Next Article in Journal
Phytochemical Characterization, Anti-Oxidant, Anti-Enzymatic and Cytotoxic Effects of Artemisia verlotiorum Lamotte Extracts: A New Source of Bioactive Agents
Next Article in Special Issue
Production, Biochemical Characterization, and Kinetic/Thermodynamic Study of Inulinase from Aspergillus terreus URM4658
Previous Article in Journal
Benzo[k,l]xanthene Lignan-Loaded Solid Lipid Nanoparticles for Topical Application: A Preliminary Study
Previous Article in Special Issue
Production and Purification of Pectinase from Bacillus subtilis 15A-B92 and Its Biotechnological Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cold-Active Enzymes and Their Potential Industrial Applications—A Review

1
Centre of Research for Development, University of Kashmir, Srinagar 190006, India
2
Department of Environmental Science, University of Kashmir, Srinagar 190006, India
3
Mountain Research Center for Field Crops, Khudwani, Sher-e-Kashmir University of Agricultural Sciences and Technology, Srinagar 190025, India
4
SIILAS Campus, Jaipur National University, Jaipur 302025, India
5
Finnish Museum of Natural History, University of Helsinki, FL10044 Helsinki, Finland
6
Department of Pharmacology, College of Pharmacy, Umm Al Qura University, Makkah 77207, Saudi Arabia
7
Department of Microbiology, PSGVP Mandal’s, S I Patil Arts, G B Patel Science & STKV Sangh Commerce College, Shahada 425409, India
8
Biology Department, Faculty of Science, King Khalid University, Abha 9004, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(18), 5885; https://doi.org/10.3390/molecules27185885
Submission received: 11 August 2022 / Revised: 7 September 2022 / Accepted: 8 September 2022 / Published: 10 September 2022

Abstract

:
More than 70% of our planet is covered by extremely cold environments, nourishing a broad diversity of microbial life. Temperature is the most significant parameter that plays a key role in the distribution of microorganisms on our planet. Psychrophilic microorganisms are the most prominent inhabitants of the cold ecosystems, and they possess potential cold-active enzymes with diverse uses in the research and commercial sectors. Psychrophiles are modified to nurture, replicate, and retain their active metabolic activities in low temperatures. Their enzymes possess characteristics of maximal activity at low to adequate temperatures; this feature makes them more appealing and attractive in biotechnology. The high enzymatic activity of psychrozymes at low temperatures implies an important feature for energy saving. These enzymes have proven more advantageous than their mesophilic and thermophilic counterparts. Therefore, it is very important to explore the efficiency and utility of different psychrozymes in food processing, pharmaceuticals, brewing, bioremediation, and molecular biology. In this review, we focused on the properties of cold-active enzymes and their diverse uses in different industries and research areas. This review will provide insight into the areas and characteristics to be improved in cold-active enzymes so that potential and desired enzymes can be made available for commercial purposes.

1. Introduction

Microorganisms are ever-present in Mother Nature and can be isolated from different environments, with features such as extreme temperatures, high salinity, water deficiency, and varying pH. Inhabiting environments with such harsh conditions, microorganisms have developed adaptive mechanisms to function under extreme circumstances. Microorganisms living and proliferating in these hostile environments are called extremophiles, and those having the ability to withstand and proliferate in multiple harsh conditions are called polyextremophiles. Microbes populating cold environs, such as bacteria, archaea, protists, unicellular algae, and fungi, are subjected to several stresses, and they have developed numerous physiological and molecular strategies to counter these circumstances [1,2,3]. These extremophiles can grow optimally in the strictest and most hostile conditions of the Earth, with the major extreme factors under which extremophiles survive and grow as follows: temperature (−2 to 20 °C, psychrophiles; 60–115 °C, thermophiles), pH (<4, acidophiles; >9, alkaliphiles), and salinity (2–5 M NaCl, halophiles) [4]. The competencies of extremophiles to thrive in these challenging environmental conditions have fascinated researchers widely. Psychrophiles were reported back in 1884, but due to the reduced focus of researchers on extreme-cold environments, and especially on pure psychrophiles, little data on psychrophiles are available from earlier times.
Microorganisms flourishing in tough cold environments have been reported from time to time; psychrophiles are habitants of environments such as the Siberian permafrost, which is considered a distinctive environment, having permanently icy grounds, limited availability of organic matter, low water activity, and additional factors [5,6,7,8,9]. Among psychrophilic microorganisms, bacteria and fungi have been isolated and investigated comprehensively as compared to yeasts, which constitute a minor part, but all forms have the potential of producing essential psychrozymes [10,11]. Cold-active enzymes retain high potential and wide implementations in food biotechnology [12]. As per recent assessments, it is forecasted that the enzyme marketplace is developing immensely and that the worldwide enzyme market may reach USD 13 to 14 billion in 2025. This is due to the specific activity of these enzymes and the high demand from different industries. Enzymes are utilized in different biotechnological and industrial areas, such as molecular biology, detergents, food and beverage processing, textile industries, and the medical and bioremediation sectors [13,14]. This review aims to cover topics relevant to prospective biotechnological uses of psychrozymes in diverse areas and to provide a better understanding of psychrophiles and their habitats.
Psychrophiles produce enzymes with about tenfold increased specific activity in this temperature range to compensate for the slow reaction rates at low temperatures [11,14]. The maximal activity for these enzymes is shifted toward the low-temperature range, showing the poor stability of these proteins and their unfolding and inactivation at mild temperatures. Such high activity at low temperatures is achieved by destabilization of the active site or maybe the whole protein, which allows the catalytic center to be more flexible at temperatures that tend to freeze molecular motions and biochemical reactions at low temperatures [14]. Psychrophilic enzymes improve the flexibility of the structural elements that are involved in the catalytic cycle, thereby resulting in an activity that is markedly heat labile.
Low temperatures are required for industrial processes that involve heat-sensitive, volatile components or in which undesirable chemical side-reactions occur at high temperatures and contamination problems must be avoided, e.g., during the manufacture of many foods, beverages, fine chemicals, and pharmaceuticals [11]. Furthermore, in addition to extremes of temperature, many industrial processes are also carried out under extremes of pH, pressure, salinity, and/or in the presence of detergents, non-aqueous solvents, among others [13].

2. Psychrophiles: Habitat and Potentiality

The diversity of microbial life inhabiting harsh or extreme environmental conditions has been extensively studied in past decades. Psychrophiles or cold-loving microorganisms belong to three domains of existence, which include archaea, bacteria, and eukaryotes; they use the extensive unpredictability of linked chemical reactions or metabolic pathways, including photosynthesis, chemoautotrophy, and heterotrophy. Different cold habitats, including both artificial and natural ones, cover about 85% of our planet, and in these habitats, a diverse range of psychrophilic microorganisms proliferates—predominantly archaea, bacteria, fungi, yeasts, and viruses—which are further categorized as psychrophiles, i.e., cold-loving microorganisms, and psychrotrophic, i.e., cold-tolerant microorganisms [9]. Psychrophilic microorganisms, especially bacteria, are considered the main source of cold-active enzymes (Figure 1a,b). Microorganisms have been differentiated based on the cold habitats in which they thrive. These environments are categorized into two groups, comprising both psychrophiles having optimum temperature for progression at about 15 °C or lower and those with maximal temperatures for progression at about 20 °C, with the minimum temperature for growth at 0 °C or lower. By contrast, the psychrotolerant ones are those microorganisms having ideal temperature for growth at about 15 °C or above, with their maximum temperature for growth being >20 °C [15,16,17]. Psychrophiles constitute about three-fourths of the biosphere and are frequently found in mountains, glacial provinces, deep-sea waters (both fresh and marine), and the Antarctic and high-altitude soils.
The majority of the biosphere of the Earth is cold, including the Arctic and Antarctica, permafrost, non-polar regions, and deep oceans [18,19]. Life has been found at temperatures as low as −32 °C and that, too, with metabolically vigorous bacteria [20]. Multiple reports are available on the isolation of psychrophiles from extreme low-temperature ecosystems. Presently the bacterium Planococcus halocryophilus Orl, isolated from Arctic permafrost, has been grown at the lowest temperature of −15 °C, and it is the lowest growth temperature authenticated to date [21]. For instance, the cold-loving yeast Cystofilobasidium capitatum SPY11 showed growth up to 20 °C isolated from and discovered in the soil of the Northern Province of India, Kashmir valley [22]. Being inhabitants of cold environments, psychrophiles have a high potential for producing industrially important enzymes, and these enzymes are moderately capable of withstanding the decrease in chemical reaction rates prompted by lower temperatures [23]. Furhan [24] reviewed current sources of important cold-adapted protease producers, their molecular adaptation, and industrial potential. The most discovered potential use of extremophiles is through their biocatalysts, organic acids, proteins, bioactive compounds, and antimicrobial agents [25,26,27,28,29]. These psychrophilic microorganisms have high potential in nutrient recycling, mineralization of organic matter, and detoxification of different metals [30,31]. With the advent of time, more focus has been given to cold-adapted microorganisms and their industrially important enzymes [32].
Psychrophiles can produce efficient and sole biocatalysts or biomolecules such as cold-active enzymes and antimicrobial compounds, and now they are receiving increasing attention from the scientific world [17,33,34,35,36]. Cold-active enzymes (CAEs) such as pectinases, proteases, amylases, lipases, cellulases, cold-shock proteins (CSPs), cold-accumulation proteins (CAPs), and ice-binding proteins (IBPs) from psychrophilic microorganisms establish a noticeable resource for biotechnological applications [37]. All components of the cell are organized correctly so that they can work at low temperatures, and the antifreeze proteins assist in avoiding ice crystal formation [38]. Long-standing adaptations of microorganisms result in genomic amendments that lead to the production of enzymes that are active at low temperatures [39]. A wide range of enzymes (pectinases, amylases, lipases, and proteases) are vastly used at the commercial level, and they have been sourced from cold-adapted microorganisms [40]. However, hydrolases are the preferred cold enzyme class discovered so far (Figure 1c). As per reports, the global market of industrially important enzymes expanded to more than USD 5.5 billion in 2018 [41]. The international marketplace for foodstuff enzymes alone expanded to USD 1.8 billion in 2017 [42]. Therefore, it is apparent that enzymes are highly used in different industrial processes and are accruing demand over time.
Enzymes from psychrophilic microorganisms are applied as additives in cleaning products; as flavoring agents in dairy and bakery products, and utilized in bioremediation processes, molecular biology, pharmaceutical, medical sciences, and biotransformation processes [2,43,44,45,46,47]. Such significant properties of psychrozymes have fascinated investigators, drawing them to work on psychrophiles and their promising metabolites.

3. Biotechnological Importance of Cold-Active Enzymes

Over the last few decades, the scientific and industrial communities have been exerting high efforts to discover novel cold-active enzymes that possess potential properties that can be used in multiple biotechnological processes. Microbial enzymes hold diversity and uniqueness in their properties, e.g., stability, high productivity, eco-friendliness, economically realistic nature, and high flexibility, among others; such properties in microbial enzymes are becoming a reason for their gaining importance within different industries [11,48]. The vital characteristic of psychrozymes is elevated activity at lower temperatures and thermolability; these features of cold-adapted enzymes can be utilized in molecular biology, detergents, and food and drink preparations [49,50,51]. Furthermore, the use of thermo-sensible enzymes permits their selective inactivation in complex mixtures [52]. These properties make them very attractive and sought after in multiple biotechnological uses, where they can carry out processes more efficiently, with notable cost-effectiveness, and eco-friendliness compared to high-temperature-adapted enzymes [53]. The prospective use of cold-loving microbes and their enzymes has been reviewed by different researchers [36,54,55,56].
Psychrophiles (including bacteria, fungi, and yeasts) that produce cold-active enzymes have found service in numerous areas (Figure 2). Psychrozymes e.g., amylases, cellulases, invertases, proteases, and lipases, are applied in different processes carried out in the food, beverage, biofuel, and detergent industries [12,33]. Cold-active β-galactosidases have been utilized in food, cosmetic, and pharmaceutical products as well as in the production of lactose-free dairy products [11]. These enzymes are characterized based on their maximum catalytic activity at lower and adequate temperatures, though they are also heated labile and get inactivated expeditiously at mild temperatures. Cold-active enzymes have shown significant biocatalytic activity at low and moderate temperatures as compared to their mesophilic and thermophilic counterparts [30,57]. The utilization of psychrozymes in lower-temperature processes preserves heat-labile compounds. The most recent omics-era trend involves enzyme bioprospecting research using ecological metagenomics [58] and functional genomics [59]. Metagenome mining, particularly when used on microbial communities in harsh conditions, has the potential to uncover new types of enzymes while avoiding the procedural difficulties of cultivating extremophiles. However, the successful discovery of potential enzymes from environmental metagenomes continues to be difficult. As a result, functional genome mining is a viable option. The number of bacterial genomes made publicly available has exploded thanks to breakthroughs in high-throughput sequencing techniques, now numbering in the tens of thousands [60]. However, to perfectly accommodate industrial uses, cold-adapted enzymes must enhance their activity, specificity, and stability. The molecular mechanisms responsible for cold-adapted enzymes’ thermostability and activity have proven significant for the implementation in protein engineering technologies that will improve some of their features.
Residue substitutions reducing the approachability of the active site in hyperthermophilic homologs are utilized to enhance the temperature constancy and activity of cold-tolerant citrate synthases [61]. For example, single-point mutation I137M in the mesophilic Bacillus subtilis lipase LipJ resulted in a 17 °C drop in the ideal activity and cold adaptation [62]. The massive aromatic remains or residues on the substrate-binding sites of a psychrophilic alkaline phosphatase were substituted with more flexible amino acids in a triple mutant, resulting in an enzyme with enhanced stability while retaining the psychrophilic nature of the wild-type enzyme. Site-directed mutagenesis of cold-adapted endo-1,5-L-arabinanase has also enabled the activity’s optimal pH to be shifted near acidic settings, allowing it to be used in the extraction of pectin and clarification of juice [63]. The use of directed evolution has been the most efficacious technique for designing innovative cold-tolerant enzymes and optimizing the properties of enzymes derived from species that live in cold settings. The results of this technique usually show that cold adaptation of enzymes can be accomplished in a variety of ways. Mutagenesis and low-temperature activity assays were persuaded chemically by the mesophilic alkaline serine protease subtilisin, which results in the generation of two different triple mutants with substitution in different protein regions, each of which improved catalytic activity due to mutation.

4. Scope of Cold-Active Enzymes in Industries

4.1. Food and Brewing Industry

The application of mesophilic enzymes in the food and brewing industries has been in practice for a long time; however, the substitutions of these enzymes by their psychrophilic counterparts are providing promising results. The application of psychrozymes in food industries is proving to be a valuable asset as they are highly energy-efficient and thus help in the conservation of energy. Lactases are used in the dairy industry; as per recent investigations, a comparative study of numerous marketable β-galactosidases showed that these enzymes are effectively operational in milk at lower temperatures that lead to lactose breakdown [45,64]. The utilization of psychrophilic β-galactosidases might streamline and ease the process and budget of commercially developed lactose-free products. An Antarctic marine bacterium, Pseudoalteromonas haloplanktis, has been reported to produce a cold-active β-galactosidase with high efficacy in the hydrolysis of lactose under refrigerated conditions; this property will address the issue of lactose intolerance [65]. Hamid et al. [66] reported two cold-loving yeast isolates producing psychrophilic lactases capable of hydrolyzing lactose at low temperatures.
The use of cold-active β-galactosidase (temperature optima 15–18 °C) produced by psychrophilic microorganisms has opened innovative areas of research in the dairy and food processing sector with diverse biotechnological applications at low-temperature processing. Enzymes are utilized in dairy industries for cheese manufacturing and other preparations of dairy products. For instance, lactase from the yeast Kluyveromyces lactis is used for the hydrolysis of milk sugar (lactose) for the production of lactose-free dairy products [67]. Due to multiple properties, β-galactosidases retain abundant potential in industrial and biotechnological sectors [45,46]. Low-temperature processing in food and brewage industries is highly favorable as it provides many fruitful advantages, e.g., the prevention of bacterial contamination and spoilage, preventing the occurrence of unwanted chemical responses that can come into the picture at advanced temperatures, superior food quality, and persistence of flavor as heat-labile flavor compounds are not degraded [68]. In addition to these benefits, a major advantage of using cold-adaptive enzymes in the food industries is the ease with which they can be separated from a reaction process by simply supplying elevated temperatures [69]. Psychrozymes thus find different uses in the food and brewing industries, as do proteases in meat tenderization, pectinases in the fruit juice industry, amylases and xylanases in the baking industries, and lactases in the dairy industry [11,30].
The use of cold-active enzymes holds great industrial potential in terms of energy savings. The kinetic parameters of cold-active enzymes have shown that their kcat at low temperatures is similar to those observed for mesophilic enzymes. They function by lowering the reaction temperature without sacrificing their activity. These enzymes also prevent undesirable chemical reactions from occurring at elevated temperatures. The Km value of cold-active hydrolases from different bacterial sources has been reported between 0.3–1.01 mg/mL kcat with various substrates. While its Vmax has been reported to be between 236–97,951 mmol/min/mg.
Many psychrophilic microorganisms produce lipases; they are abundantly used in the food industry as well as in other industrial sectors. A recombinant cold-active lipase was isolated by transferring the lipase gene from Aeromicrobium sp. into E. coli; the gene product presented high stability and catalytic activity at low temperatures [70]. Cold-active proteases extracted from Chryseobacterium sp. can be used in meat tenderization at lower temperatures [71]. Cold-active polygalacturonase isolated from Pseudoalteromonas haloplanktis can be used to degrade pectin in juice manufacturing industries [72]. A report suggests that cold-active pectinases obtained from fungi and yeasts may help promote the aromatic aspects of wines by increasing the production of volatile compounds [73]. The cold-active pectinase reduces the viscosity in juice products, making preservation for a longer period possible without any mesophilic contamination [55,74]. Psychrozyme α-amylase has been reported from Microbacterium foliorum GA2, which was isolated from the Gangotri glacier [75], and similarly, cold-active α-amylase from marine bacterium Zunongwangia profunda [76] has been reported; both of them demonstrate vital features that can be used in the food and beverage industry. Other than this, ice-binding proteins from psychrophiles decrease tissue damage caused by ice crystals so that they can be significantly used in frozen foodstuffs and for cryopreservation in the biomedical sector [46]. Many other cold-active enzymes such as esterases, chitinases, and β–galactosidase have important applications in the food and brewing industries (Table 1).
Alkaline phosphatase isolated from Vibrio has an optimal temperature of 20–40 °C, and at 37–50 °C, it loses about 80% of activity at 60 °C [115], while alkaline phosphatase from the hyperthermophilic bacterium T. maritima has a temperature optima at 65 °C [116]. However, both are not cold-active, choose a hyperthermophilic bacterial source, and state the kinetics and temperature optima in comparison with known forms.

4.2. Detergent and Cleaning Industry

A sustainable approach is needed to replace harmful chemicals with enzymes that can be used in detergent industries [117]. Proteases retrieved from microorganisms are used in different industries as they account for more than 60% of overall enzyme sales in the worldwide enzyme market. Cold-active enzymes are used as detergent flavorings and possess much scope in the commercial sector as they are eco-friendly and sustainable [118]. Proteases are found to be significant enzymes in detergents and washing powders as they possess a major share in the enzyme market [117,119]. Many cold-active proteases have shown unparalleled stability and activity in a wide-ranging alkaline pH, in addition to their compatibility with detergents [102,103,120]. Proteases isolated from Acinetobacter sp., Bacillus sp., Planococcus sp., Pseudomonas aeruginosa, and Serratia marcescens can be used as detergent additives for cold washing [103,120,121,122,123]. Mesophilic enzymes have found utility in several industries ranging from the detergent industry and food industry to the brewing industry, and many others, owing to their superior cleaning efficiency and ecological benefits in comparison to chemical agents. From long back, proteases have been comprehensively utilized in the laundry, cleansing agent, and food processing sectors [124]. Cold-adaptive enzymes are the next level of technological programs as they not only provide these benefits but are also energy efficient. The use of cold-adaptive enzymes in cleaning processes ensures reduced wash temperatures and thus leads to the conservation of energy to a good extent. A report confirms that a mere 10 °C reduction in wash temperature produced a 30% reduction in the consumption of electricity [125].
Different cold-active enzymes, such as proteases, amylases, and lipases, have proven to be highly efficient in cleaning processes at low temperatures. Solid objects that cannot be heated for washing purposes can be easily cleaned by wipes or different formulations containing psychrozymes. The utilization of lipases in detergents or cleansing agents will increase the efficiency of the detergents in removing stains [88]. In the case of food industries, cold-active enzymes are ideal for cleaning and washing purposes as they can cut the energy stipend and downtime to a great extent [126]. The psychrozymes thus also have boundless potential in cleaning and washing processes; however, genetic alterations must be made to increase their shelf life. For instance, subtilisins obtained from Antarctic Bacillus sp. were engineered to increase their shelf life and were used in cold-active detergents [127]. Implementation of cold-active enzymes in detergents has improved the quality of products.

4.3. Pharmaceutical, Medicine, and Cosmetics

Various components used in medicine, cosmetics, fragrances, and pharmaceutical products are highly thermo-labile; the psychrophilic enzymes have structural flexibility and can operate at low temperatures, thus thermo-labile ingredients can be better synthesized by using cold-active enzymes [128]. These enzymes are used in the semi-synthesis or enzymatic modification of fine chemicals and drugs [129]. Cold-adapted enzymes have strong efficacy in pharmaceuticals, e.g., cold-adapted dehalogenases, have abundant implications for the synthesis of optically pure drug intermediates such as halo-alkanoic acids [130]. In addition, cold-adapted lipases obtained from Candida antarctica retain different applications, including increasing the quality of beauty products, synthesis of optically active drug intermediates, modifying sugars and their related compounds, and resolution of racemates, such as amines and alcohols, during the formation of various cosmetic products (i.e., iso-propylmyristate), fragrance esters, and pharmaceuticals (i.e., NK1/NK2 antagonist used for asthma treatment) [131,132]. The psychrophilic enzyme β-galactosidase, is used for the synthesis of galactooligosaccharides, and tagatose is used as prebiotic and antihyperglycemic agents, respectively [133].
Moreover, cold-tolerant protease is used for the synthesis of bioactive peptides used as an antioxidant and antihypertensive agent. Due to the antimicrobial activity of cold-adapted protease, it is used against viral and bacterial infections as a therapeutic agent and is also used in cosmetics for the removal of dried and dead skin cells [134,135]. The other example of cold-adapted enzymes used in biomedicine includes nitro-reductase and α-galactosidase. The former is used as cancer pro-drug activating enzyme, whereas the latter can be used in blood transfusion therapy [136,137]. It has been shown that psychrozymes isolated from psychrophiles have great medicinal importance.

4.4. Molecular Biology

Cold-loving microorganisms and their enzymes hold colossal functions in the area of molecular biology. Researchers working on cold-active enzymes are modifying and improving features of isolated wild strains of cold-loving microorganisms to get a desired cold-active enzyme with multiple characteristics [2]. This progression in enzyme technology is only possible with the intervention of different tools of genomics. Another prominent cold-adapted enzyme used at the molecular level is alkaline phosphatase; it is employed for the removal of phosphate groups from the 5′ ends of DNA molecules and, thus, prevents self-ligation of DNA molecules. For commercial purposes, the enzyme is obtained from E. coli and calf duodenal tissue. The alkaline phosphate obtained from E. coli and calf intestinal tissue is heat stable, and hence extraction is performed via inorganic methods [138]. Cold-adapted alkaline phosphatases, both single- and double-stranded nucleases, and uracil-DNA N-glycosylases are recently being commercialized as a molecular biological tool by various companies (New England Biolabs Inc., ArcticZymes, Takara-Clontech, Affymetrix, Inc.) [139,140,141]. The Antarctic marine bacterium, Vibrio sp., was isolated and used for the production of heat-labile alkaline phosphatase. This enzyme has a higher turnover number (k(cat)) and higher apparent Michaelis–Menten factor (K(m)) as compared with enzyme from E. coli, a clear indication of cold-adaptation [108]. An Arctic shrimp heat-labile alkaline phosphatase was used for commercial purposes [142]. Two more heat-labile psychrophilic enzymes are available for their utilization in molecular biology. They are shrimp nuclease and heat-labile uracil-DNA N-glycosylase, both capable of contaminant removal and DNA helix degradation during PCR. The former is used as a recombinant form in Pichia pastoris, and the latter is obtained from Atlantic cod (Gadus morhua) and used as a recombinant form in E. coli. Both psychrozymes can be easily inactivated by moderate thermal treatment [127].
To achieve sufficient enzyme yield with efficient enzymatic properties, it is important to develop recombinant countenance in a heterologous host. Mesophilic hosts are frequently employed for recombinant expressions at the genomic level to encode psychrozymes, although these microbes’ ideal growing temperature is incompatible with the temperature at which cold-active enzymes require correct folding to maintain their arrangement and function [143]. Several novel strategies are adopted nowadays for improving the expression to improve their solubility, protein yield, and proper folding of cold-active enzymes.
The isolation of a cold-adapted bacterium with a unique enzymatic activity served as the starting point for enzyme cloning. Designing customized primers for gene intensification using the strain’s DNA as a template is the most common cloning approach. However, this approach is feasible only when the genomic sequence has already been deposited in Gene Bank and if the microbe can be easily cultured to isolate its genetic material. Another option is to create a gene with the finest codon norm for the host. The selected expression host in most cases was BL21 (DE3) as the preferred strain. Other expression hosts were Halobacterium sp., which expresses a cold-adapted hydrolase, and Pichia pastoris, which is employed as an expression host for a variety of cold-adapted proteins. For expression, the majority of these genes were cloned on plasmids from the pET system. Fusion constructs were also used for cloning many cold-adapted genes; pGEX-6P-1 allows proteins to be expressed in fusion with GST and pGEX-6P-2 in pMAL-c, which expresses protein fusion to MBP.
The chaperones Cpn60 and Cpn10 from the cold-loving bacterium Oleispira antarctica RB8 were expressed using E. coli as reported by Ferrer and colleagues in 2004 [144]. They successfully expressed a heat-labile esterase using this chaperone E. coli system, establishing a viable expression strategy for heat-sensitive proteins. They also discovered that, when compared to the enzyme extracted from a standard E. coli strain cultivated at 37 °C, lower temperature improves enzyme folding and increases specific activity 180-fold.
Agilent Technologies now sells Arctic Express, which is a competent E. coli strain that co-expresses the cold-active chaperones Cpn60 and Cpn10. In an E. coli strain, Kim et al. [145] found that PsyGroELS, a chaperonin from the psychrophilic bacterium Psychrobacter sp. PAMC21119, co-expressed a cold-active esterase. In comparison of enzyme activity with that of previously described chaperones Cpn60 and Cpn10, PsyGroELS produced better results. Using GroES/GroEL chaperones for protein production, Esteban-Torres et al. [146] copied the cold-active esterase lp 2631 into the pURI3-TEV expression vector, but the recombinant protein was expressed as inclusion bodies in E. coli BL21 (DE3). They employed the plasmid pGro7, which produces GroES/GroEL chaperones, to fix the problem. Different tools of molecular biology and recombinant DNA technology can be used to study the mechanism of cold-adaptation among psychrophiles, and this may lead to the formulation of novel cold-active enzymes with fascinating properties.

4.5. Bioremediation or Ecological Applications

Microbes have been astounding environmental cleaners from the dawn of civilization and have provided a major hand in remediating the environment. Using efficient psychrozymes in bioremediation own several advantages, as in cold climatic environments it is difficult for the whole cell to face additional challenges, so utilization of cold-active enzymes may help in meeting clean-up standards in a short time [99]. A report published in 2014 demonstrated the presence of hydroxylase genes in psychrophilic microorganisms; these genes help them to degrade crude oil and alkane into ecologically neutral compounds [147]. These psychrophiles will provide a massive advantage as they can be used easily to check the petroleum pollution in colder regions. Because of low temperatures in cold environments, it is challenging for microorganisms to meet clean-up standards for bioremediation as it consumes more time [99]. As whole bacterial cells need multiple parameters for optimal growth and metabolic activities that are lacking in groundwater, the use of enzymes could be more feasible and result-oriented than the use of whole bacterial cells [148]. In another report, psychrophilic yeasts were found to degrade phenol and other related compounds at 10 °C [149]. Many cold-adapted microbes such as Pseudomonas sp., Rhodococcus sp., and Pseudoalteromonas are capable of degrading petroleum hydrocarbons [150,151,152]. Psychrophiles, along with their bioremediation capacity, will help us in retaining soil health and crop productivity.

5. Conclusions and Future Perspectives

Microorganisms are universal, and they thrive in different environmental conditions. Microorganisms, including bacteria, fungi, and yeast, inhabiting the cold environments of the Earth are attracting the attention of microbiologists and researchers from all over the world because of their uniqueness in their metabolic activities. They have gained attention from different industries due to their cold-active enzymes. The scope of biotechnological applications of novel extremozymes from psychrophiles is expanding with time. This is because of their ability to catalyze reactions at low temperatures close to the freezing point of water and act as an energy-saving tools. The utilization of cold-active enzymes in low-temperature catalysis will help in retaining the flavor of the product and protect the process from contamination. Psychrophilic enzymes possess a vital role in industries for large-scale production, but despite great efforts, the commercialization of efficient and novel cold-active enzymes is still in its infancy. It is important to explore these potent enzymes that will be eco-friendly, economical, and efficient.

Author Contributions

B.H.: Conceptualization, writing-original draft, supervision, validation, editing, data curation. Z.B., N.M. and M.B.: writing, editing, data curation. A.M.Y.: writing, editing, review, data curation. F.M.: editing, data curation. P.P., W.H.A. and R.Z.S.: writing—original draft, review, and editing. A.A.S. and M.Y.A.: writing—original draft, review, and editing and fund acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deanship of Scientific Research at Umm Al Qura University, Makkah, Saudi Arabia, through Grant code (project code: 22UQU4310387DSR21) and The Deanship of Scientific Research at King Khalid University, Abha, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this manuscript.

Acknowledgments

The authors thank the Deanship of Scientific Research at Umm Al Qura University, Makkah, Saudi Arabia, for supporting this work through Grant code (project code: 22UQU4310387DSR21), The Deanship of Scientific Research at King Khalid University for funding this work through large Groups (Project under grant number R.G.P. 2/78/43), the Open access funding provided by University of Helsinki, Finland. Scholars are highly thankful to the Centre of Research for Development, University of Kashmir where the literature survey was done and manuscript wrote.

Conflicts of Interest

The authors declare no conflict of interest.

Ethical Approval

This article does not contain any human or animal studies performed by any of the authors.

References

  1. De Mayer, P.; Anderson, D.; Cary, C.; Cowan, D.A. Some like it cold: Understanding the survival strategies of psychrophiles. EMBO Rep. 2014, 15, 508–517. [Google Scholar] [CrossRef] [PubMed]
  2. Collins, T.; Margesin, R. Psychrophilic lifestyles: Mechanisms of adaptation and biotechnological tools. Appl. Microbiol. Biotechnol. 2019, 103, 2857–2871. [Google Scholar] [CrossRef] [PubMed]
  3. Vallesi, A.; Pucciarelli, S.; Buonanno, F.; Fontana, A.; Mangiagalli, M. Bioactive molecules from protists: Perspectives in biotechnology. Eur. J. Protistol. 2020, 75, 125720. [Google Scholar] [CrossRef] [PubMed]
  4. Yadav, A.N.; Verma, P.; Kumar, M.; Pal, K.K.; Dey, R.; Gupta, A.; Padaria, J.C.; Gujar, G.T.; Kumar, S.; Suman, A.; et al. Diversity and phylogenetic profiling of niche-specific bacilli from extreme environments of India. Ann. Microbiol. 2015, 65, 611–629. [Google Scholar] [CrossRef]
  5. Morita, R.Y. Marine psychrophilic bacteria ocean over. Mar. Biol. Ann. Rev. 1996, 4, 105–121. [Google Scholar]
  6. Farrell, J.; Rose, A.H. Temperature effects on microorganisms. In Thermobiology; Rose, A.H., Ed.; Academic Press: London, UK, 1967; pp. 147–218. [Google Scholar]
  7. Bakermans, C. Determining the limits of microbial life at sub-zero temperatures. In Psychrophiles: From Biodiversity to Biotechnology; Springer: Berlin/Heidelberg, Germany, 2017; pp. 21–38. [Google Scholar]
  8. Rivkina, E.; Abramov, A.; Spirina, E.; Petrovskaya, L.; Shatilovich, A.; Shmakova, L.; Scherbakova, V.; Vishnivetskaya, T. Earth’s perennially frozen environments as a model of cryogenic planet ecosystems. Perm. Periglac. Proc. 2018, 29, 246–256. [Google Scholar] [CrossRef]
  9. Margesin, R.; Collins, T. Microbial ecology of the cryosphere (glacial and permafrost habitats): Current knowledge. Appl. Microbiol. Biotechnol. 2019, 103, 2537–2549. [Google Scholar] [CrossRef]
  10. Margesin, R.; Miteva, V. Diversity and ecology of psychrophilic microorganisms. Res. Microbiol. 2011, 162, 346–361. [Google Scholar] [CrossRef]
  11. Mangiagalli, M.; Lotti, M. Cold-Active β-Galactosidases: Insight into Cold Adaptation Mechanisms and Biotechnological Exploitation. Mar. Drugs 2021, 19, 43. [Google Scholar] [CrossRef]
  12. Kuddus, M. Cold-active enzymes in food biotechnology: An updated mini review. J. Appl. Biol. Biotechnol. 2018, 6, 58–63. [Google Scholar]
  13. Wackett, L.P. Microbial industrial enzymes: An annotated selection of World Wide Web sites relevant to the topics in microbial biotechnology. Microb. Biotechnol. 2019, 12, 1090–1091. [Google Scholar] [CrossRef] [PubMed]
  14. Santiago, M.; Ramírez-Sarmiento, C.A.; Zamora, R.A.; Parra, L.P. Discovery, molecular mechanisms, and industrial applications of cold-active enzymes. Front. Microbiol. 2016, 7, 1408. [Google Scholar] [CrossRef]
  15. Stokes, J.L. General biology and nomenclature of psychrophilic microorganisms. In Recent Progress in Microbiology VIII; University of Toronto Press: Toronto, ON, Canada, 1963; pp. 187–192. [Google Scholar]
  16. Robinson, C.H. Cold adaptation in the Arctic and Antarctic fungi. New Phytol. 2001, 151, 341–353. [Google Scholar] [CrossRef]
  17. Troncoso, E.; Barahona, S.; Carrasco, M.; Villarreal, P.; Alcaíno, J.; Cifuentes, V.; Baeza, M. Identification and characterization of yeasts isolated from the South Shetland Islands and the Antarctic Peninsula. Polar Biol. 2017, 40, 649–658. [Google Scholar] [CrossRef]
  18. Cavicchioli, R.; Tortsen, T. Extremophilic. In Encyclopaedia of Microbiology, 2nd ed.; Lederberg, J., Ed.; Academic Press: London, UK, 2000; pp. 317–337. [Google Scholar]
  19. Buzzini, P.; Margesin, R. Cold-Adapted Yeasts: Biodiversity, Adaptation Strategies, and Biotechnological Significance; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar]
  20. Bakermans, C.; Skidmore, M.L. Microbial metabolism in ice and brine at −5 degrees C. Environ. Microbiol. 2011, 13, 2269–2278. [Google Scholar] [CrossRef]
  21. Mykytczuk, N.C.; Foote, S.J.; Omelon, C.R.; Southam, G.; Greer, C.W.; Whyte, L.G. Bacterial growth at -15 degrees C; molecular insights from the permafrost bacterium planococcushalocryophilus Or1. ISME J. 2013, 7, 1211–1226. [Google Scholar] [CrossRef]
  22. Hamid, B.; Singh, P.; Rana, R.S.; Sahay, S. Isolation and identification of psychrophilic yeast from the soil of the northern region of India. Inventi Rapid Pharm. Biotechnol. Microbiol. 2012, 1, 1–5. [Google Scholar]
  23. 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. R. Soc. 2002, 357, 917–925. [Google Scholar] [CrossRef]
  24. Furhan, J. Adaptation, production, and biotechnological potential of cold-adapted proteases from psychrophiles and psychrotrophic: A recent overview. J. Genet. Eng. Biotechnol. 2020, 18, 36. [Google Scholar] [CrossRef]
  25. Suman, A.; Verma, P.; Yadav, A.N.; Saxena, A.K. Bioprospecting for extracellular hydrolytic enzymes from culturable thermotolerant bacteria isolated from Manikaran thermal springs. Res. J. Biotechnol. 2015, 10, 33–42. [Google Scholar]
  26. Verma, P.; Yadav, A.N.; Shukla, L.; Saxena, A.K.; Suman, A. Hydrolytic enzymes produced by thermotolerant Bacillus altitudinisIARI-MB-9 and Gulbenkianiamobilis IARI-MB-18 isolated from Manikaran hot springs. Int. J. Adv. Res. 2015, 3, 1241–1250. [Google Scholar]
  27. Kumar, V.; Yadav, A.N.; Saxena, A. Unraveling rhizospheric diversity and potential of phytase-producing microbes. SM J. Biol. 2016, 2, 1009. [Google Scholar]
  28. Kumar, V.; Yadav, A.N.; Verema, P. β-Propeller phytases: Diversity, catalytic attributes, current developments, and potential biotechnological applications. Int. J. Biol. Macromol. 2017, 98, 595–609. [Google Scholar] [CrossRef] [PubMed]
  29. Kaur, R.; Saxena, A.; Sangwan, P. Production and characterization of a neutral phytase of Penicillium oxalicum EUFR-3 isolated from the Himalayan region. Nusant. Biosci. 2017, 9, 68–76. [Google Scholar] [CrossRef]
  30. Gerday, C.M.; Aittaleb, M.; Bentahier, J.P.; Chessa, P.; Claverie, T. Collins Cold-adapted enzymes: From fundamentals to biotechnology. Trends Biotechnol. 2000, 18, 103–107. [Google Scholar] [CrossRef]
  31. Connell, L.; Redman, R.; Craig, S.; Scorzetti, G.; Iszard, M.; Rodriguez, R. Diversity of soil yeasts isolated from South Victoria land, Antarctica. Microb. Ecol. 2008, 56, 448–459. [Google Scholar] [CrossRef]
  32. Joshi, S.; Satyanarayana, T. Biotechnology of cold-active proteases. Biology 2013, 2, 755–783. [Google Scholar] [CrossRef]
  33. Margesin, R.; Feller, G. Biotechnological applications of psychrophiles. Environ. Technol. 2010, 31, 835–844. [Google Scholar] [CrossRef]
  34. Duarte, A.W.F.; Dayo-Owoyemi, I.; Nobre, F.S.; Pagnocca, F.C.; Chaud, L.C.S.; Pessoa, A.; Felipe, M.G.A.; Sette, L.D. Taxonomic assessment and enzyme production by yeasts isolated from marine and terrestrial Antarctic samples. Extremophiles 2013, 17, 1023–1035. [Google Scholar] [CrossRef]
  35. Sahay, S.; Hamid, B.; Singh, P.; Ranjan, K.; Chauhan, D.; Rana, R.S.; Chaurse, V.K. Evaluation of pectinolytic activities for oenological uses from psychrotrophic yeasts. Lett. Appl. Microbiol. 2013, 57, 115–121. [Google Scholar] [CrossRef]
  36. Hamid, B.; Rana, R.S.; Chauhan, D.; Singh, P.; Mohiddin, F.A.; Sahay, S.; Abidi, I. Psychrophilic yeasts and their biotechnological applications—A review. Afr. J. Biotechol. 2014, 13, 2188–2197. [Google Scholar]
  37. Gounot, A.M. Bacterial Life at low temperature: Physiological aspects and biotechnological implications. J. Appl. Bacteriol. 1991, 71, 386–397. [Google Scholar] [CrossRef] [PubMed]
  38. Munoz, P.A.; Marquez, S.L.; Gonzalez-Nilo, F.D.; Marquez-Miranda, V.; Blamey, J.M. Structure and application of antifreeze proteins from Antarctic bacteria. Microb. Cell Fact. 2017, 16, 138. [Google Scholar] [CrossRef] [PubMed]
  39. Tribelli, P.M.; Lopez, N.I. Reporting key features in cold-adapted bacteria. Life 2018, 8, 8. [Google Scholar] [CrossRef]
  40. Kasana, R.C. Proteases from psychrotrophs: An overview. Crit. Rev. Microbiol. 2010, 36, 134–145. [Google Scholar] [CrossRef]
  41. Staff, B.R. Global Markets for Enzymes in Industrial Applications; BCC Research LLC: Wellesley, MA, USA, 2018. [Google Scholar]
  42. Chen, J. Food Enzymes: Global Markets; BCC Research: Wellesley, MA, USA, 2018. [Google Scholar]
  43. Feller, G.; Narinx, E.; Arpigny, J.L.; Aittaleb, M.; Baise, E.; Genicot, S.; Gerday, C. Enzymes from psychrophilic organisms. FEMS Microbiol. Rev. 1996, 18, 189–202. [Google Scholar] [CrossRef]
  44. Sarmiento, F.; Peralta, R.; Blamey, J.M. Cold, and hot extremozymes: Industrial relevance and current trends. Front. Bioeng. Biotechnol. 2015, 3, 148. [Google Scholar] [CrossRef] [Green Version]
  45. Bruno, S.; Coppola, D.; di Prisco, G.; Giordano, D.; Verde, C. Enzymes from Marine Polar Regions and Their Biotechnological Applications. Mar. Drugs 2019, 17, 544. [Google Scholar] [CrossRef]
  46. Mangiagalli, M.; Brocca, S.; Orlando, M.; Lotti, M. The “cold revolution”. Present and future applications of cold-active enzymes and ice-binding proteins. New Biotechnol. 2020, 55, 5–11. [Google Scholar] [CrossRef]
  47. Gupta, S.K.; Kataki, S.; Chatterjee, S.; Prasad, R.K.; Datta, S.; Vairale, M.G.; Sharma, S.; Dwivedi, S.K.; Gupta, D.K. Cold adaptation in bacteria with a special focus on cellulase production and its potential application. J. Clean. Prod. 2020, 258, 120351. [Google Scholar] [CrossRef]
  48. 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, 3, 329121. [Google Scholar] [CrossRef] [PubMed]
  49. Marx, J.C.; Collins, T.; D’Amico, S.; Feller, G.; Gerday, C. Cold-adapted enzymes from marine Antarctic microorganisms. Mar. Biotechnol. 2007, 9, 293–304. [Google Scholar] [CrossRef] [PubMed]
  50. Cavicchioli, R.; Charlton, T.; Ertan, H.; Mohd Omar, S.; Siddiqui, K.S.; Williams, T.J. Biotechnological uses of enzymes from psychrophiles. Microb. Biotechnol. 2011, 4, 449–460. [Google Scholar] [CrossRef] [PubMed]
  51. Feller, G. Psychrophilic enzymes: From folding to function and biotechnology. Scientifica 2013, 2013, 512840. [Google Scholar] [CrossRef] [PubMed]
  52. Moran, A.J.; Hills, M.; Gunton, J.; Nano, F. Heat-labile proteases in molecular biology applications. FEMS Microbiol. Ecol. 2001, 197, 59–63. [Google Scholar] [CrossRef] [PubMed]
  53. Barroca, M.; Santos, G.; Gerday, C.; Collins, T. Biotechnological Aspects of Cold-Active Enzymes. In Psychrophiles: From Biodiversity to Biotechnology; Margesin, R., Ed.; Springer International Publishing AG: Cham, Switzerland, 2017; pp. 461–475. [Google Scholar]
  54. Cavicchioli, R.; Siddiqui, K.S.; Andrews, D.; Sowers, K.R. Low-temperature extremophiles and their applications. Curr. Opin. Biotechnol. 2002, 13, 253–261. [Google Scholar] [CrossRef]
  55. Margesin, R.; Feller, G.; Gerday, C.; Russell, N. Cold-Adapted Microorganisms: Adaptation Strategies and Biotechnological Potential. In Encyclopedia of Environmental Microbiology; Bitton, G., Ed.; John Wiley & Sons: New York, NY, USA, 2002; pp. 871–885. [Google Scholar]
  56. Feller, G.; Gerday, C. Psychrophilic enzymes: Hot topics in cold adaptation. Nat. Rev. Microbiol. 2003, 1, 200–208. [Google Scholar] [CrossRef]
  57. Chattopadhyay, M.K. Mechanism of bacterial adaptation to low temperature—A review. J. Biosci. 2006, 31, 157–165. [Google Scholar] [CrossRef]
  58. Berini, F.; Casciello, C.; Marcone, G.L.; Marinelli, F. Metagenomics: Novel enzymes from non-culturable microbes. FEMS Microbiol. Lett. 2017, 364, fnx211. [Google Scholar] [CrossRef]
  59. Gerlt, J.A. Genomic enzymology: Web tools for leveraging protein family sequence-function space and genome context to discover novel functions. Biochemistry 2017, 56, 4293–4308. [Google Scholar] [CrossRef] [PubMed]
  60. Ziemert, N.; Alanjary, M.; Weber, T. The evolution of genome mining in microbes-a review. Nat. Prod. Rep. 2016, 33, 988–1005. [Google Scholar] [CrossRef] [PubMed]
  61. Gerike, U.; Danson, M.J.; Hough, D.W. Cold-active citrate synthase: Mutagenesis of active-site residues. Protein Eng. 2001, 14, 655–661. [Google Scholar] [CrossRef] [PubMed]
  62. Goomber, S.; Kumar, R.; Singh, R.; Mishra, N.; Kaur, J. Point mutation Gln121-Arg increased temperature optima of Bacilluslipase (1.4 subfamily) by fifteen degrees. Int. J. Biol. Macromol. 2016, 88, 507–514. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Q.; Hou, Y.; Xu, Z.; Miao, J.; Li, G. Optimization of cold-active protease production by the psychrophilic bacterium Colwellia sp. NJ341 with response surface methodology. Bioresour. Technol. 2008, 99, 1926–1931. [Google Scholar] [CrossRef]
  64. Horner, T.W.; Dunn, M.L.; Eggett, D.L.; Ogden, L.V. The beta-galactosidase activity of commercial lactase samples in raw and pasteurized milk at refrigerated temperatures. J. Dairy Sci. 2011, 94, 3242–3249. [Google Scholar] [CrossRef]
  65. Van de Voorde, I.; Goiris, K.; Syryn, E.; Van den Bussche, C.; Aerts, G. Evaluation of the cold-active Pseudoalteromonas haloplanktis β-galactosidase enzyme for lactose hydrolysis in whey permeate as a primary step of d-tagatose production. Process Biochem. 2014, 49, 2134–2140. [Google Scholar] [CrossRef]
  66. Hamid, B.; Singh, P.; Mohiddin, F.A.; Sahay, S. Partial characterization of cold-active β-galactosidase activity produced by Cystofilobasidium capitatum SPY11 and Rhodotorula mucilaginosa PT1. J. Endocytobiosis Cell Res. 2013, 24, 23–26. [Google Scholar]
  67. Mateo, C.; Monti, R.; Pessela, B.C.; Fuentes, M.; Torres, R.; Manuel Guisán, J.; Fernández-Lafuente, R. Immobilization of lactase from Kluyveromyceslactis greatly reduces the inhibition promoted by glucose. Full hydrolysis of lactose in milk. Biotechnol. Prog. 2004, 20, 1259–1262. [Google Scholar] [CrossRef]
  68. Qais, A.; Maqtari, A.; Waleed, A.L.; Mahdi, A.A. Cold-active enzymes and their applications in industrial fields—A review. Int. J. Res. Agric. Sci. 2019, 6, 2348–3997. [Google Scholar]
  69. Pulicherla, K.K.; Mrinmoy Ghosh, P.; Suresh, K.; Sambasiva Rao, K.R.S. Psychrozymes—The next generation industrial enzymes. J. Mar. Sci. Res. Dev. 2011, 1, 1–7. [Google Scholar] [CrossRef]
  70. Su, H.; Mai, Z.; Yang, J.; Xiao, Y.; Tian, X.; Zhang, S. Cloning, expression, and characterization of a cold-active and organic solvent tolerant lipase from Aeromicrobium sp. SCSIO 25071. J. Microbiol. Biotechnol. 2016, 26, 1067–1076. [Google Scholar] [CrossRef] [PubMed]
  71. Mageswari, A.; Subramanian, P.; Chandrasekaran, S.; Karthikeyan, S.; Gothandam, K.M. Systematic functional analysis and application of a cold-active serine protease from a novel Chryseobacterium sp. Food Chem. 2017, 217, 18–27. [Google Scholar] [CrossRef] [PubMed]
  72. Ramya, L.N.; Pulicherla, K.K. Molecular insights into cold-active polygalacturonase enzyme for its potential application in food processing. J. Food Sci. Technol. 2015, 52, 5484–5496. [Google Scholar] [CrossRef]
  73. Singh, P.; Hamid, B.; Ahmad Lone, M.; Ranjan, K.; Khan, A.; Chaurse, V. Evaluation of pectinase activity from the psychrophilic fungal strain Truncatella angustata-BPF5 for use in the wine industry. Endocytobiosis Cell Res. 2012, 22, 57–61. [Google Scholar]
  74. Coker, J.A.; Sheridan, P.P.; Lovel, C.J.; Gutshall, K.R.; Auman, A.J.; Brenchley, J.E. Biochemical characterization of a β-galactosidase with a low-temperature optimum obtained from an Antarctic Arthrobacter isolate. J. Bacteriol. 2003, 185, 5473–5482. [Google Scholar] [CrossRef] [PubMed]
  75. Kuddus, M.; Saima, R.; Ahmad, I.Z. Cold-active extracellular α-amylase production from novel bacteria Microbacterium foliorum GA2 and Bacillus cereus GA6 isolated from Gangotri glacier, Western Himalaya. J. Genet. Eng. Biotechnol. 2012, 10, 151–159. [Google Scholar] [CrossRef]
  76. Qin, Y.; Huang, Z.; Liu, Z. A novel cold-active and salt-tolerant alpha-amylase from marine bacterium Zunongwangia profunda: Molecular cloning, heterologous expression, and biochemical characterization. Extremophile 2014, 18, 271–281. [Google Scholar] [CrossRef]
  77. Wierzbicka-Wos, A.; Cieslinski, H.; Wanarska, M.; Kozlowska-Tylingo, K.; Hildebrandt, P.; Kur, J. A novel cold-active b-D-galactosidase from the Paracoccus sp. 32d-gene cloning, purification, and characterization. Microb. Cell Fact. 2011, 10, 108–120. [Google Scholar] [CrossRef]
  78. He, Z.; Zhang, L.; Mao, Y.; Gu, J.; Pan, Q.; Zhou, S.; Gao, B.; Wei, D. Cloning of a novel thermostable glucoamylase from thermophilic fungus Rhizomucor pusillus and high-level co-expression with α-amylase in Pichia pastoris. BMC Biotechnol. 2014, 14, 114. [Google Scholar] [CrossRef]
  79. He, L.; Mao, Y.; Zhang, L.; Wang, H.; Alias, S.A.; Gao, B. Functional expression of a novel α-amylase from Antarctic psychrotolerant fungus for the baking industry and its magnetic immobilization. BMC Biotechnol. 2017, 17, 22. [Google Scholar] [CrossRef] [Green Version]
  80. Arabaci, N.; Arikan, B. Isolation and characterization of a cold-active, alkaline, detergent stable α-amylase from a novel bacterium Bacillus subtilis N8. Prep. Biochem. Biotechnol. 2018, 48, 419–426. [Google Scholar] [CrossRef] [PubMed]
  81. Mao, Y.; Yin, Y.; Zhang, L.; Alias, S.S.; Gao, B.; Wei, D. Development of a novel Aspergillus uracil deficient expression system and its application in expressing a cold-adapted α-amylase gene from Antarctic fungi Geomycespannorum. Process. Biochem. 2015, 50, 1581–1590. [Google Scholar] [CrossRef]
  82. Dhaulaniya, A.S.; Balan, B.; Kumar, M.; Agrawal, P.K.; Singh, D.K. Cold survival strategies for bacteria, recent advancement, and potential industrial applications. Arch. Microbiol. 2019, 201, 1–16. [Google Scholar] [CrossRef] [PubMed]
  83. Dahiya, N.; Tewari, R.; Hoondal, G. Biotechnological aspects of chitinolytic enzymes: A review. Appl. Microbiol. Biotechnol. 2006, 71, 773–782. [Google Scholar] [CrossRef] [PubMed]
  84. Ramli, A.N.; Mahadi, N.M.; Rabu, A.; Murad, A.M.; Bakar, F.D.; Illias, R.M. Molecular cloning, expression and biochemical characterization of a cold-adapted novel recombinant chitinase from Glaciozyma antarctica PI12. Microb. Cell Fact. 2011, 10, 94. [Google Scholar] [CrossRef] [PubMed]
  85. Carrasco, M.; Rozas, J.M.; Barahona, S.; Alcaı’no, J.; Cifuentes, V.; Baeza, M. Diversity and extracellular enzymatic activities of yeasts isolated from King George Island, the sub-Antarctic region. BMC Microbiol. 2012, 12, 251. [Google Scholar] [CrossRef]
  86. Tutino, M.L.; Di Prisco, G.; Marino, G.; Pascale, D. Cold-adapted esterases and lipases: From fundamentals to application. Protein Pept. Lett. 2009, 16, 1172–1180. [Google Scholar]
  87. Kavitha, M.; Shanthi, C. Alkaline thermostable cold-active lipase from halotolerant Pseudomonas sp. VITCLP4 as detergent additive. Indian J. Biotechnol. 2017, 6, 446–455. [Google Scholar]
  88. Sahay, S.; Chouhan, D. Study on the potential of cold-active lipases from psychrotrophic fungi for detergent formulation. J. Genet. Biotechnol. 2018, 16, 319–325. [Google Scholar] [CrossRef]
  89. Wu, S.; Liu, Y.; Yan, Q.; Jiang, Z. Gene cloning, functional expression, and characterization of a novel glycogen branching enzyme from Rhizomucor miehei and its application in wheat bread making. Food Chem. 2014, 159, 85–94. [Google Scholar] [CrossRef]
  90. Huang, H.; Luo, H.; Wang, Y.; Fu, D.; Shao, N.; Yang, P. Novel low-temperature-active phytase from Erwinia carotovora var. carotovota ACCC 10276. J. Microbiol. Biotechnol. 2009, 19, 1085–1091. [Google Scholar] [CrossRef]
  91. Ranjan, K.; Sahay, S. Identification of phytase-producing yeast and optimization and characterization of extracellular phytase from Candida parapsilosis. Int. J. Sci. Nat. 2013, 4, 583–590. [Google Scholar]
  92. Yu, P.; Wang, X.T.; Liu, J.W. Purification and characterization of a novel cold-adapted phytase from strain JMUY14 isolated from the Antarctic: Characterization of a novel cold-adapted phytase. J. Basic Microbiol. 2015, 55, 1029–1039. [Google Scholar] [CrossRef] [PubMed]
  93. Nakagawa, T.; Nagaoka, T.; Miyaji, T.; Tomizuka, N. Cold-active Polygalacturonase from psychrophilic-basidiomycetous yeast Cystofilobasidium capitatum strain PPY-1. Biosci. Biotechnol. Biochem. 2005, 69, 419–421. [Google Scholar] [CrossRef]
  94. Pan, X.; Tu, T.; Wang, L.; Luo, H.; Ma, R.; Shi, P. A novel low-temperature active pectin methylesterase from Penicillium chrysogenum F46 with high efficiency in fruit firming. Food Chem. 2014, 162, 229–234. [Google Scholar] [CrossRef] [PubMed]
  95. Belda, I.; Conchillo, L.B.; Ruiz, J.; Navascue’s, E.; Marquina, D.; Santos, A. Selection and use of pectinolytic yeasts for improving clarification and phenolic extraction in winemaking. Int. J. Food Microbiol. 2016, 223, 1–8. [Google Scholar] [CrossRef]
  96. Collins, T.; Gerday, C.; Feller, G. Xylanases, xylanase families, and extremophilic xylanases. FEMS Microbiol. Rev. 2005, 29, 3–23. [Google Scholar] [CrossRef]
  97. Borges, T.A.; de Souza, A.T.; Squina, F.M.; Riaño-Pacho’n, D.M.; Corrêa dos Santos, R.A.; Machado, E.; Velasco de Castro Oliveira, J.; Dama’sio, A.R.L.; Goldman, G.H. Biochemical characterization of an endoxylanase from Pseudozymabrasiliensis sp. nov. strain GHG001 isolated from the intestinal tract of Chrysomelidae larvae associated to sugarcane roots. Process Biochem. 2014, 49, 77–83. [Google Scholar] [CrossRef]
  98. Cai, Z.W.; Hui-Hua, G.; Zhi-Wei, Y.; Run-Ying, Z.; Guang-Ya, Z. Characterization of a Novel Psychrophilic and Halophilic β-1, 3-xylanase From Deep-Sea Bacterium, Flammeovirga Pacifica Strain WPAGA1. Int. J. Biol. Macromol. 2018, 118, 2176–2184. [Google Scholar] [CrossRef]
  99. Miri, S.; Naghdi, M.; Rouissi, T.; Kaur Brar, S.; Martel, R. Recent biotechnological advances in petroleum hydrocarbons degradation under cold climate conditions: A review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 553–586. [Google Scholar] [CrossRef]
  100. Margesin, R.; Dieplinger, H.; Hofmann, J.; Sarg, B.; Lindner, H. A cold-active extracellular metalloprotease from Pedobactercryoconitis—production and properties. Res. Microbiol. 2005, 156, 499–505. [Google Scholar] [CrossRef] [PubMed]
  101. Saba, I.; Qazi, P.H.; Rather, S.A.; Dar, R.A.; Qadri, Q.A.; Ahmad, N.; Johri, S.; Taneja, S.C.; Shawl, S. Purification and characterization of a cold-active alkaline protease from Stenotrophomonas sp., isolated from Kashmir, India. World J. Microbiol. Biotechnol. 2012, 28, 1071–1079. [Google Scholar] [CrossRef] [PubMed]
  102. Furhan, J.; Awasthi, P.; Sharma, S. Biochemical characterization and homology modeling of cold-active alkophilic protease from Northwestern Himalayas and its application in detergent industry. Biocatal. Agric. Biotechnol. 2019, 17, 726–735. [Google Scholar] [CrossRef]
  103. Furhan, J.; Salaria, N.; Jabeen, M.; Qadri, J. Partial purification and characterization of cold-active metalloprotease by Bacillus sp. AP1 from Apharwat peak, Kashmir. Pak. J. Biotechnol. 2019, 16, 47–54. [Google Scholar] [CrossRef]
  104. Tamaki, H.; Hanada, S.; Kamagata, Y.; Nakamura, K.; Nomura, N.; Nakano, K.; Matsumura, M. Flavobacterium limicola sp. nov., a psychrophilic, organic-polymer-degrading bacterium isolated from freshwater sediments. Int. J. Syst. Evol. Microbiol. 2003, 53, 519–526. [Google Scholar] [CrossRef] [PubMed]
  105. Salwan, R.; Gulati, A.; Kasana, R.C. Phylogenetic diversity of alkaline protease-producing psychrotrophic bacteria from the glacier and cold environments of Lahaul and Spiti, India. J. Basic Microbiol. 2010, 50, 150–159. [Google Scholar] [CrossRef] [PubMed]
  106. Białkowska, A.M.; Szulczewska, K.M.; Krysiak, J.; Florczak, T.; Gromek, E.; Kassassir, H.; Kur, J.; Turkiewicz, M. Genetic and biochemical characterization of yeasts isolated from Antarctic soil samples. Polar Biol. 2017, 40, 1787–1803. [Google Scholar] [CrossRef]
  107. Ojha, B.K.; Singh, P.K.; Shrivastava, N. Enzymes in the Animal Feed Industry. In Enzymes in Food Biotechnology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 93–109. [Google Scholar]
  108. Hauksson, J.B.; Andresson, O.S.; Asgeirsson, B. Heat-labile bacterial alkaline phosphatase from a marine Vibrio sp. Enzyme. Microb. Technol. 2000, 27, 66–73. [Google Scholar] [CrossRef]
  109. Georlette, D.; Jonsson, Z.O.; Petegem, F.V.; Chessa, J.P.; Beeumen, J.V.; Hubscher, U.; Gerday, C. A DNA ligase from the psychrophile Pseudoalteromonas haloplanktis gives insight into the adaptation of proteins to low temperatures. Eur. J. Biochem. 2000, 267, 3502–3512. [Google Scholar] [CrossRef]
  110. Uma, S.; Jadhav, R.S.; Seshu-Kumar, G.; Shivaji, S.; Ray, M.K. An RNA polymerase with transcriptional activity at 0 °C from the Antarctic bacterium. FEBS Lett. 1999, 453, 313–317. [Google Scholar] [CrossRef]
  111. Siddiqui, K.S. Some like it hot, some like it cold: Temperature-dependent biotechnological applications and improvements in extremophilic enzymes. Biotechnol. Adv. 2015, 33, 1912–1922. [Google Scholar] [CrossRef] [PubMed]
  112. Jaafar, N.R.; Littler, D.; Beddoe, T.; Rossjohn, J.; Illias, R.M.; Mahadi, N.M.; Mackeen, M.M.; Murad, A.M.A.; Abu Bakar, F.D. Crystal structure of fuculosealdolase from the Antarctic psychrophilic yeast Glaciozyma antarctica PI12. Acta Crystallogr. F 2016, 72, 831–839. [Google Scholar] [CrossRef] [PubMed]
  113. Kobus, S.; Widderich, N.; Hoeppner, A.; Bremer, E.; Smits, S.H.J. Overproduction, crystallization and X-ray diffraction data analysis of ectoine synthase from the cold-adapted marine bacterium Sphingopyxis alaskensis. Acta Crystallogr. F 2015, 71, 1027–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Widderich, N.; Kobus, S.; Höppner, A.; Riclea, R.; Seubert, A.; Dickschat, J.S.; Heider, J.; Smits, S.H.; Bremer, E. Biochemistry and crystal structure of ectoinesynthase: A metal-containing member of the cup in the superfamily. PLoS ONE 2016, 11, e0151285. [Google Scholar] [CrossRef]
  115. Ishibashi, M.; Yamashita, S.; Tokunaga, M. Characterization of Halophilic Alkaline Phosphatase from Halomonas sp. 593, a Moderately Halophilic Bacterium. Biosci. Biotechnol. Biochem. 2005, 69, 1213–1216. [Google Scholar] [CrossRef]
  116. Wojciechowski, C.L.; Cardia, J.P.; Kantrowitz, E.R. Alkaline phosphatase from the hyperthermophilic bacterium T. maritima requires cobalt for activity. Protein Sci. 2002, 11, 903–911. [Google Scholar] [CrossRef]
  117. Kumari, U.; Singh, R.; Ray, T.; Rana, S.; Saha, P.; Malhotra, K.; Daniell, H. Validation of leaf enzymes in the detergent and textile industries: Launching of a new platform technology. Plant Biotechnol. J. 2019, 17, 1167–1182. [Google Scholar] [CrossRef]
  118. Al-Ghanayem, A.A.; Joseph, B. Current perspective in using cold-active enzymes as eco-friendly detergent additive. Appl. Microbiol. Biotechnol. 2020, 104, 2871–2882. [Google Scholar] [CrossRef]
  119. Gupta, R.; Beg, Q.; Lorenz, P. Bacterial alkaline proteases: Molecular approaches and industrial applications. Appl. Microbiol. Biotechnol. 2002, 59, 15–32. [Google Scholar]
  120. Chen, K.; Mo, Q.; Liu, H.; Yuan, F.; Chai, H.; Lu, F.; Zhang, H. Identification and characterization of a novel cold-tolerant extracellular protease from Planococcus sp. CGMCC 8088. Extremophiles 2018, 22, 473–484. [Google Scholar] [CrossRef]
  121. Tariq, A.; Reyaz, A.; Prabakaran, J.J. Purification and characterization of 56 KDa cold-active protease from Serratia marcescens. Afr. J. Microbiol. Res. 2011, 5, 5841–5847. [Google Scholar] [CrossRef]
  122. Salwan, R.; Kasana, R.C. Purification and characterization of an extracellular low temperature-active and alkaline stable peptidase from psychrotrophic Acinetobacter sp. MN 12 MTCC (10786). Indian J. Microbiol. 2013, 53, 63–69. [Google Scholar] [CrossRef]
  123. Hao, J.H.; Sun, M. Purification and characterization of a cold alkaline protease from a psychrophilic Pseudomonas aeruginosa HY1215. Appl. Biochem. Biotechnol. 2015, 175, 715–722. [Google Scholar] [CrossRef] [PubMed]
  124. Sharma, K.M.; Kumar, R.; Panwar, S.; Kumar, A. Microbial alkaline proteases: Optimization of production parameters and their properties. J. Genet. Eng. Biotechnol. 2017, 15, 115–126. [Google Scholar] [CrossRef]
  125. Nielsen, P.H. Life cycle assessment supports cold-wash enzymes. Int. J. Appl. Sci. 2005, 10, 24–26. [Google Scholar]
  126. Marshall, R.T.; Goff, H.D.; Hartel, R.W. Ice Cream; Kluwer Academic/Plenum Publishers: New York, NY, USA, 2003; pp. 1–11. [Google Scholar]
  127. Struvay, C.; Feller, G. Optimization to Low-Temperature Activity in Psychrophilic Enzymes. Int. J. Mol. Sci. 2012, 13, 11643–11665. [Google Scholar] [CrossRef]
  128. Karan, R.; Capes, M.D.; Das Sarma, S. Function and biotechnology of extremophilic enzymes in low water activity. Aquat. Biol. 2012, 8, 4. [Google Scholar] [CrossRef] [PubMed]
  129. Sondavid, K.N.; Shweta, B.B.; Jun, H.L.; Hak, J.K. Taking advantage of promiscuity of cold-active enzymes. Appl. Sci. 2020, 10, 8128. [Google Scholar]
  130. Novak, H.R.; Sayer, C.; Panning, J.; Littlechild, J.A. Characterization of an L-Haloacid Dehalogenase from the Marine Psychrophile Psychromonas ingrahamii with Potential Industrial Application. Mar. Biotechnol. 2013, 15, 695–705. [Google Scholar] [CrossRef]
  131. Huston, A.L. Biotechnological aspects of cold-adapted enzymes. In Psychrophiles: From Biodiversity to Biotechnology; Margesin, R., Schinner, F., Marx, J.C., Gerday, C., Eds.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 347–363. [Google Scholar]
  132. Trincone, A. Marine Biocatalysts: Enzymatic Features and Applications. Mar. Drugs 2011, 9, 478–499. [Google Scholar] [CrossRef]
  133. Pawlak-Szukalska, A.; Wanarska, M.; Popinigis, A.T.; Kur, J. A novel cold-active β-d-galactosidase with transglycosylation activity from the Antarctic Arthrobacter sp. 32cB—Gene cloning, purification, and characterization. Process Biochem. 2014, 49, 2122–2133. [Google Scholar] [CrossRef]
  134. Fornbacke, M.; Clarsund, M. Cold-adapted proteases as an emerging class of therapeutics. Infect. Dis. Ther. 2013, 2, 15–26. [Google Scholar] [CrossRef] [PubMed]
  135. Cazarin, C.; Lima, G.; da Silva, J.; Maro’stica, M. Enzymes in meat processing. In Enzymes in Food and Beverage Processing; Chandrasekaran, M., Ed.; CRC Press: Boca Raton, FL, USA, 2015; pp. 337–351. [Google Scholar]
  136. Balabanova, L.A.; Bakunina, I.Y.; Nedashkovskaya, O.I.; Makarenkova, I.D.; Zaporozhets, T.S.; Besednova, N.N.; Zvyagintseva, T.N.; Rasskazov, V.A. Molecular characterization and therapeutic potential of a marine bacterium Pseudoalteromonas sp. KMM 701 α-galactosidase. Mar. Biotechnol. 2010, 12, 111–120. [Google Scholar] [CrossRef] [PubMed]
  137. Çelik, A.; Yetis, G. An unusually cold-active nitro reductase for prodrug activations. Bioorg. Med. Chem. 2012, 20, 3540–3550. [Google Scholar] [CrossRef]
  138. Kobori, H.; Sullivan, C.W.; Shizuya, H. Heat-labile alkaline phosphatase from Antarctic bacteria: Rapid 5’ end-labeling of nucleic acids. Proc. Natl. Acad. Sci. USA 1984, 81, 6691–6695. [Google Scholar] [CrossRef]
  139. Lanes, O.; Leiros, I.; Smalås, A.O.; Willassen, N.P. Identification, cloning, and expression of uracil-DNA glycosylase from Atlantic cod (Gadusmorhua): Characterization and homology modeling of the cold-active catalytic domain. Extremophiles 2002, 6, 73–86. [Google Scholar] [CrossRef]
  140. Awazu, N.; Shodai, T.; Takakura, H.; Kitagawa, M.; Mukai, H.; Kato, I. Microorganism-Derived Psychrophilic Endonuclease. U.S. Patent 8,034,597 B2, 11 October 2011. [Google Scholar]
  141. Muller-Greven, J.C.; Post, M.A.; Kubu, C.J. Recombinant Colwellia psychrerythraea Alkaline Phosphatase and Uses Thereof. U.S. Patent US8486665B2, 16 July 2013. U.S. Patent US8129168B2, 3 June 2012. [Google Scholar]
  142. Rina, M.; Pozidis, C.; Mavromatis, K.; Tzanodaskalaki, M.; Kokkinidis, M.; Bouriotis, V. Alkaline phosphatase from the Antarctic strain TAB5. Properties and psychrophilic adaptations. Eur. J. Biochem. 2000, 267, 1230–1238. [Google Scholar] [CrossRef]
  143. Bjerga, G.E.; Lale, R.; Williamson, A.K. Engineering low-temperature expression systems for heterologous production of cold-adapted enzymes. Bioengineered 2016, 7, 33–38. [Google Scholar] [CrossRef]
  144. Ferrer, M.; Chernikova, T.N.; Timmis, K.N.; Golyshin, P.N. Expression of a temperature-sensitive esterase in a novel chaperone-based Escherichia coli strain. Appl. Environ. Microbiol. 2004, 70, 4499–4504. [Google Scholar] [CrossRef]
  145. Kim, H.W.; Wi, A.R.; Jeon, B.W.; Lee, J.H.; Shin, S.C.; Park, H.; Jeon, S.J. Cold adaptation of a psychrophilic chaperonin from Psychrobacter sp. and its application for heterologous protein expression. Biotechnol. Lett. 2015, 37, 1887–1893. [Google Scholar] [CrossRef]
  146. Esteban-Torres, M.; Mancheno, J.M.; de las Rivas, B.; Munoz, R. Characterization of a cold-active esterase from Lactobacillus plantarum suitable for food fermentations. J. Agric. Food Chem. 2014, 62, 5126–5132. [Google Scholar] [CrossRef] [PubMed]
  147. Bowman, J.S.; Deming, J.W. Alkane hydroxylase genes in psychrophile genomes and the potential for cold-active catalysis. BMC Genom. 2014, 15, 1120. [Google Scholar] [CrossRef]
  148. Kumar, L.; Bharadvaja, N. Chapter 6—Enzymatic bioremediation: A smart tool to fight environmental pollutants. In Smart Bioremediation Technologies; Bhatt, P., Ed.; Academic Press: Cambridge, MA, USA, 2019; pp. 99–118. [Google Scholar]
  149. Margesin, R.S.; Gander, G.; Zacke, A.M.; Gounot, A.M.; Schinner, F. Hydrocarbon degradation and enzyme activities of cold-adapted bacteria and yeasts. Extremophiles 2003, 7, 451–458. [Google Scholar] [CrossRef] [PubMed]
  150. Ruberto, L.; Vazquez, S.; Lobalbo, A.; Mac Cormack, W. Psychrotolerant hydrocarbon-degrading Rhodococcus strains isolated from polluted Antarctic soils. Antarct. Sci. 2005, 17, 47–56. [Google Scholar] [CrossRef]
  151. Aislabie, J.; Saul, D.J.; Foght, J.M. Bioremediation of hydrocarbon-contaminated polar soils. Extremophiles 2006, 10, 171–179. [Google Scholar] [CrossRef] [PubMed]
  152. Lin, X.; Yang, B.; Shen, J.; Du, N. Biodegradation of crude oil by an Arctic psychrotrophic bacterium Pseudoalteromomas sp. P29. Curr. Microbiol. 2009, 59, 341–345. [Google Scholar] [CrossRef]
Figure 1. Distribution of cold-active enzymes (a) based on organism nature, (b) on organism type, (c) and on cold-active enzymes reported.
Figure 1. Distribution of cold-active enzymes (a) based on organism nature, (b) on organism type, (c) and on cold-active enzymes reported.
Molecules 27 05885 g001
Figure 2. Application areas of cold-active enzymes (CAEs).
Figure 2. Application areas of cold-active enzymes (CAEs).
Molecules 27 05885 g002
Table 1. List of some cold-active enzymes and their potential industrial applications.
Table 1. List of some cold-active enzymes and their potential industrial applications.
Cold-Active EnzymeSource of IsolationPotential ApplicationsReferences
Lactases/β-galactosidaseParacoccus sp., Cystofilobasidium capitatum SPY11, Rhodotorula sp.Dairy industry (lactose hydrolysis)[11,45,66,77]
α-AmylasesGeomycespannorum, Bacillus subtilis N8, Geomyces pannorumFood, baking, and detergent industries[78,79,80]
CellulasesPyrococcus sp.Food and textile industry, ethanol fermentation[81,82]
ChitinasesMetschnikowia sp., Glaciozyma antarctica, Mrakia psychrophila, Sporobolomyces salmonicolorMeat tenderization, degradation of chitin rich wastes, control of phytopathogens[83,84,85]
LipasesPseudoalteromonas haloplanktis TAC125, Penicillium canesense, Pseudomonas sp. VITCLP4Animal feed, detergent, and textile industries[86,87,88]
Glycogen branching enzymeRhizomucor mieheiWheat bread manufacturing and baking industry[89]
PhytasesErwinia carotovora, Candida carpophila, Cryptococcus laurentii, Yarrowia lipolyticaFood and feed industry[90,91,92]
Pectinases (polygalacturonase and pectin-methylesterase)Cystofilobasidium capitatum PPY-1, Rhodotorula mucilaginosa PT1, Cystofilobasidium capitatum SPY11, Leucosporidium drummii, Sporobolomyces salmonicolor, Penicillium chrysogenum F46Food and fruit industries, pectin degradation, juice extraction[35,93,94,95]
XylanasesPseudoalteromonas haloplanktis TAH3A, Flavobacterium sp. MSY-2, Rhodococcus sp., Pseudomonas sp., Flammeovirga pacifica WPAGA1, Cryptococcus adeliensisBaking industry, xylan hydrolysis, osmoprotectants, biofuel production[96,97,98,99]
MetalloproteasesPedobacter cryoconitisBioremediation of wastewater at a lower temperature[100]
Alkaline proteaseStenotrophomonas sp., Bacillus subtilis WLCP1Detergent and textile industries[101,102,103]
ProteasesFlavobacterium limicola, Acinetobacter sp., Geomyces pannorum, Naganishia albidaTextile and leather industries, organic polymer mineralization in freshwater sediments, food and feed industry[104,105,106,107]
Serine proteasesPseudoalteromonas sp.Low-temperature food processing, leather industry[63]
Alkaline phosphatesVibrio sp.Molecular biology[108]
DNA ligasePseudoalteromonas haloplanktisMolecular biology and recombinant DNA technology[109]
RNA polymerasePseudomonas syringaeMolecular biology[110]
Fuculose aldolaseGlaciozyma antarctica PI12Pharmaceutical industry[111,112]
Ectoine synthaseSphingopyxis alaskensisCosmetics, biomedical industry[113,114]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hamid, B.; Bashir, Z.; Yatoo, A.M.; Mohiddin, F.; Majeed, N.; Bansal, M.; Poczai, P.; Almalki, W.H.; Sayyed, R.Z.; Shati, A.A.; et al. Cold-Active Enzymes and Their Potential Industrial Applications—A Review. Molecules 2022, 27, 5885. https://doi.org/10.3390/molecules27185885

AMA Style

Hamid B, Bashir Z, Yatoo AM, Mohiddin F, Majeed N, Bansal M, Poczai P, Almalki WH, Sayyed RZ, Shati AA, et al. Cold-Active Enzymes and Their Potential Industrial Applications—A Review. Molecules. 2022; 27(18):5885. https://doi.org/10.3390/molecules27185885

Chicago/Turabian Style

Hamid, Burhan, Zaffar Bashir, Ali Mohd Yatoo, Fayaz Mohiddin, Neesa Majeed, Monika Bansal, Peter Poczai, Waleed Hassan Almalki, R. Z. Sayyed, Ali A. Shati, and et al. 2022. "Cold-Active Enzymes and Their Potential Industrial Applications—A Review" Molecules 27, no. 18: 5885. https://doi.org/10.3390/molecules27185885

Article Metrics

Back to TopTop