Metabolitos Primarios y Secundarios
Metabolitos Primarios y Secundarios
Humana Press/Springer,
New York, 2018.
En Prensa
Chapter 1
Advancement of Biotechnology by Genetic Modifications
Abstract
One of the greatest sources of metabolic and enzymatic diversity are microorrganisms. In recent
years, emerging recombinant DNA and genomic techniques have facilitated the development of
new efficient expression systems, modification of biosynthetic pathways leading to new
metabolites by metabolic engineering, and enhancement of catalytic properties of enzymes by
directed evolution. Complete sequencing of industrially important microbial genomes is taking
place very rapidly and there are already hundreds of genomes sequenced. Functional genomics
and proteomics are major tools used in the search for new molecules and development of
higher-producing strains.
1. Introduction
Advantages of microorganisms in the production of compounds, as compared to isolation from
plants and animals or synthesis by chemists, include (a) rapid uptake of nutrients that supports
high rates of metabolism and biosynthesis, (b) capability of carrying out a wide variety of
reactions, (c) facility to adapt to a large array of different environments, (d) ease of genetic
manipulation, both in vivo and in vitro, to increase production, to modify structures and activities,
and to make entirely new products, (e) simplicity of screening procedures, and (f) a wide
diversity.
Products from microbes are very diverse, ranging from very large molecules, such as
proteins, nucleic acids, carbohydrate polymers, or even cells, to small molecules that are
usually divided into primary metabolites, i.e., those essential for vegetative growth, and
secondary metabolites, i.e., those nonessential for growth.
1
Production of a particular primary metabolite by deregulated organisms may inevitably be
limited by the inherent capacity of the particular organism to make the appropriate biosynthetic
enzymes. Recent approaches utilize the techniques of modern genetic engineering to correct
such deficiencies and develop strains overproducing primary metabolites. There are two ways
to accomplish after this (a) to increase the number of copies of structural genes coding for these
enzymes and (b) to increase the frequency of transcription.
Novel genetic technologies are important for the development of overproducers. Ongoing
genome-sequencing projects involving hundreds of genomes, the availability of sequences
corresponding to model organisms, new DNA microarray and proteomic tools, as well as the
new techniques for mutagenesis and recombination described below, will no doubt accelerate
strain improvement programs.
Genome-based strain reconstruction achieves the construction of a superior strain, which
only contains mutations crucial to hyperproduction but not other unknown mutations, which
accumulate by brute-force mutagenesis and screening (1).
The directed improvement of product formation or cellular properties via modification of
specific biochemical reactions or introduction of new ones with the use of recombinant DNA
technology is known as metabolic engineering (2, 3). Analytical methods are combined to
quantify fluxes and to control them with molecular biological techniques to implement suggested
genetic modifications. The overall flux through a metabolic pathway depends on several steps,
not just a single rate-limiting reaction. Amino acid production is one of the fields with many
examples of this approach (4). Other processes improved by this technique include vitamins,
organic acids, ethanol, 1,3-propanediol, and carotenoids.
Reverse (inverse) metabolic engineering is another approach that involves choosing a
strain which has a favorable cellular phenotype, evaluating and determining the genetic and/or
environmental factors that confer that phenotype, and finally transferring that phenotype to a
second strain via direct modifications of the identified genetic and/or environmental factors (5,
6).
Molecular breeding techniques are based on mimicking natural recombination by in vitro
homologous recombination (7). DNA shuffling not only recombines DNA fragments but also
introduces point mutations at a very low controlled rate (8). Unlike site directed mutagenesis,
this method of pooling and recombining parts of similar genes from different species or strains
has yielded remarkable improvements in a very short amount of time (9). Whole genome
shuffling is another technique that combines the advantage of multiparental crossing allowed by
DNA shuffling with the recombination of entire genomes through recursive protoplast fusion (10,
11).
Systems biology is an integrated, systemic approach to the analysis and optimization of
cellular processes by introducing a variety of perturbations and measuring the system response
(12). Altered phenotypes are created by molecular biological techniques or by altering
environments. Further characterization of the phenotype leading to maximal product formation is
analyzed and quantified through the use of genome-wide high-throughput omics data and
genome-scale computational analysis.
2
Among the amino acids, L-glutamate and L-lysine, mostly used as feed and food additives,
represent the largest products in this category. Produced by fermentation are 1.5 million tons of
L-glutamate and 850,000 tons of L-lysine HCl. The total amino acid market was about 4.5 billion
dollars in 2004 (13). High titers of amino acids are shown in Table 1.
Table 1
High amino acids levels produced by fermentation
Amino acid Titer (g L-1)
L-Alanine 120
L-Arginine 96
L-Glutamate 150
L-Glutamine 49
L-Histidine 42
L-Hydroxyproline 41
L-Isoleucine 40
L-Leucine 34
L-Lysine HCl 170
L-Methionine 25
L-Phenylalanine 51
L-Proline 108
L-Serine 65
L-Threonine 124
L-Tryptophan 60
L-Tyrosine 55
L-Valine 51
1.2.1. L –Glutamate
Glutamate was the first amino acid to be produced by fermentation because of its use as a
flavoring agent (monosodium glutamate, MSG). The process employs various species of the
genera Corynebacterium and Brevibacterium. Molar yields of glutamate from sugar are 50–60%
and broth concentrations have reached 150 g L-1. Glutamic acid overproduction is feedback
regulated. However, mutant strains with modifications in the cell membrane are able to pump
glutamate out of the cell, thus allowing its biosynthesis to proceed unabated. Introduction of the
Vitreoscilla hemoglobin gene vgb into Corynebacterium glutamicum increased cell growth and
glutamic acid and glutamine production via increased oxygen uptake (14). Workers at Ajinomoto
Co., Inc. increased glutamate production from glucose by 9% by suppressing CO2 liberation in
the pyruvate dehydrogenase reaction (15). They did this by cloning the xfp gene-encoding
phosphoketolase from Bifidobacterium animalis into C. glutamicum and overexpressing it.
1.2.2. L-Lysine
Since the bulk of the cereals consumed in the world are deficient in L-lysine, this essential
amino acid became an important industrial product. Although the lysine biosynthetic pathway is
controlled very tightly in an organism such as Escherichia coli and in lysine-producing
organisms (e.g., mutants of C. glutamicum and its relatives), there is only a single aspartate
kinase (AK), which is regulated via concerted feedback inhibition by threonine plus lysine.
Metabolic engineering has been used in C. glutamicum to improve L–lysine production (4). A
chimeric AK, composed of the N-terminal catalytic region from Bacillus subtilis AKII and the C-
terminal region from Thermus thermophilus, was evolved through random mutagenesis and
then screened using a high-throughput synthetic RNA device which comprises an L-lysine-
sensing riboswitch and a selection module. Of three evolved aspartate kinases, the best mutant
(BT3) showed 160 % increased in vitro activity compared to the wild-type enzyme (16).
3
Comparative genome analysis between a wild-type strain and an L-lysine-producing C.
glutamicum strain identified three mutations that increased L-lysine production when introduced
into the wild-type strain. Introduction of the 6-phosphogluconate-dehydrogenase gnd mutation
increased yield by 15% (17). Further improvement was achieved by introducing the mqo
mutation (malate: quinone oxidoreductase) resulting in an increased titer of 95 g L-1 in a fed-
batch culture (18). Transcriptome analysis revealed that strain B-6, as compared to the wild
type, is upregulated in both the pentose phosphate pathway genes and the amino acid
biosynthetic genes and downregulated in TCA cycle genes. Lysine HCl titers have reached 170
g L-1 (19). Metabolic flux studies of wild-type C. glutamicum and improved lysine-producing
mutants showed that yield increased from 1.2% to 24.9% relative to the glucose flux. Different
approaches have been used to increase lysine production by C. glutamicum mutants including
(1) deletion of different genes from the glycolytic pathway (20), (2) improving the availability of
pyruvate by eliminating pyruvate dehydrogenase activity (21), (3) overexpression of pyruvate
carboxylase or DAP dehydrogenase genes, and (4) overexpression of gene NCg10855
encoding a methyltransferase or the amtA-ocd-soxA operon (22). A genetically defined C.
glutamicum L-lysine overproducer strain was developed by system metabolic engineering of the
wild type. Implementation of only 12 defined genome-based changes in genes encoding central
metabolic enzymes redirected major carbon fluxes as desired towards the optimal pathway
usage predicted by in silico modeling. The final engineered C. glutamicum strain was able to
produce lysine with a high yield of 0.55 g per gram of glucose, a titer of 120 g L-1 of lysine and a
productivity of 4.0 g L-1 h-1 in fed-batch culture (23).
1.2.3. L –Threonine
Overproduction of L-threonine has been achieved in Serratia marcescens by transductional
crosses, which combined several feedback control mutations in a single organism. This resulted
in titers up to 25 g L-1 (24). Combination in a single strain of another six regulatory mutations
derived by resistance to amino acid analogs led to desensitization and derepression of the key
enzymes of threonine synthesis. The resultant transductant produced 40 g L-1 of threonine (25).
The use of recombinant DNA technology led to strains that made 63 g L-1 threonine (26). An E.
coli strain was developed via mutation and genetic engineering and optimized by the
inactivation of threonine dehydratase (TD) resulting in a process yielding 100 g L-1 of L-
threonine in 36 h of fermentation (27).
In C. glutamicum ssp. lactofermentum, L-threonine production reached 58 g L-1 when a
strain producing both threonine and lysine was transformed with a plasmid carrying its own
hom, thrB, and thrC genes (28). A recombinant isoleucine auxotrophic strain of E. coli (carrying
extra copies of the thrABC operon, an inactivated tdh gene and mutated to resist high
concentrations of L–threonine and L-homoserine) produced 80 g L-1 L-threonine in 1.5 days
with a yield of 50% (29). Cloning extra copies of threonine export genes into E. coli also
increased threonine production (30).
Systems metabolic engineering was used to develop an L- threonine1-overproducing E. coli
strain (31). Feedback inhibition of aspartokinase I and III (encoded by thrA and lysC,
respectively) and transcriptional attenuation regulation (thrL) were removed. Deletion of tdh and
mutation of ilvA avoid threonine degradation. The metA and lysA genes were deleted to make
more precursors available for threonine biosynthesis. Further target genes to be engineered
were identified by transcriptome profiling combined with in silico flux response analysis. The
final engineered E. coli strain was able to reach a high yield of 0.393 g of threonine per gram of
glucose and 82.4 g L-1 in a fed-batch culture. With the use of an L-isoleucine leaky (ILEL) and α-
amino-β-hydroxyvaleric acid resistant (AHVr) E. coli TRFC strain, carrying a pTHR101 plasmid
containing a Thr operon, a threonine production of 118 g L-1 was achieved after 38 h with a
productivity of 3.1 g L-1 h-1 (46% conversion ratio from glucose to threonine) (32Chen et al.,
2009). A further increase in amino acid production to 124.57 g L-1 was obtained in TRFC strain
4
by decreasing acetic acid production using combined feeding strategies (33). Acetic acid is an
inhibitor of growth and threonine production.
1.2.4. L-Valine
Combined rational modification, transcriptome profiling, and systems-level in silico analysis was
used to develop an E. coli strain for the production of this amino acid (34). The ilvA, leuA, and
panB genes were deleted to make more precursors available for L–valine biosynthesis. This
engineered strain, harboring a plasmid overexpressing the ilvBN genes, produced 1.3 g L-1 L-
valine. Overexpression of the lrp and ygaZH genes (encoding a global regulator Lrp and L-
valine exporter, respectively), and amplification of the lrp, ygaZH, and lrp-ygaZH genes
enhanced production of L-valine by 21.6, 47.1, and 113%, respectively. Further improvement
was achieved by using in silico gene knockout simulation, which identified the aceF, mdh, and
pfkA genes as knockout targets. The VAMF strain (Val aceF mdh pfkA) was able to
produce 7.5 g L-1 L–valine from 20 g L-1 glucose in batch culture, resulting in a high yield of
0.378 g of L-valine per gram of glucose. Production by mutant strain VAL1 of C. glutamicum
amounted to 31 g L-1 (35). The mutant was constructed by overexpressing biosynthetic
enzymes via a plasmid, eliminating ilvA encoding threonine dehydratase, and deleting two
genes encoding enzymes of pantothenate biosynthesis. The culture was grown with limiting
concentrations of isoleucine and pantothenate. By applying metabolic engineering, a C.
glutamicum strain (WCC003 harboring pJYW-4-ilvBNC1-lrp1-brnFE) deleted in the genes aceE,
alaT and ilvA and overexpressing the ilvB, ilvN, ilvC, lrp1, brnF and brnE genes, produced 51 g
L-1 L-valine in fed-batch fermentation, with almost no detectable amino-acid by-products such as
L-alanine and L-isoleucine (36).
1.2.5. L–Isoleucine
Isoleucine processes have been devised in various bacteria such as S. marcescens, C.
glutamicum ssp. flavum, and C. glutamicum. In S. marcescens, resistance to isoleucine
hydroxamate and α-aminobutyric acid led to derepressed L-threonine deaminase (TD) and
acetohydroxyacid synthase AHAS) and production of 12 g L-1 of isoleucine (37). Further work
involving transductional crosses into a threonine-overproducer yielded isoleucine at 25 g L-1
(24).
Feedback regulation in C. glutamicum was eliminated (38) by replacing the native threonine
dehydratase gene ilvA with the feedback resistant gene from E. coli. By introducing additional
copies of genes encoding branched amino acid biosynthetic enzymes, lysine- or threonine-
producing strains were converted into L-isoleucine producers with improved titers (39– 41).
Amplification of the wild-type threonine dehydratase gene ilvA in a threonine-producing strain of
Corynebacterium lactofermentum led to isoleucine production (42).
A threonine-overproducing strain of C. glutamicum was sequentially mutated to resistance
to thiaisoleucine, azaleucine, and aminobutyric acid; it produced 10 g L-1 of isoleucine (43).
Metabolic engineering studies involving overexpression of biosynthetic genes were useful in
improving isoleucine production by this species. Colon et al. (42) obtained an isoleucine-
producing strain by cloning multiple copies of hom (encoding HDI), and wild-type ilvA (encoding
TD) into a lysine-overproducer, and by increasing HK (encoded by thrB); a titer of 15 g L-1
isoleucine was obtained. Independently, Morbach et al. (44) cloned three copies of the
feedback-resistant HD gene (hom) and multicopies of the deregulated TD gene (ilvA) in a
deregulated lysine producer of C. glutamicum, yielding an isoleucine producer (13 g L-1) with no
threonine production and reduced lysine production. Application of a closed loop control fed-
batch strategy raised production to 18 g L-1 (45). Further metabolic engineering work involving
amplification of feedback inhibition-insensitive biosynthetic enzymes converted lysine
5
overproducers and threonine overproducers into C. glutamicum strains yielding 30 g L-1 of
isoleucine (46).
C. glutamicum ssp. flavum studies employed resistance to α-amino- β-hydroxyvaleric acid
and the resultant mutant produced 11 g L-1 (47). D-Ethionine resistance was used by Ikeda et
al. (48) to yield a mutant producing 33 g L-1 in a fermentation continuously fed with acetic acid.
Accumulated L-isoleucine in the cell is excreted via a two-component export system BrnFE,
which is regulated by the global regulator Lrp. Overexpression of Lrp and BrnFE in C.
glutamicim led to 27 g L-1 production in fed-batch fermentation (49). A C. glutamiiun strain
overexpressing the threonine dehydratase (ilvA1), and acetohydroxy acid synthase (ilvBN1)
feed-back resistant genes, as well as the ppnk1 gene, encoding NAD kinase (Strain
IWJ001/pDXW-8-ilvBN1-ilvA1-ppnk1), led to a 32 g L-1 isoleucine production in a 72 h fed-batch
fermentation (50). Recently in a Brevibacterium flavum strain KM011 (Met-+Ethr+α-
ABr+LysL+AECr), betaine, vitamin B12 and Vitamin B5 concentrations were optimized for
isoleucine production in shake flasks. Under these conditions, a máximum of 13.35 g L-1
isoleucine were produced. However, a production of 35.26 g L-1 was obtained with optimized
conditions using a 5-L fermenter (51).
1.2.6. L-Alanine
Lee et al. (52) introduced into an E. coli double mutant (lacking genes encoding a protein of the
pyruvate dehydrogenase complex [aceF] and lactate dehydrogenase [ldhA], a plasmid
containing the Bacillus sphaericus alanine dehydrogenase gene (alaD). The strain produced L-
alanine in 27 h with a yield on glucose of 0.63 g g-1 and a maximum productivity of 2 g L-1 h-1.
Further work has raised the titer to 114 g L-1. A genetically engineered E. coli W (strain, XZ132),
with a D-lactate dehydrogenase replaced by an alanine dehydrogenase from Geobacillus
stearothermophilus and deletions in the methylglyoxal synthase (mgsA) and the catabolic
alanine racemase (dadX) genes, produced 1,279 mmol alanine (53). Deletion of both enzymes
reduced low levels of lactate and conversion of L- to D-alanine. A thermo-regulated genetic
switch designed to dynamically control the expression of L-alanine dehydrogenase (alaD) from
G. stearothermophilus on the E. coli B0016-060BC chromosome, led to an L-alanine titer of
120.8 g L-1 with higher overall and oxygen-limited volumetric productivities of 3.09 and 4.18 g L-1
h-1, respectively, using glucose as the sole carbon source (54).
1.2.7. L –Proline
Proline-hyperproducing strains of bacteria, exhibiting reduced proline-mediated feedback
inhibition of -glutamyl kinase (GK) activity (a result of single-base pair substitutions in the
bacterial proB gene coding region), have been isolated based on their resistance to toxic proline
analogs (L-azetidine-2-carboxylic acid and 3,4-dehydro-DL-proline), compounds which inhibit
GK activity while not interfering with protein synthesis. Cloning of the three genes of proline
biosynthesis in E. coli on multicopy plasmids and selection of mutants of such plasmid-
containing strains to resistance to 3,4-dehydroproline led to a process producing 20 g L-1 proline
(55).
A mutant of S. marcescens resistant to 3,4-dehydroproline, thiazolidine-4-carboxylate, and
azetidine-2-carboxylate and unable to utilize proline produced 50–55 g L-1 L-proline (56).
Cloning of a gene bearing the dehydroproline-resistance locus on a plasmid yielded a
recombinant strain of S. marcescens producing 75 g L-1 (57). Further development work
increased the production to over 100 g L-1 (58).
A sulfaguanidine-resistant mutant of C. glutamicum ssp. flavum produced 35 g L-1proline
(59). When a glutamate-producing strain of C. glutamicum was grown under modified
6
conditions, it made 48 g L-1 (60). A strain of Corynebacterium acetoacidophilum produced 108 g
L-1 proline when grown in the presence of glutamate (61).
1.2.8. L-Hydroxyproline
Introduction of the proline 4-hydroxylase gene from Dactylosporangium sp. into recombinant E.
coli producing L-proline at 1.2 g L-1 led to a new strain producing 25 g L-1 of hydroxyproline
(trans -4-hydroxy-L-proline) (62). When proline was added, the hydroxyproline titer reached 41 g
L-1, with a yield of 87% based on the amount of proline added. Optimization of the gene codons
and vectors of proline-4-4-hydroxylase (P4H) in a recombinant E. coli BL21 strain and chemo-
physical combination mutagenesis, allowed isolation of strain NA45, able to grow in glycerol as
a sole carbon source. With further systemic optimization of nutritional elements, the strain
produced 25.4 g L−1 trans-4-hydroxy-L-proline at 48 h in fed batch mode in a 5 L fermentor (63).
1.3. Nucleotides
Nucleotide fermentations became commercially important due to the activity of two purine
ribonucleoside 5’-monophosphates, namely guanylic acid (GMP) and inosinic acid (IMP), as
enhancers of flavor. Titers of IMP and GMP have reached 30 g L-1 (64). The techniques used to
achieve such production are similar to those used for amino acid fermentations. In Japan, 2,500
tons of GMP and IMP are produced annually with a market of $350 million. GMP can also be
made by bioconversion of xanthylic acid (XMP). Genetic modification of Corynebacterium
ammoniagenes involving transketolase, an enzyme of the nonoxidative branch of the pentose
phosphate pathway, resulted in the accumulation of 39 g L-1 of XMP (65).
1.4. Vitamins
Riboflavin. Production of riboflavin (vitamin B2) reached over 6,000 tons per year and a titer of
20 g L-1 by overproducers such as the yeast-like molds, Eremothecium ashbyii and Ashbya
gossypii. A bacterial process using C. ammoniagenes (previously Brevibacterium
ammoniagenes) was developed by cloning and overexpressing the organisms own riboflavin
biosynthetic genes (66) and its own promoter sequences. The resulting culture produced 15 g L-
1
riboflavin in 3 days.
Genetic engineering of a Bacillus subtilis strain, already containing purine analog-resistance
mutations, led to improved production of riboflavin (67). The industrial strain of B. subtilis was
produced by (1) making purine analog-resistance mutations to increase guanosine triphosphate
(GTP; a precursor) production and (2) using a riboflavin analog (roseflavin)-resistance mutation
in ribC that deregulated the entire pathway (68). Resultant production was over 25 g L-1. A
genome-wide transcript expression analysis (69) was successfully used to discover new targets
for further improvement of the fungus A. gossypii (70). The authors identified 53 genes of known
function, some of which could clearly be related to riboflavin production. A metabolic
engineering approach overexpressing the ribA (encoding 3,4-dihydroxy-2-butanone 4-
phosphate synthase) gene in B. subtilis RB50::[pRF69]n[pRF93] strain, led to productions of 15
g L-1 riboflavin (71).
Biotin has traditionally been made by chemical synthesis but recombinant microbes have
approached a competitive economic position. Cloning of a biotin operon (bioABFCD) on a
multicopy plasmid allowed E. coli to produce 10,000 times more biotin than did the wild-type
strain (72). Sequential mutation of S. marcescens to resistance to the biotin antimetabolite
acedomycin (= actithiazic acid) led to mutant strain SB412, which produced 20 mg L-1 of biotin
(73). Further improvements were made by mutating selected strains to ethionine resistance
(strain ET2, 25 mg L-1), then mutating ET2 to S-2-aminoethylcysteine resistance (strain ETA23,
33 mg L-1) and finally cloning in the resistant bio operon yielding a strain able to produce 500
mg L-1in a fed-batch fermention along with 600 mg L-1 of biotin vitamers. Later advances led to
production by recombinant S. marcescens of 600 mg L-1 of biotin (74). A process using an E.
7
coli mutant resistant to -hydroxynorvaline (a threonine antimetabolite) yielding 970 mg L-1 has
been patented (75). Biotin production was further increased to over 1 g L-1 by the use of a B.
subtilis strain resistant to 5-(2-thenyl) pentanoic acid (a biotin analog) and overexpressing
several bio genes.
Vitamin C (ascorbic acid) has traditionally been made in a five step chemical process by
first converting glucose to 2-keto-L–gulonic acid (2-KGA) with a yield of 50% and then
converting the 2-KGA by acid or base to ascorbic acid (L-AA). Annual production is 110,000
tons generating revenue of over 600 million dollars. A novel process for vitamin C synthesis
involved the use of a genetically engineered Erwinia herbicola strain containing a gene from
Corynebacterium sp. The engineered organism converted glucose into 1 g L-1 of 2-KGA (76). A
better process was devised independently which converted 40 g L-1 of glucose into 20 g L-1 of 2-
KGA (77). This process involved cloning and expressing the gene encoding 2, 5-diketo-D-
gluconate reductase from Corynebacterium sp. into Erwinia citreus. Another process used a
recombinant strain of Gluconobacter oxydans containing genes encoding L-sorbose
dehydrogenase and L-sorbosone dehydrogenase from G. oxydans T-100. The new strain was
an improved producer of 2-KGA (78). Further mutation to suppress the L-idonate pathway and
to improve the promoter led to the production of 130 g L-1 of 2-KGA from 150 g L-1 of sorbitol.
Further improvement of the strain was possible by suppressing the L-idonate pathway (79). An
Erwinia herbicola transformed with the 2,5-diketo-d-gluconate reductase gene from
Corynebacterium sp., produced up to 120 g L-1 2-KGA in less than 120 h with the help of
glucose dehydrogenase, gluconate dehydrogenase, ketogluconate dehydrogenase, and 2,5-
diketo-d-gluconate reductase (80). Similar to the chemical conversion of 2-KGA to vitamin C by
lactonization, X. campestris probably lactonizes 2-KGA under oxidative stress to form L-AA.
Using this method, 20.4 g L-1 L-AA were found in the extracellular broth which corresponds to
the 14-fold amount of that detected in yeast cells directly synthesizing L-AA from D-glucose
(81).
Vitamin B12 production depends on avoidance of feedback repression by vitamin B12. The
vitamin is industrially produced by Propionibacterium shermanii or Pseudomonas denitrificans.
The early stage of the P. shermanii fermentation is conducted under anaerobic conditions in the
absence of the precursor 5,6-dimethylbenzimidazole. These conditions prevent vitamin B12
synthesis and allow for the accumulation of the intermediate, cobinamide. Then, the culture is
aerated and dimethylbenzimidazole is added, converting cobinamide to the vitamin. In the P.
denitrificans fermentation, the entire process is carried out under low levels of oxygen. A high
level of oxygen results in an oxidizing intracellular environment, which represses formation of
the early enzymes of the pathway. Production of vitamin B12 has reached levels of 150 mg L-1,
10 tons per year, and a world market of $71 million.
Other vitamins. Recombinant E. coli, transformed with genes encoding pantothenic acid
(vitamin B5) biosynthesis, and resistant to salicylic and/or other acids, produce 65 g L-1 of D–
pantothenic acid from glucose using alanine as precursor (82). A total of 7,000 tons per year are
made chemically and microbiologically. Thiamine (vitamin B1) is produced synthetically at 4,000
tons per year. Pyridoxine (vitamin B6) is made chemically at 2,500 tons per year.
1.5. Carotenoids
Carotenoid production processes have been extensively studied (83), but they have had
difficulty in economically challenging chemical methods. Of over 600 microbial carotenoids, only
-carotene and astaxanthin are produced industrially by fermentation (84). A semi-industrial -
carotene process was developed using mated cultures of Blakeslea trispora plus and minus
strains. -Carotene was produced at 1 g L-1 in the early 1960s (85). By the addition of
carotogenic chemicals and antioxidants, the titer was raised to over 3 g L-1 (86). Processes in
development include those yielding -carotene, lycopene, zeaxanthin, and astaxanthin. Some
have been improved by metabolic engineering and directed evolution (87–89). Metabolic
8
engineering of E. coli has led to strains forming 0.2 g L-1 of lycopene (90). Lutein, a xanthophyll
carotenoid with antioxidant properties, had sales as a food colorant of $150 million in the USA
(91). It is thought that this carotenoid prevents age-related macular degeneration and cataracts.
It is made from petals of marigold, but microalgae are a potential new source.
Chemical production of trans-astaxanthin has a selling price of $2,000 kg (92) and a market
of over $100 million per year. It is mainly used for pigmentation of salmonids raised in
aquaculture, a multibillion dollar industry (93). Astaxanthin can be made by the yeast Phaffia
rhodozyma (Xanthophyllomyces dendrorhous) and the microalga Haematococcus pluvialis.
Genetically improved strains of P. rhodozyma produce 10 mg g-1 cells in industrial fermentors.
Recent improvements in astaxanthin production have been published by de la Fuente et al. (94)
and Rodríguez-Sáiz et al. (95) to get maximal astaxathin titers of 420 mg L-1 when X.
dendrorhous is fermented under continuous white light. Recently astaxanthine production has
been reported with an engineered C. glutamicum strain. Volumetric productivities of up to about
0.4 mg L-1 h-1 reported in simple shaking flask cultures by the recombinant strain compare
favorably with those reported for the commercially used production hosts such as the green
microalga H. pluvialis, the red yeast X. dendrorhous and the recombinant E. coli. (96).
9
tolerance (160–200 g L-1 glucose), while simultaneously enhancing L-lactic acid production by
71% as compared to the wild type. Shuffling of a mutant strain of Lactobacillus delbrueckii NCIM
2025 and Bacillus amyloliquefaciens ATCC 23842 produced a fusant that could utilize liquefied
cassava bagasse starch directly to yield a titer of 40 g L-1 of lactic acid with a 96% conversion of
starch to lactic acid (106).
Although lactobacilli make more lactic acid than Rhizopus oryzae, they produce mixed
isomers. The fungus, however, produces L-(+) lactic acid exclusively. The yield is about 60–
80% of added glucose, the remainder going to ethanol. By increasing lactic dehydrogenase
levels via plasmid transformation with ldhA, more lactate could be made from pyruvate and
production was increased to 78 g L-1, whereas the undesirable coproduct ethanol was reduced
from 10.6 to 8.7 g L-1 (107). A recombinant Saccharomyces cerevisiae strain containing six
copies of bovine L-lactate dehydrogenase produced 122 g L-1 from sugar cane with an optical
purity of 99.9% or higher (108). Expression of the same bovine enzyme and a deletion of the
pyruvate decarboxylase gene in Kluyveromyces lactis produced 109 g L-1 (109).
A recombinant E. coli strain was constructed that produced optically active pure D-lactic
acid from glucose at virtually the theoretical maximum yield, e.g., two molecules from one
molecule of glucose (110). D-Lactic acid has also been produced at 61 g L-1 by a recombinant
strain of S. cerevisiae containing the D-lactic dehydrogenase gene from Leuconostoc
mesenteroides (111).
1.7. Alcohols
Ethanol. Fermentation of sugars by S. cerevisiae in the case of hexoses, and Kluyveromyces
fragilis or Candida species with lactose or a pentose, results in the production of ethanol. Under
optimum conditions, approximately 120 g L-1 ethanol can be obtained. Such a high
concentration slows down growth and the fermentation ceases.
10
A S. cerevisiae fusant library obtained by genome shuffling was screened for growth at 35,
40, 45, 50, and 55°C on agar plates containing different concentrations of ethanol (119). After
three rounds of genome shuffling, a strain was obtained which was able to grow on plates up to
55°C, completely utilized 20% (w/v) glucose at 45–48°C, produced 99 g L-1 ethanol, and
tolerated 25% (v/v) ethanol stress.
In silico metabolic models have been used to overcome the redox imbalance in S.
cerevisiae engineered with the Xyl1 and Xyl2 genes from Pichia stipitis (120, 121).
Overexpression of both genes led to an accumulation of NADH and a shortage of NADPH.
Deletion of NADP+-dependent glutamate dehydrogenase (GGH1) and overexpression of NAD+-
dependent GDH2 led to an increase in ethanol production using xylose as fermentation
substrate. An in silico genome-scale gene insertion strategy was used to improve ethanol
production and decrease the production of byproducts glycerol and xylitol (122). Introduction of
glyceraldehyde-3-phosphate dehydrogenase in S. cerevisiae led to a 58% reduction in glycerol,
a 33% reduction in xylitol, and a 24% increase in ethanol production.
When biomass is used as a carbon source for ethanol production, its breakdown results in
liberation of acetic acid. The acid interferes with ethanol production. Tolerance of Candida
krusei GL560 to acetic acid was improved by genome shuffling (123). A mutant, S4-3, which
was isolated and selected after four rounds, had a higher viability in different media containing
acetic acid than did the parent strain GL560. The mutant also improved its multiple stress
tolerance to ethanol, H2O2, heat, and freeze-thawing.
E. coli was converted into an ethanol producer (43 g L-1) by cloning the alcohol
dehydrogenase II and pyruvate decarboxylase genes from Zymomonas mobilis (124). By
cloning and expressing the same two genes in Klebsiella oxytoca, the recombinant was able to
convert crystalline cellulose to ethanol in high yield when fungal cellulase was added (125).
Maximum theoretical yield was 81–86% and a titer of 47 g L-1 of ethanol was produced from 100
g L-1 of cellulose.
Recombinant strains of E. coli, Zymomonas, and Saccharomyces can convert corn fiber
hydrolysate to 21–35 g L-1 ethanol with yields of 0.41–0.50 ethanol per gram of sugar consumed
(126). For a recombinant E. coli strain making 35 g L-1, time was 55 h and yield was 0.46 g
ethanol per gram of available sugar, which is 90% of the attainable maximum.
1,3-Propanediol. A strain of Clostridium butyricum converts glycerol to 1,3-propanediol
(PDO) at a yield of 0.55 g g-1glycerol consumed (127). A major metabolic engineering feat was
carried out in E. coli leading to a culture growing on glucose and producing PDO at 135 g L-1,
with a yield of 51% and a rate of 3.5 g L-1 h-1 (128). To do this, eight new genes were introduced
to convert dihydroxyacetone phosphate (DHAP) into PDO. Production was further improved by
modifying 18 E. coli genes, including regulatory genes. PDO is the monomer used to chemically
synthesize industrial polymers such as polyurethanes and the polyester fiber Sorono™ by
DuPont. This bioplastic is polytrimethylene terephthalate (3GT polyester) made by reacting
terephthalic acid with PDO (129). PDO is also used as a polyglycol-like lubricant and as a
solvent.
D-Mannitol is a naturally occurring polyol, widely used in the food, chemical, and
pharmaceutical industries. A whole cell bioconversion of D-fructose to D-mannitol was
developed by metabolic engineering of E. coli (130). The mdh gene encoding mannitol
dehydrogenase from Leuconostoc pseudomesenteroides and the fdh gene encoding formate
dehydrogenase from Mycobacterium vaccae were coexpressed in E. coli along with the glf gene
encoding the glucose facilitator protein of Z. mobilis. The process yielded 75–91 g L-1 of
D-mannitol, a specific productivity of 3.1–4.1 g g-1 h-1 and a molar yield of 84–92% with no by-
products. An improved bioconversion process was developed with a recombinant E. coli strain
in the presence of added glucose isomerase yielding 145 g L-1 of D-mannitol from 180 g L-1
glucose (131). Supplementation of the medium used for mannitol production by Candida
magnolia with Ca2+ and Cu2+ increased production up to 223 g L-1 (132).
11
Sorbitol, also called D-glucitol, is 60% as sweet as sucrose and is used in the food,
pharmaceutical, and other industries. Its worldwide production is estimated to be higher than
500,000 tons per year, and it is made chemically by catalytic hydrogenation of D-glucose or
syrup with a 50:50 mixture of glucose and fructose. It is also produced by extraction from
seaweed as a by-product of alginate and iodine manufacture. However, excellent microbial
processes have been developed (133). Toluenized (permeabilized) cells of Z. mobilis produce
290 g L-1 of sorbitol and 283 g L-1 of gluconic acid from a glucose and fructose mixture in 16 h
with yields near 95% for both products (134). Other leading organisms are recombinant C.
glutamicum at 285 g L-1, Lactobacillus intermedius at 227 g L-1, C. magnoliae at 223 g L-1, and
many others producing between 100 and 200 g L-1. Metabolic engineering of Lactobacillus
plantarum for high sorbitol production was successfully achieved by a simple two-step strategy
overexpressing the two sorbitol-6-phosphate dehydrogenase genes (srlD1 and srlD2) identified
in the genome sequence (135).
n-Butanol is a good alternative fuel additive as it has two more carbons than ethanol, which
results in an energy content about 40% higher. Also, automobile engines do not require
modification until the percentage of butanol reaches over 40% of the total automobile fuel. In
contrast, modification is required when ethanol is added to gasoline at levels exceeding 15%.
Butanol can be obtained from the acetone–butanol–ethanol fermentation of Clostridium
beijerinkii or Clostridium acetobutylicum. Butanol-resistant mutants showed increased
production of butanol and acetone (136). Biochemical engineering modifications were able to
increase total acetone, butanol, and ethanol production (ABE) to 69 g L-1 (137). A mutant in the
presence of added acetate was able to produce almost 21 g L-1 butanol and 10 g L-1 of acetone
from glucose (138).
Because butanol’s octane number is lower than that of ethanol and the octane number
increases with methyl branching and double bonds, other higher alcohols are also being
considered as biofuels, e.g., branched C 4 and C 5 alcohols. They are also desirable because
of their higher energy density, lower vapor pressure, and lower hygroscopicity as compared to
ethanol (139). They include isopropanol, 1-propanol, 1-butanol (n-butanol), isobutanol (2-
methyl-propanol), 3-methyl-1-butanol, 2-methyl-1-butanol, isopentanol (3-methyl-1 butanol), and
isopentenol (3-methyl-3-buten-1-ol). A novel screening method based on overcoming the
toxicity associated with the accumulation of prenyl diphosphate was used to screen a library of
19,000 clones harboring fragments of Bacillus genomic DNA (140). Two genes, yhfR and nudF,
coding for proteins capable of overcoming the toxicity associated with accumulating IPP and
DMAPP were isolated. Both protein products have an affinity for IPP and DMAPP, converting
them into isopentenol. Clostridium beijerinckii (Clostridium butylicum) produces 20 g L-1 of 1-
butanol and 2 g L of isopropanol as part of a mixed product. Recombinant E. coli can produce
4.9 g L-1 of isopropanol. Recently, a new strategy for the production of these alcohols has been
reported (141). This approach is based on the diversion of 2-keto acid intermediates from the
endogenous amino acid pathway to alcohol biosynthesis especially that of isobutanol. As a
result, engineered E. coli can produce 22 g L-1 of isobutanol in 110 h with a yield of 86% of the
theoretical maximum.
Other alcohols. The noncariogenic, noncaloric, and diabetic-safe sweetener erythritol has
70–80% of the sweetness of sucrose. It can be produced by Aureobasidium sp. (165 g L-1), an
acetate-negative mutant of Yarrowia lipolytica (170 g L-1), a C. magnoliae osmophilic mutant
(187 g L-1), the osmophile Trichosporon sp. (88 g L-1), Torula sp. (200 g L-1), and the yeast
Pseudozyma tsukubaensis (245 g L-1) (142).
Xylitol is a naturally occurring sweetener with anti-cariogenic properties, which is used for
some diabetes patients. It can be produced by chemical reduction of D-xylose or by
fermentation. Xylitol production at 150 g L-1 was obtained with Candida guilliermondii 2,581 at
pH 6.0 and shaking at 60 rpm (143).
12
2. Secondary Metabolites
As a group that includes antibiotics, pesticides, pigments, toxins, pheromones, enzyme
inhibitors, immunomodulating agents, receptor antagonists and agonists, pesticides, antitumor
agents, immunosuppressants, cholesterol-lowering agents, plant protectants, and animal and
plant growth factors, these metabolites have tremendous economic importance. This
remarkable group of compounds is produced by certain restricted taxonomic groups of
organisms and is usually formed as mixtures of closely related members of a chemical family.
2.1. Antibiotics
Antibiotics are the most well-known secondary metabolites and have a tremendous importance.
The most well known are the β-lactams, tetracyclines, aminoglycosides, chloramphenicol,
macrolides, and other polyketides, polyenes, glycopeptides, among others. They have been
crucial in the increase in average life expectancy in the USA from 47 years in 1900 to 74 for
males and 80 for women in 2000 (144). More than 350 agents have reached the market as
antimicrobials. The global market for finished antibiotics has reached $35 billion.
2.1.1. β–Lactams
The β-lactams are the most important class of antibiotics in terms of use. Included are the
penicillins, cephalosporins, cephamycins, clavulanic acid, and the carbapenems. Many of the
current penicillins and cephalosporins are semisynthetic. All of the above are of great
importance in chemotherapy of bacterial infections.
β-Lactamases of pathogenic bacteria are the major cause of resistance development and
there are over 450 such enzymes. However, β-lactams are still very useful due to the discovery
of β-lactamase inhibitors. Although clavulanic acid is a β-lactam compound, it has only low
antibacterial activity, but is used widely as an inhibitor of β-lactamase, in combination with β-
lactam antibiotics. Conventional strain improvement by protoplast fusion of auxotrophic strains
yielded a fusant producing 30-fold more clavulanic acid than the wild type (145). Inactivation of
two G-3-P dehydrogenases, encoded by gap1 and gap2 by targeted gene disruption, doubled
clavulanic acid production (146). Also, increased dosage of biosynthetic genes ceas and cs2
(147) or overexpression of positive regulatory genes increased production two- to threefold
(148, 149).
Yield improvements have been achieved through different strategies. Thus, protoplast
fusion between strains of Penicillium chrysogenum yielded a higher-producing strain of penicillin
G (150). Metabolic engineering was used to replace the normal promoter with the ethanol
dehydrogenase promoter (151), increasing penicillin production up to 30-fold.
Protoplast fusion was also been carried out with strains of Acremonium chrysogenum to
obtain a strain that produced 40% more cephalosporin C than the parent (152). Production was
also improved by cloning multiple copies of cyclase (153) or the pcbC and the cefEF genes
(154).
Two improved cephamycin C-producing strains from Nocardia lactamdurans were fused to
obtain cultures, which produced 10–15% more antibiotic (155). Overexpression of lat, encoding
lysine-aminotransferase also led to an overproducing strain (156). High-level expression of
ccaR, a positive regulatory gene in Streptomyces clavuligerus (157) led to a two- to threefold
increase in antibiotic production.
Chemical methods had traditionally been used to produce 7-aminocephalosporanic acid (7-
ACA) and 7-aminodeacetoxycepalosporanic acid (7-ADCA), chemical precursors of
semisynthetic cephalosporins. These processes are being replaced by safer microbiological
processes. Transformation of P. chrysogenum with bacterial cefD and cefE genes allowed the
production of deacetoxycephalosporin C (DAOC) (158), another key intermediate in the
commercial production of semisynthetic cephalosporins. Also, cloning of cefE from S.
clavuligerus or cefEF and cefG from Acremonium chrysogenum into P. chrysogenum fed with
13
adipic acid as side-chain precursor (159) resulted in the formation of several adipyl-6/7-
intermediates. Enzymatic removal of the adipoyl side chain led to the production of 7-ADCA.
Disruption and one-step replacement of the cefEF gene of an industrial strain of A.
chrysogenum yielded strains accumulating up to 20 g L-1 of penicillin N. Cloning and expression
of the cefE gene from S. clavuligerus into those high-producing strains yielded recombinant
strains producing high titers of DAOC (160). An E. coli strain containing the D-amino acid
oxidase gene from Trigonopsis variabilis and the glutaryl-7-aminocephalosporanic acid acylase
gene from Pseudomonas sp. was able to convert cephalosporin C directly to 7-ACA (161).
Natural carbapenems, such as thienamycin, are made by Streptomyces cattleya , Erwinia
carotovora subsp. carotovora , Serratia sp., and Photorhabdus luminescens (162).
Carbapenems are resistant to attack by most -lactamases. The commercial carbapenems are
made synthetically and include imipenem, meropenem, and ertapenem. Thienamycin is one of
the most potent and broadest in spectrum of all antibiotics known today. Although a -lactam, it
is not a member of the penicillins or cephalosporins. It is active against aerobic and anaerobic
bacteria, both Gram positive and Gram negative, including Pseudomonas. This novel structure
was isolated in Spain from a new soil species, which was named S. cattleya (163). Interestingly,
this culture also produces penicillin N and cephamycin C. Also used to combat β -lactamase
containing pathogens are new carbapenems, such as the recently approved doripenem (S-
4661) which has broad spectrum activity against resistant bacteria including Pseudomonas
aeruginosa.
14
Their use became severely limited by an increase in multidrug resistance. One group of narrow-
spectrum compounds are the streptogramins which are synergistic pairs of antibiotics made by
a single microbial strain. The pairs are constituted by a (Group A) polyunsaturated macrolactone
containing an unusual oxazole ring and a dienylamide fragment and a (Group B) cyclic
hexadepsipeptide possessing a 3-hydroxypicolinoyl exocyclic fragment. Such streptogramins
include virginiamycin and pristinamycin (179). Pristinamycin, made by Streptomyces
pristinaespiralis, is a mixture of a cyclodepsipeptide (pristinamycin I) and a polyunsaturated
macrolactone (pristinamycin II). Increasing resistance to pristinamycin in the pristinamycin
producer S. pristinaespiralis was combined with genome shuffling to increase pristinamycin
production ninefold (180).
An important new strategy to improve the discovery of new antibiotics is genome mining, which
has come about due to advances in microbial genomics (181). Mining of whole genome
sequences and genome scanning allows the rapid identification of more than 450 clusters of
genes in antibiotic-producing cultures encoding biosynthesis of new bioactive products and the
prediction of structure based on gene sequences (182, 183). These efforts include mining of
whole genome sequences, genome scanning, heterologous expression, and discovery of novel
chemistry. Genomics will also provide a huge group of new targets against which natural
products can be screened (184). Up to February 2017, more than 16,000 complete microbial
genome sequences were available in one or two scaffolds. Of them, 411 belong to fungi, and
699 to actinobacteria. Currently, 69 streptomycete genomes have been sequenced (185). The
smaller genome corresponds to Stretomyces xiamenensis with 5.95 Mb size and the biggest is
Streptomyces rapamyonicus with 12.7 Mb.
Consensus sequences can be obtained through alignments from the enzyme domains
and used to construct Hidden Markov Models (HMM) of the chemical structure to predict
SMILES. For this purpose, many online and stand-alone sources are freely available. Some of
them include antiSMASH 2.0 (186), which detects 24 types of secondary metabolites.
CLUSEAN, NaPDoS, NP.search, Bagel2 are used for lantipeptides, while PKS/NRPS Analysis,
SBSPKS (187) and SMURF pipeline were developed for fungal systems. AntiSMASH is one of
the most popular softwares since it is friendly and easy going.
15
Cloning of the actI, actIV and actVII genes from Streptomyces coelicolor into the 2-
hydroxyaklavinone producer, S. galilaeus 31,671 yielded novel hybrid metabolites,
desoxyerythrolaccin and 1- O -methyl-desoxyerythrolaccin (193). Similar studies yielded the
novel metabolite aloesaponarin II (194).
Epirubicin (4´-epidoxorubicin) is a semisynthetic anthracycline with less cardiotoxicity than
doxorubicin (195). Genetic engineering of a blocked S. peucetius strain provided a new method
to produce it (196). The gene introduced was avrE of the avermectin-producing Streptomyces
avermitilis or the eryBIV genes of the erythromycin producer, Saccharopolyspora erythrea.
These genes and the blocked gene in the recipient are involved in deoxysugar biosynthesis.
Taxol ®, a diterpene alkaloid, is approved for breast and ovarian cancer and acts by
blocking depolymerization of microtubules. In addition, taxol promotes tubulin polymerization
and inhibits rapidly dividing mammalian cancer cells (197). Taxol was originally isolated from the
bark of the Pacific yew tree (Taxus brevifolia), but it took six trees of 100 years of age to treat
one cancer patient (198). It is now produced by plant cell culture or by semisynthesis from
taxoids made by Taxus species. Early genetic engineering of S. cerevisiae yielded no taxadiene
(the taxol precursor) because too little of the intermediate, geranylgeranyl diphosphate, was
formed. When the Taxus canadensis geranylgeranyl diphosphate synthase gene was
introduced, 1 mg L-1 of taxadiene was obtained (199). More recent metabolic engineering
studies (200) yielded a S. cerevisiae strain producing over 8 mg L-1 taxadiene and 33 mg L-1
geranylgeraniol. Taxol has sales of $1.6 billion per year.
Epothilones are an important group of new anticancer agents produced by the
myxobacterium, Sorangium cellulosum (201). Rounds of classical mutation and screening
followed by recursive protoplast fusion resulted in fusants able to produce 130-fold more
epothilone B compared to the starting strain. Epothilones have a mode of action similar to Taxol
and, very importantly, are active against Taxol-resistant tumors.
2.4. Antihelmintics
Microbially produced polyethers such as monensin, lasalocid, and salinomycin dominate the
coccidiostat market.
The avermectins, a family of secondary metabolites having both antihelmintic and
insecticidal activities, produced by S. avermitilis, have a market of over 1 billion dollars per year.
Despite their macrolide structure, avermectins lack antibiotic activity, do not inhibit protein
synthesis nor are they ionophores; instead, they interfere with neurotransmission in many
invertebrates.
16
Although the Merck Laboratories had earlier developed a commercially useful synthetic
product, thiobenzole, they had enough foresight to also examine microbial broths for
antihelmintic activity and found a nontoxic fermentation broth which killed the intestinal
nematode, Nematospiroides dubius, in mice. The S. avermitilis culture, which was isolated by
Omura and coworkers at the Kitasato Institute in Japan (207), produced a family of secondary
metabolites having both antihelmintic and insecticidal activities. These were discovered by
Merck scientists and named “avermectins.” They are disaccharide derivatives of macrocyclic
lactones with exceptional activity against parasites, i.e., at least ten times higher than any
synthetic antihelmintic agent known. They have activity against both nematode and arthropod
parasites in sheep, cattle, dogs, horses, and swine.
Incorporation of multiple copies of afsR2, a global regulatory gene, from S. lividans into
wild-type S. avermitilis increased avermectin production by 2.3-fold (208). Transposon
mutagenesis was used to eliminate production of the troublesome toxic oligomycin in S.
avermitilis (209). DNA shuffling of the ave C gene of S. avermitilis gave an improved ratio of the
undesirable CHC-B2 to the desirable CHC-B1 of 0.07:1. This was an improvement of 21-fold
over the ratio with the starting strain (210).
A semisynthetic derivative, 22, 23-dihydroavermectin B1 (Ivermectin) is 1,000 times more
active than thiobenzole and is a commercial veterinary product. Ivermectin is made by
hydrogenation at C22–C23 of avermectin B1a and B1b with rhodium chloride acting as catalyst.
By genetic engineering of S. avermitilis, in which certain PKS genes were replaced by genes
from the PKS of S. venezuelae (the pikromycin producer), ivermectin could be made directly by
fermentation, thus avoiding semisynthesis (211).
A new avermectin, called Doramectin (=cyclohexyl avermectin B1), was developed at Pfizer
by the technique of mutational biosynthesis (212). Indeed, it was the first commercially
successful example of mutational biosynthesis.
2.6. Bioinsecticides
The insecticidal bacterium, Bacillus thuringiensis (BT), owes its activity to its crystal protein
produced during sporulation. Crystals plus spores had been applied to plants for years to
protect them against lepidopteran insects. BT preparations are highly potent, some 300 times
more active on a molar basis than synthetic pyrethroids and 80,000 times more active than
17
organophosphate insecticides. In 1993, BT represented 90% of the biopesticide market and had
annual sales of $125 million.
A very important insecticide is Spinosad (Naturalyte ®) produced by Saccharopolyspora
spinosa and used for protection of crops and feedstock animals. Spinosad is a mixture of two
tetracyclic macrolides containing forosamine and tri- O -methyl rhamnose with different levels of
methylation on the polyketide moiety. The two components are spinosyns A and D, which differ
by a methyl group at position 6 of the polyketide. Spinosad is an excellent nontoxic agricultural
insecticide. Genome shuffling has been used for strain improvement (218).
3.1. Bacteria
Bacterial systems are used to make somatostatin, insulin, bovine growth hormone for veterinary
applications, −1 antitrypsin, interleukin-2, tumor necrosis factor, -interferon, and -interferon.
E. coli is one of the earliest and most widely used hosts for the production of heterologous
proteins (224). As early as 1993, recombinant processes of E. coli were responsible for almost
$5 billion worth of products, i.e., insulin, human growth hormone, interferons, and G-CSF.
Advantages of E. coli include rapid growth, rapid expression, ease of culture and genome
modifications, low cost, and high product yields (225). It is used for massive production of many
commercialized proteins. This system is excellent for functional expression of non-glycosylated
proteins.
E. coli genetics are far better understood than those of any other microorganism. Recent
progress in the fundamental understanding of transcription, translation, and protein folding in E.
coli, together with the availability of improved genetic tools, are making this bacterium more
valuable than ever for the expression of complex eukaryotic proteins. Its genome can be quickly
and precisely modified with ease, promotor control is not difficult, and plasmid copy number can
be readily altered. This system also features alteration of metabolic carbon flow, avoidance of
incorporation of amino acid analogs, formation of intracellular disulfide bonds, and reproducible
18
performance with computer control. E. coli can accumulate recombinant proteins up to 80% of
its dry weight and survives a variety of environmental conditions. Recombinant protein
production in E. coli can be increased by mutations which eliminate acetate production (226).
Avecia Biologics achieved a titer of 14 g L-1 of recombinant protein using E. coli (227).
The value of transcriptome analysis in process improvement, was shown by Choi et al.
(228). They analyzed an E. coli process yielding human insulin-like growth factor 1 fusion
protein (IGF-If) in a high density culture. Of 200 or so genes whose expression was
downregulated after induction, the ones involved in biosynthesis of amino acids or nucleotides
were studied. Amplification of two of these, prsA (encoding PRPP synthetase) and glpF
(encoding the glycerol transporter), raised product formation from 1.8 to 4.3 g L-1.
Bacilli have yields as high as 3 g L-1. The organisms used are usually Bacillus megaterium,
B. subtilis, and Bacillus brevis. Staphylococcus carnosus can produce 2 g L-1 of secreted
mammalian protein.
An improved Gram-negative host for recombinant protein production has been developed
using Ralstonia eutropha (229). The system appears superior to E. coli with respect to inclusion
body formation. Organophosphohydrolase, a protein prone to inclusion body formation with a
production of less than 100 mg L-1 in E. coli, was produced at 10 g L-1 in R. eutropha. Another
useful bacterium is Pseudomonas fluorescens MB101 developed by Dowpharma (230). This
system has produced 4 g L-1 of TNF-α.
3.2. Yeasts
Yeasts, the single-celled eukaryotic fungal organisms, are often used to produce recombinant
proteins that are not produced well in E. coli because of problems dealing with folding or the
need for glycosylation. The major advantages of yeast expression systems are listed in Table 2.
The yeast strains are genetically well characterized and are known to perform many
posttranslational modifications. They are easier and less expensive to work with than insect or
mammalian cells and are easily adapted to fermentation processes.
The two most utilized yeast strains are S. cerevisiae and the methylotrophic yeast P.
pastoris. Glucose oxidase from A. niger is produced by S. cerevisiae at 9 g L-1. Recombinant
products on the market which are made in S. cerevisiae are insulin, hepatitis B surface antigen,
urate oxidase, glucagons, granulocyte macrophage colony-stimulating factor (GM-CSF), hirudin,
and platelet-derived growth factor.
P. pastoris has the desirable qualities of dense growth and methanol-induced expression
and secretion of recombinant proteins. It is used for the commercial production of non-
glycosylated human serum albumin and glycosylated vaccines. Strains have been developed
which are capable of human type N-glycosylation and such products are already in clinical
testing. High recombinant protein yields can be obtained with P. pastoris, e.g., 10 g L-1 of tumor
necrosis factor, (231) 14.8 g L-1 of gelatin, 15 g L-1 of mouse collagen (232), and E. coli phytase
and Candida parapsilosis lipase/acetyltransferase at 6 g L-1. Bacterial proteins such as
intracellular tetanus fragment C were produced as 27% of protein with a titer of 12 g L-1 (233).
Production of serum albumin in S. cerevisiae amounted to 0.15 g L-1, whereas in P. pastoris the
titer was 10 g L-1 (234). Indeed, claims have been made that P. pastoris can make 20–30 g L-1 of
recombinant proteins (235).
Table 2
Advantages of yeast expression systems
Stable strains
High density of growth
Durability
High production titers and yields
Protein glycosylation
19
Reasonable cost
Product processing similar to that of mammalian cells
Suitable for isotopically labeled proteins
Suitable for S–S-rich proteins
Protein folding assisted
Chemically defined media supporting rapid growth
20
from 5–50 mg L-1 in 1985 to 50–500 mg L-1 1in 1995 and to 5 g L-1 in 2005 (248). A number of
mammalian processes are producing 3–5 g L-1 of recombinant protein (249) and in some cases,
protein titers have reached 10 g L-1 in industry (250), including antibodies (251). A rather new
system is that of a human cell line known as PER.C6 of Crucell Holland BV, which, in
cooperation with DSM Biologics, was reported to produce 15 g L-1 (252) and then later, 27 g L-1
of a monoclonal antibody (253). Protein production of over 20 g L-1 has been achieved in serum-
free medium, but the production of 2–3 g L-1 in such media is more usual.
4. Enzymes
Over the years, high titers of enzymes were obtained using “brute force” mutagenesis and
random screening of microorganisms. Recombinant DNA technology acted as a boon for the
enzyme industry in the following ways (261): (1) plant and animal enzymes could be made by
microbial fermentations, e.g., chymosin; (2) enzymes from organisms difficult to grow or handle
genetically were now produced by industrial organisms such as species of Aspergillus and
Trichoderma, and K. lactis, S. cerevisiae, Y. lipolytica, and Bacillus licheniformis (e.g.,
thermophilic lipase was produced by A. oryzae and Thermoanaerobacter cyclodextrin glycosyl
transferase by Bacillus); (3) enzyme productivity was increased by the use of multiple gene
copies, strong promoters, and efficient signal sequences; (4) production of a useful enzyme
from a pathogenic or toxin-producing species could now be done in a safe host; and (5) protein
engineering was employed to improve the stability, activity, and/or specificity of an enzyme.
Genes encoding many microbial enzymes have been cloned and the enzymes expressed at
levels hundreds of times higher than those naturally produced. Over 60% of the enzymes used
in different applications including detergent, food and starch processing industry are
21
recombinant proteins (262). Recombinant molds are one of the main sources of enzymes for
industrial applications. Yields as high as 4.6 g L-1 have been reached for several hosts including
A. niger, A. oryzae, A. awamori, C. lucknowense, and A. chrysogenum.
Plant phytase (263), produced in recombinant A. niger was used as a feed for 50% of all
pigs in Holland. A 1,000-fold increase in phytase production was achieved in A. niger by the use
of recombinant technology (239). Mammalian chymosin was cloned and produced by A. niger or
E. coli and recombinant chymosin was approved in the USA; its price was half that of natural
calf chymosin.
Three fungal recombinant lipases are currently used in the food industry. They are from
Rhizomucor miehi, Thermomyces lanuginosus, and Fusarium oxysporum and are produced in
A. oryzae. They are used for laundry cleaning, inter-esterification of lipids and esterification of
glucosides producing glycolipids which have applications as biodegradable nonionic surfactants
for detergents, skin care products, contact lenses and as food emulsifiers. Washing powders
have been improved in activity and low temperature operation has been achieved by the
application of recombinant DNA technology and site-directed mutagenesis of genes encoding
proteases and lipases (264, 265). The first commercial recombinant lipase used in a detergent
was from Humicola lanuginose. The gene was cloned into the A. oryzae genome.
A multicopy plasmid of B. subtilis was used to increase by 2,500- fold the production of an
α-amylase from B. amyloliquefaciens (266). An exoglucanase from Cellulomonas fimi was
overproduced in E. coli to a level of over 20% of cell protein (267). The same host has also
been used to clone the endo--glucanase components from Thermomonospora and Clostridium
thermocellum as well as the cellobiohydrolase I gene of Trichoderma reesei (268). P. pastoris
was engineered to produce and excrete S. cerevisiae invertase into the medium at 100 mg L-1
(269). Self-cloning of the xylanase gene in S. lividans resulted in a sixfold overproduction of the
enzyme (270).
The properties of many enzymes have been modified by random mutagenesis and
screening of microorganisms over the years leading to changes in substrate specificity,
feedback inhibition, kinetic parameters (Vmax, Km or Ki), pH optimum, thermostability, and carbon
source inhibition. Based on this information, more rational techniques such as site-directed
mutagenesis were used to introduce single changes in amino acid sequences yielding similar
types of changes in a large variety of enzymes. Modification of eight amino acids increased heat
tolerance and temperature stability at 100°C of a protease from Bacillus stearothermophilus
(271). Interestingly, all mutations were far from the active site of the enzyme.
Molecular breeding techniques (e.g., DNA shuffling and DNA family shuffling) are being
currently used to generate enzymes with improved properties such as activity and stability at
different pH values and temperatures (272), increased or modified enantioselectivity (273),
altered substrate specificity (274), stability in organic solvents (275), novel substrate specificity
and activity (276), increased biological activity of protein pharmaceuticals and biological
molecules (277) as well as novel vaccines (278, 279). Two proteins from directed evolution work
were already on the market by year 2,000 (280). These were green fluorescent protein of
Clontech (281) and Novo Nordisk’s LipoPrime ® lipase.
5. Closing Remarks
The fermentation industry developed slowly from the beginning of the twentieth century to the
early 1970s using brute force mutagenesis followed by screening or selection. However, the
birth of the era of recombinant DNA (rDNA) in 1971–1973 catalyzed a major change in the way
useful processes could be developed. Production of primary metabolites was markedly
improved by modern genetic techniques. Environmentally friendly fermentations replaced
chemical synthesis to a great extent. Of great interest has been the application of rDNA
technology to the production of secondary metabolites and to the elucidation of their
biosynthetic pathways. Tools include transposition mutagenesis, targeted deletions, genetic
22
recombination via combinatorial biosynthesis, transcriptome analysis, proteomics,
metabolomics, metabolic engineering, etc. Genes encoding many enzymes have been cloned
and the enzymes have been expressed at levels hundreds of times higher than those naturally
produced. Random redesign techniques have generated enzymes with improved properties
including activity, stability, increased or modified enantioselectivity, altered substrate specificity,
etc. An entirely new field of industrial microbiology has arisen out of rDNA, i.e., the
biopharmaceutical industry which is the most rapidly expanding segment of the biological
industry, especially that of monoclonal antibodies. The best is yet to come from the fantastic
combination of industrial microbiology and rDNA technology. Many of the new techniques are
carried out by small companies and academic groups who could play a major role in rescuing us
from the antibiotic crisis that we are now experiencing. Furthermore, we look forward to its role
in eventually replacing the environmentally dangerous energy sources that we live with today,
i.e., petroleum, coal, etc., with future biofuels made from agricultural and forest biomass.
References
1. Ohnishi J, Mitsuhashi S, Hayashi M, Ando S, Yokoi H, Ochiai K et al (2002) A novel
methodology employing Corynebacterium glutamicum genome information to generate a
new L-lysine-producing mutant. Appl Microbiol Biotechnol 58:217–223
2. Stephanopoulos G (1999) Metabolic fluxes and metabolic engineering. Metab Eng 1:1–11
3. Nielsen J (2001) Metabolic engineering. Appl Microbiol Biotechnol 55:263–283
4. Sahm H, Eggeling L, de Graaf AA (2000) Pathway analysis and metabolic engineering in
Corynebacterium glutamicum. Biol Chem 381:899–910
5. Santos CSS, Stephanopoulos G (2008) Combinatorial engineering of microbes for optimizing
cellular phenotype. Curr Opin Chem Biol 12:168–176
6. Bailey JE, Sburlati A, Hatzimanikatis V, Lee K, Renner WA, Tsai PS (1996) Inverse metabolic
engineering: a strategy for directed genetic engineering of useful phenotypes. Biotechnol
Bioeng 52:109–121
7. Ness JE, del Cardayre SB, Minshull J, Stemmer WP (2000) Molecular breeding: the natural
approach to protein design. Adv Protein Chem 55:261–292
8. Zhao H, Arnold FH (1997) Optimization of DNA shuffling for high fidelity recombination.
Nucleic Acids Res 25:1307–1308
9. Patten PA, Howard RJ, Stemmer WP (1997) Applications of DNA shuffling to
pharmaceuticals and vaccines. Curr Opin Biotechnol 8:724–733
10. Zhang YX, Perry K, Vinci VA, Powell K, Stemmer WP, del Cardayre SB (2002) Genome
shuffling leads to rapid phenotypic improvement in bacteria. Nature 415: 644–646
11. Hou L (2009) Novel methods of genome shuffling in Saccharomyces cerevisiae.
Biotechnol Lett 31:671–677
12. Stephanopoulos G, Alper H, Moxley J (2004) Exploiting biological complexity for strain
improvement through systems biology. Nat Biotechnol 22:1261–1267
13. Leuchtenberger W, Huthmacher K, Drauz K (2005) Biotechnological production of amino
acids and derivatives: current status and prospects. Appl Microbiol Biotechnol 69:1–8
14. Liu Q, Zhang J, Wei X-X, Ouyang S-P, Wu Q, Chen G-Q (2008) Microbial production of
L-glutamate and L-glutamine by recombinant Corynebacterium glutamicum harboring
Vitreoscilla hemoglobin gene vgb. Appl Microbiol Biotechnol 77:1297–1304
15. Chinen A, Kozlov YI, Hara Y, Izui H, Yasueda H (2007) Innovative metabolic pathway
design for efficient L-glutamate production by suppreassing CO2 emission. J Biosci Bioeng
103:262–269
16. Wang J, Gao D, Yu X, Li W, Qi Q (2015) Evolution of a chimeric aspartate kinase for
L-lysine production using a synthetic RNA device. Appl Microbiol Biotechnol 99:8527-8536.
17. Ohnishi J, Katahira R, Mitsuhashi S, Kakita S, Ikeda M (2005) A novel gnd mutation leading
to increased L-lysine production in Corynebacterium glutamicum. FEMS Microbiol Lett
23
242:265–274
18. Ikeda M, Ohnishi J, Hayashi M, Mitsuhashi S (2006) A genome-based approach to create a
minimally mutated Corynebacterium glutamicum strain for efficient L-lysine production. J Ind
Microbiol Biotechnol 33:610–615
19. Hiyashi M, Ohnishi J, Mitsuhashi S, Yonetani Y, Hashimoto S, Ikeda M (2006)
Transcriptome analysis reveals global expression changes in an industrial L-lysine producer
of Corynebacterium glutamicum. Biosci Biotechnol Biochem 70:546–550
20. Radmacher E, Eggeling L (2007) The three tricarboxylate synthase activities of
Corynebacterium glutamicum and increase of L-lysine synthesis. Appl Microbiol Biotechnol
76:587–595
21. Blombach B, Schreiner ME, Moch M, Oldiges M, Eikmanns BJ (2007) Effect of pyruvate
dehydrogenase complex deficiency on L-lysine production with Corynebacterium
glutamicum. Appl Microbiol Biotechnol 76:615–623
22. Sindelar G, Wendisch VF (2007) Improving lysine production by Corynebacterium
glutamicum through DNA microarray-based identification of novel target genes. Appl
Microbiol Biotechnol 76:677–689
23. Becker J, Zelder O, Häfner S, Schröder H, Wittmann C (2011) From zero to hero-design-
based systems metabolic engineering of Corynebacterium glutamicum for L-lysine
production. Metab Eng 13:159-168
24. Komatsubara S, Kisumi M, Chibata I (1979) Transductional construction of a threonine
producing strain of Serratia marcescens. Appl Environ Microbiol 38:1045–1051
25. Komatsubara S, Kisumi M, Chibata I (1983) Transductional construction of a threonine
hyperproducing strain of Serratia marcescens: lack of feedback controls of three
aspartokinases and two homoserine dehydrogenases. Appl Environ Microbiol 45:1445–1452
26. Sugita T, Komatsubara S (1989) Construction of a threonine-hyperproducing strain of
Serratia marcescens by amplifying the phosphoenolpyruvate carboxylase gene. Appl
Microbiol Biotechnol 30:290–293
27. Debabov VG (2003) The threonine story. Adv Biochem Eng Biotechnol 79:113–136
28. Ishida M, Kawashima H, Sato K, Hashiguchi K, Ito H, Enei H et al (1994) Factors improving
L-threonine production by a three L-threonine biosynthetic genes-amplified recombinant
strain of Brevibacterium lactofermentum. Biosci Biotechnol Biochem 58:768–770
29. Eggeling L, Sahm H (1999) Amino acid production: principles of metabolic engineering. In:
Lee SY, Papoutsakis ET (eds) Metabolic engineering. Marcel Dekker, New York, pp 153–
176
30. Kruse D, Kraemer R, Eggeling L, Rieping M, Pfefferle W, Tchieu JH et al (2002) Influence
of threonine exporters on threonine production in Escherichia coli. Appl Microbiol Biotechnol
59:205–210
31. Lee KH, Park JH, Kim TY, Kim HU, Lee SY (2007) Systems metabolic engineering of
Escherichia coli for L-threonine production. Mol Syst Biol 3:149–157
32. Chen N, Huang J, Feng Z, Yu L, Xu Q-y, Wen T-y (2009) Optimization of fermentation
conditions for the biosynthesis of L-threonine by Escherichia coli. Appl Biochem Biotechnol
158: 595
33. Wang J, Cheng LK, Chen N (2014) High-level production of L-threonine by recombinant
Escherichia coli with combined feeding strategies. Biotechnol Biotechnol Equip 4: 28: 495–
501
34. Park JH, Lee KH, Kim TY, Lee SY (2007) Metabolic engineering of Escherichia coli for the
production of L-valine based on transcriptome analysis and in silico gene knockout
stimulation. Proc Natl Acad Sci USA 104:7797–7802
35. Lange C, Rittmann D, Wendisch VF, Bott M, Sahm H (2003) Global expression profiling
and physiological characterization of Corynebacterium glutamicum grown in the presence of
L-valine. Appl Environ Microbiol 69:2521–2532
24
36. Chen C, Li Y, Hu J, Dong X, Wang X (2015) Metabolic engineering of Corynebacterium
glutamicum ATCC13869 for L-valine production. Metab Eng 29:66–75
37. Kisumi M, Komatsubara S, Chibata I (1977) Enhancement of isoleucine hydroxamate-
mediated growth inhibition and improvement of isoleucine-producing strains of Serratia
marcescens. Appl Environ Microbiol 34: 647–653
38. Guillouet S, Rodal AA, An G-H, Lessard PA, Sinskey AJ (1999) Expression of the
Escherichia coli catabolic threonine dehydratase in Corynebacterium glutamicum and its
effect on isoleucine production. Appl Environ Microbiol 65:3100–3107
39. Morbach S, Sahm H, Eggeling L (1996) L-Isoleucine production with Corynebacterium
glutamicum: further flux increase and limitation of export. Appl Environ Microbiol 62: 4345–
4351
40. Hashiguchi K, Takesada H, Suzuki E, Matsui H (1999) Construction of an L-isoleucine
overproducing strain of Escherichia coli K-12. Biosci Biotechnol Biochem 63:672–679
41. Eggeling L, Morbach S, Sahm H (1977) The fruits of molecular physiology: engineering the
L-isoleucine biosynthesis pathway in Corynebacterium glutamicum. J Biotechnol 56:167–
182
42. Colon GE, Nguyen TT, Jetten MSM, Sinskey A, Stephanopoulos G (1995) Production of
isoleucine by overexpression of ilvA in Corynebacterium lactofermentum threonine
producer. Appl Microbiol Biotechnol 43: 482–488
43. Kase H, Nakayama K (1977) L-Isoleucine induction by analog-resistant mutants derived
from threonine-producing strain of Corynebacterium glutamicum. Agric Biol Chem 41:109–
116
44. Morbach S, Sahm H, Eggeling L (1995) Use of feedback-resistant threonine dehydratases
of Corynebacterium glutamicum to increase carbon flux towards L-isoleucine. Appl Environ
Microbiol 61:4315–4320
45. Morbach S, Kelle R, Winkels S, Sahm H, Eggeling L (1996) Engineering the homoserine
dehydrogenase and threonine dehydratase control points to analyse flux towards
L-isoleucine in Corynebacterium glutamicum. Appl Microbiol Biotechnol 45:612–620
46 Sahm H, Eggeling L, Morbach S, Eikmanns B (1999) Construction of L-isoleucine
overproducing strains of Corynebacterium glutamicum. Naturwissenschaften 86:33–38
47. Shiio I, Sasaki A, Nakamori S, Sano K (1973) Production of L-isoleucine by AHV resistant
mutants of Brevibacterium flavum. Agric Biol Chem 37:2053–2061
48. Ikeda S, Fujita I, Hirose Y (1976) Culture conditions of L isoleucine fermentation from
acetic acid. Agric Biol Chem 40:517–522
49. Yin L, Shi F, Hu X, Chen C, Wang X (2013) Increasing L-isoleucine production in
Corynebacterium glutamicum by overexpressing global regulator Lrp and two-component
export system BrnFE. J Appl Microbiol 114(5):1369-1377
50. Yin L, Zhao J, Chen C, Hu X, Chen C, Wang X (2014) Enhancing the carbon flux and
NADPH supply to increase L-isoleucine production in Corynebacterium glutamicum.
Biotechnol Biopro Eng 19: 132-142
51. Peng L, Lijun L, Wang J (2015) The Optimization of L-isoleucine fermentation conditions by
Brevibacterium flavum KM011. Microb Microbial Tech 1(2): 002
52. Lee M, Smith GM, Eiteman MA, Altman E (2004) Aerobic production of alanine by
Escherichia coli aceF IdhA mutants expressing the Bacillus sphaericus alaD gene. Appl
Microbiol Biotechnol 65:56–60
53. Zhang X, Jantama K, Moore JC, Shanmugam KT, Ingram LO (2007) Production of L-alanine
by metabolically engineered Escherichia coli. Appl Microbiol Biotechnol 77(2):355-366.
54. Zhou L, Deng C, Cui W-J, Liu Z-M, Zho Z-M (2015) Efficient L-alanine production by a
thermo-regulated switch in Escherichia coli. Appl Biochem Biotechnol 178:324–337
55. Bloom F, Smith CJ, Jessee J, Veileux B, Deutch AH (1984) The use of genetically
engineered strains of Escherichia coli for the overproduction of free amino acids: proline as
25
a model system. In: Downey K, Voellmy RW (eds) Advances in gene technology: molecular
genetics of plants and animals. Academic, New York, pp 383–394
56. Sugiura M, Takagi T, Kisumu M (1995) Proline production by regulatory mutants of Serratia
marcescens. Appl Microbiol Biotechnol 21:213–239
57. Sugiura M, Imai Y, Takagi T, Kisumi M (1985) Improvement of a proline-producing strain of
Serratia marcescens by subcloning of a mutant allele of the proline gene. J Biotechnol 3:47–
58
58. Masuda M, Takamatu S, Nishimura N, Komatsubara S, Tosa T (1993) Improvement of
culture conditions for L-proline production by a recombinant strain of Serratia marcescens.
Appl Biochem Biotechnol 43:189–197
59. Tsuchida T, Kubota K, Yoshinaga F (1986) Improvement of L-proline production by
sulfaguanidine resistant mutants derived from L-glutamic acid-producing bacteria. Agric
Biol Chem 50:2201–2207
60. Nakanishi T, Yokote Y, Taketsugu Y (1973) Conversion of L-glutamic acid fermentation to a
L-proline fermentation by Corynebacterium glutamicum. J Ferment Technol 51:742–749
61. Nakanishi T, Hirao T, Azuma T, Sakurai M, Hagino H (1987) Application of L-glutamate
to L-proline fermentation by Corynebac terium acetoacidophilum. J Ferment Technol 65: 139–
144
62. Shibasaki T, Hashimoto S, Mori H, Ozaki A (2000) Construction of a novel hydroxyproline-
producing recombinant Escherichia coli by introducing a proline 4-hydroxylase gene. J
Biosci Bioeng 90:522–525
63. Wang J, Zhang Z, Liu H, Sun FF, Yue C, Hu J, Wang C (2016) Construction and
optimization of trans-4-hydroxy-L-proline production recombinant Escherichia coli strain
taking the glycerol as carbon source. J Chem Technol Biotechnol 91:2389-2398
64. Vandamme EJ (2007) Microbial gems: microorganisms without frontiers. SIM News 57:81–
90
65. Kamada N, Yasuhara A, Takano Y, Nakano T (2001) Effect of transketolase modifications
on carbon flow to the purine-nucleotide pathway in Corynebacterium ammoniagenes. Appl
Microbiol Biotechnol 56:710–717
66. Koizumi S, Yonetani Y, Maruyama A, Teshiba S (2000) Production of riboflavin by
metabolically engineered Corynebacterium ammoniagenes. Appl Microbiol Biotechnol 51:
674–679
67. Perkins JB, Pero J (1993) Biosynthesis of riboflavin, biotin, folic acid, and cobalamin. In:
Sonenshein AL, Hoch JA, Losick R (eds) Bacillus subtilis and other Gram-positive bacteria:
biochemistry, physiology and molecular genetics. American Society for Microbiology,
Washington, DC, pp 319–334
68. Perkins JB, Sloma A, Hermann T, Theriault K, Zachgo E, Erdenberger T et al (1999)
Genetic engineering of Bacillus subtilis for the commercial production of riboflavin. J Ind
Microbiol Biotechnol 22:8–18
69. Brenner S, Johnson M, Bridgham J, Golda G, Lloyd DH, Johnson D et al (2000) Gene
expression analysis by massively parallel signature sequencing (MPSS) on microbead
arrays. Nat Biotechnol 18:630–634
70. Karos M, Vilarino C, Bollschweiler C, Revuelta JL (2004) A genome wide transcription
analysis of a fungal riboflavin overproducer. J Biotechnol 113:69–76
71. Shi SB, Chen T, Zhang ZG, Chen X, Zhao XM (2009) Transcriptome analysis guided
metabolic engineering of Bacillus subtilis for riboflavin production. Metab Eng 11:243–252
72. Levy-Schil S, Debussche L, Rigault S, Soubrier F, Bacchette F, Lagneaux D et al (1993)
Biotin biosyntheric pathway in a recombinant strain of Escherichia coli overexpressing bio
genes: evidence for a limiting step upstream from KAPA. Appl Microbiol Biotechnol
38:755–762
26
73. Sakurai N, Imai Y, Masuda M, Komatsubara S, Tosa T (1994) Improvement of a D-biotin
hyperproducing recombinant strain of Serratia marcescens. J Biotechnol 36:63–73
74. Masuda M, Takahashi K, Sakurai N, Yanagiya K, Komatsubara S, Tosa T (1995) Further
improvement of D-biotin production by a recombinant strain of Serratia marcescens .
Process Biochem 30:553–562
75. Matsui J, Ifuku O, Kanzaki N, Kawamoto T, Nakahama K (2001) Microorganism resistant
to threonine analogue and production of biotin. US patent 6284500
76. Anderson S, Marks CB, Lazarus R, Miller J, Stafford K, Seymour J et al (1985) Production of
2-keto-L-gulonate: an intermediate in L-ascorbate synthesis by a genetically modified
Erwinia herbicola. Science 230:144–149
77. Grindley JF, Payton MA, van de Pol H, Hardy KG (1988) Conversion of glucose to 2-keto-
L-gulonate, an intermediate in L–ascorbate synthesis, by a recombinant strain of Erwinia
citreus. Appl Environ Microbiol 54:1770–1775
78. Saito Y, Ishii Y, Hayashi H, Imao Y, Akashi T, Yoshikawa K et al (1997) Cloning of genes
coding for L-sorbose and L-sorbosone dehydrogenases from Gluconobacter oxydans and
microbial production of 2-keto-gulonate, a precursor of L-ascorbic acid, in a recombinant
G. oxydans strain. Appl Environ Microbiol 63:454–460
79. Saito Y, Ishii Y, Hayashi K, Yoshikawa K, Noguchi Y, Yoshida S, Soeda S, Yoshida M
(1998) Direct fermentation of 2-keto-L-gulonic acid in recombinant Gluconobacter oxydans.
Biotechnol Bioeng 58:309–315
80. Chotani G, Dodge T, Hsu A, Kumar M, LaDuca R, Trimbur D, Weyler W, Sanford K (2000)
The commercial production of chemicals using pathway engineering. Biochimica et
Biophysica Acta 1543:434-455
81. Rao YM, Sureshkumar GK (2000) Direct biosynthesis of ascorbic acid from glucose by
Xanthomonas campestris through induced free-radicals. Biotechnoil Lett 22:407-411
82. DeBaets S, Vandedrinck S, Vandamme EJ (2000) Vitamins and related biofactors,
microbial production. In: Lederberg J (ed) Encyclopedia of microbiology, vol 4, 2nd
edn. Academic, New York, pp 837–853
83. Johnson EA, Schroeder WA (1995) Microbial carotenoids. Adv Biochem Eng Biotechnol
53:119–178
84. López-Nieto MJ, Costa J, Peiro E, Méndez E, Rodríguez-Sáiz M, de la Fuente JL, Cabri
W, Barredo JL (2004) Biotechnological lycopene production by mated fermentation of
Blakeslea trispora. Appl Microbiol Biotechnol 66:153–159
85. Ciegler A (1965) Microbial carotenogenesis. Adv Appl Microbiol 7:1–34
86. Ninet L, Renaut J (1979) Carotenoids. In: Peppler HJ, Perlman D (eds) Microbial
technology, vol 1, 2nd edn. Academic, New York, pp 529–544
87. Barkovich R, Liao JC (2001) Metabolic engineering of isoprenoids. Metab Eng 3:27–39
88. Lee PC, Schmidt-Dannert C (2002) Metabolic engineering towards biotechnological
production of carotenoids in microorganisms. Appl Microbiol Biotechnol 60:1–11
89. Tao L, Jackson RE, Cheng Q (2005) Directed evolution of copy number of a broad host
range plasmid for metabolic engineering. Metab Eng 7:10–17
90. Alper H, Miyaoku K, Stephanopoulos G (2006) Characterization of lycopene-overproducing
E. coli strains in high cell density fermentations. Appl Microbiol Biotechnol 72:968–974
91. Fernández-Sevilla JM, Acien Fernández FG, Molina Grima E (2010) Biotechnological
production of lutein and its applications. Appl Microbiol Biotechnol 86:27–40
92. Margalith PZ (1999) Production of ketocarotenoids by microalgae. Appl Microbiol Biotechnol
51:431–438
93. Johnson EA (2003) Phaffia rhodozyma: a colorful odyssey. Int Microbiol 6:169–174
94. de la Fuente JL, Rodríguez-Sáiz M, Schleissner C, Díez B, Peiro E, Barredo JL (2010) High
titer production of astaxanthin by the semiindustrial fermentation of Xanthophyllomyces
dendrorhous. J Biotechnol 148:144–146
27
95. Rodríguez-Sáiz M, de la Fuente JL, Barredo JL (2010) Xanthophyllomyces dendrorhous for
the industrial production of astaxanthin. Appl Microbiol Biotechnol 88:645–658
96. Henke NA, Heider SAE, Peters-Wendisch P, Wendisch VF (2016) Production of the marine
carotenoid astaxanthin by metabolically engineered Corynebacterium glutamicum. Mar.
Drugs 14:124
97. Sauer M, Porro D, Mattanovich D, Branduardi P (2008) Microbial production of organic
acids: expanding the markets. Trends Biotechnol 26:100–108
98. Sánchez S, Demain AL (2008) Metabolic regulation and overproduction of primary
metabolites. Microb Biotechnol 1:283–319
99. Ward OP, Qin WM, Hanjoon JD, Singh EJYA (2006) Physiology and biotechnology of
Aspergillus. Adv Appl Microbiol 58:1–75
100. Kubicek CP, Roehr M (1986) Citric acid fermentation. Crit Rev Biotechnol 3:331–373
101. Anastasiadis S, Rehm HJ (2005) Continuous citric acid secretion by a high specific pH
dependent active transport system in yeast Candida oleophila ATCC 20177. Electronic J
Biotechnol 8(2):26-42
102. Causey TB, Zhou S, Shanmugam KT, Ingram LO (2003) Engineering the metabolism of
Escherichia coli W3110 for the conversion of sugar to redox-neutral and oxidized products:
homoacetate production. Proc Natl Acad Sci USA 100:825–832
103. Parekh SR, Cheryan M (1994) High concentrations of acetate with a mutant strain of C.
thermoaceticum. Biotechnol Lett 16:139–142
104. Fukaya M, Tayama K, Tamaki T, Tagami H, Okumura H, Kawamura Y et al (1989)
Cloning of the membrane-bound aldehyde dehydrogenase gene of Acetobacter
polyoxogenes and improvement of acetic acid production by use of the cloned gene. Appl
Environ Microbiol 55:171–176
105. Patnaik R, Louie S, Gavrilovic V, Perry K, Stemmer WPC, Ryan CM et al (2002)
Genome shuffling of Lactobacillus for improved acid tolerance. Nat Biotechnol 20:707–712
106. John RP, Gangadharan D, Madhavan Nampoothiri K (2008) Genome shuffling of
Lactobacillus delbrueckii mutant and Bacillus amyloliquuefaciens through protoplastic fusion
for L-latic acid production from starchy wastes. Bioresour Technol 99:8008–8015
107. Skory CD (2004) Lactic acid production by Rhizopus oryzae transformants with modified
lactate dehydrogenase activity. Appl Microbiol Biotechnol 64:237–242
108. Saito S, Ishida N, Onishi T, Tokuhiro K, Nagamori E, Kitamoto K et al (2005) Genetically
engineered wine yeast produces a high concentration of L-lactic acid of extremely high
optical purity. Appl Environ Microbiol 71:2789–2792
109. Porro D, Bianchi MM, Brambilla L, Menghini R, Bolzani D, Carrera V et al (1999)
Replacement of a metabolic pathway for large scale production of lactic acid from
engineered yeasts. Appl Environ Microbiol 65:4211–4215
110. Zhou S, Yomano LP, Shanmugam KT, Ingram LO (2003) Fermentation of 10% (w/v) sugar
to D(-)-lactate by engineered Escherichia coli B. Biotechnol Lett 27:1891–1896
111. Ishida N, Suzuki T, Tokuhiro K, Nagamori E, Onishi T, Saitoh S et al (2006) D-Lactic acid
production by metabolically engineered Saccharomyces cerevisiae. J Biosci Bioeng 101:
172–177
112. Sánchez AM, Bennett GN, San KY (2005) Novel pathway engineering design of the
anaerobic central metabolic pathway in Escherichia coli. Metab Eng 7:229–239
113. Lee SJ, Song H, Lee SY (2006) Genome-based metabolic engineering of Mannheimia
succiniciproducens for succinic acid production. Appl Environ Microbiol 72:1939–1948
114. Lin H, Bennett GN, San KY (2005) Fedbatch culture of a metabolically engineered
Escherichia coli strain designed for high-level succinate production and yield under aerobic
conditions. Biotechnol Bioeng 90:775–779
115. Vemuri GN, Eiteman MA, Altman E (2002) Succinate production in dual-phase Escherichia
coli fermentations depends on the time of transition from aerobic to anaerobic conditions.
28
J Ind Microbiol Biotechnol 28: 325–332
116. Liu X, Yang S-T (2005) Metabolic engineering of Clostridium tyrobutyricum for butyric acid
fermentation. Proceedings of the 229th ACS National Meeting, San Diego,
Abstract 70
117. Neufeld RJ, Peleg Y, Rokem JS, Pines O, Goldberg I (1991) L-Malic acid formation by
immobilized Saccharomyces cerevisiae amplified for fumarase. Enzyme Microb Technol
13:991–996
118. Picataggio S, Rohver T, Deander K, Lanning D, Reynolds R, Mielenz J et al (1992)
Metabolic engineering of Candida tropicalis for the production of long-chain dicarboxylic
acids. Nat Biotechnol 10:894–898
119. Shi DJ, Wang CL, Wang KM (2009) Genome shuffling to improve thermotolerance, ethanol
tolerance and ethanol productivity of Saccharomyces cerevisiae. J Ind Microbiol Biotechnol
36:139–147
120. Chu BCH, Lee H (2007) Genetic improvement of Saccharomyces cerevisiae for xylose
fermentation. Biotechnol Adv 25:425–441
121. Jeffries TW (2006) Engineering yeast for xylose metabolism. Curr Opin Biotechnol
17:320–326
122. Bro C, Regenberg B, Forster J, Nielsen J (2006) In silico aided metabolic engineering
of Saccharomyces cerevisiae for improved bioethanol production. Metab Eng 8:102–111
123. Wei P, Li Z, He P, Lin Y, Jiang N (2008) Genome shuffling in the ethanologenic yeast
Candida krusei to improve acetic acid tolerance. Biotechnol Appl Biochem 49:113–120
124. Ingram LO, Conway T, Clark DP, Sewell GW, Preston JF (1987) Genetic engineering of
ethanol production in Escherichia coli. Appl Environ Microbiol 53:2420–2425
125. Doran JB, Ingram LO (1993) Fermentation of crystalline cellulose to ethanol by Klebsiella
oxytoca containing chromosomally integrated Zymomonas mobilis genes. Biotechnol Prog
9:533–538
126. Dien BS, Nichols NN, O’Bryan PJ, Bothast RJ (2000) Development of new ethanologenic
Escherichia coli strains for fermentation of lignocellulosic biomass. Appl Biochem
Biotechnol 84:181–196
127. Papanikolaou S, Ruiz-Sánchez P, Pariset B, Blanchard F, Fick M (2000) High production
of 1,3-propanediol from industrial glycerol by a newly isolated Clostridium butyricum strain.
J Biotechnol 77:191–208
128. Sanford K, Valle F, Ghirnikar R (2004) Pathway engineering through rational design.
Tutorial: designing and building cell factories for biobased production. Genet Eng News
24:44–45
129. Nakamura C, Whited G (2003) Metabolic engineering for the microbial production of 1,3
propanediol. Curr Opin Biotechnol 14:454–459
130. Kaup B, Bringer-Meyer S, Sahm H (2004) Metabolic engineering of Escherichia coli:
construction of an efficient biocatalyst for D–mannitol formation in a whole-cell
biotransformation. Appl Microbiol Biotechnol 64:333–339
131. Kaup B, Bringer-Meyer S, Sahm H (2005) D-Mannitol formation from D-glucose in a whole-
cell biotransformation with recombinant Escherichia coli . Appl Microbiol Biotechnol 69:397–
403
132. Lee J-K, Oh D-K, Song H-Y, Kim I-W (2007) Ca2+ and Cu2+ supplementation increases
mannitol production by Candida magnoliae. Biotechnol Lett 29:291–294
133. Song SH, Vieille C (2009) Recent advances in the biological production of mannitol. Appl
Microbiol Biotechnol 84:55–62
134. Chun UH, Rogers PL (1988) The simultaneous production of sorbitol and gluconic acid
by Zymomonas mobilis . Appl Microbiol Biotechnol 29:19–24
135. Ladero V, Ramos A, Wiersma A, Goffin P, Schanck A, Kleerbezem M (2007) High-level;
production of the low-calorie sugar sorbitol by Lactobacillus plantarum through metabolic
29
engineering. Appl Environ Microbiol 73:1864–1872
136. Hermann M, Fayolle F, Marchal R, Podvin L, Sebald M, Vandecasteele J-P (1985)
Isolation and characterization of butanol-resistant mutants of Clostridium acetobutylicum.
Appl Environ Microbiol 50:1238–1243
137. Qureshi N, Maddox IS, Freidl A (1992) Application of continuous substrate feeding to the
ABE fermentation: relief of product inhibition using extraction, perstraction, stripping and
pervaporation. Biotechnol Prog 8:382–390
138. Chen C-K, Blaschek HP (1999) Acetate enhances solvent production and prevents
degeneration in Clostridium beijerinckii BA101. Appl Microbiol Biotechnol 52:
170–173
139. Connor MR, Liao JC (2009) Microbial production of advanced transportation fuels in
non-natural hosts. Curr Opin Biotechnol 20:307–315
140. Whiters ST, Gottlieb SS, Lieu B, Newman JD, Keasling JD (2007) Identification of
isopentenol biosynthetic genes from Bacillus subtilis by a screening method based on
isoprenoid precursor toxicity. Appl Environ Microbiol 73:6277–6283
141. Atsumi S, Hanai T, Liao JC (2008) Nonfermentative pathways for synthesis of
branched-chain higher alcohols as biofuels. Nature 451:86–89
142. Jeya M, Lee K-M, Tiwari MK, Kim J-S, Gunasekaran P, Kim S-Y, Kim I-W, Lee J-K
(2009) Isolation of a novel high erythritol producing Pseudomonas tsukubaensis and scaleup
of erythritol fermentation to industrial level. Appl Microbiol Biotechnol 83:225–231
143. Zagustina NA, Rodionova NA, Mestechkina NM, Shcherbukhin VD, Bezborodov AM
(2001) Xylitol production by a culture of Candida guilliermondii 2581. Appl Biochem
Microbiol 37:489–492
144. Lederberg J (2000) Pathways of discovery: infectious history. Science 288:287–293
145. Choi KP, Kim KH, Kim JW (1997) Strain improvement of clavulanic acid producing
Streptomyces clavuligerus. Proc 10th Internat Symp Biol Actinomycetes (ISBA), Beijing,
Abstr 12, 9
146. Li R, Townsend CA (2006) Rational strain improvement for enhanced clavulanic acid
production by genetic engineering of the glycolytic pathway in Streptomyces clavuligerus.
Metab Eng 8:240–252
147. Pérez-Redondo RA, Rodríguez-García A, Martín JF, Liras P (1999) Deletion of the pyc
gene blocks clavulanic acid biosynthesis except in glycerol-containing medium: evidence for
two different genes in formation of the C3 unit. J Bacteriol 181:6922–6928
148. Paradkar AS, Aiodoo KA, Jensen SE (1998) A pathway-specific transcriptional activator
regulates late steps of clavulanic acid biosynthesis in Streptomyces clavuligerus. Mol
Microbiol 27:831–843
149. Pérez-Llarena FJ, Liras P, Rodríguez-García A, Martín JF (1997) A regulatory gene (ccaR)
required for cephamycin and clavulanic acid production in Streptomyces clavuligerus:
amplification results in overproduction of both -lactam compounds. J Bacteriol 179:
2053–2059
150. Lein J (1986) The Panlabs penicillin strain improvement program. In: Vanek Z, Hostalek
Z (eds) Overproduction of microbial metabolites; strain improvement and process control
strategies. Butterworth, Boston, pp 105–139
151. Kennedy J, Turner G (1996) -(L--Aminoadipyl)-L-cysteinyl-D-valine synthetase is a rate
limiting enzyme for penicillin production in Aspergillus nidulans. Mol Gen Genet 253:189–
197
152. Hamlyn PF, Ball C (1979) Recombination studies with Cephalosporium acremonium. In:
Sebek OK, Laskin AI (eds) Genetics of industrial microorganisms. American Society for
Microbiology, Washington, DC, pp 185–191
153. Skatrud PL, Fisher DL, Ingolia TD, Queener SW (1987) Improved transformation of
Cephalosporium acremonium. In: Alacevic M, Hranueli D, Toman Z (eds) Genetics of
30
industrial microorganisms, part B. Zagreb, Pliva, pp 111–119
154. Skatrud PL, Tietz AJ, Ingolia TD, Cantwell CA, Fisher DL, Chapman JL et al (1989) Use
of recombinant DNA to improve production of cephalosporin C by Cephalosporium
acremonium. Nat Biotechnol 7:477–485
155. Wesseling AC, Lago B (1981) Strain improvement by genetic recombination of cephamycin
producers, Nocardia lactamdurans and Streptomyces griseus. Dev Ind Microbiol
22:641–651
156. Chary VK, de la Fuente JL, Leitao AL, Liras P, Martín JF (2000) Overexpression of the lat
gene in Nocardia lactamdurans from strong heterologous promoters results in very high
levels of lysine-6-aminotransferase and up to a two-fold increase in cephamycin C
production. Appl Microbiol Biotechnol 53:282–288
157. Bignell DRD, Tahlan K, Colvin KR, Jensen SE, Leskiw BK (2005) Expression of ccaR,
encoding the positive activator of cephamycin C and clavulanic acid production in
Streptomyces clavuligerus, is dependent on bldG. Antimicrob Agents Chemother
49:1529–1541
158. Cantwell C, Beckmann R, Whiteman P, Queener SW, Abraham EP (1992) Isolation of
deacetoxycephalosporin C from fermentation broths of Penicillium chrysogenum
transformants: construction of a new fungal biosynthetic pathway. Proc R Soc Lond (Biol)
248:283–289
159. Crawford L, Stepan AM, Mcada PC, Rambosek JA, Conder MJ, Vinci VA et al (1995)
Production of cephalosporin intermediates by feeding adipic acid to recombinant
Penicillium chrysogenum strains expressing ring expansion activity. Nat Biotechnol 13:58–
62
160. Velasco J, Adrio JL, Moreno MA, Díez B, Soler G, Barredo JL (2000) Environmentally safe
production of 7-aminodeacetoxycephalosporanic acid (7ADCA) using recombinant strains of
Acremonium chrysogenum. Nat Biotechnol 18:857–861
161. Luo H, Yu H, Qiang L, Shen Z (2004) Cloning and co-expression of D-amino acid oxidase
and glutaryl-7-aminocephalosporanic acid acylase genes in Escherichia coli. Enzyme Microb
Technol 35:514–518
162. Coulthurst SJ, Barnard AM, Salmond GP (2005) Regulation and biosynthesis of
carbapenem antibiotics in bacteria. Nat Rev Microbiol 3:295–306
163. Kahan JS, Kahan FM, Goegelman R, Currie SA, Jackson M, Stapley EO, Miller TW, Miller
AK, Hendlin D, Mochales S, Hernandez S, Woodruff HB, Birnbaum J (1979) Thienamycin, a
new -lactam antibiotic. 1. Discovery, taxonomy, isolation and physical properties. J Antibiot
32:1–12
164. Park SR, Han AR, Ban Y-H, Yoo YJ, Kim EJ et al (2010) Genetic engineering of macrolide
biosynthesis: past advances, current state and future prospects. Appl Microbiol Biotechnol
85:1227–1239
165. Reeves AR, Cernota WH, Brikun IA, Wesley RK, Weber JM (2004) Engineering precursor
flow for increased erythromycin production in Aeromicrobium erythreum. Metab Eng
6:300–312
166. Solenberg PJ, Cantwell CA, Tietz AJ, McGilvray D, Queener SW, Baltz RH (1996)
Transposition mutagenesis in Streptomyces fradiae: identification of a neutral site for the
stable insertion of DNA by transposon exchange. Gene 16:67–72
167. Brautaset T, Sletta H, Nedal A, Borgos SEF, Degnes KF, Bakke I, Volokhan O,
Sekurova ON, Treshalin ID, Mirchink EP, Dikiy A, Ellingsen TE, Zotchev SB (2008)
Improved antifungal polyene macrolides via engineering of the nystatin biosynthetic
genes in Streptomyces noursei. Chem Biol 15:1198–1206
168. Galm U, Shen B (2006) Expression of biosynthetic gene clusters in heterologous
hosts for natural product production and combinatorial biosynthesis. Expert Opin Drug
Discov 1:409–437
31
169. Méndez C, Salas JA (2003) On the generation of novel anticancer drugs by recombinant
DNA technology: the use of combinatorial biosynthesis to produce novel drugs. Comb
Chem High Throughput Screen 6:513–526
170. Rodríguez E, McDaniel R (2001) Combinatorial biosynthesis of antimicrobials and other
natural products. Curr Opin Microbiol 4: 526–534
171. Trefzer A, Blanco G, Remsing L, Kunzel E, Rix U, Lipata F et al (2002) Rationally
designed glycosylated premithramycins: hybrid aromatic polyketides using genes from
three different biosynthetic pathways. J Am Chem Soc 124:6056–6062
172. Gomi S, Ikeda D, Nakamura H, Naganawa H, Yamashita F, Hotta K et al (1984) Isolation
and structure of a new antibiotic, indolizomycin, produced by a strain SK2-52 obtained by
interspecies fusion treatment. J Antibiot 37:1491–1494
173. Traxler P, Schupp T, Wehrli W (1982) 16, 17-dihydrorifamycin S and 16,17-dihydro-17-
hydroxyrifamycin S, two novel rifamycins from a recombinant strain C5/42 of Nocardia
mediterranei. J Antibiot 35:594–601
174. Okanishi M, Suzuki N, Furuta T (1996) Variety of hybrid characters among recombinants
obtained by interspecific protoplast fusion in streptomycetes. Biosci Biotechnol
Biochem 60:1233–1238
175. Zhou L, Ahlert J, Xue Y, Thorson JS, Sherman DH, Liu H-W (1999) Engineering a
methymycin/pikromycin-calicheamicin hybrid: construction of two new macrolides carrying
a designed sugar moiety. J Am Chem Soc 121:9881–9882
176. Méndez C, Salas JA (2001) Altering the glycosylation pattern of bioactive compounds.
Trends Biotechnol 19:449–456
177. Decker H, Hutchinson CR (1993) Transcriptional analysis of the Streptomyces
glaucescens tetracenomycin biosynthesis gene cluster. J Bacteriol 175:3887–3892
178. Wohlert S-E, Blanco G, Lombo F, Fernández E, Brana AF, Reich S, Udvarnoki G, Méndez
C, Decker H, Frevert J et al (1998) Novel hybrid tetracenomycins through combinatorial
biosynthesis using a glycosyltransferase encoded by the elm genes in cosmid 16 F4 which
shows a very broad sugar substrate specificity. J Am Chem Soc 120: 10596–10601
179. Barriere JC, Berthaud N, Beyer D, Dutka-Malen S, Paris JM, Desnottes JF (1998) Recent
developments in streptogramin research. Curr Pharm Des 4:155–180
180. Xu B, Jin Z, Wang H, Jin Q, Jin X et al (2008) Evolution of Streptomyces pristinaespiralis
for resistance and production of pristinamycin by genome shuffling. Appl Microbiol
Biotechnol 80:261–267
181. Van Lanen SG, Shen B (2006) Microbial genomics for the improvement of natural product
discovery. Curr Opin Microbiol 9:252–260
182. Jenke-Kodama H, Sandmann A, Müller R, Dittmann E (2005) Evolutionary implications
of bacterial polyketide synthases. Mol Biol Evol 22:2027–2039
183. Zazopoulos E, Hwang K, Staffa A, Liu W, Bachmann BO, Nonaka K et al (2003) A
genomics-guided approach for discovering and expressing cryptic metabolic pathways.
Nat Biotechnol 21:187–190
184. Moir DT, Shaw KJ, Hare RS, Vovis GF (1999) Genomics and antimicrobial drug discovery.
Antimicrob Agents Chemother 43:439–446
185. NCBI (National Center for Biotechnology Information) (2017) Microbial genomes. Complete
Eukaryotic and Prokariotic genomes. http://www.ncbi.nlm.nih.gov/genome/browse/.
Accessed 15 Feb 2017).
186. Blin K, Medema MH, Kazempour D et al (2013) antiSMASH 2.0—a versatile platform for
genome mining of secondary metabolite producers. Nucleic Acids Res 41:W204–W212.
187. Anand S, Prasad MV, Yadav G (2010) SBSPKS: structure based sequence analysis of
polyketide synthases. Nucleic Acids Res 2010; 38(Web Server issue): W487-96
188. Newman DJ, Shapiro S (2008) Microbial prescreens for anticancer activity. SIM News
58:132–150
32
189. Hwang CK, Kim HS, Hong YS, Kim YH, Hong SK, Kim SJ, Lee JJ (1995) Expression of
Streptomyces peucetius genes for doxorubicin resistance and aklavinone 11-hydroxylase
in Streptomyces galilaeus ATCC 31133 and production of a hybrid aclacinomycin.
Antimicrob Agents Chemother 39:1616–1620
190. Kim HS, Hong YS, Kim YH, Yoo OJ, Lee JJ (1996) New anthracycline metabolites
produced by the aklavinone 11-hydroxylase gene in Streptomyces galilaeus ATCC 3113.
J Antibiot 49:355–360
191. Niemi J, Mäntäslä P (1995) Nucleotide sequences and expression of genes from
Streptomyces purpurascens that cause the production of new anthracyclines. J Bacteriol
177:2942–2945
192. Ylihonko K, Hakala J, Kunnari T, Mäntsälä P (1996) Production of hybrid anthracycline
antibiotics by heterologous expression of Streptomyces nogalater nogalamycin biosynthesis
genes. Microbiology 142:1965–1972
193. Strohl WR, Bartel PL, Li Y, Connors NC, Woodman RH (1991) Expression of polyketide
biosynthesis and regulatory genes in heterologous streptomycetes. J Ind Microbiol 7:163–
174
194. Bartel PL, Zhu CB, Lampel JS, Dosch DC, Connors NC, Strohl WR, Beale JM Jr, Floss
HG (1990) Biosynthesis of anthraquinones by interspecies cloning of actinorhodin genes
in streptomycetes: clarification of actinorhodin gene functions. J Bacteriol 172:
4816–4826
195. Arcamone F, Penco S, Vigevani A, Redaelli S, Franchi G, Di Marco A, Casazza AM,
Dasdia T, Formelli F, Necco A, Soranzo C (1975) Synthesis and antitumor properties of new
glycosides of daunomycinone and adriamycinone. J Med Chem 18:703–707
196. Madduri K, Kennedy J, Rivola G, Inventi-Solari A, Filippini S, Zanuso G, Colombo AL,
Gewain KM, Occi JL, MacNeil DJ, Hutchinson CR (1998) Production of the antitumor drug
epirubicin (4´-epidoxorubicin) and its precursor by a genetically engineered strain of
Streptomyces peucetius. Nat Biotechnol 16:69–74
197. Manfredi JJ, Horowitz SB (1984) Taxol: an antimitotic agent with a new mechanism of
action. Pharmacol Ther 25:83–125
198. Horwitz SB (1994) Taxol (paclitaxel): mechanisms of action. Ann Oncol 5(Suppl 6): S3–S6
199. Dejong JM, Liu Y, Bollon AP, Long RM, Jennewein S, Williams D, Croteau RB (2005)
Genetic engineering of taxol biosynthetic genes in Saccharomyces cerevisiae. Biotechnol
Bioeng 93:212–224
200. Engels B, Dahm P, Jennewein S (2008) Metabolic engineering of taxadiene biosynthesis
in yeast as a first step towards Taxol (Paclitaxel) production. Metab Eng 10: 201–206
201. Borzlleri RM, Vite GD (2002) Epothilones: new tubulin polymerization agents in preclinical
and clinical development. Drugs Future 27:1149–1163
202. Brown AG, Smale TC, King TJ, Hasenkamp R, Thompson RH (1976) Crystal and
molecular structure of compactin: a new antifungal metabolite from Penicillium
brevicompactum. J Chem Soc Perkin Trans I (11):1165–1170
203. Endo A, Kuroda M, Tsujita Y (1976) ML-236A, ML-236B and ML-236C, new inhibitors of
cholesterolgenesis produced by Penicillium citrinun. J Antibiot 29:1346–1348
204. Endo A (1979) Monacolin K, a new hypocholesterolemic agent produced by Monascus
species. J Antibiot 32:852–854
205. Alberts AW, Chen J, Kuron G, Hunt V, Huff J, Hoffman C et al (1980) Mevinolin: a highly
potent competitive inhibitor of hydroxymethylglutaryl-coenzyme A reductase and a
cholesterol-lowering agent. Proc Natl Acad Sci USA 77:3957–3961
206. Askenazi M, Driggers EM, Holtzman DA, Norman TC, Iverson S, Zimmer DP et al (2003)
Integrating transcriptional and metabolite profiles to direct the engineering of lovastatin-
producing fungal strains. Nat Biotechnol 21:150–156
33
207. Stapley EO (1982) Avermectins, antiparasitic lactones produced by Streptomyces
avermitilis isolated from a soil in Japan. In: Umezawa H, Demain AL, Hata R, Hutchinson CR
(eds) Trends in antibiotic research. Japan Antibiotic Research Association, Tokyo, pp 154–
170
208. Lee J-Y, Hwang Y-S, Kim S-S, Kim E-S, Choi C-Y (2000) Effect of a global regulatory
gene, afsR2, from Streptomyces lividans on avermectin production in Streptomyces
avermitilis. Biosci Bioeng 89:606–608
209. Ikeda H, Takada Y, Pang C-H, Tanaka H, Omura S (1993) Transposon mutagenesis by
Tn4560 and applications with avermectin-producing Streptomyces avermitilis. J Bacteriol
175:2077–2082
210. Stutzman-Engwall K, Conlon S, Fedechko R, McArthur H, Pekrun K, Chen Y, Jenne S, La
C, Trinh N, Kim S, Zhang Y-X, Fox R, Gustafsson C, Krebber A (2005) Semisynthetic DNA
shuffling of aveC leads to improved industrial scale production of doramectin by
Streptomyces avermitilis. Metab Eng 7:27–37
211. Zhang X, Chen Z, Li M, Wen Y, Song Y, Li J (2006) Construction of ivermectin producer
by domain swaps of avermectin polyketide synthase in Streptomyces avermitilis. Appl
Microbiol Biotechnol 72:986–994
212. McArthur HIA (1998) The novel avermectin, Doramectin-a successful application of
mutasynthesis. In: Hutchinson CR, McAlpine J (eds) Developments in industrial
microbiology-BMP 97. Society for Industrial Microbiology, Fairfax, pp 43–48
213. Vezina C, Kudelski A, Sehgal SN (1975) Rapamycin (AY 22,989), a new antifungal
antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle.
J Antibiot 28:721–726
214. Kino T, Hatanaka H, Hashimoto M, Nishiyama M, Goto T, Okuhara M, Kohsaka M, Aoki H,
Imanaka H (1987) FK-506, a novel immunosuppressant isolated from Streptomyces. I.
Fermentation, isolation and physico-chemical and biological characteristics. J Antibiot
40:1249–1255
215. Jung S, Moon S, Lee K, Park Y-J, Yoon S et al (2009) Strain development of Streptomyces
sp. for tacrolimus production using sequential adaptation. J Ind Microbiol Biotechnol
36:1467–1471
216. Chen X, Wei P, Fan L, Yang D, Zhu X, Shen W et al (2009) Generation of high-yield
rapamycin-producing strains through protoplast-related techniques. Appl Microbiol
Biotechnol 83:507–512
217. Kuscer E, Coates N, Challis I, Gregory M, Wilkinson B, Sheridan R, Petkovic H (2007)
Roles of rapH and rapG in positive regulation of rapamycin biosynthesis in Streptomyces
hygroscopicus . J Bacteriol 189:4756–4763
218. Jin ZH, Xu B, Lin SZ, Jin QC, Cen PL (2009) Enhanced production of spinosad in
Saccharopolyspora spinosa by genome shuffling. Appl Biochem Biotechnol 159:655–663
219. Leader B, Baca QJ, Golan DE (2008) Protein therapeutics: a summary and
pharmacological classification. Nat Rev Drug Discov 7:21–39
220. Demain AL, Vaishnav P (2009) Production of recombinant proteins by microbes and higher
organisms. Biotechnol Adv 27:297–306
221. Rayder RA (2008) Expression systems for process and product improvement. Bioprocess
Int 6:4–9
222. Choi BK, Bobrowicz P, Davidson RC, Hamilton SR, Kung DH, Li H et al (2003) Use of
combinatorial genetic libraries to humanize N-linked glycosylation in the yeast Pichia
pastoris . Proc Natl Acad Sci USA 100:5022–5027
223. Yuan L, Kurek I, English J, Keenan R (2005) Laboratory-directed protein evolution.
Microbiol Mol Biol Rev 69:373–392
224. Terpe K (1996) Overview of bacterial expression systems for heterologous protein
production: from molecular and biochemical fundamentals to commercial systems. Appl
34
Microbiol Biotechnol 72:211–223
225. Swartz J (1996) Escherichia coli recombinant DNA technology. In: Neidhardt FC (ed)
Esherichia coli and Salmonella: cellular and molecular biology, 2nd edn. American
Society of Microbiology, Washington, DC, pp 1693–1771
226. Wong MS, Wu S, Causey TB, Bennett GN, San K-Y (2008) Reduction of acetate
accumulation in Escherichia coli cultures for increased recombinant protein production.
Metab Eng 10:97–108
227. Morrow KJ (2009) Grappling with biologic manufacturing concerns. Genet Eng Biotechnol
News 29(5):54–55
228. Choi JH, Lee SJ, Lee SJ, Lee SY (2003) Enhanced production of insulin-like growth factor I
fusion protein in Escherichia coli by coexpression of the down-regulated genes identified by
transcriptome profiling. Appl Environ Microbiol 69:4737–4742
229. Barnard GC, Henderson GE, Srinivasan S, Gerngross TU (2004) High level recombinant
T7 RNA polymerase based amplification. Protein Expr Purif 38:264–271
230. Squires CH, Lucy P (2008) Vendor voice: a new paradigm for bacterial strain engineering.
Bioprocess Int 6:22–27
231. Sreekrishana K, Nelles L, Potenz R, Cruze J, Mazzaferro P et al (1989) High level
expression, purification, and characterization of recombinant human tumor necrosis factor
synthesized in the methylotrophic yeast Pichia pastoris. Biochemistry 28:4117–4125
232. Werten MWT, van den Bosch TJ, Wind RD, Mooibroek H, De Wolf FA (1999) High yield
secretion of recombinant gelatins by Pichia pastoris. Yeast 15:1087–1096
233. Clare JJ, Rayment FB, Ballantine SP, Sreekrishna K, Romanos MA (1991) High-level
expression of tetanus toxin fragment C in Pichia pastoris strains containing multiple
tandem integrations of the gene. Nat Biotechnol 9:455–460
234. Nevalainen KMH, Te’o VSJ, Bergquist PL (2005) Heterologous protein expression in
filamentous fungi. Trends Biotechnol 23: 468–474
235. Morrow KJ (2007) Strategic protein production. Genet Eng Biotechnol News 27:50–54
236. Gellison G, Janowicz ZA, Weydemann U, Melber K, Strasser AWM, Hollenberg CP
(1992) High-level expression of foreign genes in Hansenula polymorpha. Biotechnol Adv
10:179–189
237. Meyer V (2008) Genetic engineering of filamentous fungi-progress, obstacles and future
trends. Biotechnol Adv 26:177–185
238. Lubertozzi D, Keasling JD (2009) Developing Aspergillus as a host for heterologous
expression. Biotechnol Adv 27:53–75
239. Van Hartinsveldt W, van Zeijl CM, Harteeld GM, Gouka RJ, Suykerbuyk M, Luiten RG et al
(1993) Cloning, characterization and overexpression of the phytase-encoding gene (phyA) of
Aspergillus niger. Gene 127:87–94
240. Ward PP, Piddlington CS, Cunningham GA, Zhou X, Wyatt RD, Conneely OM (1995) A
system for production of commercial quantities of human lactoferrin: a broad spectrum
natural antibiotic. Nat Biotechnol 13:498–503
241. Christensen T, Woeldike H, Boel E, Mortensen SB, Hjortshoej K, Thim L et al (1988) High
level expression of recombinant genes in Aspergillus oryzae. Nat Biotechnol 6:1419–1422
242. Headon DR, Wyatt RD (1995) Human lactoferrin from Aspergillus spp. SIM News 45:113–
117
243. Verdoes JC, Punt PJ, Burlingame R, Bartels J, van Dijk R, Slump E, Meens M, Joosten R,
Emalfarb M (2007) A dedicated vector for efficient library construction and high throughput
screening in the hyphal fungus Chrysosporium lucknowense. Ind Biotechnol 3:48–57
244. Andersen DC, Krummen L (2002) Recombinant protein expression for therapeutic
applications. Curr Opin Biotechnol 13:117–123
245. Wrotnowski C (1998) Animal cell culture; novel systems for research and production.
Genet Eng News 18(3):13–37
35
246. Griffin TJ, Seth G, Xie H, Bandhakavi S, Hu W-S (2007) Advancing mammalian cell culture
engineering using genome-scale technologies. Trends Biotechnol 25:401–408
247. Decaria P, Smith A, Whitford W (2009) Many considerations in selecting bioproduction
culture media. Bioprocess Int 7:44–51
248. Scott C, Montgomery SA, Rosin LJ (2007) Genetic engineering leads to microbial, animal
cell, and transgenic expression systems. BIO Internat Convention, pp. 27–34
249. Morrow KJ (2007) Improving protein production processes. Genet Eng Biotechnol News
27:44–47
250. Ryll T (2008) Antibody production using mammalian cell culture—how high can we push
productivity? SIM Annual Meeting Program & Abstract, San Diego, S146, p.101
251. Meyer HP, Biass J, Jungo C, Klein J, Wenger J, Mommers R (2008) An emerging star for
therapeutic and catalytic protein production. Bioprocess Int 6:10–21
252. CocoMartin JM, Harmsen MM (2008) A review of therapeutic protein expression by
mammalian cells. Bioprocess Int 6:28–33
253. Jarvis LM (2008) A technology bet. DSM’s pharma product unit leverages its biotech
strength to survive in a tough environment. Chem Eng News 86:30–31
254. Agathos SN (1991) Production scale insect cell culture. Biotechnol Adv 9:51–68
255. Luckow VA, Summers MD (1988) Trends in the development of baculovirus expression
vectors. Nat Biotechnol 6:47–55
256. Miller LK (1988) Baculoviruses as gene expression vectors. Annu Rev Microbiol 42:177–
199
257. Wilkinson BE, Cox M (1998) Baculovirus expression system: the production of proteins for
diagnostic, human therapeutic or vaccine use. Genet Eng News 18, 35(Nov)
258. Maiorella B, Harano D (1988) Large scale insect cell culture for recombinant protein
production. Nat Biotechnol 6:1406–1409
259. Morrow KJ Jr (2007) Improving protein production processes. Genet Eng News 27:50–54
260. Knight P (1991) Baculovirus vectors for making proteins in insect cells. ASM News 57:567–
570
261. Falch E (1991) Industrial enzymes-developments in production and application. Biotechnol
Adv 9:643–658
262. Cowan D (1996) Industrial enzyme technology. Trends Biotechnol 14:177–178
263. Vohra A, Satyanarayana T (2003) Phytases: microbial sources, production, purification,
and potential biotechnological applications. Crit Rev Biotechnol 23:29–60
264. Vaishnav P, Demain AL (2009) Industrial biotechnology overview. In: Schaechter M,
Lederberg J (eds) Encyclopedia of microbiology, 3rd edn. Elsevier, Oxford, p 335
265. Wackett LP (1997) Bacterial biocatalysis: stealing a page from nature’s book. Nat
Biotechnol 15:415–416
266. Palva I (1982) Molecular cloning of alpha-amylase gene from Bacillus amyloliquefaciens
and its expression in B. subtilis . Gene 19: 81–87
267. O’Neill GP, Kilburn DG, Warren RAJ, Miller RC (1986) Overproduction from a cellulase
gene with a high guanosine-plus-cytosine content in Escherichia coli. Appl Environ
Microbiol 52:737–743
268. Shoemaker S, Schweickart V, Ladner M, Gelfand D, Kwok S, Myambo K et al (1983)
Molecular cloning of exo-cellobiohydrolase I derived from Trichoderma reesei strain L27.
Nat Biotechnol 1:691–696
269. Van Brunt J (1986) Fungi: the perfect hosts? Biotechnology 4:1057–1062
270. Mondou F, Shareck F, Morosoli R, Kleupfel D (1986) Cloning of the xylanase gene of
Streptomyces lividans. Gene 49:323–329
271. Van den Burg B, Vriend G, Veltman O, Venema G, Eijsink VGH (1998) Engineering
an enzyme to resist boiling. Proc Natl Acad Sci USA 95:2056–2060
272. Ness JE, Welch M, Giver L, Bueno M, Cherry JR, Borchert TV et al (1999) DNA shuffling
36
of subgenomic sequences of subtilisin. Nat Biotechnol 17:893–896
273. Jaeger KE, Reetz MT (2000) Directed evolution of enantioselective enzymes for organic
chemistry. Curr Opin Chem Biol 4:68–73
274. Suenaga H, Mitsokua M, Ura Y, Watanabe T, Furukawa K (2001) Directed evolution of
biphenyl dioxygenase: emergence of enhanced degradation capacity for benzene, toluene,
and alkylbenzenes. J Bacteriol 183: 5441–5444
275. Song JK, Rhee JS (2001) Enhancement of stability and activity of phospholipase A(1) in
organic solvents by directed evolution. Biochim Biophys Acta 1547:370–378
276. Raillard S, Krebber A, Chen Y, Ness JE, Bermudez E, Trinidad R et al (2001) Novel
enzyme activities and functional plasticity revealed by recombining highly homologous
enzymes. Chem Biol 8:891–898
277. Kurtzman AL, Govindarajan S, Vahle K, Jones JT, Heinrichs V, Patten PA (2001)
Advances in directed protein evolution by recursive genetic recombination: applications to
therapeutic proteins. Curr Opin Biotechnol 12:361–370
278. Marshall SH (2002) DNA shuffling: induced molecular breeding to produce new generation
long-lasting vaccines. Biotechnol Adv 20:229–238
279. Locher CP, Soong NW, Whalen RG, Punnonen J (2004) Development of novel vaccines
using DNA shuffling and screening strategies. Curr Opin Mol Ther 6:34–39
280. Tobin MB, Gustafsson C, Huisman GW (2000) Directed evolution: the ‘rational’ basis for
‘irrational’ design. Curr Opin Struct Biol 10:421–427
281. Crameri A, Whitehorn A, Stemmer WPC (1996) Improved green fluorescent protein by
molecular evolution using DNA shuffling. Nat Biotechnol 14:315–319
37