosing the complete picture of process manufacturing to the purchaser may come to be of use for any anticipated regulatory queries. Previously, chemistry seemed to lead the growth in the life-sciences sector. But now the tables are slowly turning with changing times. It is the biosciences now, which will drive groundbreaking innovation in the forthcoming century, including synthetic approaches to small molecules. To design, develop and execute a synthetic route incorporating biotransformation techniques in it, interdisciplinary scientific skills are required. Specialization in areas like organic chemistry, analytical chemistry, (bio)chemical engineering, fermentation techniques and molecular biology have become essential for the successful implementation of any such project. Very few companies have these prerequisite scientific skills and plant capabilities integrated in-house. Most companies outsource these complex procedures to contract research organizations. CROs such as transgene Biotek Ltd, etc. which have their own propriety enzymes also work closely with commercial enzyme suppliers like Almac, Codexis, Johnson-Matthey, Libragen, Syncozymes and Enzymeworks to supply enzymes or cells to be used in organic synthesis.
Biocatalysis is executed for the generation of chiral active pharmaceutical ingredients (APIs) in the fine chemical industry. But its utilization is not seen as much in the pharmaceutical industry. Presently a significant change is observed in this pattern as more and more pharmaceuticals are adapting biocatalysis. Certain business reports showed that the industrial enzyme market stood at ca. $4.4 billion, in 2015.177 It also predicted a promising growth rate of ca. 6% per year. These numbers provide hope for the future of biocatalysis. It looks like different business sectors are becoming more welcoming towards this change, which was very much resisted in the past. Moreover, some big names in the pharmaceutical industry like to encourage their employees by rewarding them for excellent executed work. ‘The Pfizer Award in Enzyme chemistry’, established in 1945, recognizes efforts put in by scientists to stimulate research in enzyme chemistry. Merck likes to lead by example – their green chemistry team was awarded, in 2012, for re-formatting their manufacturing process in certain sectors. Furthermore, US Presidential Green Chemistry Awards are also awarded to companies for extensive research work in biocatalysis. Some of the recipients of this award are Bristol Myers Squibb, Metabolix, Merck & Codexis, Eastman Chemical Company, etc. Pfizer also received ‘The AstraZeneca Award’ for excellence in green chemistry and engineering for its success in the pregabalin project.178
Thus, it is important for us to bring a change in the age-old procedures of manufacturing, and green chemistry can be the accelerating biocatalyst in this change-over. Industries have to be self-inspired and motivated to encourage their teams to think differently and make the green revolution a success!
API | Active pharmaceutical ingredient |
ATCC | American Type Culture Collection |
BASF | Badische Anilin und Soda Fabrik |
BCA | Bicinchoninic acid assay |
CAL-A | Candida antarctica lipase A |
CAL-B | Candida antarctica lipase B |
cGMP | Current good manufacturing practices |
CLEAs | Cross-linked enzyme aggregates |
CLECs | Cross-linked enzyme crystals |
DE | Dextrose equivalent |
DNA | Deoxyribonucleic acid |
DSM | Deutsche Sammlung von Mikroorganismen |
DP | Degree of polymerization |
FAD | Flavin adenine dinucleotide |
HFCS | High fructose corn syrup |
HPLC | High performance liquid chromatography |
LPS | Lipopolysaccharides |
OD | Optical density |
NAD+ | Nicotinamide adenine dinucleotide |
NADH | Nicotinamide adenine dinucleotide reduced form |
NADP+ | Nicotinamide adenine dinucleotide phosphate |
NADPH | Nicotinamide adenine dinucleotide phosphate reduced form |
PS | Pseudomonas cepacia |
US | United States |
API: active pharmaceutical ingredient.
Biocatalyst: biological agents such as microorganisms or enzymes that activate or speed-up a chemical reaction.
Bioconversion: chemical conversion of a substance using biological methods (enzymes or whole cells, biocatalysis or biotransformation).
Biocatalysis: chemical conversion of a substance (defined starting material) into a desired product with the aid of a free or an immobilized enzyme.
Biotransformation: chemical conversion of a substance (defined starting material) into a desired product with the aid of a (usually) living whole cell, containing the necessary enzyme(s).
Biosynthesis: De novo production of an entire molecule by a living organism. Unlike biotransformation, which acts on a starting substance, the biosynthesis is not dependent on starting substances, but only on nutrients.
Building blocks: compounds that can combine with other compounds to form a new molecule.
Clone: a population of genetically identical microorganisms derived from a single parent cell.
Cloning (of a gene): Technical term to describe the transfer of genes from one (micro)organism to another. Inserting a population of DNA molecules, known to contain the DNA of interest, into a population of vector DNA molecules in such a way that each vector molecule contains only a single DNA molecule from the original population.
DNA: Acronym for deoxyribonucleic acid, usually 2-deoxy-5-ribonucleic acid. DNA contains the genetic information of an organism and the code in the genes contain the blueprints to form the different proteins.
Enzyme: Proteins that act as biological catalysts (biocatalysts), initiating all biochemical reactions. Enzymes are classified into six enzyme classes according to their mechanism.
Fermentation: Process for the cultivation of microorganisms in special vessels (fermenters) that allow the ‘monoseptique’ propagation of a desired microorganism – only the desired microorganism is allowed to grow using sterile fermentation technology. Thus the result of a fermentation is microbial biomass, which contains the desired enzyme used in the following biotransformation. The biotransformation can take place during fermentation or after the fermentation in a separate vessel, which often does not need to be operated in a sterile manner.
Fine chemicals: value-added intermediates and active substances used, for example, in pharmaceuticals.
Gene: Basic unit of the hereditary material DNA, an ordered sequence of nucleotide bases that encodes a product. The product of a gene finally is a protein. Genes are the blueprints for the enzymes used in biotransformation.
Genome: the entire complement of genetic material in a (micro)organism.
Green chemistry: The term for sustainable (chemical) industrial manufacturing processes striving for, for example, minimal waste production and energy consumption. Biosynthesis and biotransformation are assumed to play a key role in green chemistry in the future.
Host: Microorganism that is used for the expression of foreign DNA, and consequently the production of a foreign protein (enzyme), which is encoded in a gene. The expression of a foreign gene for the production of a foreign protein is also called heterologous expression, because the gene does not belong naturally to the producing organism (host).
Immobilized enzymes: the half-life and stability of enzymes can be prolonged by ‘associating’ enzymes using cross-linking, covalently binding them to, for example, polymer resins, by encapsulating them.
Metagenomics: Environmental genomics is the study of genomes recovered from environmental samples as opposed to clonal cultures. This relatively new field of genetic research allows the genomic study of organisms that are not easily cultured in a laboratory.
ISPR (in situ product recovery): Removal of the inhibiting biotransformation product from the solution during the reaction. By removing the product, which inhibits cellular growth and/or an enzyme activity used for biotransformation, the productivity can be maximized.
NCE: New chemical entity.
Product inhibition: finely tuned enzyme activities are often inhibited at higher concentrations of the resulting product of their catalytic activity.
Secondary metabolite: A product of the secondary metabolism. Secondary metabolites are typically produced in tiny amounts, very complex (sometimes with ‘dozens’ of chiral centers) and biologically very active. Secondary metabolites (e.g. vancomycin) are not essential for normal growth, development or reproduction of a microorganism. On the other hand, the products of the primary metabolism (e.g. ethanol sugar fermentation) are essential for the survival of the microorganism.
Sequence: The order of amino acids in a protein or the order of nucleotides of a gene in DNA. Knowing the sequence allows the cloning, heterologous expression and production of a protein in different hosts or microorganisms.
Strain: A genetically homogeneous population of (micro)organisms of common origin. A strain can be biochemically or morphologically differentiated from other strains. Microbial strains are able to produce enzymes with distinctive chemo-, regio-, and enantioselectivity. Another characteristic of microbial strains is that they divide and grow very quickly, and are well suited for rapid production of large amounts of enzymes.
Strain collection: Microorganisms can be stored for several years in small ampoules in frozen form (−80 °C) or lyophilized. Individual strains can be revived easily for tests. The larger a collection of different strains, the higher the success rate for a new biotransformation candidate.
System biology (in silico biology): Computational models of biological systems, where varying streams of biochemical information are integrated and modeled in silico. System biology is applied from drug discovery to metabolic pathway simulation.
White biotechnology: a term for industrial biotechnology, which includes a vast area of products, processes, and industries.
1. J. Sylvestre, H. Chautard, F. Cedrone and M. Delcourt, Org. Process Res. Dev., 2006, 10, 562.
2. A. S. Wells, G. S. Finch, P. C. Michels and J. W. Wong, Org. Process Res. Dev., 2012, 16, 1986.
3. M. Butters, D. Catterick, A. Craig, A. Curzons, D. Dale, A. Gilmore, S. P. Green, I. Marziano, J.-P. Sherlock and W. White, Chem. Rev., 2006, 106, 3002.
4. V. Sharma, Applicability of Green Chemistry in Pharmaceutical Processes, Pharma Bio world, Piramal enterprises, 2015.
5. R. Wohlgemuth, Curr. Opin. Microbiol., 2010, 13(3), 283.
6. R. J. Fox and G. W. Huisman, Trends Biotechnol., 2008, 26, 132.
7. D. J. Pollard and J. M. Woodley, Trends Biotechnol., 2007, 25, 66.
8. H. E. Schoemaker, D. Mink and M. G. Wubbolts, Science, 2003, 299, 1694.
9. A. J. Straathof, S. Panke and A. Schmid, Curr. Opin. Biotechnol., 2002, 13(6), 548.
10. A. Schmid, J. S. Dordick, B. Hauer, A. Kiener, M. Wubbolts and B. Witholt, Nature, 2001, 401, 258.
11. A. J. Bridges, P. S. Raman, G. S. Y. Ng and J. B. Jones, J. Am. Chem. Soc., 1984, 106, 1461.
12. J. Tao, L. Zhao and N. Ran, Org. Process Res. Dev., 2007, 11, 259.
13. O. Kirk, T. V. Borchert and C. C. Fuglsang, Curr. Opin. Biotechnol., 2002, 13(4), 345.
14. P. Anastas and N. Eghbali, Chem. Soc. Rev., 2010, 39, 301.
15. J. L. Tucker, Org. Process Res. Dev., 2006, 10, 315.
16. N. Ran, L. Zhao, Z. Chen and J. Tao, Green Chem., 2008, 10, 361.
17. R. N. Patel, Coord. Chem. Rev., 2008, 252(5–7), 659.
18. A. M. Rouhi, Chem. Eng. News, 2003, 81, 37.
19. J. J. Chen, T. C. Nugent, C. V. Lu, S. Kondapally, P. Giannousis, Y. Wang and J. T. Wilmot, Org. Process Res. Dev., 2003, 7, 313.
20. H. R. Bjorsvik, L. Liguori and J. A. V. Merinero, J. Org. Chem., 2002, 67, 7493.
21. C. Jamieson, M. S. Congreve, D. F. Emiabata-Smith, S. V. Ley and J. J. Scicinski, Org. Process Res. Dev., 2002, 6, 823.
22. J. L. Navarette-Bolaños, H. Jimenez-Islas, E. Botello-Alvárez and R. Rico-Martínez, Org. Process Res. Dev., 2002, 6, 841.
23. M. P. J. Donners, M. C. Hermsis, J. P. A. Custers, J. Meueldijk, J. A. J. M. Vekemans and L. A. Hulshof, Org. Process Res. Dev., 2002, 6, 606.
24. P. J. Hogan, P. A. Hopes, W. O. Moss, G. E. Robinson and I. Patel, Org. Process Res. Dev., 2002, 6, 225.
25. H. R. Bjorsvik and T. Engell, Org. Process Res. Dev., 2002, 6, 113.
26. D. J. Gavin and C. A. Mojica, Org. Process Res. Dev., 2001, 5, 659.
27. J. M. Hawkins and T. W. Makowski, Org. Process Res. Dev., 2001, 5, 328.
28. M. Owen and S. Godbert, Org. Process Res. Dev., 2001, 5, 324.
29. M. R. Owen, C. Luscombe, L.-W. Lai, S. Godbert, D. L. Crookes and D. Emiabata-Smith, Org. Process Res. Dev., 2001, 5, 308.
30. V. W. Rosso, J. L. Pazdan and J. J. Venit, Org. Process Res. Dev., 2001, 5, 294.
31. A. Illanes and A. Fajardo, J. Mol. Catal. B: Enzym., 2001, 11, 587.
32. M. C. Hersmis, A. J. H. Spiering, R. J. M. Waterval, J. Meueldijk, J. A. J. M. Vekemans and L. A. Hulshof, Org. Process Res. Dev., 2001, 5, 54.
33. E. W. Kirchchoff, D. R. Anderson, S. Zhang, C. S. Cassidy and M. T. Flavin, Org. Process Res. Dev., 2001, 5, 50.
34. C. Jamieson, M. S. Congreve, D. F. Emiabata-Smith and S. V. Ley, Synlett, 2000, 11, 1603.
35. P. Subra and P. Jestin, Ind. Eng. Chem. Res., 2000, 39(11), 4178.
36. J. Zhang, E. W. Kirchhoff, D. E. Zembower, N. Jiménez, P. Sen, Z.-Q. Xu and M. T. Flavin, Org. Process Res. Dev., 2000, 4, 577.
37. C. Simms and J. Singh, Org. Process Res. Dev., 2000, 4, 554.
38. H. R. Bjorsvik, L. Liguori and F. Minisci, Org. Process Res. Dev., 2000, 4, 534.
39. R. Brereton, Chem. Br., 2000, 36(5), 28.
40. J. Siro, A. Ramos, J. J. Vaquero, J. Alvarez-Builla and J. L. García-Navío, Tetrahedron, 2000, 56, 2469.
41. G. Annadurai, T. Sivakumar and S. Rajesh Babu, Bioprocess Eng., 2000, 23(2), 167.
42. C. J. Shieh and Y. F. Lai, J. Agric. Food Chem., 2000, 48(4), 1124.
43. R. Ramos, M. Menéndez and J. Santamaría, Catal. Today, 2000, 56, 239.
44. H. R. Bjorsvik and F. Minisci, Org. Process Res. Dev., 1999, 3, 330.
45. H. R. Bjorsvik and K. Norman, Org. Process Res. Dev., 1999, 3, 341.
46. R. Lazaro, P. Bouchet, J. Elguero, M. Sergent, R. Luu and T. Phan, ACH-Models Chem., 1999, 136(5–6), 525.
47. D. P. Guimaraes, F. A. A. Costa, M. I. Rodrigues and F. Maugeri, Braz. J. Chem. Eng., 1999, 16(2), 129.
48. D. R. Philipauskas, Med. Res. Rev., 1999, 19(5), 463.
49. D. R. Philipauskas, in Process Chemistry in the Pharmaceutical Industry, ed. K. G. Gadamasetti, Marcel Dekker, New York, 1999, p. 411.
50. H. J. M. Gijsen, H. C. C. K. van Bakel, W. Zwaan and L. A. Hulshof, Org. Process Res. Dev., 1999, 3, 38.
51. H. J. Yang and C. H Yang, J. Appl. Polym. Sci., 1998, 69(3), 551.
52. S. Raharja, L. Rigal, L. Barre and P. F. Vidal, Can. J. Chem. Eng., 1997, 75, 913.
53. S. Bouttemy, J.-M. Aubry, M. Sergent and R. P. Than-Luu, New J. Chem., 1997, 21, 1073.
54. H. R. Bjorsvik, P. Bravo, M. Crucianelli, A. Volonterio and M. Zanda, Tetrahedron: Asymmetry, 1997, 8, 2817.
55. M. L. Boys, K. J. Cain-Janicki, W. W. Doubleday, P. N. Faried, M. Kar, S. T. Nugent, J. R. Behling and D. R. Pilipauskas, Org. Process Res. Dev., 1997, 1, 233.
56. J. R. Inga and B. I. Morsi, Can. J. Chem. Eng., 1997, 75, 872.
57. J. E. Burks Jr, L. Espinosa, E. S. LaBell, J. M. McGill, A. R. Ritter, J. L. Speakman, M. Williams, D. A. Bradley, M. G. Haehl and C. R. Schmid, Org. Process Res. Dev., 1997, 1, 198.
58. L. M. Osborne and T. W. Miyakawa, J. Liq. Chromatogr. Relat. Technol., 1997, 20, 501.
59. K. E. Saranteas and G. D. Botsaris, in Separation and Purification by Crystallization, ed. G. D. Botsaris and K. Toyokura, American Chemical Society, Washington DC, ACS Symposium Series, 1997, vol. 667, p. 150.
60. C. Liu, J. S. Ng, J. E. Behling, C. H. Yen, A. L. Campbell, K. S. Fuzail, E. E. Yonan and D. V. Mehrotra, Org. Process Res. Dev., 1997, 1, 45.
61. L. Xu and S. L. Gipson, Polyhedron, 1996, 15, 2211.
62. P. W. Araujo and R. G. Brereton, Trends Anal. Chem., 1996, 15, 156.
63. P. W. Araujo and R. G. Brereton, Trends Anal. Chem., 1996, 15, 63.
64. V. L. Mylroie, L. Valente, L. Fiorella and M. A. Osypian, in Catalysis of Organic Reactions, ed. R. E. Malz Jr, Chemical Industries Series, Marcel Dekker, New York, 1996, vol. 68, pp. 301–312.
65. A. J. Solodar, E. D. Sall and L. E. E. Stout Jr, in Catalysis of Organic Reactions, ed. R. E. Malz, Chemical Industries Series, Marcel Dekker, New York, 1996, vol. 68, pp. 119–132.
66. P. M. Anderson, T. Lundstedt and L. Abramo, J. Chemom., 1996, 10, 379.
67. H. R. Bjorsvik and A. Aukrust, Acta Chem. Scand., 1995, 49, 115.
68. M. Arroyo and J. V. Sinisterra, J. Org. Chem., 1994, 59, 4410.
69. H. R. Bjorsvik, Acta Chem. Scand., 1994, 48, 445.
70. H. R. Bjorsvik, A. W. Aabye, R. Carlson and P. Carlsen, Acta Chem. Scand., 1994, 48, 582.
71. J. Drouin, S. Gauthier, O. Patricola, P. Lanteri and R. Longeray, Synlett, 1993, 10, 791.
72. R. Carlson and A. Nordahl, Topics in Current Chemistry, Vol. 166, Springer-Verlag, Berlin, Heidelberg, 1993, p. 1.
73. P. Tufvesson, J. L. Ramos, M. Nordblad and J. M. Woodley, Org. Process Res. Dev., 2011, 15, 266.
74. A. J. J. Straathof, S. Panke and A. Schmid, Curr. Opin. Biotechnol., 2002, 13(6), 548.
75. S. G. Burton, D. A. Cowan and J. M. Woodley, Nat. Biotechnol, 2002, 20, 37.
76. J. E. Leresche and H. P. Meyer, Org. Process Res. Dev., 2006, 10, 572.
77. ICH Harmonised Guideline, Guideline for Elemental Impurities Q3D, 2014.
78. C. Joséa and R. D. Bonettoa, J. Mol. Catal. B: Enzym., 2011, 71(3–4), 95.
79. K. A. Powell and S. W. Ramer, Angew. Chem., Int. Ed., 2001, 40, 3948.
80. M. T. Reetz, J. Am. Chem. Soc., 2013, 135, 12480.
81. M. T. Reetz, Sci. Prog., 2000, 83, 157.
82. P. H. Nielsen and K. M. Oxenbøll, J. Life Cycle Assess., 2007, 12, 432.
83. G. A. Behrens and A. Hummel, Adv. Synth. Catal., 2011, 353, 2191.
84. G. Carrea and S. Riva, Angew. Chem., Int. Ed., 2000, 39, 2226.
85. L. Cao, F. V. Rantwijk and R. A. Sheldon, Org. Lett., 2000, 2, 1361.
86. N. L. St. Clair and M. A. Navia, J. Am. Chem. Soc., 1992, 114, 7314.
87. A. M. Collins, C. Maslin and R. J. Davies, Org. Process Res. Dev., 1998, 2, 400.
88. A. M. Thayer, Chem. Eng. News, 2001, 79, 27.
89. S. M. Daluge, EP0434450, 1992.
90. G. Bold and A. Fässler, J. Med. Chem., 1998, 41, 3387.
91. W. Greenberg, K. Wong, A. Varvak and M. Burk, WO2004027075, 2004.
92. R. Richard and E. P. Oliveto, US3164618, 1965.
93. E. O. Stapley and J. M. Mata, US3914157, 1975.
94. H. C. Murray and D. H. Peterson, US2602769, 1952.
95. K. Yokozeki, H. Shirae, K. Kobayashi, H. Shirigami and Y. Irie, US4970148, 1990.
96. K. Klaus and R. Gerhard, US3232839, 1966.
97. L. Kanerva, O. Sundholm, P. Kairisalo and M. Hytonen, US5514589, 1996.
98. R. A. Holt and S. R. Rigby, US5580764, 1996.
99. K. Petzoldt, R. Wiechert, H. Laurent, K. Nikisch and D. Bittler, US4416985, 1983.
100. R. Holt, C. Reeve and S. Taylor, WO9617076, 1996.
101. L. F. Solaresa and M. Díaza, Tetrahedron: Asymmetry, 2002, 13, 2577.
102. M. J. Homann and W. B. Morgan, US5919672, 1999.
103. G. Hempel, M. Kennecke, B. Krieger, R. Phillippson, H. Triem and A. Weber, US4898693, 1990.
104. G. Rosenkraez, US3124571, 1964.
105. J.-L. Moutou, F. Mouton, G. Pellegrino, J.-M. Dillenschneider and J. Lafay, EP1586579A, 2011.
106. H. A. Drake, K. M. Mann and D. E. Rayman, US2794816, 1957.
107. J. Manosroi, Y. Chisti and A. Manosroi, Appl. Biochem. Microbiol., 2006, 42, 479.
108. L. A. Mitscher and D. Lednicer, US Pat. 4471052, 1984.
109. B. C. Buckland, S. W. Drew, N. C. Connors, M. M. Chartrain, C. Lee, P. M. Salmon, K. Gbewonyo, W. Zhou, P. Gaillot, R. Singhvi, R. C. Olewinski, W.-J. Sun, J. Reddy, J. Zhang, B. A. Jackey, C. Taylor, K. E. Goklen, B. Junker and R. L. Greasham, Metab. Eng., 1999, 1, 63.
110. J. A. V. Coates, I. M. Mutton, C. R. Penn, R. Storer and C. Williamson, WO9117159, 1991.
111. K. Min, D. H. Park and Y. J. Yoo, J. Biotechnol., 2010, 146, 40.
112. O. K. Sebek and J. B. Spero, US Pat. 2897218, 1959.
113. R. N. Patel, Annu. Rev. Microbiol., 1998, 52, 361.
114. A. Terahara and M. Tanaka, US Pat. 4346227, 1982.
115. A. Nobile, US Pat. 3134718, 1964.
116. R. M. Hindley and S. N. Woroniecki, WO Pat. 9310254, 1993.
117. G. Biffi, L. Feliciani, A. D'Alfonso, A. Porta and G. Zanoni, EP Pat. 2294209, 2011.
118. L. J. Giver, L. M. Newman, E. Mundorff, G. W. Huisman, S. J. Jenne, J. Zhu, B. Behrouzian, J. H. Grate and J. Lalonde, US Pat. 7879585, 2011.
119. M. P. Burns and J. W. Wong, US Pat. 0118962, 2008.
120. S. C. Davis, J. H. Grate, D. R. Gray, J. M. Gruber, G. W. Huisman, S. K. Ma, L. M. Newman, R. Sheldon and L. A. Wang, US Pat. 7807423, 2010.
121. J. Liang, S. J. Jenne, E. Mundorff, C. Ching, J. M. Gruber, A. Kreber and G. W. Huisman, WO Pat. 2009/036404, 2009.
122. B. Mijts, S. Muley, J. Liang, L. M. Newman, X. Zhang, J. Lalonde, M. D. Clay, J. Zhu, J. M. Gruber, J. Colbeck, J. D. Munger Jr, J. Mavinahalli and R. Sheldon, WO Pat. 2010/008828, 2010.
123. E. Mundorff and E. de Vries, WO Pat. 2010/025085, 2010.
124. Y. K. Bong, M. Vogel, S. J. Collier, V. Mitchell and J. Mavinahalli, WO Pat. 2011/005527, 2011.
125. R. V. MacAllister, Adv. Carbohydr. Chem. Biochem., 1979, 36, 15.
126. T. Komaki and N. Taji, Agric. Biol. Chem., 1968, 32, 860.
127. J. H. Pazur and K. Kleppe, J. Biol. Chem., 1962, 237, 1002.
128. T. A. Kuracina, K. Lorenz and K. Kulp, Cereal Chem., 1987, 64, 182.
129. N. B. Havewala and W. H. Pitcher, Enzyme Eng., 1974, 2, 315.
130. N. E. Lloyd and K. Khaleeluddin, Cereal Chem., 1976, 53, 270.
131. Y. Takasaki, Agric. Biol. Chem., 1967, 31, 309.
132. W. J. Hoover, Food Processing Catalogue, 1964, p. 65.
133. L. Hobbes, in Starch: Chemistry and Technology, ed. J. BeMiller and R. Whistler, Elsevier, Amsterdam, 3rd edn, 2009, vol. 21, p. 797.
134. A. Tonkova, Enzyme Microb. Technol., 1998, 22, 678.
135. M. K. Sethi, S. R. Bhandya, R. Shukla, A. Kumar, N. Maddur, V. S. N. Jayalakshmi Mittapalli, V. S. Rawat and R. K. Yerramalla, J. Mol. Catal. B: Enzym., 2014, 108, 77.
136. K. G. Au Eong, S. Beatty and S. J. Charles, Singapore Med. J., 2000, 41(6), 298.
137. I. G. Sia and R. Patel, Clin. Microbiol. Rev., 2000, 13, 83.
138. P. G. Tranos, I. Georgalas, P. Founti and I. Ladas, Clin. Ophthalmol., 2008, 2, 961.
139. I. Yust and Z. Fox, Eur. J. Clin. Microbiol. Infect. Dis., 2004, 23(7), 550.
140. M. Curran and S. Noble, Drugs, 2001, 61, 1145.
141. J. C. Martin, M. A. Tippe, D. P. McGee and J. P. Verheyden, J. Pharm. Sci., 1987, 76, 180.
142. M. K. Sethi, S. R. Bhandya, A. Kumar, N. Maddur, R. Shukla and V. S. N. Jayalakshmi Mittapalli, Tetrahedron: Asymmetry, 2013, 24, 374.
143. M. K. Sethi, S. R. Bhandya, A. Kumar, N. Maddur, R. Shukla and V. S. N. Jayalakshmi Mittapalli, J. Mol. Catal. B: Enzym., 2013, 91, 87.
144. W. J. Rutter and J. R. Hunsley, Methods Enzymol., 1966, 9, 480.
145. S. C. Davis, J. H. Grate, D. R. Gray, J. M. Gruber, G. W. Huisman, S. K. Ma, L. M. Newman, R. Sheldon and L. A. Wang, US Pat. 7132267, 2006.
146. S. C. Davis, J. H. Grate, D. R. Gray, J. M. Gruber, G. W. Huisman, S. K. Ma, L. M. Newman, R. Sheldon and L. A. Wang, US Pat. 7125693, 2006.
147. W. Hummel and B. Riebel, US Pat. 6413750, 2002.
148. C.-H. Wong and C. W. Bradshaw, US Pat. 5225339, 1992.
149. R. N. Patel, A. Banerjee, C. G. McNamee and L. J. Szarka, US Pat. 5391495, 1995.
150. J. C. Moore, M. G. Sturr, K. McLaughlin and J. Kim, US Pat. 7109004, 2006.
151. V. Niddam-Hildesheim, WO Pat. 2009045507, 2009.
152. E. Keinan, S. C. Sinha and A. S. Bagchi, J. Org. Chem., 1992, 57, 3631.
153. R. N. Patel, L. Chu and R. Mueller, Tetrahedron: Asymmetry, 2003, 14, 3105.
154. B. N. Zhou, A. S. Gopalan, F. VanMiddlesworth, W.-R. Shieh and C. J. Sih, J. Am. Chem. Soc., 1983, 105, 5925.
155. H. Gröger and W. Hummel, Org. Lett., 2003, 5, 173.
156. M. D. Truppo and D. Pollard, Org. Lett., 2007, 9, 335.
157. M. J. Kim, Y. Ahn and J. Park, Curr. Opin. Biotechnol., 2002, 13, 578.
158. O. Pàmies and J. E. Bäckvall, Chem. Rev., 2003, 103, 3247.
159. M. K. Srinivasu, B. M. Rao, S. S. Reddy, P. R. Kumar, K. B. Chandrasekhar and P. K. Mohakhud, J. Pharm. Biomed. Anal., 2005, 38, 320.
160. G. M. Reddy, V. V. N. K. V. P. Raju, K. Satyanarayana, V. Ravindra, S. Vyala and G. Venkateshwarlu, Chem. Biol. Interface, 2012, 2, 251.
161. T. D. Butters, R. A. Dwek and F. M. Platt, Chem. Rev., 2000, 100, 4683.
162. I. S. Kim, H. Y. Lee and Y. H. Jung, Heterocycles, 2007, 71, 1787.
163. M. Ravinder, T. N. Reddy, B. Mahendar and V. J. Rao, ARKIVOK (ix), 2012, 287.
164. H. Iida, N. Yamazaki and C. Kibayashi, J. Org. Chem., 1987, 52, 3337.
165. T. Schroder and M. Stubbe, US Pat. 4806650, 1989.
166. G. Kinast and M. Schedel, US Pat. 4266025, 1981.
167. R. M. Braga and L. Kato, J. Braz. Chem. Soc., 2003, 14, 822.
168. G. Kinast and M. Schedel, Angew. Chem., Int. Ed. Engl., 1981, 20, 805.
169. J. Behling and P. Farid, Synth. Commun., 1991, 21, 1383.
170. H. Setoi, H. Takeno and M. Hashimoto, Chem. Pharm. Bull., 1986, 34, 2642.
171. H. Setoi, H. Takeno and M. Hashimoto, Tetrahedron Lett., 1985, 26, 4617.
172. C. R. R. Matos, R. S. C. Lopes and C. C. Lopes, Synthesis, 1999, 4, 571.
173. P. Somfai and P. Marchand, Tetrahedron, 2003, 59, 1293.
174. H. Han, Tetrahedron Lett., 2003, 44, 1567.
175. D. L. Comins and A. B. Fulp, Tetrahedron Lett., 2001, 42, 6839.
176. Guidance for Industry (Nonclinical Studies for the Safety Evaluation of Pharmaceutical Excipients), U.S. Department of Health and Human Services, Food and Drug Administration, 2005, https://www.fda.gov/ohrms/dockets/98fr/2002d-0389-gdl0002.pdf.
177. J. Whittall and P. Sutton, Practical Methods for Biocatalysis and Biotransformations, John Wiley & Sons, Inc., Hoboken, 2010.
178. J. Tao, G. Q. Lin and A. Liese, Biocatalysis for the Pharmaceutical Industry: Discovery, Development, and Manufacturing, Wiley-VCH Verlag GmbH, Weinheim, 2008.