Abstract
Abstract: Genomic and proteomic advances in extremophile microorganism studies are increasingly demonstrating their ability to produce a variety of enzymes capable of converting biomass into bioenergy. Such microorganisms are found in environments with nutritional restrictions, anaerobic environments, high salinity, varying pH conditions and extreme natural environments such as hydrothermal vents, soda lakes, and Antarctic sediments. As extremophile microorganisms and their enzymes are found in widely disparate locations, they generate new possibilities and opportunities to explore biotechnological prospecting, including biofuels (biogas, hydrogen and ethanol) with an aim toward using multi-omics tools that shed light on biotechnological breakthroughs.
Keywords: Biodiversity, microorganisms, enzymes, molecular methods, multi-omics, extremophiles
1. MICROBIAL COMMUNITIES IN EXTREME ENVIRONMENTS – AN INTRODUCTION
From a human point of view, some habitats seem so harsh and inhospitable for life that they are commonly thought to be sterile. The > 85°C hot springs in the Yellowstone Park (USA), the ultra-arid Atacama Desert (Chile), and the permanent glacier ice of South Pole (Antarctica) represent some examples. These “extreme” environments, however, are dominated by microbial cells from all three domains of life: Bacteria, Archaea, and Eukarya. Microorganisms inhabiting such environments are called “extremophiles” because of their endemic (or restricted) nature to survive at the edge of life [1]. Extremophiles exhibit unusual biochemical and ecological adaptations not found by organisms dwelling in no-extreme habitats. Physical and chemical parameters in the extreme environments include a broad range of temperature (from –40 to >130°C), salinity (from rainwater to 5 M NaCl), pH (from virtually 0 to 13), pressure (from 0.3 atm at Mount Everest to 1200 atm in the deepest point of Mariana Trench), desiccation (from wet to the complete absence of water), and radiation (wavelengths of ultraviolet and gamma rays) [2, 3].
Extremophiles are classified in categories based on environmental parameters, with sub-definitions created to accommodate organisms that present moderate, extreme, hyper-extreme and also obligate extremophile (Table 1).
Comparative studies between extremophiles and their non-extreme counterparts have led to systematic descriptions of how growth and survival are possible under extreme conditions. In general, extremophiles adjust their membranes, proteins and nucleic acids to preserve the functional stability of their structures and molecular components [4, 5]. To achieve this stability, several strategies evolved as microorganisms were exposed to environmental stresses. For example, to maintain their metabolism under extreme cold, psychrophiles must resist freezing and, at the same time, conserve membrane flexibility. These cold-adapted microorganisms enriched their membranes with unsaturated fatty acids that provide flexibility, and present enzymes enriched in hydrophilic amino acids (e.g., glycine) that help retains liquid water for metabolism [6].
Virtually all extremophiles present enzymatic adaptations to cope with their unusual environments. As enzymatic function depends on temperature, pH and water availability, for example, all intra- and extracellular reactions rely on adaptations of extremophilic enzymes-also known as “extremozymes”. While performing the same enzymatic functions as their non-extreme counterparts, extremozymes catalyze chemical reactions in conditions that would completely inhibit or even denature other enzymes [7]. In fact, some of these enzymes display a phenomenon known as polyextremophilicity, i.e., they present activity in several extreme conditions. For example, biochemical analysis of an alcohol dehydrogenase from the hyperthermophilic archaea Thermococcus sibiricus showed activity at high temperatures (optimal close to 100°C) and high salt concentrations (4 M NaCl) [8, 9]. These unique features from extremophiles and their enzymes opened a myriad of applications in bioprocesses and bioenergy research [7, 10].
2. EXTREMOZYMES AND MULTI-OMIC TOOLS
Extremozymes are enzymes obtained from extremophile microorganisms that have several ecological roles in environment and succession, being a rich niche to be explored considering enzymes and other molecules [11]. The biodegradability and high stability of extremozymes have been exploited in a large range of industrial applications with various tolerances, including acid, cold, salt and alkali-tolerant enzymes [12, 13]. Due to its catalytic capability, enzymes have been widely applied for industrial purposes, the prospection of enzymes market is in increasing growth. In 2004, it was estimated at $2 billion dollars; in 2010, the value increased to $3.3 billion; and in 2018, the estimated value of this market was $7.1 million [14-16].
With the advancement of “omics” technologies, the development of new bioinformatics tools is possible to estimate the microbial composition communities and also understand the functional aspects of these communities and prospecting genes and enzymes with a desirable characteristic or functionality [17-19]. Thus, metagenomic [20, 21], meta-transcriptomic [22] and metaproteomic [23, 24] are interesting strategies to prospect and study these enzymes and their role in extremophile communities. This variety of strategies can be used to identify molecular mechanisms or metabolic routes with possible industrial applications of microorganisms and enzymes (Fig. 1).
Using metagenomic analysis is possible to evaluate the thermophile community composition by sequencing markers like 16S rRNA [25, 26], however, analyzing the whole DNA of these communities by shotgun metagenome strategy, the potential functionality of the microbiome may be addressed and also identify coding genes that play an important role in metabolic route with biotechnological interest or genes with unknown function [19, 27, 28].
Besides metagenomics data and complete genome sequences from different species widely available and led to discover new genes, transcriptome and proteome techniques are very useful not only to identify the expressed genes but also to understand their expression with certain conditions and the physiological responses of these conditions [29-31]. Although the metatranscriptome is very useful to provide insights about the transcriptional activity of certain environmental conditions [18, 32], the mRNA and protein expressions could not be directly correlated because of the post-transcription regulations [33]. However, metaproteomic analysis is usually less restrictive and can offer rapid physiological responses since protein expression is modulated very fast. In addition, metaproteomics analysis has some important bias, which depends on the protein extraction and separation methods [34].
Though the combination of metagenome, metatranscriptome and metaproteomics could bring more evidence of the functionality and expression of certain enzymes with biotechnological interest [18], these studies with extremophile microbial communities are scarce in the literature, especially because this is still a developing field. A study of Mandelli et al. [35] performed a multi-omic approach with the extremophile Thermus filiformis using DNA sequencing, RNA-Seq and mass spectrometry. They considered that the multi-omic analyses were complementary to each other and allowed them to identify the main changes in the physiological state of T. filiformis under varying temperatures. The authors also discovered several thermostable macromolecules produced by T. filiformis, including amylases, pyrophosphatases, glucosidases, and galactosidases, all of which have a wide range of biotechnological and industrial applications.
More recently, Chignell, De Long, and Reardon [23] used label-free metaproteomics and 16S rRNA gene-based community analysis approaches to compare protein expression in acetate-fed anode biofilms before and after the onset of robust energy generation. They discovered a significant enrichment of protein categories specific to membrane and transport functions among proteins from electricity-producing biofilms. They discovered that proteins detected only in electricity-producing biofilms were associated with gluconeogenesis, the glyoxylate cycle, and fatty acid β-oxidation, as well as with denitrification and competitive inhibition.
2.1. Thermophilic Enzymes
Thermophilic enzymes (also called thermozymes) present industrial advantages, including resistance to chemical denaturants, thermostability, thermal activation and high catalytic activity, capable of optimal operation even in high-temperature processes [36, 37]. Thermozymes have low adhesiveness and contamination risk and also greater solubility of substrates [38]. These advantages can be attributed to a small number of specific modifications, including a high content of alanine and arginine acids, barrel folding, hydrophobic interactions and disulfide bonds [39, 40]. In addition, the introduction of disulfide and hydrogen bonds, along with hydrolytic changes, are being explored as mechanisms of improving these enzymes' performance [41-43].
When compared to mesophilic counterparts, thermozymes show more specific activity, better accessibility to the active site and compact oligomers [44]. In addition, these enzymes have characteristics suitable for industrial exploration, such as the maintenance of its thermal properties even when expressed in mesophilic hosts, and without of phylogenetic relationships among them. It suggests that these characteristics are genetically encoded and thus, these organisms are appropriated metabolic engineering platforms [45, 46]. Among the range of known thermozymes a wide number of enzymes have been characterized, such as pullulanase, amylases, xylanases, chitinases, pectinases, esterases, phytases, mannanase, proteases, lipase and cellulases [43-47]. Such diversity enables different applications on biotechnological demands, or even for genetic modifications with a wide range of purposes.
2.2. Psychrophilic Enzymes
Most ecosystems on Earth are exposed to temperatures permanently below 5°C, distributed among marine and terrestrial areas. Permanently frozen formations such as glacier ice and permafrost are a vast depository of ancient viable cells [48]. Successful survival of these microorganisms in cold environments depends on a number of adaptations in their metabolism, including cold-shock proteins, ice-binding molecules, and cold-adapted enzymes.
Low temperatures impose a strong negative effect on most biological reactions. As described by the Arrhenius equation, enzymatic reaction rates decrease exponentially with any decrease in the temperature. Psychrophilic and psychrotolerant microorganisms evolved their enzymes to cope with this thermodynamic problem [49]. Several structural adaptations are observed in cold-adapted enzymes, most of them increasing molecular flexibility. For example, reduced number of hydrogen bonds and salt bridges [48, 50], reduced use of proline and arginine [13, 51, 52] and increased surface-loaded residues [53] provide greater flexibility and less enthalpic contribution to stability [51]. Other important structural adaptations, such as loop extensions, provide greater accessibility of the active site for substrate and cofactor binding [54, 55], and also improve electrostatics in the vicinity of the reactive center [54, 56], ensuring the efficiency of these enzymes in the cold.
According to the type of catalyzed reaction, enzymes are classified into six main classes: oxidoreductases, transferases, hydrolases, lyases, isomerases and ligases [57]. Cold-adapted enzymes are found among all these classes (Table 2).
In recent years, the search for cold-adapted enzymes was improved by metagenomic approaches. Arctic and Antarctic glaciers, snow, lakes, soil and marine sediments had their microbial community characterized and cold-adapted enzyme producing organisms were identified. Oh et al. [115] used metagenomics in Antarctic soils to search for microorganisms able to produce cold-adapted lignolytic and cellulolytic enzymes. The metagenomics analysis found that 162 (1.42%) of a total of 11,436 genes were annotated as active carbohydrate enzymes. Actinobacteria, the dominant phylum in that soil metagenome, harbored most of the genes related to catabolism of lignocellulose, including glycoside hydrolases (GH13, GH26 and GH5). On the other hand, degradation pathways of lignocellulose in the Antarctic soil metagenome have shown a synergistic role for several bacterial genera, including Streptomyces, Streptosporangium and Amycolatopsis.
There are several advantages to using cold-adapted enzymes in biotechnology. These enzymes are generally thermolabile and may be inactivated by heat-controlled procedures [11]. They catalyze the desired reactions at lower temperatures, present high catalytic efficiency at small amounts of the enzyme, and have lower activation energy cost [57].
2.3. Piezophilic Enzymes
Naturally, the Earth shows several environments with high-pressure, with 62% of the total biosphere characterized by pressures greater than 10 bars [116]. In this context, several evolutionary events occurred independently, selecting piezophilic organisms in the sea and vulcanos [117-120]. Initially, they were referred to as barophilic, later changed to piezophilic (piezo in Greek means pressure); they are characterized by several adaptations [121, 122]. The mechanisms of adaptation include polyunsaturated acid fatty acids, Docosahexaenoic Acid (DHA) and Eicosapentaenoic Acid (EPA), both being necessary to the maintenance and fluidity of lipid membranes, presence of electrostatic and hydrogen bond in place, beyond that expression of genes regulated by pressure such as ompH and ompL present in Photobacterium spp. [123, 124]. Some piezophilic organisms have no saline channels, distinguishing them from thermopiezophiles that are adapted to low temperature and pressure [125, 126].
Pressure usually has been assumed as denaturant of proteins, whereas structural characteristics can ensure resistance to different pressures. However, the activity and/or stability of enzymes can be potentialized utilizing hydrostatic pressure, as thermolysin and hydrogenase [127, 128]. Sulfolobus solfataricus is an example of a thermoacidophilic archaebacterial isolated from vulcanos that produce a ribonuclease with piezophilic adaptation, with possibilities of industry application. Thus, piezophilic enzymes show potential for biotechnology and industry applications. They can optimize processes such as the production of trisaccharide instead of maltobiose and tetrasaccharide using less energy of substrate or high efficiency as a detergent in chemical products [11, 71, 102, 129, 130].
2.4. Acidophilic Enzymes
Acidophilic microorganisms can be found in several environments containing pH between 5-3 or less, showing optimal growth in this range of pH. Among these microorganisms, there are prokaryotes and eukaryotes found growing in acid lakes. Acid tolerance differs from acidophilia; acid-tolerant organisms can resist variation of pH and do not necessarily depend on it, whereas acidophilia depends on a range of pH to growth [131, 132].
Several adaptations are required to permit their survival; the intracellular pH is neutral in general and cytoplasmic enzymes show an optimal function at pH 7 [94, 133]. Beyond that, the extracellular enzymes and redox-active proteins located at periplasm of gram-negative acidophilic show stability at low pH. These organisms are distributed into three domains. The major groups of archaeal acidophilic are Euryarchaeota and Crenarchaeota, within bacterial phyla exist some groups known by containing acidophiles, including Firmicutes, Actinobacteria, Proteobacteria, Nitrospira and Aquifex [132].
There are some industrial applications utilizing these microorganisms; one of the possibilities of biotechnology using acidophilic microorganisms includes biomining utilizing organisms to solubilize and extract minerals from ore [134, 135]. A comparison between archaea (Acidianus brierleyi) and bacteria (Acidithiobacillus ferrooxidans) demonstrated the solubilization of metals (chalcopyrite, sphalerite and pyrite), demonstrating a better result for archaea at 65ºC/pH 1.2 than for bacteria at 30°C/pH 2 [136]. Another application involves cellulolytic and xylanolytic enzymes combined with high temperature and pH promote processes of hydrolysis [94, 114].
2.5. Halophilic Enzymes
Halophilic microorganisms are found in the most saline environments, from natural salt lakes to solar salterns, found even in the Dead Sea [137]. They can be classified according to the optimal saline concentration, including extreme halophiles, (developing at 2,500-5,200 mM NaCl), moderate halophiles (developing at 500-2,500 mM NaCl) and the slight halophiles (developing at 200-500 mM NaCl) [138]. Halotolerant bacteria maintain the osmotic balance, excluding salt from the cytoplasm (‘salt-out’ strategy) and using organic solutes in energy costly strategies [139]. On the other hand, haloarchaea sustain osmotic balance in high salinities via equimolar salt accumulation in the cytoplasm (‘salt-in’ strategy), using potassium due to its lower hydrophilicity when compared to sodium [140, 141]. Although being aerobic heterotrophs, some haloarchaea have the potential for anaerobic growth [142].
Some of the most studied halophiles genus include Natronococcus, Haloferax, Halobacterium, Halobacillus, Halorhabdus, Halothermothrix, Micrococcus, Marinococcus, Acinetobacter, Bacillus, producing different proteins, such as amylases, lipases, proteases, and xylanases [143-145]. Haloenzyme properties include their stability in the presence of organic solvents, high functionality through a large pH variation (3.0 to 10.5) and even high temperatures (80°C) [101, 146]. Halophilic proteins are maintained in solutions through a network of hydrated cations coordinated by negative charges, and they also possess small hydrophilic residues on protein surfaces, unlike the common large hydrophobic residues [147, 148]. However, these characteristics give these enzymes low solubility in aqueous/organic and non-aqueous media [149, 150].
3. POTENTIAL USE OF EXTREMOZYMES IN BIOENERGY CHAIN
The use of extremozymes has provided the industry with a wide range of resistant biomolecules with various applications, meeting the main objectives in enzyme research, with novel extreme activities and improved stability [13, 38]. The large biotechnological interest and applications can be attributed to their consistency, selectivity, efficacy, reproducibility, diversity and the low production of by-products, contributing to environmental impact decreases as a greener solution for many industrial purposes [58, 151, 152].
There are various types of extremozymes suitable for the energy production process, including lipases, esterases, cellulases and xylanases [153-155]. The use of extremophile enzymes offers the advantages of working at high temperatures and eliminating some procedural steps, increasing sugar production, decreasing production cost and improving process efficiency [156].
Among the biotechnological applications of extremozymes, biofuels production has been consistently increasing. Crosby et al. [46] discuss the utilization of extremophile microbes in industrial processes. Focusing on extreme thermophiles from the genera Caldicellulosiruptor, Pyrococcus, Thermococcus, Thermotoga and Sulfolobus, the authors were able to engineer the metabolism of these microorganisms to increase the production of valuable industrial products, including biohydrogen, bioethanol, butanol and others. The main setback is the scalability of such systems is that these studies have been conducted in laboratory conditions and have yet to be tested on an industrial scale. Thus, the enzymes of extremophiles and themselves can be applied to the production of second-generation ethanol, microorganisms such as Zymomonas mobilis, Bacillus subtilis, Geobacillus thermoglucosidasius, Clostridium thermocellum, Thermoanaerobacter ethanolicus and Thermoanaerobacter mathranii have been investigated [157-163]. Their adaptations allow the fermentation and distillation concomitant, eliminating the step of cool utilized before the distillation [164]. Beyond that, some advantages such as shown by Clostridium thermocellum that can metabolize the pentose and hexose sugars resulting in the fermentation of cellulose to ethanol contain potential to application in industry or the tolerance of Zymomonas mobilis to the production of ethanol [161, 165, 166].
Other extremozymes that are under the interest of industry applications are glucosidases capable of hydrolyzing steryl glucosides that harm the ethanol production due to the formation of insoluble precipitates and thus cause aggregation, crystallization and precipitation of other molecules [167, 168]. These consequences can clog or block oil filters, hamper cold-flow conditions and cause engine failures, increasing production costs [169, 170]. In this sense, thermostable glucosidases capable of hydrolyzing steryl glucosides in biodiesel would produce free sugars and sterols (soluble in biodiesel) enabling water washing steps, due to the enzyme’s thermostability and solvent tolerance [168, 171].
CONCLUSION
Extremophile microorganisms can produce enzymes capable of converting biomass into bioenergy (biogas, biohydrogen, and bioethanol). In this context, combined genomics, transcriptomics, and proteomics technologies open enormous possibilities for prospecting new extremophile microorganism and their noble extremozymes for biotechnological applications.
Fig. (1).
“Multi-omics” strategies and basic workflows. Based on metagenomic, meta-transcriptomic and metaproteomic approaches to prospect genes and proteins for biotechnological application. (A higher resolution / colour version of this figure is available in the electronic copy of the article).
Table 1.
Classification of extremophiles according to environmental parameters.
| Environmental Parameter | Classification | Definition | Examples |
|---|---|---|---|
| Temperature | Hyperthermophile | Optimal growth > 85°C | Pyrolobus fumarii, (113°C) |
| Thermophile | Optimal growth 45–85°C | Synechococcus lividis | |
| Psychrophile | Optimal growth < 15°C | Psychrobacter sp., Polaromonas vacuolata | |
| Salinity | Halophile | Requires 2–5 M NaCl | Halobacterium salinarum, Danaliella salina |
| pH | Alkaliphile | pH > 8 |
Natronobacterium, Bacillus firmus OF4, Spirulina spp. (all pH 10.5) |
| Acidophile | pH < 5 | Cyanidium caldarium, Ferroplasma sp. (both pH 0) | |
| Pressure | Piezophile | Growth > 400 atm | Thermococcus piezophilus, Shewanella oneidensis |
| Desiccation | Xerophile | Aw < 0.8 | Trichosporonoides nigrescens |
| Radiation | Radiotolerant | Up to 60 Gy/hour | Deinococcus radiodurans |
Adapted from Duarte et al. [3].
Table 2.
Extremozymes isolated from different microorganisms.
| Enzyme Name | Adaptation | Host Organism | References |
|---|---|---|---|
| Oxidoreductases | |||
| Alcohol dehydrogenase | Hot | Thermococcus sibiricus | [8, 9] |
| Catalase | Cold | Bacillus sp. N2a | [58, 59] |
| Glutathione reductase | Cold | Colwellia psychrerythraea | [60] |
| Transferases | |||
| Aspartate aminotransferase | Cold | Pseudoalteromonas haloplanktis TAC125 | [61] |
| Glutathione s-transferase | Cold | Pseudoalteromonas sp. ANT506 | [62] |
| Glutathione s-transferase | Cold | Halomonas sp. ANT108 | [63] |
| Hydrolases | |||
| α-galactosidase | Hot | Geobacillus stearothermophilus | [44] |
| β-galactosidase | Cold | Pseudoalteromonas sp. 22b | [64, 65] |
| β-galactosidase | Cold | Rahnella sp.R3 | [66] |
| β-galactosidase | Acid | Teratosphaeria acidotherma AIU BGA-1 | [67] |
| α-amylase | Hot | Anoxybacillus sp. GXS-BL | [68] |
| α-amylase | Cold | Pseudoalteromonas sp. M175 | [69] |
| α-amylase | Cold | Luteimonas abyssi XH031T | [70] |
| Xylanase | Hot |
Caldicellulosiruptor sp. Tok7B |
[47] |
| Xylanase | Hot | Sulfolobus solfataricus strain MT4 | [71] |
| Xylanase | Cold | Flavobacterium frigidarium | [72] |
| Xylanase | Cold | Flavobacterium johnsoniae | [73] |
| Xylanase | Acid | Aureobasidium pullulans XynI | [74] |
| Chitinase | Hot | Humicola grisea | [75] |
| Chitinase | Hot | Paenibacillus barengoltzii CAU904 | [76] |
| Protease | Hot | Anoxybacillus kamchatkensis M1V | [77] |
| Protease | Cold | Pseudoalteromonas sp. NJ276 | [78] |
| Protease | Cold | Flavobacterium xanthum ANS4-15 | [79] |
| Protease | High Salt | Bacillus licheniformis TD4 | [80] |
| Asparaginase | High Salt | Bacillus aryabhattai GA5 | [81] |
| Aminopeptidase | Cold | Pseudoalteromonas haloplanktis TAC125 | [82] |
| Lipase | Hot | Pyrococcus furiosus | [83] |
| Lipase | Cold | Bacillus pumilus ArcL5 | [84] |
| Lipase | Cold | Pseudomonas sp. LSK25 | [85] |
| Lipase | Cold | Pseudomonas vancouverensis | [86] |
| Enzyme Name | Adaptation | Host Organism | References |
| Phytase | Hot | Sporotrichum thermophile BJTLR50 | [87] |
| Phytase | Cold | Rhodotorula mucilaginosa JMUY14 | [88] |
| Phytase | Cold | Bacillus licheniformis ATCC14580 | [89] |
| Esterase | Hot | Thermus thermophilus HB27 | [90] |
| Esterase | Cold | Thalassospira sp. | [91] |
| Esterase | Cold | Oleispira antarctica | [92] |
| Esterase | Cold | Pseudomonas sp.TB11 | [93] |
| Esterase | Acid | Ferroplasma acidiphilum DSM 12658 | [94] |
| α-glucosidase | Cold | Leucosporidium antarcticum | [65] |
| β-glucosidase | Cold | Micrococcus antarcticus | [95] |
| Cellulase | Hot | Myceliophthora thermophila ATCC 42464 | [96] |
| Cellulase | Hot | Geobacillus sp. HTA426 | [97] |
| Cellulase | Cold | Flavobacterium sp. AUG42 | [98] |
| Cellulase | Cold | Bacillus sp. K-11 | [99] |
| Cellulase | Alkaline | Streptomyces thermoalkaliphilus 4-2-13 | [100] |
| Cellulase | High Salt | Paenibacillus tarimensis L88 | [101] |
| Ribonuclease | Hot | Sulfolobus solfataricus | [102] |
| Ribonuclease | Cold | Psychrobacter sp. ANT206 | [103] |
| Phosphatase | Cold | Cobetia marina | [104] |
| Lyases | |||
| Alginate lyase | Hot | Nitratiruptor sp. SB155-2 | [105] |
| Γ-carbonic anhydrase | Cold | Colwellia psychrerythraea | [106] |
| Glutamic acid decarboxylase | Cold | Colwellia psychrerythraea | [107, 108] |
| Pectate lyase | Cold | Massilia eurypsychrophila | [109] |
| Pectate lyase | Acid-Alkaline | Bacillus amyloliquefaciens S6 | [110] |
| Isomerases | |||
| Sedoheptulose 7 phosphate isomerase | Cold | Colwellia psychrerythraea 34H | [107] |
| Triose phosphate isomerase | Cold | Pseudomonas sp. π9 | [111] |
| Ligases | |||
| Glutathione synthetase | Cold | Pseudoalteromonas haloplanktis | [112] |
| DNA ligase | Cold | Pseudoalteromonas haloplanktis TAE72 | [113] |
| DNA ligase | Acid | Ferroplasma acidarmanus Fer1 | [114] |
Acknowledgements
Declared none.
Consent for Publication
Not applicable.
Funding
The UFSC group is supported by the CAPES/PRINT Project (project number 88881.310783/2018-01) and scholarships by CNPq and CAPES agencies.
CONFLICT OF INTEREST
The authors declare no conflict of interest, financial or otherwise.
REFERENCES
- 1.Rothschild L.J., Mancinelli R.L. Life in extreme environments. Nature. 2001;409(6823):1092–1101. doi: 10.1038/35059215. [DOI] [PubMed] [Google Scholar]
- 2.Durvasula R., Rao D.V.S. In: Extremophiles: From Biology to Biotechnology; Durvasula, R.V. Rao D.V.S., editor. CRC Press; 2018. pp. 1–18. [Google Scholar]
- 3.Duarte R.T.D., Nóbrega F., Nakayama C.R., Pellizari V.H. Brazilian research on extremophiles in the context of astrobiology. Int. J. Astrobiol. 2012;11(4):325–333. doi: 10.1017/S1473550412000249. [DOI] [Google Scholar]
- 4.Morozkina E.V., Slutskaia E.S., Fedorova T.V., Tugaĭ T.I., Golubeva L.I., Koroleva O.V. Extremophilic microorganisms: biochemical adaptation and biotechnological application. Prikl. Biokhim. Mikrobiol. 2010;46(1):5–20. doi: 10.1134/S0003683810010011. [review]. [DOI] [PubMed] [Google Scholar]
- 5.Brininger C., Spradlin S., Cobani L., Evilia C. The more adaptive to change, the more likely you are to survive: Protein adaptation in extremophiles. Semin. Cell Dev. Biol. 2018;84:158–169. doi: 10.1016/j.semcdb.2017.12.016. [DOI] [PubMed] [Google Scholar]
- 6.Violot S., Aghajari N., Czjzek M., Feller G., Sonan G.K., Gouet P., Gerday C., Haser R., Receveur-Bréchot V. Structure of a full length psychrophilic cellulase from Pseudoalteromonas haloplanktis revealed by X-ray diffraction and small angle X-ray scattering. J. Mol. Biol. 2005;348(5):1211–1224. doi: 10.1016/j.jmb.2005.03.026. [DOI] [PubMed] [Google Scholar]
- 7.Dumorné K., Córdova D.C., Astorga-Eló M., Renganathan P. Extremozymes: a potential source for industrial applications. J. Microbiol. Biotechnol. 2017;27(4):649–659. doi: 10.4014/jmb.1611.11006. [DOI] [PubMed] [Google Scholar]
- 8.Stekhanova T.N., Mardanov A.V., Bezsudnova E.Y., Gumerov V.M., Ravin N.V., Skryabin K.G., Popov V.O. Characterization of a thermostable short-chain alcohol dehydrogenase from the hyperthermophilic archaeon Thermococcus sibiricus. Appl. Environ. Microbiol. 2010;76(12):4096–4098. doi: 10.1128/AEM.02797-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Bezsudnova E.Y., Boyko K.M., Polyakov K.M., Dorovatovskiy P.V., Stekhanova T.N., Gumerov V.M., Ravin N.V., Skryabin K.G., Kovalchuk M.V., Popov V.O. Structural insight into the molecular basis of polyextremophilicity of short-chain alcohol dehydrogenase from the hyperthermophilic archaeon Thermococcus sibiricus. Biochimie. 2012;94(12):2628–2638. doi: 10.1016/j.biochi.2012.07.024. [DOI] [PubMed] [Google Scholar]
- 10.Rampelotto P.H. 2016. Biotechnology of Extremophiles: Advances and Challenges, Grand Challenges in Biology and Biotechnology. [Google Scholar]
- 11.Cavicchioli R., Amils R., Wagner D., McGenity T. Life and applications of extremophiles. Environ. Microbiol. 2011;13(8):1903–1907. doi: 10.1111/j.1462-2920.2011.02512.x. [DOI] [PubMed] [Google Scholar]
- 12.Jaenicke R., Schurig H., Beaucamp N., Ostendorp R. In: Structure and stability of hyperstable proteins: glycolytic enzymes from hyperthermophilic bacterium Thermotoga maritima. Advances in protein chemistry and structural biology: Enzymes and proteins from hyperthermophilic microorganisms; Richards, F.M.; Eisenberg, D.S. Kim P.S., editor. 1996. pp. 181–269. [DOI] [PubMed] [Google Scholar]
- 13.Cavicchioli R., Siddiqui K.S., Andrews D., Sowers K.R. Low-temperature extremophiles and their applications. Curr. Opin. Biotechnol. 2002;13(3):253–261. doi: 10.1016/S0958-1669(02)00317-8. [DOI] [PubMed] [Google Scholar]
- 14.Joseph B., Ramteke P.W., Thomas G. Cold active microbial lipases: some hot issues and recent developments. Biotechnol. Adv. 2008;26(5):457–470. doi: 10.1016/j.biotechadv.2008.05.003. [DOI] [PubMed] [Google Scholar]
- 15.Sarrouh B. Santos, T.M.; Miyoshi, A.; Dias, R.; Azevedo, V. Up-to-date insight on industrial enzymes applications and global market. J. Bioprocess. Biotech. 2012;S4(002):1–10. doi: 10.4172/2155-9821.S4-002. [DOI] [Google Scholar]
- 16.Dewan S.S. Global Markets for Enzymes in Industrial Applications. https://www.bccresearch.com/market- research/biotechnology/enzymes-industrial-applications-bio030h.html
- 17.Madhavan A., Sindhu R., Parameswaran B., Sukumaran R.K., Pandey A. Metagenome analysis: a powerful tool for enzyme bioprospecting. Appl. Biochem. Biotechnol. 2017;183(2):636–651. doi: 10.1007/s12010-017-2568-3. [DOI] [PubMed] [Google Scholar]
- 18.Hassa J., Maus I., Off S., Pühler A., Scherer P., Klocke M., Schlüter A. Metagenome, metatranscriptome, and metaproteome approaches unraveled compositions and functional relationships of microbial communities residing in biogas plants. Appl. Microbiol. Biotechnol. 2018;102(12):5045–5063. doi: 10.1007/s00253-018-8976-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Ngara T.R., Zhang H. Recent advances in function-based metagenomic screening. Genomics Proteomics Bioinformatics. 2018;16(6):405–415. doi: 10.1016/j.gpb.2018.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Dávila-Ramos S., Castelán-Sánchez H.G., Martínez-Ávila L., Sánchez-Carbente M.D.R., Peralta R., Hernández-Mendoza A., Dobson A.D.W., Gonzalez R.A., Pastor N., Batista-García R.A. A review on viral metagenomics in extreme environments. Front. Microbiol. 2019;10:2403. doi: 10.3389/fmicb.2019.02403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Gómez-Silva B., Vilo-Muñoz C., Galetović A., Dong Q., Castelán-Sánchez H.G., Pérez-Llano Y., Sánchez-Carbente M.D.R., Dávila-Ramos S., Cortés-López N.G., Martínez-Ávila L., Dobson A.D.W., Batista-García R.A. Metagenomics of Atacama lithobiontic extremophile life unveils highlights on fungal communities, biogeochemical cycles and carbohydrate-active enzymes. Microorganisms. 2019;7(12):1–25. doi: 10.3390/microorganisms7120619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Macklaim J.M., Gloor G.B. 2018. From RNA-seq to biological inference: using compositional data analysis in meta-transcriptomics. [DOI] [PubMed] [Google Scholar]
- 23.Chignell J.F., De Long S.K., Reardon K.F. Meta-proteomic analysis of protein expression distinctive to electricity-generating biofilm communities in air-cathode microbial fuel cells. Biotechnol. Biofuels. 2018;11(1):121. doi: 10.1186/s13068-018-1111-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Hart E.H., Creevey C.J., Hitch T., Kingston-Smith A.H. Meta-proteomics of rumen microbiota indicates niche compartmentalisation and functional dominance in a limited number of metabolic pathways between abundant bacteria. Sci. Rep. 2018;8(1):10504. doi: 10.1038/s41598-018-28827-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Chan C.S., Chan K-G., Tay Y-L., Chua Y-H., Goh K.M. Diversity of thermophiles in a Malaysian hot spring determined using 16S rRNA and shotgun metagenome sequencing. Front. Microbiol. 2015;6(177):177. doi: 10.3389/fmicb.2015.00177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Mardanov A.V., Gumerov V.M., Beletsky A.V., Ravin N.V. Microbial diversity in acidic thermal pools in the Uzon Caldera, Kamchatka. Antonie van Leeuwenhoek. 2018;111(1):35–43. doi: 10.1007/s10482-017-0924-5. [DOI] [PubMed] [Google Scholar]
- 27.López-López O., Cerdán M.E., González Siso M.I. New extremophilic lipases and esterases from metagenomics. Curr. Protein Pept. Sci. 2014;15(5):445–455. doi: 10.2174/1389203715666140228153801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Ferrer M., Martínez-Martínez M., Bargiela R., Streit W.R., Golyshina O.V., Golyshin P.N. Estimating the success of enzyme bioprospecting through metagenomics: current status and future trends. Microb. Biotechnol. 2016;9(1):22–34. doi: 10.1111/1751-7915.12309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Schneider T., Riedel K. Environmental proteomics: analysis of structure and function of microbial communities. Proteomics. 2010;10(4):785–798. doi: 10.1002/pmic.200900450. [DOI] [PubMed] [Google Scholar]
- 30.Wilmes P., Bond P.L. Metaproteomics: studying functional gene expression in microbial ecosystems. Trends Microbiol. 2006;14(2):92–97. doi: 10.1016/j.tim.2005.12.006. [DOI] [PubMed] [Google Scholar]
- 31.Kleiner M. Metaproteomics: much more than measuring gene expression in microbial communities. mSystems. 2019;4(3):1–6. doi: 10.1128/mSystems.00115-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Shi Y., Tyson G.W., Eppley J.M., DeLong E.F. Integrated metatranscriptomic and metagenomic analyses of stratified microbial assemblages in the open ocean. ISME J. 2011;5(6):999–1013. doi: 10.1038/ismej.2010.189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Pradet-Balade B., Boulmé F., Beug H., Müllner E.W., Garcia-Sanz J.A. Translation control: bridging the gap between genomics and proteomics? Trends Biochem. Sci. 2001;26(4):225–229. doi: 10.1016/S0968-0004(00)01776-X. [DOI] [PubMed] [Google Scholar]
- 34.Leary D.H., Hervey W.J., IV, Deschamps J.R., Kusterbeck A.W., Vora G.J. Which metaproteome? the impact of protein extraction bias on metaproteomic analyses. Mol. Cell. Probes. 2013;27(5-6):193–199. doi: 10.1016/j.mcp.2013.06.003. [DOI] [PubMed] [Google Scholar]
- 35.Mandelli F., Couger M.B., Paixão D.A.A., Machado C.B., Carnielli C.M., Aricetti J.A., Polikarpov I., Prade R., Caldana C., Paes Leme A.F., Mercadante A.Z., Riaño-Pachón D.M., Squina F.M. Thermal adaptation strategies of the extremophile bacterium Thermus filiformis based on multi-omics analysis. Extremophiles. 2017;21(4):775–788. doi: 10.1007/s00792-017-0942-2. [DOI] [PubMed] [Google Scholar]
- 36.Zheng L., Pan L., Miao J., Lin Y., Wu J. Application of a series of biomarkers in Scallop Chlamys farreri to assess the toxic effects after exposure to a priority hazardous and noxious substance (HNS)-Acrylonitrile. Environ. Toxicol. Pharmacol. 2018;64:122–130. doi: 10.1016/j.etap.2018.10.002. [DOI] [PubMed] [Google Scholar]
- 37.Akram F., Haq I.U., Imran W., Mukhtar H. Insight perspectives of thermostable endoglucanases for bioethanol production: A Review. Renew. Energy. 2018;122:225–238. doi: 10.1016/j.renene.2018.01.095. [DOI] [Google Scholar]
- 38.Raddadi N., Cherif A., Daffonchio D., Neifar M., Fava F. Biotechnological applications of extremophiles, extremozymes and extremolytes. Appl. Microbiol. Biotechnol. 2015;99(19):7907–7913. doi: 10.1007/s00253-015-6874-9. [DOI] [PubMed] [Google Scholar]
- 39.Sammond D.W., Kastelowitz N., Himmel M.E., Yin H., Crowley M.F., Bomble Y.J. Comparing residue clusters from thermophilic and mesophilic enzymes reveals adaptive mechanisms. PLoS One. 2016;11(1):e0145848. doi: 10.1371/journal.pone.0145848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Farnoosh G., Latifi A.M. A review on engineering of Organophosphorus Hydrolase (OPH) enzyme. J. Appl. Biotechnol. Reports. 2014;1(1):1–10. [Google Scholar]
- 41.Rakotoarivonina H., Hermant B., Aubry N., Rémond C. Engineering the hydrophobic residues of a GH11 xylanase impacts its adsorption onto lignin and its thermostability. Enzyme Microb. Technol. 2015;81:47–55. doi: 10.1016/j.enzmictec.2015.07.009. [DOI] [PubMed] [Google Scholar]
- 42.Liu T., Wang Y., Luo X., Li J., Reed S.A., Xiao H., Young T.S., Schultz P.G. Enhancing protein stability with extended disulfide bonds. Proc. Natl. Acad. Sci. USA. 2016;113(21):5910–5915. doi: 10.1073/pnas.1605363113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Wahab M.K.H. bin A.; Jonet, M.A. bin; Illias, R.M. Thermostability Enhancement of Xylanase Aspergillus Fumigatus RT-1. J. Mol. Catal., B Enzym. 2016;134:154–163. doi: 10.1016/j.molcatb.2016.09.020. [DOI] [Google Scholar]
- 44.Merceron R., Foucault M., Haser R., Mattes R., Watzlawick H., Gouet P. The molecular mechanism of thermostable α-galactosidases AgaA and AgaB explained by X-ray crystallography and mutational studies. J. Biol. Chem. 2012;287(47):39642–39652. doi: 10.1074/jbc.M112.394114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Vieille C., Zeikus G.J. Hyperthermophilic enzymes: sources, uses, and molecular mechanisms for thermostability. Microbiol. Mol. Biol. Rev. 2001;65(1):1–43. doi: 10.1128/MMBR.65.1.1-43.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Crosby J.R., Laemthong T., Lewis A.M., Straub C.T., Adams M.W., Kelly R.M. Extreme thermophiles as emerging metabolic engineering platforms. Curr. Opin. Biotechnol. 2019;59:55–64. doi: 10.1016/j.copbio.2019.02.006. [DOI] [PubMed] [Google Scholar]
- 47.Sunna A., Bergquist P.L. A gene encoding a novel extremely thermostable 1,4-β-xylanase isolated directly from an environmental DNA sample. Extremophiles. 2003;7(1):63–70. doi: 10.1007/s00792-002-0296-1. [DOI] [PubMed] [Google Scholar]
- 48.Feller G., Gerday C. Psychrophilic enzymes: hot topics in cold adaptation. Nat. Rev. Microbiol. 2003;1(3):200–208. doi: 10.1038/nrmicro773. [DOI] [PubMed] [Google Scholar]
- 49.Collins T., Roulling F., Piette F., Marx J-C., Feller G., Gerday C., D’Amico S. Fundamentals of Cold-Adapted Enzymes. Psychrophiles: from Biodiversity to Biotechnology. Berlin, Heidelberg: Springer Berlin Heidelberg; 2008. pp. 211–227. [Google Scholar]
- 50.Giordano D., Coppola D., Russo R., Tinajero-Trejo M., di Prisco G., Lauro F., Ascenzi P., Verde C. 2013. [DOI] [PubMed] [Google Scholar]
- 51.D’Amico S., Collins T., Marx J-C., Feller G., Gerday C. Psychrophilic microorganisms: challenges for life. EMBO Rep. 2006;7(4):385–389. doi: 10.1038/sj.embor.7400662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Latip M.A.A., Hamid A.A.A., Nordin N.F.H. Microbial hydrolytic enzymes: in silico studies between polar and tropical regions. Polar Sci. 2019;20:9–18. doi: 10.1016/j.polar.2019.04.003. [DOI] [Google Scholar]
- 53.D’Amico S., Claverie P., Collins T., Georlette D., Gratia E., Hoyoux A., Meuwis M-A., Feller G., Gerday C. Molecular basis of cold adaptation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2002;357(1423):917–925. doi: 10.1098/rstb.2002.1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Leiros H-K.S., Pey A.L., Innselset M., Moe E., Leiros I., Steen I.H., Martinez A. Structure of phenylalanine hydroxylase from Colwellia psychrerythraea 34H, a monomeric cold active enzyme with local flexibility around the active site and high overall stability. J. Biol. Chem. 2007;282(30):21973–21986. doi: 10.1074/jbc.M610174200. [DOI] [PubMed] [Google Scholar]
- 55.Sonan G.K., Receveur-Brechot V., Duez C., Aghajari N., Czjzek M., Haser R., Gerday C. The linker region plays a key role in the adaptation to cold of the cellulase from an Antarctic bacterium. Biochem. J. 2007;407(2):293–302. doi: 10.1042/BJ20070640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Brandsdal B.O., Heimstad E.S., Sylte I., Smalås A.O. Comparative molecular dynamics of mesophilic and psychrophilic protein homologues studied by 1.2 ns simulations. J. Biomol. Struct. Dyn. 1999;17(3):493–506. doi: 10.1080/07391102.1999.10508380. [DOI] [PubMed] [Google Scholar]
- 57.Bruno S., Coppola D., di Prisco G., Giordano D., Verde C. Enzymes from marine polar regions and their biotechnological applications. Mar. Drugs. 2019;17(10):1–36. doi: 10.3390/md17100544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Sarmiento F., Peralta R., Blamey J.M. Cold and hot extremozymes: industrial relevance and current trends. Front. Bioeng. Biotechnol. 2015;3(148):148. doi: 10.3389/fbioe.2015.00148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Wang W., Sun M., Liu W., Zhang B. Purification and characterization of a psychrophilic catalase from Antarctic Bacillus. Can. J. Microbiol. 2008;54(10):823–828. doi: 10.1139/W08-066. [DOI] [PubMed] [Google Scholar]
- 60.Ji M., Barnwell C.V., Grunden A.M. Characterization of recombinant glutathione reductase from the psychrophilic Antarctic bacterium Colwellia psychrerythraea. Extremophiles. 2015;19(4):863–874. doi: 10.1007/s00792-015-0762-1. [DOI] [PubMed] [Google Scholar]
- 61.Birolo L., Tutino M.L., Fontanella B., Gerday C., Mainolfi K., Pascarella S., Sannia G., Vinci F., Marino G. Aspartate aminotransferase from the Antarctic bacterium Pseudoalteromonas haloplanktis TAC 125. Cloning, expression, properties, and molecular modelling. Eur. J. Biochem. 2000;267(9):2790–2802. doi: 10.1046/j.1432-1327.2000.01299.x. [DOI] [PubMed] [Google Scholar]
- 62.Shi Y., Wang Q., Hou Y., Hong Y., Han X., Yi J., Qu J., Lu Y. Molecular cloning, expression and enzymatic characterization of glutathione S-transferase from Antarctic sea-ice bacteria Pseudoalteromonas sp. ANT506. Microbiol. Res. 2014;169(2-3):179–184. doi: 10.1016/j.micres.2013.06.012. [DOI] [PubMed] [Google Scholar]
- 63.Hou Y., Qiao C., Wang Y., Wang Y., Ren X., Wei Q., Wang Q. Cold-adapted glutathione S-transferases from antarctic psychrophilic bacterium Halomonas sp. ANT108: heterologous expression, characterization, and oxidative resistance. Mar. Drugs. 2019;17(3):1–13. doi: 10.3390/md17030147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Cieśliński H., Kur J., Białkowska A., Baran I., Makowski K., Turkiewicz M. Cloning, expression, and purification of a recombinant cold-adapted β-galactosidase from antarctic bacterium Pseudoalteromonas sp. 22b. Protein Expr. Purif. 2005;39(1):27–34. doi: 10.1016/j.pep.2004.09.002. [DOI] [PubMed] [Google Scholar]
- 65.Turkiewicz M., Kur J., Białkowska A., Cieśliński H., Kalinowska H., Bielecki S. Antarctic marine bacterium Pseudoalteromonas sp. 22b as a source of cold-adapted β-galactosidase. Biomol. Eng. 2003;20(4-6):317–324. doi: 10.1016/S1389-0344(03)00039-X. [DOI] [PubMed] [Google Scholar]
- 66.Fan Y., Yi J., Hua X., Zhang Y., Yang R. Preparation and characterization of gellan gum microspheres containing a cold-adapted β-galactosidase from Rahnella sp. R3. Carbohydr. Polym. 2017;162:10–15. doi: 10.1016/j.carbpol.2017.01.033. [DOI] [PubMed] [Google Scholar]
- 67.Isobe K., Yamada M. 2019. [Google Scholar]
- 68.Liao S-M., Liang G., Zhu J., Lu B., Peng L-X., Wang Q-Y., Wei Y-T., Zhou G-P., Huang R-B. Influence of calcium ions on the thermal characteristics of α-amylase from thermophilic Anoxybacillus sp. GXS-BL. Protein Pept. Lett. 2019;26(2):148–157. doi: 10.2174/0929866526666190116162958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Wang X., Kan G., Ren X., Yu G., Shi C., Xie Q., Wen H., Betenbaugh M. Molecular cloning and characterization of a novel α-amylase from Antarctic sea ice Bacterium Pseudoalteromonas sp. M175 and its primary application in detergent. BioMed Res. Int. 2018;2018:3258383. doi: 10.1155/2018/3258383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Song Q., Wang Y., Yin C., Zhang X-H. Laa A, a novel high-active alkalophilic alpha-amylase from deep-sea bacterium Luteimonas abyssi XH031(T). Enzyme Microb. Technol. 2016;90:83–92. doi: 10.1016/j.enzmictec.2016.05.003. [DOI] [PubMed] [Google Scholar]
- 71.Cannio R., Di Prizito N., Rossi M., Morana A. A xylan-degrading strain of Sulfolobus solfataricus: isolation and characterization of the xylanase activity. Extremophiles. 2004;8(2):117–124. doi: 10.1007/s00792-003-0370-3. [DOI] [PubMed] [Google Scholar]
- 72.Humphry D.R., George A., Black G.W., Cummings S.P. Flavobacterium frigidarium sp. nov., an aerobic, psychrophilic, xylanolytic and laminarinolytic bacterium from Antarctica. Int. J. Syst. Evol. Microbiol. 2001;51(Pt 4):1235–1243. doi: 10.1099/00207713-51-4-1235. [DOI] [PubMed] [Google Scholar]
- 73.Chen S., Kaufman M.G., Miazgowicz K.L., Bagdasarian M., Walker E.D. Molecular characterization of a cold-active recombinant xylanase from Flavobacterium johnsoniae and its applicability in xylan hydrolysis. Bioresour. Technol. 2013;128:145–155. doi: 10.1016/j.biortech.2012.10.087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Tanaka H., Okuno T., Moriyama S., Muguruma M., Ohta K. Acidophilic xylanase from Aureobasidium pullulans: efficient expression and secretion in Pichia pastoris and mutational analysis. J. Biosci. Bioeng. 2004;98(5):338–343. doi: 10.1016/S1389-1723(04)00292-0. [DOI] [PubMed] [Google Scholar]
- 75.Kumar M., Brar A., Vivekanand V., Pareek N. 2017.
- 76.Yang S., Fu X., Yan Q., Guo Y., Liu Z., Jiang Z. Cloning, expression, purification and application of a novel chitinase from a thermophilic marine bacterium Paenibacillus barengoltzii. Food Chem. 2016;192:1041–1048. doi: 10.1016/j.foodchem.2015.07.092. [DOI] [PubMed] [Google Scholar]
- 77.Mechri S., Bouacem K., Zaraî J.N., Rekik H., Ben Elhoul M., Omrane B.M., Hacene H., Bejar S., Bouanane-Darenfed A., Jaouadi B. Identification of a novel protease from the thermophilic Anoxybacillus kamchatkensis M1V and its application as laundry detergent additive. Extremophiles. 2019;23(6):687–706. doi: 10.1007/s00792-019-01123-6. [DOI] [PubMed] [Google Scholar]
- 78.Wang Q-F., Hou Y-H., Xu Z., Miao J-L., Li G-Y. Purification and properties of an extracellular cold-active protease from the psychrophilic Bacterium Pseudoalteromonas sp. NJ276. Biochem. Eng. J. 2008;38(3):362–368. doi: 10.1016/j.bej.2007.07.025. [DOI] [Google Scholar]
- 79.Matsui M., Kawamata A., Kosugi M., Imura S., Kurosawa N. Diversity of proteolytic microbes isolated from antarctic freshwater lakes and characteristics of their cold-active proteases. Polar Sci. 2017;13:82–90. doi: 10.1016/j.polar.2017.02.002. [DOI] [Google Scholar]
- 80.Suganthi C., Mageswari A., Karthikeyan S., Anbalagan M., Sivakumar A., Gothandam K.M. Screening and optimization of protease production from a halotolerant Bacillus Licheniformis isolated from saltern sediments. J. Genet. Eng. Biotechnol. 2013;11(1):47–52. doi: 10.1016/j.jgeb.2013.02.002. [DOI] [Google Scholar]
- 81.Shirazian P., Asad S., Amoozegar M.A. The potential of halophilic and halotolerant bacteria for the production of antineoplastic enzymes: L-asparaginase and L-glutaminase. EXCLI J. 2016;15:268–279. doi: 10.17179/excli2016-146. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Pascale D., Giuliani M., De Santi C., Bergamasco N., Amoresano A., Carpentieri A., Parrilli E., Tutino M.L. PhAP protease from Pseudoalteromonas haloplanktis TAC125: gene cloning, recombinant production in E. Coli and enzyme characterization. Polar Sci. 2010;4(2):285–294. doi: 10.1016/j.polar.2010.03.009. [DOI] [Google Scholar]
- 83.Alquéres S.M.C., Branco R.V., Freire D.M.G., Alves T.L.M., Martins O.B., Almeida R.V. Characterization of the recombinant thermostable lipase (Pf2001) from Pyrococcus furiosus: effects of thioredoxin fusion tag and Triton X-100. Enzyme Res. 2011;2011316939 doi: 10.4061/2011/316939. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Wi A.R., Jeon S-J., Kim S., Park H.J., Kim D., Han S.J., Yim J.H., Kim H-W. Characterization and a point mutational approach of a psychrophilic lipase from an arctic bacterium, Bacillus pumilus. Biotechnol. Lett. 2014;36(6):1295–1302. doi: 10.1007/s10529-014-1475-8. [DOI] [PubMed] [Google Scholar]
- 85.Salwoom L., Raja Abd Rahman R.N.Z., Salleh A.B., Mohd S.F., Convey P., Pearce D., Mohamad A.M.S. Isolation, characterisation, and lipase production of a cold-adapted bacterial strain Pseudomonas sp. LSK25 isolated from Signy Island, Antarctica. Molecules. 2019;24(4):1–14. doi: 10.3390/molecules24040715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Gupta G., Prakash V. Isolation and identification of a novel, cold active lipase producing Psychrophilic Bacterium Pseudomonas vancouverensi. Trends Biosci. 2014;7(22):3708–3711. [Google Scholar]
- 87.Ranjan B., Satyanarayana T., Recombinant H.A.P. Phytase of the thermophilic mold Sporotrichum thermophile: expression of the codon-optimized phytase gene in Pichia pastoris and applications. Mol. Biotechnol. 2016;58(2):137–147. doi: 10.1007/s12033-015-9909-7. [DOI] [PubMed] [Google Scholar]
- 88.Yu P., Wang X-T., Liu J-W. Purification and characterization of a novel cold-adapted phytase from Rhodotorula mucilaginosa strain JMUY14 isolated from Antarctic. J. Basic Microbiol. 2015;55(8):1029–1039. doi: 10.1002/jobm.201400865. [DOI] [PubMed] [Google Scholar]
- 89.Borgi M.A., Khila M., Boudebbouze S., Aghajari N., Szukala F., Pons N., Maguin E., Rhimi M. The attractive recombinant phytase from Bacillus licheniformis: biochemical and molecular characterization. Appl. Microbiol. Biotechnol. 2014;98(13):5937–5947. doi: 10.1007/s00253-013-5421-9. [DOI] [PubMed] [Google Scholar]
- 90.Fuciños P., Atanes E., López-López O., Solaroli M., Cerdán M.E., González-Siso M.I., Pastrana L., Rúa M.L. Cloning, expression, purification and characterization of an oligomeric His-tagged thermophilic esterase from Thermus Thermophilus HB27. Process Biochem. 2014;49(6):927–935. doi: 10.1016/j.procbio.2014.03.006. [DOI] [Google Scholar]
- 91.De Santi C., Leiros H-K.S., Di Scala A., de Pascale D., Altermark B., Willassen N-P. Biochemical characterization and structural analysis of a new cold-active and salt-tolerant esterase from the marine bacterium Thalassospira sp. Extremophiles. 2016;20(3):323–336. doi: 10.1007/s00792-016-0824-z. [DOI] [PubMed] [Google Scholar]
- 92.Lemak S., Tchigvintsev A., Petit P., Flick R., Singer A.U., Brown G., Evdokimova E., Egorova O., Gonzalez C.F., Chernikova T.N., Yakimov M.M., Kube M., Reinhardt R., Golyshin P.N., Savchenko A., Yakunin A.F. Structure and activity of the cold-active and anion-activated carboxyl esterase OLEI01171 from the oil-degrading marine bacterium Oleispira antarctica. Biochem. J. 2012;445(2):193–203. doi: 10.1042/BJ20112113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Dong J., Zhao W., Gasmalla M.A.A., Sun J., Hua X., Zhang W., Han L., Fan Y., Feng Y., Shen Q. A novel extracellular cold-active esterase of Pseudomonas sp. TB11 from Glacier No.1: differential induction, purification and characterisation. J. Mol. Catal., B Enzym. 2015;121:53–63. doi: 10.1016/j.molcatb.2015.07.015. [DOI] [Google Scholar]
- 94.Golyshina O.V., Timmis K.N. Ferroplasma and relatives, recently discovered cell wall-lacking archaea making a living in extremely acid, heavy metal-rich environments. Environ. Microbiol. 2005;7(9):1277–1288. doi: 10.1111/j.1462-2920.2005.00861.x. [DOI] [PubMed] [Google Scholar]
- 95.Miao L-L., Hou Y-J., Fan H-X., Qu J., Qi C., Liu Y., Li D-F., Liu Z-P. Molecular structural basis for the cold adaptedness of the Psychrophilic β-glucosidase BglU in Micrococcus antarcticus. Appl. Environ. Microbiol. 2016;82(7):2021–2030. doi: 10.1128/AEM.03158-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Wang J., Gong Y., Zhao S., Liu G. 2018.
- 97.Potprommanee L., Wang X-Q., Han Y-J., Nyobe D., Peng Y-P., Huang Q., Liu J.Y., Liao Y-L., Chang K-L. Characterization of a thermophilic cellulase from Geobacillus sp. HTA426, an efficient cellulase-producer on alkali pretreated of lignocellulosic biomass. PLoS One. 2017;12(4):e0175004. doi: 10.1371/journal.pone.0175004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Herrera L.M., Braña V., Franco Fraguas L., Castro-Sowinski S. Characterization of the cellulase-secretome produced by the Antarctic bacterium Flavobacterium sp. AUG42. Microbiol. Res. 2019;223-225:13–21. doi: 10.1016/j.micres.2019.03.009. [DOI] [PubMed] [Google Scholar]
- 99.Caf Y., Valipour E., Arikan B. Study on cold-active and acidophilic cellulase (CMCase) from a novel psychrotrophic isolate Bacillus sp. K-11. Int. J. Curr. Microbiol. Appl. Sci. 2014;3(5):16–25. [Google Scholar]
- 100.Wu H., Liu B., Ou X., Pan S., Shao Y., Huang F. Streptomyces thermoalkaliphilus sp. nov., an alkaline cellulase producing thermophilic actinomycete isolated from tropical rainforest soil. Antonie van Leeuwenhoek. 2018;111(3):413–422. doi: 10.1007/s10482-017-0964-x. [DOI] [PubMed] [Google Scholar]
- 101.Raddadi N., Cherif A., Daffonchio D., Fava F. Halo-alkalitolerant and thermostable cellulases with improved tolerance to ionic liquids and organic solvents from Paenibacillus tarimensis isolated from the Chott El Fejej, Sahara desert, Tunisia. Bioresour. Technol. 2013;150:121–128. doi: 10.1016/j.biortech.2013.09.089. [DOI] [PubMed] [Google Scholar]
- 102.Fusi P., Grisa M., Mombelli E., Consonni R., Tortora P., Vanoni M. Expression of a synthetic gene encoding P2 ribonuclease from the extreme thermoacidophilic archaebacterium Sulfolobus solfataricus in mesophylic hosts. Gene. 1995;154(1):99–103. doi: 10.1016/0378-1119(94)00828-G. [DOI] [PubMed] [Google Scholar]
- 103.Wang Y., Hou Y., Nie P., Wang Y., Ren X., Wei Q., Wang Q. A novel cold-adapted and salt-tolerant RNase R from Antarctic sea-ice Bacterium Psychrobacter sp. ANT206. Molecules. 2019;24(12):1–12. doi: 10.3390/molecules24122229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Golotin V., Balabanova L., Likhatskaya G., Rasskazov V. Recombinant production and characterization of a highly active alkaline phosphatase from marine bacterium Cobetia marina. Mar. Biotechnol. (NY) 2015;17(2):130–143. doi: 10.1007/s10126-014-9601-0. [DOI] [PubMed] [Google Scholar]
- 105.Inoue A., Anraku M., Nakagawa S., Ojima T. Discovery of a novel alginate lyase from Nitratiruptor sp. SB155-2 thriving at deep-sea hydrothermal vents and identification of the residues responsible for its heat stability. J. Biol. Chem. 2016;291(30):15551–15563. doi: 10.1074/jbc.M115.713230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.De Luca V., Vullo D., Del Prete S., Carginale V., Osman S.M., Al-Othman Z., Supuran C.T., Capasso C. Cloning, characterization and anion inhibition studies of a γ-carbonic anhydrase from the Antarctic bacterium Colwellia psychrerythraea. Bioorg. Med. Chem. 2016;24(4):835–840. doi: 10.1016/j.bmc.2016.01.005. [DOI] [PubMed] [Google Scholar]
- 107.Do H., Kim S.J., Lee C.W., Kim H-W., Park H.H., Kim H.M., Park H., Park H., Lee J.H. Crystal structure of UbiX, an aromatic acid decarboxylase from the psychrophilic bacterium Colwellia psychrerythraea that undergoes FMN-induced conformational changes. Sci. Rep. 2015;5(1):8196. doi: 10.1038/srep08196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Do H., Yun J.S., Lee C.W., Choi Y.J., Kim H.Y., Kim Y.J., Park H., Chang J.H., Lee J.H. Crystal structure and comparative sequence analysis of GmhA from Colwellia psychrerythraea strain 34H provides insight into functional similarity with DiaA. Mol. Cells. 2015;38(12):1086–1095. doi: 10.14348/molcells.2015.0191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Tang Y., Wu P., Jiang S., Selvaraj J.N., Yang S., Zhang G. A new cold-active and alkaline pectate lyase from Antarctic bacterium with high catalytic efficiency. Appl. Microbiol. Biotechnol. 2019;103(13):5231–5241. doi: 10.1007/s00253-019-09803-1. [DOI] [PubMed] [Google Scholar]
- 110.Bekli S., Aktas B., Gencer D., Aslim B. Biochemical and molecular characterizations of a novel pH- and temperature-stable Pectate Lyase from Bacillus amyloliquefaciens S6 for Industrial application. Mol. Biotechnol. 2019;61(9):681–693. doi: 10.1007/s12033-019-00194-2. [DOI] [PubMed] [Google Scholar]
- 111.See Too W.C., Few L.L. Cloning of triose phosphate isomerase gene from an antarctic psychrophilic Pseudomonas sp. by degenerate and splinkerette PCR. World J. Microbiol. Biotechnol. 2010;26(7):1251–1259. doi: 10.1007/s11274-009-0295-9. [DOI] [PubMed] [Google Scholar]
- 112.Albino A., Marco S., Di Maro A., Chambery A., Masullo M., De Vendittis E. Characterization of a cold-adapted glutathione synthetase from the psychrophile Pseudoalteromonas haloplanktis. Mol. Biosyst. 2012;8(9):2405–2414. doi: 10.1039/c2mb25116g. [DOI] [PubMed] [Google Scholar]
- 113.Georlette D., Jónsson Z.O., Van Petegem F., Chessa J., Van Beeumen J., Hübscher U., Gerday C. A DNA ligase from the psychrophile Pseudoalteromonas haloplanktis gives insights into the adaptation of proteins to low temperatures. Eur. J. Biochem. 2000;267(12):3502–3512. doi: 10.1046/j.1432-1327.2000.01377.x. [DOI] [PubMed] [Google Scholar]
- 114.Jackson B.R., Noble C., Lavesa-Curto M., Bond P.L., Bowater R.P. Characterization of an ATP-dependent DNA ligase from the acidophilic archaeon “Ferroplasma acidarmanus” Fer1. Extremophiles. 2007;11(2):315–327. doi: 10.1007/s00792-006-0041-2. [DOI] [PubMed] [Google Scholar]
- 115.Oh H.N., Park D., Seong H.J., Kim D., Sul W.J. Antarctic tundra soil metagenome as useful natural resources of cold-active lignocelluolytic enzymes. J. Microbiol. 2019;57(10):865–873. doi: 10.1007/s12275-019-9217-1. [DOI] [PubMed] [Google Scholar]
- 116.Prieur D., Marteinsson V.T. Prokaryotes Living under Elevated Hydrostatic Pressure. In: Antranikian G., editor. Biotechnology of Extremophiles.Advances in Biochemical Engineering/ Biotechnology. Berlin, Heidelberg: Springer; 1998. pp. 23–35. [Google Scholar]
- 117.Kato C., Nogi Y., Arakawa S. In: Isolation, cultivation, and diversity of deep-sea piezophiles. High-Pressure Microbiology; Michiels, C.; Bartlett, D.H. Aersten A., editor. American Society of Microbiology; 2008. pp. 203–217. [Google Scholar]
- 118.Lauro F.M., Chastain R.A., Blankenship L.E., Yayanos A.A., Bartlett D.H. The unique 16S rRNA genes of piezophiles reflect both phylogeny and adaptation. Appl. Environ. Microbiol. 2007;73(3):838–845. doi: 10.1128/AEM.01726-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Fang J., Zhang L., Bazylinski D.A. Deep-sea piezosphere and piezophiles: geomicrobiology and biogeochemistry. Trends Microbiol. 2010;18(9):413–422. doi: 10.1016/j.tim.2010.06.006. [DOI] [PubMed] [Google Scholar]
- 120.Daniel I., Oger P., Winter R. Origins of life and biochemistry under high-pressure conditions. Chem. Soc. Rev. 2006;35(10):858–875. doi: 10.1039/b517766a. [DOI] [PubMed] [Google Scholar]
- 121.Yayanos A.A. Microbiology to 10,500 meters in the deep sea. Annu. Rev. Microbiol. 1995;49(1):777–805. doi: 10.1146/annurev.mi.49.100195.004021. [DOI] [PubMed] [Google Scholar]
- 122.Zobell C.E., Johnson F.H. The influence of hydrostatic pressure on the growth and viability of terrestrial and marine bacteria. J. Bacteriol. 1949;57(2):179–189. doi: 10.1128/JB.57.2.179-189.1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Horikoshi K. Barophiles: deep-sea microorganisms adapted to an extreme environment. Curr. Opin. Microbiol. 1998;1(3):291–295. doi: 10.1016/S1369-5274(98)80032-5. [DOI] [PubMed] [Google Scholar]
- 124.Jaenicke R., Závodszky P. Proteins under extreme physical conditions. FEBS Lett. 1990;268(2):344–349. doi: 10.1016/0014-5793(90)81283-T. [DOI] [PubMed] [Google Scholar]
- 125.Gomes J., Gomes I., Terler K., Gubala N., Ditzelmüller G., Steiner W. Optimisation of culture medium and conditions for α-l-Arabinofuranosidase production by the extreme thermophilic eubacterium Rhodothermus marinus. Enzyme Microb. Technol. 2000;27(6):414–422. doi: 10.1016/S0141-0229(00)00229-5. [DOI] [PubMed] [Google Scholar]
- 126.Gomes J., Steiner W. Production of a high activity of an extremely thermostable B-Mannanase by the Thermophilic Eubacterium Rhodothermus marinus, grown on locust bean gum. Biotechnol. Lett. 1998;20(8):729–733. doi: 10.1023/A:1005330618613. [DOI] [Google Scholar]
- 127.Fukuda M., Kunugi S. Pressure dependence of thermolysin catalysis. Eur. J. Biochem. 1984;142(3):565–570. doi: 10.1111/j.1432-1033.1984.tb08323.x. [DOI] [PubMed] [Google Scholar]
- 128.Konisky J., Michels P.C., Clark D.S. Pressure stabilization is not a general property of thermophilic enzymes: the adenylate kinases of Methanococcus voltae, Methanococcus maripaludis, Methanococcus thermolithotrophicus, and Methanococcus jannaschii. Appl. Environ. Microbiol. 1995;61(7):2762–2764. doi: 10.1128/AEM.61.7.2762-2764.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Giuliano M., Schiraldi C., Marotta M.R., Hugenholtz J., De Rosa M. Expression of Sulfolobus solfataricus α-glucosidase in Lactococcus lactis. Appl. Microbiol. Biotechnol. 2004;64(6):829–832. doi: 10.1007/s00253-003-1493-2. [DOI] [PubMed] [Google Scholar]
- 130.Mombelli E., Shehi E., Fusi P., Tortora P. Exploring hyperthermophilic proteins under pressure: theoretical aspects and experimental findings. Biochim. Biophys. Acta. 2002;1595(1-2):392–396. doi: 10.1016/S0167-4838(01)00361-2. [DOI] [PubMed] [Google Scholar]
- 131.Schneider S.C., Fosholt M.T., Hessen D.O., Kaste Ø. Juncus bulbosus nuisance growth in oligotrophic freshwater ecosystems: different triggers for the same phenomenon in rivers and lakes? Aquat. Bot. 2013;104:15–24. doi: 10.1016/j.aquabot.2012.10.001. [DOI] [Google Scholar]
- 132.Johnson D.B. In: Physiology and ecology of acidophilic microorganisms. Physiology and biochemistry of extremophiles; Gerday, C. Glansdorff N., editor. Washington, D.C.: ASM Press; 2007. pp. 257–271. [Google Scholar]
- 133.Poole R.K. Novartis Foundation Symposium 221 ‐ Bacterial Responses to pH. John Wiley and Sons; 2007. Introduction. pp. 1–3. [PubMed] [Google Scholar]
- 134.Rawlings D.E. Heavy metal mining using microbes. Annu. Rev. Microbiol. 2002;56(1):65–91. doi: 10.1146/annurev.micro.56.012302.161052. [DOI] [PubMed] [Google Scholar]
- 135.Rohwerder T., Gehrke T., Kinzler K., Sand W. Bioleaching review part A: progress in bioleaching: fundamentals and mechanisms of bacterial metal sulfide oxidation. Appl. Microbiol. Biotechnol. 2003;63(3):239–248. doi: 10.1007/s00253-003-1448-7. [DOI] [PubMed] [Google Scholar]
- 136.Konishi Y., Tokushige M., Asai S. 1999. [DOI] [PubMed] [Google Scholar]
- 137.Mullakhanbhai M.F., Larsen H. Halobacterium volcanii spec. nov., a Dead Sea halobacterium with a moderate salt requirement. Arch. Microbiol. 1975;104(3):207–214. doi: 10.1007/BF00447326. [DOI] [PubMed] [Google Scholar]
- 138.Pikuta E.V., Hoover R.B., Tang J. Microbial extremophiles at the limits of life. Crit. Rev. Microbiol. 2007;33(3):183–209. doi: 10.1080/10408410701451948. [DOI] [PubMed] [Google Scholar]
- 139.Leigh J.A., Albers S-V., Atomi H., Allers T. Model organisms for genetics in the domain Archaea: methanogens, halophiles, Thermococcales and Sulfolobales. FEMS Microbiol. Rev. 2011;35(4):577–608. doi: 10.1111/j.1574-6976.2011.00265.x. [DOI] [PubMed] [Google Scholar]
- 140.Christian J.H.B., Waltho J.A. Solute concentrations within cells of halophilic and non-halophilic bacteria. Biochim. Biophys. Acta. 1962;65(3):506–508. doi: 10.1016/0006-3002(62)90453-5. [DOI] [PubMed] [Google Scholar]
- 141.Oren A. Microbial life at high salt concentrations: phylogenetic and metabolic diversity. Saline Syst. 2008;4(2):2. doi: 10.1186/1746-1448-4-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Falb M., Müller K., Königsmaier L., Oberwinkler T., Horn P., von Gronau S., Gonzalez O., Pfeiffer F., Bornberg-Bauer E., Oesterhelt D. Metabolism of halophilic archaea. Extremophiles. 2008;12(2):177–196. doi: 10.1007/s00792-008-0138-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Sutrisno A., Ueda M., Abe Y., Nakazawa M., Miyatake K. A chitinase with high activity toward partially N-acetylated chitosan from a new, moderately thermophilic, chitin-degrading bacterium, Ralstonia sp. A-471. Appl. Microbiol. Biotechnol. 2004;63(4):398–406. doi: 10.1007/s00253-003-1351-2. [DOI] [PubMed] [Google Scholar]
- 144.Taylor I.N., Brown R.C., Bycroft M., King G., Littlechild J.A., Lloyd M.C., Praquin C., Toogood H.S., Taylor S.J.C. Application of thermophilic enzymes in commercial biotransformation processes. Biochem. Soc. Trans. 2004;32(Pt 2):290–292. doi: 10.1042/bst0320290. [DOI] [PubMed] [Google Scholar]
- 145.Wołosowska S., Synowiecki J. Thermostable β-glucosidase with a broad substrate specifity suitable for processing of lactose-containing products. Food Chem. 2004;85(2):181–187. doi: 10.1016/S0308-8146(03)00104-3. [DOI] [Google Scholar]
- 146.Zaccai G. The effect of water on protein dynamics. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2004;359(1448):1269–1275. doi: 10.1098/rstb.2004.1503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Lanyi J.K. Salt-dependent properties of proteins from extremely halophilic bacteria. Bacteriol. Rev. 1974;38(3):272–290. doi: 10.1128/MMBR.38.3.272-290.1974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Mevarech M., Frolow F., Gloss L.M. Halophilic enzymes: proteins with a grain of salt. Biophys. Chem. 2000;86(2-3):155–164. doi: 10.1016/S0301-4622(00)00126-5. [DOI] [PubMed] [Google Scholar]
- 149.Suzuki T., Nakayama T., Kurihara T., Nishino T., Esaki N. Cold-active lipolytic activity of psychrotrophic Acinetobacter sp. strain no. 6. J. Biosci. Bioeng. 2001;92(2):144–148. doi: 10.1016/S1389-1723(01)80215-2. [DOI] [PubMed] [Google Scholar]
- 150.Serour E., Antranikian G. Novel thermoactive glucoamylases from the thermoacidophilic Archaea Thermoplasma acidophilum, Picrophilus torridus and Picrophilus oshimae. Antonie van Leeuwenhoek. 2002;81(1-4):73–83. doi: 10.1023/A:1020525525490. [DOI] [PubMed] [Google Scholar]
- 151.Rozzell J.D. Commercial scale biocatalysis: myths and realities. Bioorg. Med. Chem. 1999;7(10):2253–2261. doi: 10.1016/S0968-0896(99)00159-5. [DOI] [PubMed] [Google Scholar]
- 152.Gurung N., Ray S., Bose S., Rai V. A broader view: microbial enzymes and their relevance in industries, medicine, and beyond. BioMed Res. Int. 2013;2013:329121. doi: 10.1155/2013/329121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Litchfield C.D. Potential for industrial products from the halophilic Archaea. J. Ind. Microbiol. Biotechnol. 2011;38(10):1635–1647. doi: 10.1007/s10295-011-1021-9. [DOI] [PubMed] [Google Scholar]
- 154.Pei J., Pang Q., Zhao L., Fan S., Shi H. Thermoanaerobacterium thermosaccharolyticum β-glucosidase: a glucose-tolerant enzyme with high specific activity for cellobiose. Biotechnol. Biofuels. 2012;5(1):31. doi: 10.1186/1754-6834-5-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Schreck S.D., Grunden A.M. Biotechnological applications of halophilic lipases and thioesterases. Appl. Microbiol. Biotechnol. 2014;98(3):1011–1021. doi: 10.1007/s00253-013-5417-5. [DOI] [PubMed] [Google Scholar]
- 156.Wyman C.E., Dale B.E., Elander R.T., Holtzapple M., Ladisch M.R., Lee Y.Y. Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 2005;96(18):1959–1966. doi: 10.1016/j.biortech.2005.01.010. [DOI] [PubMed] [Google Scholar]
- 157.Romero S., Merino E., Bolívar F., Gosset G., Martinez A. Metabolic engineering of Bacillus subtilis for ethanol production: lactate dehydrogenase plays a key role in fermentative metabolism. Appl. Environ. Microbiol. 2007;73(16):5190–5198. doi: 10.1128/AEM.00625-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Chang T., Yao S. Thermophilic, lignocellulolytic bacteria for ethanol production: current state and perspectives. Appl. Microbiol. Biotechnol. 2011;92(1):13–27. doi: 10.1007/s00253-011-3456-3. [DOI] [PubMed] [Google Scholar]
- 159.Jarboe L.R., Liu P., Kautharapu K.B., Ingram L.O. Optimization of enzyme parameters for fermentative production of biorenewable fuels and chemicals. Comput. Struct. Biotechnol. J. 2012;3(4):e201210005. doi: 10.5936/csbj.201210005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Tripetchkul S., Tonokawa M., Ishizaki A. Ethanol production by Zymomonas mobilis using natural rubber waste as a nutritional source. J. Ferment. Bioeng. 1992;74(6):384–388. doi: 10.1016/0922-338X(92)90036-T. [DOI] [Google Scholar]
- 161.Zverlov V.V., Velikodvorskaya G.A., Schwarz W.H. A newly described cellulosomal cellobiohydrolase, CelO, from Clostridium thermocellum: investigation of the exo-mode of hydrolysis, and binding capacity to crystalline cellulose. Microbiology. 2002;148(Pt 1):247–255. doi: 10.1099/00221287-148-1-247. [DOI] [PubMed] [Google Scholar]
- 162.Cripps R.E., Eley K., Leak D.J., Rudd B., Taylor M., Todd M., Boakes S., Martin S., Atkinson T. Metabolic engineering of Geobacillus thermoglucosidasius for high yield ethanol production. Metab. Eng. 2009;11(6):398–408. doi: 10.1016/j.ymben.2009.08.005. [DOI] [PubMed] [Google Scholar]
- 163.Yao S., Mikkelsen M.J. Metabolic engineering to improve ethanol production in Thermoanaerobacter mathranii. Appl. Microbiol. Biotechnol. 2010;88(1):199–208. doi: 10.1007/s00253-010-2703-3. [DOI] [PubMed] [Google Scholar]
- 164.Otohinoyi D.A., Omodele I. Prospecting microbial extremophiles as valuable resources of biomolecules for biotechnological applications. Int. J. Sci. Res. (Ahmedabad) 2015;4(1):1042–1059. [Google Scholar]
- 165.Rogers P.L., Lee K.J., Skotnicki M.L., Tribe D.E. In: Ethanol production by Zymomonas mobilis. Advances in Biochemical Engineering. Fiechter A., editor. Springer-Verlag; 1982. pp. 37–84. [Google Scholar]
- 166.Demain A.L., Newcomb M., Wu J.H.D. Cellulase, clostridia, and ethanol. Microbiol. Mol. Biol. Rev. 2005;69(1):124–154. doi: 10.1128/MMBR.69.1.124-154.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Lacoste F., Dejean F., Griffon H., Rouquette C. Quantification of free and esterified steryl glucosides in vegetable oils and biodiesel. Eur. J. Lipid Sci. Technol. 2009;111(8):822–828. doi: 10.1002/ejlt.200800297. [DOI] [Google Scholar]
- 168.Aguirre A., Eberhardt F., Hails G., Cerminati S., Castelli M.E., Rasia R.M., Paoletti L., Menzella H.G., Peiru S. The production, properties, and applications of thermostable steryl glucosidases. World J. Microbiol. Biotechnol. 2018;34(3):40. doi: 10.1007/s11274-018-2423-x. [DOI] [PubMed] [Google Scholar]
- 169.Menzella H., Peiru S., Vetcher L. Enzymatic Removal of Steryl Glycosides. WO. 2012;2013/138671:A1. [Google Scholar]
- 170.Soe J. Method. WO. 2010;2010004423:A2. [Google Scholar]
- 171.Elleuche S., Schröder C., Sahm K., Antranikian G. Extremozymes--biocatalysts with unique properties from extremophilic microorganisms. Curr. Opin. Biotechnol. 2014;29:116–123. doi: 10.1016/j.copbio.2014.04.003. [DOI] [PubMed] [Google Scholar]

