Abstract
The current upper thermal limit for life as we know it is approximately 120°C. Microorganisms that grow optimally at temperatures of 75°C and above are usually referred to as ‘extreme thermophiles’ and include both bacteria and archaea. For over a century, there has been great scientific curiosity in the basic tenets that support life in thermal biotopes on earth and potentially on other solar bodies. Extreme thermophiles can be aerobes, anaerobes, autotrophs, heterotrophs, or chemolithotrophs, and are found in diverse environments including shallow marine fissures, deep sea hydrothermal vents, terrestrial hot springs – basically, anywhere there is hot water. Initial efforts to study extreme thermophiles faced challenges with their isolation from difficult to access locales, problems with their cultivation in laboratories, and lack of molecular tools. Fortunately, because of their relatively small genomes, many extreme thermophiles were among the first organisms to be sequenced, thereby opening up the application of systems biology-based methods to probe their unique physiological, metabolic and biotechnological features. The bacterial genera Caldicellulosiruptor, Thermotoga and Thermus, and the archaea belonging to the orders Thermococcales and Sulfolobales, are among the most studied extreme thermophiles to date. The recent emergence of genetic tools for many of these organisms provides the opportunity to move beyond basic discovery and manipulation to biotechnologically relevant applications of metabolic engineering.
Graphical Abstract

Introduction
Extreme thermophiles are distinct from other organisms due to their ability to subsist optimally at temperatures in excess of 75°C. Their survival in these harsh environments piqued the interest of curious microbiologists as far back as the turn of the 20th century. In fact, one of the earliest reports of thermophiles occurred in 1903 describing bacterial samples taken from pools in Yellowstone National Park1. Although this drew interest and debate about the limits of life and our evolutionary history, the study of thermophiles did not begin in earnest until the 1960’s. Around this time, extensive sampling projects in Yellowstone lead to the isolation of Thermus aquaticus2 (an aerobic bacterium with optimal growth between 70–75°C), known for its DNA polymerase that revolutionized the field of molecular biology through its use in the Polymerase Chain Reaction (PCR). This enzyme in particular represented one of the earliest uses of thermally-stable enzymes for a biotechnological application. The next few decades yielded the discovery of thermophiles around the globe in extremely diverse environments, ranging from volcanoes and calderas to deep sea smoker vents to terrestrial mud pools3–7.
The apparent diversity and novelty of these microbes likely drove early research in this field to uncover the molecular machinery central to their survival. In fact, some of the earliest sequenced genomes were extremophiles8–10, furthering efforts to understand the molecular and genetic basis for thermophily and the evolution of life. However, a lack of genetics tools has impeded the extensive study of these organisms by traditional approaches (i.e., gene deletions to understand the consequences of loss of function). In lieu of more traditional methods, the availability of genomic data for many extreme thermophiles supported ‘omics’-based approaches to ascertain the function of specific genes and their roles in the unique biochemistry of these organisms. As such, the merger of systems biology (e.g., transcriptomics and genomics), traditional microbiological studies, and newly emerging genetic systems11 are opening the door for metabolic engineering opportunities to bring extreme thermophiles into the technological limelight. This will allow for these organisms to be utilized as sources of uniquely functioning enzymes, optimized niche industrial strains, and novel metabolic engineering platforms. Such opportunities for biotechnological application are already being pursued for members of the bacterial genera Caldicellulosiruptor and Thermotoga and for archaea belonging to the orders Thermococcales and Sulfolobales. Here we present a brief overview of these extremely thermophilic organisms, with the intention of highlighting potential biotechnological applications, which exploit their distinctive metabolisms.
Sulfolobales
Perhaps the most distinctive subject matter for this review focuses on the extreme thermoacidophiles from order Sulfolobales. The Sulfolobales comprise an order of archaea taxonomically defined within the class Thermoprotei, within the phylum Crenarchaeota. These organisms inhabit environments characterized by both extreme temperatures (65–90 °C) and low pH (1.0–3.5), such as terrestrial solfatara or mud pools, often closely associated with volcanic activity and laden with inorganic materials12. In fact, the first species of the order to be isolated, Sulfolobus acidocaldarius (from Locomotive Spring in Yellowstone National Park), was reported to oxidize sulfur to fuel autotrophic growth, leading to the name Sulfolobales7. However, this phenotype has not been observed in the currently studied S. acidocaldarius type strains, although many isolates from the genera Sulfolobus, Metallosphaera and Acidianus utilize S0 as an electron donor5,13–21. Thus, reports that S. acidocaldarius strains from culture collections cannot5 oxidize S° suggests that repeated passages on rich media have led to the loss of this ability or that inherent difficulties exist in isolating pure cultures from environmental enrichments. Beyond sulfur oxidation, several species, especially those from the genus Acidianus, Sulfurisphaera, and Stygiolobus, are capable of sulfur reduction, and often utilize hydrogen to produce hydrogen sulfide as a metabolic end-product13,18,19,21–24.
While many members of the order grow lithotrophically, most known species exhibit modes of either strict heterotrophy or mixotrophy. Most members of the genera Sulfolobus and Metallosphaera are capable of utilizing protein-rich substrates, such as yeast extract or tryptone, under aerobic conditions. Furthermore, several species, such as Sulfolobus solfataricus, Sulfolobus shibitae25, and Sulfolobus islandicus26, use a wide variety of sugars in catabolic metabolism. In addition, members of the order, particularly in the genera Metallosphaera and Acidianus, are capable of oxidizing metal sulfides, a trait that is particularly useful for bioleaching of base, precious and strategic metals from mineral ores27–30. Finally, some members of the genus Acidianus are capable of using ferric iron as an electron acceptor under anaerobic conditions21,31.
Carbon Dioxide Fixation
Interestingly, the natural habitats of many Sulfolobales (solfatara/calderas) are limited or devoid of complex carbon sources, necessitating the process of autotrophy. The ability of organisms to fix carbon dioxide from the atmosphere is considered by many to explain the early formation of the multi-carbon molecules required to fuel life, explaining their retention in species across all three domains of life32,33. As it stands, 6 major routes exist for the fixation of carbon dioxide: the Calvin-Benson cycle (present in most plants), the reductive citric acid cycle (green sulfur bacteria), the reductive acetyl-CoA cycle (acetogens/methanogens), the hydroxypropionate bi-cycle (Chloroflexus), and the 3-hydroxypropionate/4-hydroxybutyrate (3-HP/4-HB) or dicarboxylate/4-hydroxybutyrate pathways (both from the Crenarchaeota)34. The habitats from which the Sulfolobales were isolated, unlike many other organisms, is characterized by copious oxygen and inorganic electron donors12. While the reactions driving this cycle are some of the most energy-demanding for autotrophic carbon assimilation, their advantage may lie in their relative insensitivity to oxygen, avoidance of side-reactions, direct utilization of bicarbonate, and thermal stability34,36.
Briefly, the 3-HP/4-HB cycle has two major products that enter into cellular metabolism. The first portion involves the addition of two bicarbonate molecules to acetyl-CoA to form succinyl-CoA, which is subsequently reduced in the second half of the cycle to 4-hydroxybutyrate and eventually dissociated to two molecules of acetyl-CoA (Figure 1)37. The key enzyme in the cycle is a biotinylated acetyl-CoA/propionyl-CoA carboxylase that is bi-functional and efficient in substrate turnover38. Metabolic analysis of the cycle has revealed that the major product of the cycle is not acetyl-CoA (as originally hypothesized), but rather succinyl-CoA (roughly two-thirds of the carbon flux), yielding malate and oxaloacetate in subsequent oxidation reactions39. This requires a turn and a half of the cycle to maintain acetyl-CoA levels and generate succinyl-CoA.
Figure 1. 3-hydroxypropionate/4-hydroxybutyrate (3-HP/4-HB) cycle from Metallosphaera sedula.
The cycle consists of two major portions: carbon incorporation (via bicarbonate) occurs in the first half (blue) of the cycle and is followed by subsequent reduction and reformation of two acetyl-CoA molecules in the second half (red).. Enzymes listed and their references: acetyl-coA carboxylase38,196 (ACC), acetoacetyl-CoA β-ketothiolase37 (ACCT), acryloyl-CoA reductase197 (ACR), crotonyl-CoA hydratase198 (CCH), 4-hydroxybutyrate-CoA dehydratase37 (HBCD), 4-hydroxybutyrate-CoA synthase199 (HBCS), 3-hydroxypropionate-CoA dehydratase197 (HPCD), 3-hydroxypropionate-CoA synthase200 (HPCS), methylmalonyl-CoA epimerase201 (MCE), methylmalonyl-CoA mutase201 (MCM), malonyl-CoA/succinyl-CoA reductase202 (MCR), malonate semialdehyde reductase202 (MSR), succinate semialdehyde reductase202 (SSR).
From an application-oriented point-of-view, it may be possible to use this pathway to sustainably produce high-value specialty chemicals, such as 3-HP or succinate. The former is used industrially in polymer production and the latter is used to produce solvents and polymers40. For this reason, several attempts have been made to utilize these genes in metabolically-engineered hosts. For instance, the first three enzymes have been expressed in P. furiosus to introduce a temperature-shift-responsive metabolic mode for the production of 3-HP41. Further work with this metabolically engineered strain has demonstrated that the addition of a biotin protein ligase can improve 3-hydroxy-propionate titers more than eight-fold42. This dramatic improvement is likely due to the presence of a biotinylated subunit on the key acetyl-CoA carboxylase enzyme from the cycle37. Thus, this well-studied pathway has opportunities to be utilized and improved upon.
Sulfur Utilization
In contrast to carbon metabolism, sulfur metabolic pathway discovery is hampered by the tendency of elemental sulfur and its derivatives to react non-enzymatically, masking the true identity of an enzyme’s substrate or products43,44. Because S0 is sparingly soluble in water under standard conditions, the true substrate for microbial growth on sulfur is likely soluble polysulfides and polythionates, introduced by non-enzymatic reactions45–47. However, these solubilizing reactions only occur at near-neutral pH, since under acidic conditions the equilibrium strongly favors elemental sulfur44. In order to overcome these solubility issues, it has been proposed that acidophiles may physically associate with sulfur particles or that at elevated temperatures sulfur becomes sufficiently soluble to support growth47. Regardless, these organisms have a suite of proteins capable of manipulating the initial elemental sulfur from the environment, as well as many of its derivatives.
The sulfur oxygenase reductase (SOR), first identified in a member of the Sulfolobales, Acidianus ambivalens48, appears to be key to acidophilic sulfur oxidation. This intracellular enzyme is active on elemental sulfur, indicating transport of elemental sulfur or one of its derivatives to the cytoplasm by some still unidentified mechanism. SOR appears to be limited to the Sulfolobales and a few extremophilic bacteria49; this makes sense given that other sulfur lithotrophs grow closer to neutral pH where more soluble sulfur species are abundant. SOR acts on elemental sulfur by disproportionating it equally between oxidized (sulfite; SO32−) and reduced (hydrogen sulfide; H2S) products. Further, the production of thiosulfate observed in early studies of the SOR48 is now believed to be the result of a non-enzymatic reaction50. SOR requires oxygen to function but uses no additional co-factors, suggesting its ability to conserve cellular energy. Instead, it “activates” long, unwieldy hydrophobic sulfur chains into smaller intermediates that can be used by other enzymes to generate energy.
Acidianus ambivalens has served as the model organism for sulfur metabolism studies in the Sulfolobales, since measurements of sulfur-active enzymes from cell extracts were used to construct a conceptual model of its sulfur oxidation pathways51. While the SOR enzyme has been the most thoroughly characterized enzyme with respect to sulfur metabolism48,49,52, studies of enzymes purified or detected from A. ambivalens cell extracts provide some insights into the complete oxidation pathway. The A. ambivalens genome remains unsequenced, so many of its enzymes are identifiable only by their activity in cell-extracts. This presents a challenge for systems biology-based efforts to understand the details of sulfur oxidation in other Sulfolobales, or even how this process contributes to A. ambivalens energetics and metabolism. Regardless, there appears to be two parallel processes by which A. ambivalens (and presumably the other sulfur-oxidizing members of the Sulfolobales) gain energy while oxidizing elemental sulfur to sulfate (SO42−). One pathway uses the membrane-associated oxidoreductases, TQO and SAOR, to reduce an electron carrier (such as quinone)53, thereby generating proton motive force via the terminal oxidase54,55, while the other pathway produces one high-energy phosphate bond (ADP from AMP) by direct substrate level phosphorylation via APSR and APAT, generating sulfate in the process51 (Figure 2). While the sulfur metabolism has been examined in bioleaching applications (see next section), the ability of S. metallicus to remove toxic H2S from high-temperature gas streams represents a potentially important technological use of sulfur oxidation56.
Figure 2. Chemolithotrophic pathways in the Sulfolobales.
The first half of the figure (blue) shows the hypothetical pathways for sulfur utilization in the Sulfolobales, including both oxidizing and reducing pathways, beginning with elemental sulfur. Sulfur reducing complexes: hydrogenase (Hyd), sulfur reductase (SR)57. Sulfur oxidizing enzymes: sulfur oxygenase reductase (SOR)48, thiosulfate:quinone oxidoreductase (TQO)53, Sulfite:acceptor oxidoreductase (SAOR), adenylylsulfate reductase (APSR), adenylylsulfate:phosphate adenyltransferase (APAT)51. The second panel shows a hypothetical pathway for the oxidation of ferrous iron using several fox stimulon proteins as well as some iron-responsive respiratory proteins. Ferrous-Oxidation (Fox), rusticyanin (Rus), cystathionine-β-synthase containing protein subunits A and B (CbsAB), sulfur oxidation (Sox), NADH dehydrogenase (NAD)62,64.
A. ambivalens has also served as the model system for anaerobic sulfur reduction among the Sulfolobales57. The enzyme pathway for sulfur reduction in A. ambivalens appears to be much simpler than for oxidation, possibly involving only two closely associated membrane complexes. A membrane hydrogenase passes electrons (via a quinone molecule) from H2 to a sulfur reductase, where they are used to reduce elemental sulfur to H2S57. The cycling of quinones between the two enzymes – forming a “redox loop” similar to the one used in Escherichia coli during growth on nitrate58 – is likely the way protons are transported across the membrane, coupling sulfur reduction to energy conservation (Figure 2).
Metal Oxidation
Along with interest in sulfur metabolism, some of the earliest work in determining the mechanism of metal oxidation in the Sulfolobales (and acidophiles, in general) involved the spectroscopic identification of unique cytochromes from iron-oxidizing cultures59. This original research led to the intensive study and eventual development of a model in the mesoacidophile Acidithiobacillus ferroxidans, involving the shuttling of electrons from the outer membrane of the cell to the inner membrane, driving a terminal oxidase to maintain pH homeostasis and the production of reducing power for intracellular metabolic needs60. Not surprisingly, the spectroscopic data from S. metallicus demonstrated early on that key differences exist between bacterial and archaeal metal oxidation, particularly in the presence of cytochromes and the roles of various protein complexes in transporting electrons59,61. However, some of the same systems-based approaches were utilized to detect the transcriptomic response of known iron-oxidizers, including S. metallicus62, M. yellowstonensis63 and M. sedula64, in the presence of iron. Interestingly, these experiments suggest the importance of merging several systems biology techniques in order to ascertain new pathways and information. While all three species contain the fox stimulon (an assortment of ferrous-responsive genes A–J) and key related genes (such as rusticyanin and the cystathionin-β-synthase subunits A and B), their regulation varies dramatically among the species with both constitutive and inducible expression observed during iron supplemented growth62–64. Yet, the merger of this transcriptomic data and genomics analysis yielded a hypothesized pathway for metal biooxidation in these organisms (Figure 2), which relies on a cytochrome b (as opposed to a cytochrome c), bifurcating rusticyanin(s), and two terminal complexes: an NADH dehydrogenase (generating reducing power) and a putative cytochrome c oxidase (driving pH homeostasis). Although similarities exist between the two systems in A. ferroxidans and Metallosphaera/Sulfolobus spp., distinctive co-factors and apparent differences in organization suggest that these systems are evolutionarily divergent modes of biooxidation65.
These differences, as well as the major phenotypical differences between these two classes of metal mobilizers, relates to their use in metal bioleaching applications. For example, many mesophilic organisms are ill-suited to bioleaching of highly gangue (i.e., high sulfur content) ores due to the extremely exothermic nature of sulfur oxidation chemistry. The build-up of heat can be problematic in large heap operations that rely on mesophiles alone66,67. This physiological trait cannot be undervalued given that the removal of elemental sulfur can improve cyanidation (a form of chemically-driven mineralization), which is commonly used in gold mining. Furthermore, extreme thermophiles appear to have some niche advantages over mesophiles for bioleaching of several specific types of copper ores, including the enhanced dissolution of copper from recalcitrant primary ores (such as chalcopyrite)68–70, selective mobilization of copper over molybdenum in copper-bearing molybdenite71,72, and the unassisted mineralization of arsenic in the form of arsenate from enargite ores73. Bioleaching operations targeting copper have increased dramatically and currently account for more than 15% of the global output74. Thus organisms that present an inherent propensity for copper solubilisation such as A. brierleyi, S. metallicus, or M. sedula deserve more investigation for their potential industrial application.
Pyrococcus furiosus
Pyrococcus furiosus, the type strain of the genus, was first isolated in 1986 from a hydrothermal vent off of the coast of Vulcano Island (Italy) and has been one of the most studied hyperthermophiles to date, due to its intriguing phenotypical characteristics75. Exhibiting optimal growth at 100°C and a pH near 7, it was the second genus, after the autotrophic sulfur-oxidizing Pyrodictium, capable of growth at temperatures at or above 100°C75. As a heterotrophic organism, P. furiosus is capable of utilizing hexose oligosaccharides such as cellobiose and laminarin76, chitin77,78, and peptides79. Efforts over the past three decades have elucidated many unique features of this organism, including various novelties in metabolic pathways, regulatory mechanisms and proteins and enzymes.
Central Glycolytic Metabolism
P. furiosus grows well on disaccharides (maltose and cellobiose) and glucans (laminarin and starch), but not on monosaccharides nor pentoses80. The reasons for this anomaly are unknown but monosaccharides may not be available externally in these high temperature environments as they are susceptible to the Maillard reactions, in which sugars react with available amino acids to form glycosylamines; this problem is especially exacerbated in peptide-rich media. Disaccharide and polysaccharide transport may also be more efficient energetically.
P. furiosus derives no net substrate level phosphorylation from glucose to pyruvate conversion, unlike the traditional Embden-Meyerhof (EM) pathway that provides two ATP per glucose and the Entner-Doudoroff (ED) pathway which yields one ATP per glucose (see Figure 3). Thus, the only net substrate level phosphorylation gains are a result of ATP-forming hydrolysis of acetyl-CoA, produced from pyruvate via pyruvate oxidoreductase, by acetyl-CoA synthetase81. P. furiosus contains a non-traditional variation of Embden-Meyerhof glycolysis, in which glucokinase and phosphofructokinase utilize ADP as the phosphoryl group donor, generating AMP80. In the early 1990s, these were the first reported ADP-dependent kinases80,82.
Figure 3.
Comparison of traditional Embden-Meyerhof-Parnas pathway with the modified pathway in the archaeon P. furiosus. Included are three fermentative pathways which utilize the reduced ferredoxin produced via glycolysis. Enzyme abbreviations: hydrogenase (hyd), ferredoxin:NADP oxidoreductase (FNOR), glutamate deaminase (GD), alanine aminotransferase (AT), and NADP:sulfur oxidoreductase (NSOR).
The absence of an energy-conserving step in the glycolytic pathway is due to the absence of a 1,3-biphosphoglycerate intermediate, which is found in both the EM and ED pathways. As shown in Figure 3, this direct conversion from glyceraldehyde-3-phosphate (GAP) to 3-phosphoglycerate (3PG) does result in production of a reducing equivalent in the form of reduced ferredoxin, but does not result in substrate level phosphorylation. The phosphate group is released via hydrolysis without capture of this high-energy bond. The enzyme responsible for the conversion of GAP to 3PG, GAP ferredoxin oxidoreductase (GAPOR), is unusual in that it requires tungsten, an element rarely found in biology83. The absence of other, more traditional glycolytic enzymes makes GAPOR’s function critical to sugar utilization. Thus, tungsten levels have a significant impact on the growth of P. furiosus in the presence of maltose83.
Fermentation Pathways
P. furiosus produces reduced ferredoxin through central glycolysis and, as an obligate anaerobe, must have a route to dispose of any excess reducing power. Two primary routes exist for this purpose depending on hydrogen partial pressures, the availability of elemental sulfur and nitrogen, and other regulatory factors. The primary route of regenerating oxidized ferredoxin is through a membrane-bound hydrogenase (MBH) that produces a ion gradient that allows ATP production via ATP synthase84. The hydrogenase is thought to exchange the proton gradient generated by hydrogen production for a Na+ gradient and this is utilized by a Na+-dependent ATP synthase85. This energy conserving hydrogenase therefore constitutes a single-step electron transport chain, and has been proposed as an evolutionary precursor to the complicated, multi-step electron transport chains that are common in present day microbes81. While the exact mechanism coupling proton transfer and hydrogen production is unknown, it is estimated that 0.3–0.4 molecules of ATP are produced per two electrons transferred81. Thus approximately 1.2 moles of ATP are produced for every mole of glucose converted to acetate via glycolytically produced reducing equivalents. Given the low energy production resulting from glycolysis, this fermentative process is particularly critical81.
When elemental sulfur is present, P. furiosus produces hydrogen sulfide rather than hydrogen gas86. As with H2 production, a proton gradient is formed by a membrane bound oxidoreductase (MBX)87. MBX is highly homologous to the membrane-bound hydrogenase and is thought to oxidize ferredoxin87. However, it is not known if MBX reduces sulfur directly or generates NADPH that is then used by a cytoplasmic NADPH- and CoA-dependent enzyme87. The reason for the preference for sulfur over proton reduction is not clear but it is strong since the switch from H2 to H2S production begins only minutes after the addition of sulfur to a growing culture88. The shift is mediated by SurR, a redox-responsive transcriptional regulator that has been well characterized89,90.
Another method for disposing of reductant during fermentation involves the transformation of pyruvate to alanine with the addition of available nitrogen91. This results in a major energetic penalty, however, as the pyruvate is not used to produce acetyl-CoA, which is responsible for the majority of ATP production. Thus, alanine pathway is only utilized when sulfur is absent and the hydrogen partial pressures are high91.
Applied Biocatalysis and Metabolic Engineering
Prior to detailed knowledge of the P. furiosus genome and development of genetic manipulation methodology, early work focused on characterizing its novel enzymes, with an eye towards industrial applications. While the DNA polymerase from Thermus aquaticus (Taq) is the most widely known and utilized thermostable polymerase in PCR reactions, the P. furiosus DNA polymerase is considerably more thermostable and of higher fidelity, yet two-to three-fold lower processivity. Due to its 3′ to 5′ exonuclease proofreading activity, the polymerase exhibits a ten-fold reduction in error rate compared to the Taq polymerase92. Additionally, the NADP(H)-dependent hydrogenase (SH1) from P. furiosus is extremely thermostable and has a temperature optimum of 95°C84. It has been utilized in a renewable H2 production in vitro system in which sugars are completely oxidized to CO2 and H2. A combination of pure enzymes comprising the pentose phosphate pathway (PPP) were used to convert sugars to CO2 and the NADPH that is then produced is oxidized and H2 is produced by SHI 84. Many other enzymes of interest from P. furiosus have been purified and characterized93, and include carbohydrate hydrolyzing enzymes, (e.g., α-amylase94, amylopullulanase94, endoglucanase95, and β-glucosidase96, and chitinase77) and proteases97,98.
P. furiosus now has a facile genetic system which has led to efforts directed at metabolic engineering99. Earlier work on a related extreme thermophile, Thermococcus kodakarensis, a member of the same order as P. furiosus, the Thermococcales, paved the way for the P. furiosus genetic tools100. For P. furiosus, its high growth temperature and tolerance to cold shock opens up its use for hosting metabolic pathways from much less extreme thermophiles101–103. In fact, a novel temperature-shift strategy has been demonstrated that minimizes P. furiosus metabolism at sub-optima temperatures to direct energy to heterologous production formation104. As mentioned above, P. furiosus produces soluble hydrogenases, which can regenerate reducing equivalents from hydrogen gas41. These hydrogenases could allow metabolically engineered P. furiosus to use electrons from H2 to produce highly reduced chemical products41. The insertion of three genes from the M. sedula 3-HP/4-HB carbon fixation cycle into P. furiosus demonstrated production of 3-hydroxyproprionate (3-HP) utilizing sugars and sequestering carbon dioxide for a portion of the molecule41. Another heterologous pathway expressed in P. furious utilized genes from three thermophilic organisms with optimal temperatures ranging from 65–75°C for the production of n-butanol105. With this alcohol pathway, significant diversion to ethanol was shown due to promiscuity of the aldehyde dehydrogenase enzymes105. The use of less thermophilic enzymes in heterologous biosynthetic pathways has provided insights into P. furiosus native metabolism at lower temperatures. For example, at 70–80°C, acetoin is produced as a major metabolic product. Along these lines, it was shown that the removal of acetolactate synthase in P. furiosus generates small amounts of ethanol as a metabolic end product, as pyruvate was directed toward acetate and eventually ethanol rather than acetoin103. In addition to its native abilities to utilize sugars, P. furiosus was engineered with a 16 gene cluster to oxidize carbon monoxide to carbon dioxide, producing H2 and energy in the process101. Overall, the ability to engineer utilization of unique energy sources, manipulate temperatures to optimize enzyme activities, and insert genes from a variety of organisms with different optimal growth temperature provide tools not typically available in model mesophilic hosts.
Caldicellulosiruptor spp
Caldicellulosiruptor is a bacterial genus containing the most thermophilic, cellulolytic microorganisms known to date. Isolated worldwide and having optimal growth temperatures between 70–78°C, these Gram-positive, asporogenic, obligate anaerobes have the ability to degrade unpretreated lignocellulosic biomass, a highly sought after phenotype for consolidated bioprocessing of fuels and chemicals. Many well-studied cellulolytic microbes are known to either secrete individual enzymes or large cellulolytic enzyme complexes (e.g., the cellulosome106) into their environment. In contrast, Caldicellulosiruptor species instead use an array of multi-modular enzymes to breakdown plant biomass107–112. These carbohydrate active enzymes (CAZymes) are composed of both catalytic (e.g., glycoside hydrolases [GH]) and non-catalytic (e.g., carbohydrate binding modules [CBM]) domains. All Caldicellulosiruptor species are able to utilize fructose, galactose, glucose, xylose, and pectin via a classical EMP pathway113–115. However, arabinose, rhamnose, and fucose utilization, is not conserved throughout all species113,115–118. Some of the sugars, such as xylose and arabinose, are broken down via the non-oxidative pentose phosphate pathway (PPP) and then piped into the EMP pathway as intermediates119. Although Caldicellulosiruptor species lack the oxidative PPP, which generally is responsible for NADPH production, they are still capable of generating NADPH; the exact enzymatic mechanism for this process is currently unknown120. Members of the Caldicellulosiruptor genus also contain an incomplete tricarboxylic acid (TCA) cycle, consisting of a reductive branch leading to fumarate and an oxidative branch producing succinyl-CoA.
Carbohydrate Utilization
Caldicellulosiruptor spp. produce many highly versatile and efficient multi-modular carbohydrate-degrading enzymes, made up of combinations of glycoside-hydrolase (GH) and carbohydrate binding module (CBM) domains. For example, CelA, a lignocellulosic CAZyme, is composed of five carbohydrate-specific domains: GH9-CBM3-CBM3-CBM3-GH48 connected by linker regions121–123. These different segments allow proteins to have multiple functions: simultaneously binding to its substrate (via the CBM3s), as well as cleavage of specific bonds (via the GH9 and GH48 domains); these GH9 and GH48 domains are capable of endo- and exoglucanase activity. CelA is present in only the most cellulolytic species of the genus and its two GH domains provide these species with a unique ‘drilling’ mode of action during biomass deconstruction121. While CBMs allow the CAZyme complex to adhere to the biomass, surface layer homology (SLH) domains are also found in the Caldicellulosiruptor multi-modular scheme124,125. Instead of being freely transported out of the cell, CAZymes with SLH domains are tethered the cell’s S-layer. As such, the enzymes can breakdown and bind, if they contain CBMs, substrates in close proximity to the microbe, providing better access to available sugar moieties. It also has been recently found that Caldicellulosiruptor species have a novel method of attaching themselves to crystalline cellulose. Structurally unique proteins, called tāpirins, are expressed on the cell surface, and contain a binding domain specific to insoluble cellulose126. Present in every member of the genus and highly expressed, tāpirins are thought to play an important role in how plant matter is deconstructed by microbes this genus.
Both methods of attachment, along with the large inventory of glycolytic enzymes, give this genus its impressive ability to degrade a wide variety of lignocellulosic substrates106. Caldicellulosiruptor species are capable of breaking down cellulose and hemicellulose (hexoses and pentoses), both as simple monosaccharides and complex biomasses108–110,116,127–131. Unlike many cellulolytic organisms, they do not exhibit carbon catabolite repression, a process by which certain sugars are preferentially metabolized, while excluding the usage of others106,111,114. This is especially advantageous in an industrial process involving lignocellulose conversion to fuels and chemicals, as these microbes can utilize multiple sugars simultaneously, with numerous points of entry to central carbon metabolism (see figure 4). Although Caldicellulosiruptor saccharolyticus was shown to grow well on a variety of sugars (arabinose, fructose, galactose, glucose, mannose, and xylose) simultaneously, the extent of which each monosaccharide was digested varied, with fructose being the most utilized111. In the absence of an apparent carbon utilization regulatory system, variation in sugar utilization among Caldicellulosiruptor species is likely due to presence or absence of certain metabolic pathways, e.g. the oxidative PPP, and/or essential transporters; the latter has only recently been better understood for a few Caldicellulosiruptor species with transcriptomics analysis of growth on substrates, such as simple sugars, crystalline cellulose (Avicel), and complex biomasses like switchgrass107,108,111,120,132.
Figure 4. Conserved metabolic pathways in all Caldicellulosiruptor species.
This includes sugar uptake, glycolytic, and fermentative pathways. The figure includes only the major steps, or start & end products. P is the abbreviation for phosphate, NADH for reduced nicotinamide adenine dinucleotide, ATP for adenosine triphosphate, Fdred for reduced ferrodoxin, and GDP for guanosine diphosphate.
One option to improve degradation of lignocellulose is to increase the CAZyme inventory of a microbe. While generally highly conserved in the Caldicellulosiruptor genus, the SLH domain xylanase from Caldicellulosiruptor kronotskyensis, Calkro_0402, is not present in C. bescii and, thus, was inserted into the genome to improve its ability to utilize xylan125. The manipulated strain successfully expressed the protein on the S-layer of the cell and improved xylan utilization significantly by doubling xylose release into the supernatant from oat spelt xylan. Growth on washed and unwashed birch xylan was improved, while dilute acid-pretreated switchgrass solubilization remained unaffected; this indicated that there are still other hurdles to lignocellulosic degradation that must be overcome. However, on substrates with high xylan content, the engineered strain showed improved solubilisation, possibly due to increased substrate attachment.
Fermentation
The major fermentation products of the Caldicellulosiruptor genus are hydrogen, carbon dioxide, and acetate116,133,134. Lactate production has also been measured, but only a trace amount of ethanol has been detected in wild-type cultures113,117,118,135. By far, the most studied product of all of these is molecular hydrogen, especially with C. saccharolyticus114,120,136–149. H2 generation is completed via hydrogenases utilizing reducing equivalents (Fdred and NADH) from central carbon metabolism (see figure 4). C. saccharolyticus, along with several other extreme thermophiles, is considered to be an ideal ‘biohydrogen factory’, as reported yields are close to the so-called ‘Thauer limit’ of 4 moles H2 per mol glucose114,116,150. Decreased H2 production results in the accumulation of NADH and Fdred, while increased H2 can instead push metabolic flux toward lactate production. At high levels of molecular hydrogen, product inhibition occurs via increased dissolved H2 levels151. NADH and Fdred are simultaneously oxidized by a bifurcating [FeFe]-hydrogenase, which uses both electron donors at the same time152, while Fdred is also oxidized by a membrane-bound [NiFe]-hydrogenase that is related to that found in P. furiosus85.
With the recent development of a genetic engineering system in Caldicellulosiruptor bescii, based on auxotrophic selection, targeted manipulations of the Caldicellulosiruptor genome and consequently metabolism are now possible153–156. The first directed demonstration of these methods actually involved the deletion of the single lactase dehydrogenase (ldh - Athe_1918) present in C. bescii to halt lactate production157. While the wild type and parent strains produced less hydrogen than the well-studied C. saccharolyticus (1.8 and 1.7 vs 2.5 mol H2/mol of glucose, respectively), the ldh knockout produced significantly more H2 on switchgrass, closer to the theoretical goal (3.4 mol H2/mol of glucose). As lactate formation ceased, acetate production increased by 38–40% over the wild type and parent strains.
More recently, ethanol production was demonstrated in C. bescii through the addition of an NADH-dependent alcohol dehydrogenase gene from Clostridium thermocellum (adhE – Cthe_0423) into the strain lacking lactate formation158; C. bescii does not possess a native alcohol or acetaldehyde dehydrogenase, and thus a representative gene was recruited from another thermophilic Firmicute. Due to the lower thermostability of the C. thermocellum protein, growth of the engineered strain was done at a maximum of 65°C. Strain growth on cellobiose, Avicel and switchgrass, and resulted in 14.8 mM, 14 mM, and 12.8 mM ethanol, respectively. Acetate production was also lowered, ranging from ~4–5 mM compared to the wild type (~6 mM) and parent (~8–9 mM) on all tested substrates. Another attempt at ethanol production was completed by individually inserting two bi-functional alcohol dehydrogenase genes from Thermoanaerobacter pseudethanolicus 39E, adhB (Teth39_0218) and adhE (Teth39_0206), into the C. bescii lactase dehydrogenase knockout159. Growing the modified strain at 75°C with cellobiose, ethanol was produced at reported levels of 1.4 mM and 2.9 mM, acetate at 15.5 mM and 14.1 mM, and H2 at 23.2 mM and 22.5 mM for the AdhB and AdhE knock-in strains, respectively; similar ethanol levels were also measured on Avicel and switchgrass.
With a genetic system now in place, gene ‘knockouts’ in C. bescii can be strategies to understand Caldicellulosiruptor metabolism and physiology. For instance, an uncharacterized [Ni-Fe] hydrogenase maturation gene cluster (hypABFCDE – Athe_1088–Athe_1093) was deleted from the aforementioned, modified ethanol-producing C. bescii strain containing adhE from C. thermocellum160. The resulting strain produced 20% less H2 than its parent, yet its H2 yield per mol of cellobiose increased 63% (3.58 vs 2.19 mol H2/mol cellobiose). Fermentation patterns on Avicel, cellobiose, and switchgrass showed that the engineered strain also produced acetate (1.6–5.7 mM – 34% less than parent) and ethanol (1.9–2.7 mM – 73% less than parent). Additionally, the knockout had reduced growth and a longer lag phase, which could result from the deleted gene acting as an ATP-generating protein pump. Other genetic manipulations with C. bescii include the deletion of CelA (Athe_1867)123, a predicted pectin lysase, and a putative AraC family transcriptional regulator genes (pecABCR – Athe_1853–1856)161. Continued efforts with the newly established genetics tools in C. bescii, and eventually other Caldicellulosiruptor species, will help reveal the basis for its ability to grow on lignocellulosic substrates.
Thermotoga spp
The bacterial genus Thermotoga contains nine named species that are obligate anaerobes capable of growth at optimal temperatures between 65–80°C, mostly isolated from submarine geothermal features162–164. These rod-shaped, Gram-negative, eubacteria were originally identifiable by their distinctive ‘toga’-like outer sheath and absence of an outer membrane. Beyond their unique appearance, the species in the genus Thermotoga share a remarkably large number of homologs (roughly 24% of the genome) with sequenced archaea165. This curious result has led some research into the evolutionary divergence/convergence of this bacterial lineage, suggesting the genomic features that may be critical in defining thermophily, such as the discovery of genes associated with biosynthesis of di-myo-inositol-phosphate, which may serve as a critical thermoprotectant compatible solute166. Further phylogenetic analysis has even suggested that mesophily may have developed from thermophily (within the order Thermotogales), given the ancestral sequence reconstruction of more thermally stable myo-inositol-phosphate synthase (MIPS)167 and emergence of ‘mesotoga’ species168. Also of interest is the presence of a system for catabolizing myo-inositol that provides utilization of compatible solutes but cannot provide a complete source for carbon utilization169. Within the genus, Thermotoga maritima has served as a model species for studying evolution, biomass deconstruction, and biohydrogen production170.
Carbohydrate Utilization
All Thermotoga species are chemoheterotrophs, although the range of substrate usage varies and includes numerous pentoses, hexoses, disaccharides, and polysaccharides, as well as yeast extract, acetate, methanol, and pectin163,164,171–175. This ability to utilize a broad array of carbohydrates appears to be supported by bioinformatics and transcriptomics suggesting a substrate-specific regulation and function of large numbers of ABC-transporters176, as well as many α- and β-glycoside hydrolases (GH)112. Intriguingly, T. maritima grows faster on complex carbohydrates than on monosaccharides, suggesting an adaptation to the breakdown of biomass in their natural environments177. The metabolism of carbohydrates by these organisms results in the formation of some typical fermentation products, such as acetate, carbon dioxide, and lactate, but also the generation of molecular hydrogen, and small amounts of ethanol, butanol, and butyrate178. It is worth noting that Thermotoga species utilize the traditional EMP and ED pathways179,180 for carbon utilization. However, they also contain, in some cases, unique enzymes that are adapted to optimizing the use of reducing power and energy generated from biomass deconstruction for the synthesis of fermentation products.
The initial genome annotation of T. maritima suggested a prevalence of mono- and polysaccharide utilization proteins (as much as 7% of identified genes)165. In contrast to organisms that produce large complexes for carbohydrate degradation (i.e., cellulosomes), T. maritima utilizes a broad array of both extra- and intra-cellular glycoside hydrolases, which have been detailed in previous reviews112. More recent examination of the pan genome, as well as transcriptomic data, suggests that Thermotoga species vary with respect to specific ABC sugar transporters and glycoside hydrolases181. Overall, the preponderance of thermally-stable, polysaccharide-degrading enzymes makes members of the genus and their enzymes intriguing candidates for the deconstruction of complex carbohydrates in industrial applications182. However, one of the limiting factors is the absence of any apparent capacity for growth on crystalline cellulose, suggesting a lack of cellulolytic enzymes in Thermotoga species183. In fact, to address this issue, efforts were directed at the ectopic expression of cellulases from C. saccharolyticus fused with T. maritima signal peptides. The resulting plasmids were used for Thermotoga sp. strain RQ2 transformations, where enhanced exoglucanase activity was observed, but eventually was lost due to poor plasmid maintenance184. However, a stable genetic system for T. maritima and T. sp. RQ7 was recently reported, based around a cryptic plasmid isolated from the latter185.
Fermentation
Of the major fermentation products from Thermotoga spp., H2 production is particularly interesting from a biotechnological perspective. High yields (3.8 mol H2/mol glucose) reported by Thermotoga neapolitana under anaerobic and microaerobic growth conditions186–188 approach the Thauer limit150. The production of H2 is most efficient when the balance of fermentation products is skewed toward acetate production as compared to lactate production, given that the enzymes identified in acetate production, phosphate acetyl-transferase and acetate kinase, avoid the re-oxidation of NADH and instead produce Fdred and ATP, respectively. In contrast, the production of lactate is driven by a lactate dehydrogenase that uses reducing equivalents (NADH) generated in the glycolytic process. Another possible key to efficient hydrogen production in these organisms, as in the Caldicellulosiruptor, is the coupling of Fdred and NADH oxidation by a bifurcating, [FeFe] hydrogenase189, in which Fdred likely drives the less favorable oxidation of NADH and improves the overall thermodynamics for producing hydrogen. This enzyme complex, first identified in T. maritima, appears to have a homolog in T. neapolitana, which has the highest reported H2 yields within the genus. There is also evidence that the build-up of molecular hydrogen and a possible inhibition mechanism can be alleviated through the co-culturing of T. maritima with Methanococcus jannaschii; the latter oxidizes H2 and generates methane190. This results in significant up-regulation of carbohydrate-active enzymes and growth-phase enzymes, as well as denser cultures191 (see Figure 5).
Figure 5.
Co-culture of Thermotoga maritima yellow/green rods) and Methanocaldococcus jannaschii (red cocci)191 – permission pending.
Besides molecular hydrogen, several species have been reported to produce ethanol as a fermentation product172,192. This result was not expected given the lack of detectable pyruvate decarboxylase activity. However, more recent work has identified the presence of both an alcohol dehydrogenase (from Thermotoga hypogea)193 and a bi-functional pyruvate ferredoxin oxidoreductase-pyruvate decarboxylase enzyme194. Additionally, butyrate (an odiferous organic used primarily as a perfume or food additive) production has been linked with hydrogen biosynthesis in studies involving T. neapolitana195; however, the mechanism of butyrate synthesis is still unknown in these organisms.
Thermus spp
The genus Thermus was among the first bacteria to be studied with respect to thermophily (Oshima and Imahori 1974), and the DNA polymerase from Thermus aquaticus2 was used in early efforts with the polymerase chain reaction (PCR) (ref). While not all Thermus species (at least 11 have been named and characterized) grow optimally at 70°C and above (extremely thermophilic), there are some that meet this thermal threshold. Thermus species are typically non-motile, non-sporulating, and are not naturally capable of fermentation. Efforts directed at understanding T. thermophilus metabolism revealed that this bacterium uses glycolysis and the TCA cycle to drive carbon flux and bioenergetics (Lee et al., Microb Cell fact 2014; Swarup et al. Metab Eng 2014). Molecular genetics tools were developed for Thermus thermophilus (Koyama Hoshino 1986 J Bact), based on its natural competence, which opened up opportunities for it to be examined as a model thermophile. In fact, the relative stability of thermophilic enzymes and early interest in the genus sparked the undertaking of crystallization projects aimed at characterizing recombinant versions of all the identified coding ORFs from Thermus thermophilus (Yokoyama et al. 2000; Sazanov and Hinchliffe 2006; Selmer et al. 2006; Severinov 2000; Yusupov et al. 2001). The overall goal of such projects was to provide a comprehensive database of structural characteristics that aid in the determination of protein function and domain architecture representing all of the major classes of proteins identified to date.
Enzyme and metabolic engineering efforts with Thermus
Although more thermophilic microorganisms have become available, Thermus species can be sources of thermostable enzymes for biotechnological applications. For example, enzymes from Thermus were included in an in vitro pathway that converted glucose into n-butanol (Krutsakorn Metab Engr 2013) and a xylose isomerse from this species was used to enable a recombinant Saccharomyces cerevisiae strain to grow on xylose (Karhumaa et al. Yeast 2005). Although genetics are relatively facile for these organisms, metabolic engineering pursuits have been limited. One of the earliest examples of metabolic engineering of the organism involved the transfer of nitrification genes among two members of the genus, allowing an aerobic Thermus species to grow anaerobically (Ramirez-Arcos, 1998). Additionally, a few attempts have been made at overexpression of native genes for the purpose of biotechnological applications involving specific enzymes such as DNA polymerase and Mn-dependent catalases (Hidalgo 2004; Moreno 2005). More recently, a strain of T. thermophilus HB8 was generated that could co-utilize xylose and glucose at temperatures up to 81°C, with a view towards processing lignocellulose, although this strain could not deconstruct biomass nor ferment the C5/C6 sugars (Cordova et al., 2016).
Conclusion
Although the study of extreme thermophiles has only gained traction in the past few decades, there are numerous metabolic and physiological features that distinguish these organisms from the other major groups of life and justify continued research endeavours. Much of this information has been ascertained via the use of genomics, transcriptomics, and proteomics in conjunction with traditional microbiological/biochemical techniques. Furthermore, this synthesis has led to the development of metabolic and physiological models in extreme thermophiles that are beginning to rival better characterized mesophilic systems. With the advent of next-generation sequencing technologies, it seems likely that previous work will be furthered by large-scale comparative genomics and metagenomics projects; this should further the discovery of novel metabolic features (i.e. enzymes and native biological pathways) with vital importance to our fundamental understanding of biology.
Beyond the scientific merit of studying extreme thermophiles, numerous opportunities exist to utilize these organisms for biotechnological advancement. As previously emphasized, the extreme conditions under which these organisms subsist has led to evolutionarily distinct metabolic and physiological features. In general, thermally stable proteins and heat-tolerant metabolic hosts could provide a major economic benefit to industrial processes. In the case of upstream processes, it may be possible to eliminate or minimize the energy costs associated with cooling or sterilizing bioreactors; while downstream processes may benefit from simple techniques -- such as heat pre-treatment -- to select for thermophilic enzymes produced recombinantly in mesophilic hosts, eliminating costly purification steps. Additionally, the increase in available genetic systems in these organisms will open many avenues for metabolic engineering. In fact, these organisms could have vital roles in the future of bioprocessing ranging from sustainable biochemical engineering to specialty chemical production to the deconstruction of inorganic and organic raw materials and even the recovery of base, precious and strategic metals.
Acknowledgments
This work was supported in part by the US National Science Foundation (CBET-1264052, CBET-1264053), the BioEnergy Science Center (BESC), a U.S. Department of Energy Bioenergy Research Center supported by the Office of Biological and Environmental Research in the DOE Office of Science, and the US Air Force Office of Sponsored Research (FA9550-13-1-0236). JAC and BMZ acknowledge support from NIH Biotechnology Traineeships (NIH T32GM008776-16). LLL acknowledges support from U.S. NSF Graduate Fellowship. CTS acknowledges support from a U.S. DoEd GAANN Molecular Biotechnology Fellowship.
Contributor Information
James A. Counts, Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27595-7905
Benjamin M. Zeldes, Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27595-7905
Laura L. Lee, Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27595-7905
Christopher T. Straub, Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27595-7905
Michael W.W. Adams, Department of Biochemistry and Molecular Biology, University of Georgia, Athens, GA 30602
Robert M. Kelly, Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27595-7905
References
- 1.Setchell WA. The upper temperature limits of life. Science. 1903;17:934–937. doi: 10.1126/science.17.441.934. [DOI] [PubMed] [Google Scholar]
- 2.Brock TD, Freeze H. Thermus aquaticus gen n and sp n., a non-sporulating extreme thermophile. J Bacteriol. 1969;98:289–297. doi: 10.1128/jb.98.1.289-297.1969. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Zillig W, et al. The Sulfolobus-‘Caldariella’ group: taxonomy on the basis of the structure of DNA-dependent RNA polymerases. Arch Microbiol. 1980;125:259–269. [Google Scholar]
- 4.Stetter KO, Lauerer G, Thomm M, Neuner A. Isolation of extremely thermophilic sulfate reducers: evidence for a novel branch of archaebacteria. Science. 1987;236:822–824. doi: 10.1126/science.236.4803.822. [DOI] [PubMed] [Google Scholar]
- 5.Huber G, Spinnler C, Gambacorta A, Stetter KO. Metallosphaera sedula gen and sp nov represents a new genus of aerobic, metal-mobilizing, thermoacidophilic archaebacteria. Syst Appl Microbiol. 1989;12:38–47. [Google Scholar]
- 6.Jannasch HW, Huber R, Belkin S, Stetter KO. Thermotoga neapolitana sp nov of the extremely thermophilic, eubacterial genus. Thermotoga Arch Microbiol. 1988;150:103–104. [Google Scholar]
- 7.Brock TD, Brock KM, Belly RT, Weiss RL. Sulfolobus: a new genus of sulfur-oxidizing bacteria living at low pH and high temperature. Arch Mikrobiol. 1972;84:54–68. doi: 10.1007/BF00408082. [DOI] [PubMed] [Google Scholar]
- 8.Bult CA, et al. Complete Genome Sequence of the Methanogenic Archaeon, Methanococcus jannaschii. Science. 1996;273:1058–1073. doi: 10.1126/science.273.5278.1058. [DOI] [PubMed] [Google Scholar]
- 9.Kelnk HP, et al. The complete genome sequence of the hyperthermophilic, sulphate-reducing archaeon Archaeoglobus fulgidus. Nature. 1997;390:364–370. doi: 10.1038/37052. [DOI] [PubMed] [Google Scholar]
- 10.Kawarabayasi Y, et al. Complete sequence and gene organization of the genome of a hyperthermophilic archaebacterium, Pyrococcus horikoshii OT3. DNA Res. 1998;5:55–76. doi: 10.1093/dnares/5.2.55. [DOI] [PubMed] [Google Scholar]
- 11.Zeldes BM, et al. Extremely thermophilic microorganisms as metabolic engineering platforms for production of fuels and industrial chemicals. Front Microbiol. 2015;6:1–17. doi: 10.3389/fmicb.2015.01209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Albers S-V, Siebers B. In: The Prokaryotes - other major lineages of Bacteria and the Archaea. Rosenberg E, DeLong EF, Lory S, Stackebrandt E, Thompson F, editors. Springer-Verlag; 2006. pp. 323–346. [DOI] [Google Scholar]
- 13.Giaveno MA, Urbieta MS, Ulloa JR, González Toril E, Donati ER. Physiologic versatility and growth flexibility as the main characteristics of a novel thermoacidophilic Acidianus strain isolated from Copahue Geothermal Area in Argentina. Microb Ecol. 2013;65:336–346. doi: 10.1007/s00248-012-0129-4. [DOI] [PubMed] [Google Scholar]
- 14.Grogan D, Palm P, Zillig W. Isolate B12, which harbours a virus-like element, represents a new species of the archaebacterial genus Sulfolobus, Sulfolobus shibatae, sp nov. Arch Microbiol. 1990;154:594–599. doi: 10.1007/BF00248842. [DOI] [PubMed] [Google Scholar]
- 15.Segerer A, Neuner A, Kristjansson JK, Stetter KO. Acidianus infernus, gen nov., sp nov and Acidianus brierleyi comb nov.: facultatively aerobic, extremely acidophilic thermophilic sulfur-metabolizing archaebacteria. Int J Syst Bacteriol. 1986;36:559–564. [Google Scholar]
- 16.Takayanagi S, et al. Sulfolobus hakonensis sp nov., a novel species of acidothermophilic archaeon. Int J Syst Bacteriol. 1996;46:377–382. doi: 10.1099/00207713-46-2-377. [DOI] [PubMed] [Google Scholar]
- 17.Brierley CL, Brierley JA. A chemoautotrophic and thermophilic microorganism isolated from an acid hot spring. Can J Microbiol. 1973;19:183–188. doi: 10.1139/m73-028. [DOI] [PubMed] [Google Scholar]
- 18.Ding J, et al. A novel acidophilic, thermophilic iron and sulfur-oxidizing archaeon isolated from a hot spring of Tengchong, Yunnan, China. Brazilian J Microbiol. 2011;42:514–525. doi: 10.1590/S1517-838220110002000016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.He ZG, Zhong H, Li Y. Acidianus tengchongensis sp nov., a new species of acidothermophilic archaeon isolated from an acidothermal spring. Curr Microbiol. 2004;48:159–163. doi: 10.1007/s00284-003-4155-9. [DOI] [PubMed] [Google Scholar]
- 20.Fuchs T, Huber H, Teiner K, Burggraf S, Stetter KO. Metallosphaera prunae, sp nov., a novel metal-mobilizing, thermoacidophilic archaeum, isolated from a uranium mine in Germany. Syst Appl Microbiol. 1995;18:560–566. [Google Scholar]
- 21.Plumb JJ, Haddad CM, Gibson JaE, Franzmann PD. Acidianus sulfidivorans sp. nov., an extremely acidophilic, thermophilic archaeon isolated from a solfatara on Lihir Island, Papua New Guinea, and emendation of the genus description. Int J Syst Evol Microbiol. 2007;57:1418–1423. doi: 10.1099/ijs.0.64846-0. [DOI] [PubMed] [Google Scholar]
- 22.Urbieta MS, et al. Draft genome sequence of the novel thermoacidophilic archaeon Acidianus copahuensis strain ALE1, isolated from the Copahue Volcanic Area in Neuquén, Argentina. Genome Announc. 2014;2 doi: 10.1128/genomeA.00259-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Segerer AH, Trincone A, Gahrtz M, Stetter KO. Stygiolobus azoricus gen nov., sp nov represents a novel genus of anaerobic, extremely thermoacidophilic archaebacteria of the order Sulfolobales. Int J Syst Bacteriol. 1991;41:495–501. [Google Scholar]
- 24.Kurosawa N, et al. Sulfurisphaera ohwakuensis gen. nov., sp nov., a novel extremely thermophilic acidophile of the order Sulfolobales. Int J Syst Bacteriol. 1998;48:451–456. doi: 10.1099/00207713-48-2-451. [DOI] [PubMed] [Google Scholar]
- 25.Grogan DW. Phenotypic characterization of the archaebacterial genus Sulfolobus: Comparison of five wild-type strains. J Bacteriol. 1989;171:6710–6719. doi: 10.1128/jb.171.12.6710-6719.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Zillig W, et al. Screening for Sulfolobales, their plasmids and their viruses in Icelandic solfataras. Systematic and Applied Microbiology. 1994;16:609–628. [Google Scholar]
- 27.Dew DW, Buuren C, Van McEwan K, Bowker C. Bioleaching of base metal sulphide concentrates: A comparison of high and low temperature bioleaching. J South African Inst Min Metall. 2000;100:409–414. [Google Scholar]
- 28.Norris PR, Burton NP, Clark DA. Mineral sulfide concentrate leaching in high temperature bioreactors. Miner Eng. 2013;48:10–19. [Google Scholar]
- 29.Wheaton G, Counts J, Mukherjee A, Kruh J, Kelly R. The confluence of heavy metal biooxidation and heavy metal resistance: implications for bioleaching by extreme thermoacidophiles. Minerals. 2015;5:397–451. [Google Scholar]
- 30.Brierley CL, Brierley JA. Progress in bioleaching: part B: applications of microbial processes by the minerals industries. Appl Microbiol Biotechnol. 2013;97:7543–7552. doi: 10.1007/s00253-013-5095-3. [DOI] [PubMed] [Google Scholar]
- 31.Yoshida N, et al. Acidianus manzaensis sp nov., a novel thermoacidophilic archaeon growing autotrophically by the oxidation of H2 with the reduction of Fe3+ Curr Microbiol. 2006;53:406–411. doi: 10.1007/s00284-006-0151-1. [DOI] [PubMed] [Google Scholar]
- 32.Wächtershäuser G. On the chemistry and evolution of the pioneer organism. Chem Biodivers. 2007;4:584–602. doi: 10.1002/cbdv.200790052. [DOI] [PubMed] [Google Scholar]
- 33.Fuchs G. Alternative pathways of carbon dioxide fixation: insights into the early evolution of life? Annu Rev Microbiol. 2011;65:631–658. doi: 10.1146/annurev-micro-090110-102801. [DOI] [PubMed] [Google Scholar]
- 34.Berg IA. Ecological aspects of the distribution of different autotrophic CO2 fixation pathways. Appl Environ Microbiol. 2011;77:1925–1936. doi: 10.1128/AEM.02473-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Martin W, Baross J, Kelley D, Russell M. Hydrothermal vents and the origin of life. Nat Rev Microbiol. 2008;6:805–814. doi: 10.1038/nrmicro1991. [DOI] [PubMed] [Google Scholar]
- 36.Bar-Even A, Flamholz A, Noor E, Milo R. Thermodynamic constraints shape the structure of carbon fixation pathways. Biochim Biophys Acta - Bioenerg. 2012;1817:1646–1659. doi: 10.1016/j.bbabio.2012.05.002. [DOI] [PubMed] [Google Scholar]
- 37.Berg IA, et al. A 3-hydroxypropionate/4-hydroxybutyrate autotrophic carbon dioxide assimilation pathway in archaea. Science. 2007;318:1782–1786. doi: 10.1126/science.1149976. [DOI] [PubMed] [Google Scholar]
- 38.Hügler M, Krieger RS, Jahn M, Fuchs G. Characterization of acetyl-CoA/propionyl-CoA carboxylase in Metallosphaera sedula: carboxylating enzyme in the 3-hydroxypropionate cycle for autotrophic carbon fixation. Eur J Biochem. 2003;270:736–744. doi: 10.1046/j.1432-1033.2003.03434.x. [DOI] [PubMed] [Google Scholar]
- 39.Estelmann S, et al. Labeling and enzyme studies of the central carbon metabolism in Metallosphaera sedula. J Bacteriol. 2011;193:1191–1200. doi: 10.1128/JB.01155-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Werpy T, Petersen G. Top value added chemicals from biomass volume I — results of screening for potential candidates from sugars and synthesis gas. 2004 doi: 10.2172/15008859. [DOI] [Google Scholar]
- 41.Keller MW, et al. Exploiting microbial hyperthermophilicity to produce an industrial chemical, using hydrogen and carbon dioxide. Proc Natl Acad Sci U S A. 2013;110:5840–5845. doi: 10.1073/pnas.1222607110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Lian H, et al. Ancillary contributions of heterologous biotin protein ligase and carbonic anhydrase for CO2 incorporation into 3-hydroxypropionate by metabolically engineered Pyrococcus furiosus. Biotechnol Bioeng. 2016 doi: 10.1002/bit.26033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Suzuki I. Oxidation of inorganic sulfur compounds: Chemical and enzymatic reactions. Can J Microbiol. 1999;45:97–105. [Google Scholar]
- 44.Kletzin A. In: Archaea. Garrett RA, Klenk H-P, editors. Blackwell Publishing Ltd; 2006. pp. 261–274. [Google Scholar]
- 45.Blumentals II, Itoh M, Olson GJ, Kelly RM. Role of polysulfides in reduction of elemental sulfur by the hyperthermophilic archaebacterium Pyrococcus furiosus. Appl Environ Microbiol. 1990;56:1255–1262. doi: 10.1128/aem.56.5.1255-1262.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Rabus R, Hansen TA, Widdel F. In: The Prokaryotes. Dworkin M, Falkow S, Rosenberg E, Schleifer K-H, Stackebrandt E, editors. Springer; New York: 2006. pp. 659–768. [Google Scholar]
- 47.Schauder R, Müller E. Polysulfide as a possible substrate for sulfur-reducing bacteria. Arch Microbiol. 160:377–382. [Google Scholar]
- 48.Kletzin A. Coupled enzymatic production of sulfite, thiosulfate, and hydrogen sulfide from sulfur: purification and properties of a sulfur oxygenase reductase from the facultatively anaerobic archaebacterium Desulfurolobus ambivalens. J Bacteriol. 1989;171:1638–1643. doi: 10.1128/jb.171.3.1638-1643.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Veith A, et al. Substrate pathways and mechanisms of inhibition in the sulfur oxygenase reductase of Acidianus ambivalens. Front Microbiol. 2011;2:1–12. doi: 10.3389/fmicb.2011.00037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Kletzin A. In: Microbial Sulfur Metabolism. Dahl DC, Friedrich DCG, editors. Springer; Berlin Heidelberg: 2008. pp. 184–201. [Google Scholar]
- 51.Zimmermann P, Laska S, Kletzin A. Two modes of sulfite oxidation in the extremely thermophilic and acidophilic archaeon Acidianus ambivalens. Arch Microbiol. 1999;172:76–82. doi: 10.1007/s002030050743. [DOI] [PubMed] [Google Scholar]
- 52.Kletzin A. Molecular characterization of the sor gene, which encodes the sulfur oxygenase/reductase of the thermoacidophilic archaeum Desulfurolobus ambivalens. J Bacteriol. 1992;174:5854–5859. doi: 10.1128/jb.174.18.5854-5859.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Müller FH, et al. Coupling of the pathway of sulphur oxidation to dioxygen reduction: characterization of a novel membrane-bound thiosulphate:quinone oxidoreductase. Mol Microbiol. 2004;53:1147–1160. doi: 10.1111/j.1365-2958.2004.04193.x. [DOI] [PubMed] [Google Scholar]
- 54.Purschke WG, Schmidt CL, Petersen A, Schäfer G. The terminal quinol oxidase of the hyperthermophilic archaeon Acidianus ambivalens exhibits a novel subunit structure and gene organization. J Bacteriol. 1997;179:1344–1353. doi: 10.1128/jb.179.4.1344-1353.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Anemüller S, et al. A cytochrome aa3-type quinol oxidase from Desulfurolobus ambivalens, the most acidophilic archaeon. FEMS Microbiol Lett. 1994;117:275–280. [Google Scholar]
- 56.Morales P, Gentina J, Morales M, Aroca G, Silva J. Biofiltration of hydrogen sulfide by Sulfolobus metallicus at high temperatures. Water Sci Technol. 2012;66:1958–1961. doi: 10.2166/wst.2012.402. [DOI] [PubMed] [Google Scholar]
- 57.Laska S, Lottspeich F, Kletzin A. Membrane-bound hydrogenase and sulfur reductase of the hyperthermophilic and acidophilic archaeon Acidianus ambivalens. Microbiology. 2003;149:2357–2371. doi: 10.1099/mic.0.26455-0. [DOI] [PubMed] [Google Scholar]
- 58.Jormakka M, Byrne B, Iwata S. Protonmotive force generation by a redox loop mechanism. FEBS Lett. 2003;545:25–30. doi: 10.1016/s0014-5793(03)00389-2. [DOI] [PubMed] [Google Scholar]
- 59.Barr DW, Ingledew WJ, Norris PR. Respiratory chain components of iron-oxidizing acidophilic bacteria. FEMS Microbiol Lett. 1990;70:85–89. [Google Scholar]
- 60.Quatrini R, et al. Extending the models for iron and sulfur oxidation in the extreme acidophile Acidithiobacillus ferrooxidans. BMC Genomics. 2009;10:394–413. doi: 10.1186/1471-2164-10-394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Blake RC, Shute Ea, Greenwood MM, Spencer GH, Ingledew WJ. Enzymes of aerobic respiration on iron. FEMS Microbiol Rev. 1993;11:9–18. doi: 10.1111/j.1574-6976.1993.tb00261.x. [DOI] [PubMed] [Google Scholar]
- 62.Bathe S, Norris PR. Ferrous iron- and sulfur-induced genes in Sulfolobus metallicus. Appl Environ Microbiol. 2007;73:2491–2497. doi: 10.1128/AEM.02589-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Kozubal Ma, et al. Terminal oxidase diversity and function in ‘Metallosphaera yellowstonensis’: gene expression and protein modeling suggest mechanisms of Fe(II) oxidation in the Sulfolobales. Appl Environ Microbiol. 2011;77:1844–1853. doi: 10.1128/AEM.01646-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Auernik KS, Kelly RM. Identification of components of electron transport chains in the extremely thermoacidophilic crenarchaeon Metallosphaera sedula through iron and sulfur compound oxidation transcriptomes. Appl Environ Microbiol. 2008;74:7723–7732. doi: 10.1128/AEM.01545-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Ilbert M, Bonnefoy V. Insight into the evolution of the iron oxidation pathways. Biochim Biophys Acta. 2013;1827:161–175. doi: 10.1016/j.bbabio.2012.10.001. [DOI] [PubMed] [Google Scholar]
- 66.Clark ME, Batty JD, van Buuren CB, Dew DW, Eamon Ma. Biotechnology in minerals processing: Technological breakthroughs creating value. Hydrometallurgy. 2006;83:3–9. [Google Scholar]
- 67.van Staden PJ, Gericke M, Craven PM. Minerals biotechnology: trends, opportunities and challenges. Hydrometallurgy 2008: Proceedings of the Sixth International Symposium; Society for Mining Metallurgy and Exploration; 2008. pp. 6–13. [Google Scholar]
- 68.Gericke M, Pinches A, van Rooyen J. Bioleaching of a chalcopyrite concentrate using an extremely thermophilic culture. Int J Miner Process. 2001;62:243–255. [Google Scholar]
- 69.Gericke M, Govender Y, Pinches A. Tank bioleaching of low-grade chalcopyrite concentrates using redox control. Hydrometallurgy. 2010;104:414–419. [Google Scholar]
- 70.Norris PR, Calvo-Bado La, Brown CF, Davis-Belmar CS. Ore column leaching with thermophiles: I, copper sulfide ore. Hydrometallurgy. 2012;127–128:62–69. [Google Scholar]
- 71.Romano P, et al. Comparative study on the selective chalcopyrite bioleaching of a molybdenite concentrate with mesophilic and thermophilic bacteria. FEMS Microbiol Lett. 2001;196:71–75. doi: 10.1111/j.1574-6968.2001.tb10543.x. [DOI] [PubMed] [Google Scholar]
- 72.Abdollahi H, et al. Mesophilic and thermophilic bioleaching of copper from a chalcopyrite-containing molybdenite concentrate. Int J Miner Process. 2014;128:25–32. [Google Scholar]
- 73.Takatsugi K, Sasaki K, Hirajima T. Mechanism of the enhancement of bioleaching of copper from enargite by thermophilic iron-oxidizing archaea with the concomitant precipitation of arsenic. Hydrometallurgy. 2011;109:90–96. [Google Scholar]
- 74.Johnson DB. Biomining-biotechnologies for extracting and recovering metals from ores and waste materials. Curr Opin Biotechnol. 2014;30:24–31. doi: 10.1016/j.copbio.2014.04.008. [DOI] [PubMed] [Google Scholar]
- 75.Fiala G, Stetter KO. Pyrococcus furiosus sp nov represents novel genus of marine heterotrophic archaebacteria growing optimally at 100°C. Arch Microbiol. 1986;145:56–61. [Google Scholar]
- 76.Kengen SWM, Luesink EJ, Stams AJM, Zenhder AJB. Purification and characterization of an extremely thermostable beta-glucosidase from the hyperthermophilic archaeon Pyrococcus furiosus. Eur J Biochem. 1993;213:305–312. doi: 10.1111/j.1432-1033.1993.tb17763.x. [DOI] [PubMed] [Google Scholar]
- 77.Gao J, Bauer MW, Shockley KR, Pysz MA, Kelly RM. Growth of hyperthermophilic archaeon Pyrococcus furiosus on chitin involves two family 18 chitinases. Appl Environ Microbiol. 2003;69:3119. doi: 10.1128/AEM.69.6.3119-3128.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Kreuzer M, Schmutzler K, Waege I, Thomm M, Hausner W. Genetic engineering of Pyrococcus furiosus to use chitin as a carbon source. BMC Biotechnol. 2013;13:1–10. doi: 10.1186/1472-6750-13-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Snowden LJ, Blumentals II, Kelly RM. Regulation of proteolytic activity in the hyperthermophile Pyrococcus furiosus. Appl Environ Microbiol. 1992;58:1134–1141. doi: 10.1128/aem.58.4.1134-1141.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Kengen SWM, et al. Evidence for the operation of a novel Embden-Meyerhof pathway that involves ADP-dependent kinases during sugar fermentation by Pyrococcus furiosus. J Biol Chem. 1994;269:17537–17541. [PubMed] [Google Scholar]
- 81.Sapra R, Bagramyan K, Adams MWW. A simple energy-conserving system: proton reduction coupled to proton translocation. Proc Natl Acad Sci U S A. 2003;100:7545–7550. doi: 10.1073/pnas.1331436100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Mai X, Adams MW. Purification and characterization of two reversible and ADP-dependent acetyl coenzyme A synthetases from the hyperthermophilic archaeon Pyrococcus furiosus. J Bacteriol. 1996;178:5897–5903. doi: 10.1128/jb.178.20.5897-5903.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Schicho RN, et al. Influence of tungsten on metabolic patterns in Pyrococcus furiosus, a hyperthermophilic archaeon. Arch Microbiol. 1993;159:380–385. [Google Scholar]
- 84.Bryant FO, Adams MW. Characterization of hydrogenase from the hyperthermophilic archaebacterium, Pyrococcus furiosus. J Biol Chem. 1989;264:5070–5079. [PubMed] [Google Scholar]
- 85.Schut GJ, Boyd ES, Peters JW, Adams MWW. The modular respiratory complexes involved in hydrogen and sulfur metabolism by heterotrophic hyperthermophilic archaea and their evolutionary implications. FEMS Microbiol Rev. 2013;37:182–203. doi: 10.1111/j.1574-6976.2012.00346.x. [DOI] [PubMed] [Google Scholar]
- 86.Adams MWW, et al. Key role for sulfur in peptide metabolism and in regulation of three hydrogenases in the hyperthermophilic archaeon Pyrococcus furiosus. J Bacteriol. 2001;183:716–724. doi: 10.1128/JB.183.2.716-724.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Bridger SL, et al. Deletion strains reveal metabolic roles for key elemental sulfur-responsive proteins in Pyrococcus furiosus. J Bacteriol. 2011;193:6498–6504. doi: 10.1128/JB.05445-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Schut GJ, Bridger SL, Adams MWW. Insights into the metabolism of elemental sulfur by the hyperthermophilic archaeon Pyrococcus furiosus: characterization of a coenzyme A-dependent NAD(P)H sulfur oxidoreductase. J Bacteriol. 2007;189:4431–4441. doi: 10.1128/JB.00031-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Lipscomb GL, et al. SurR: A transcriptional activator and repressor controlling hydrogen and elemental sulphur metabolism in Pyrococcus furiosus. Mol Microbiol. 2009;71:332–349. doi: 10.1111/j.1365-2958.2008.06525.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Yang H, et al. SurR regulates hydrogen production in Pyrococcus furiosus by a sulfur-dependent redox switch. Mol Microbiol. 2010;77:1111–1122. doi: 10.1111/j.1365-2958.2010.07275.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Kengen SWM, Stams AJM. Formation of L-alanine as a reduced end product in carbohydrate fermentation by the hyperthermophilic archaeon Pyrococcus furiosus. Arch Microbiol. 1994;161:168–175. [Google Scholar]
- 92.Lundberg KS, et al. High-fidelity amplification using a thermostable DNA polymerase isolated from Pyrococcus furiosus. Gene. 1991;108:1–6. doi: 10.1016/0378-1119(91)90480-y. [DOI] [PubMed] [Google Scholar]
- 93.Niehaus F, et al. Extremophiles as a source of novel enzymes for industrial application. Appl Microbiol Biotechnol. 1999;51:711–729. doi: 10.1007/s002530051456. [DOI] [PubMed] [Google Scholar]
- 94.Dong GQ, Vieille C, Zeikus JG, Savchenko A, Zeikus JG. Cloning, sequencing, and expression of the gene encoding extracellular alpha-amylase from Pyrococcus furiosus and biochemical characterization of the recombinant enzyme. Appl Env Microbiol. 1997;63:3569–3576. doi: 10.1128/aem.63.9.3569-3576.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Bauer MW, Driskill LE, Kelly RM. An endoglucanase, EglA, from the hyperthermophilic archaeon Pyrococcus furiosus hydrolyzes β-1,4 bonds in mixed-linkage (1-->3),(1-->4)-β-D-glucans and cellulose. J Bacteriol. 1999;181:284. doi: 10.1128/jb.181.1.284-290.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Bauer MW, Kelly RM. The family 1 beta-glucosidases from Pyrococcus furiosus and Agrobacterium faecalis share a common catalytic mechanism. Biochemistry. 1998;37:17170–17178. doi: 10.1021/bi9814944. [DOI] [PubMed] [Google Scholar]
- 97.Blumentals II, Robinson AS, Kelly RM. Characterization of sodium dodecyl sulfate-resistant proteolytic activity in the hyperthermophilic archaebacterium Pyrococcus furiosus. Appl Environ Microbiol. 1990;56:1992–1998. doi: 10.1128/aem.56.7.1992-1998.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Bauer M, Bauer S, Kelly R. Purification and characterization of a proteasome from the hyperthermophilic archaeon Pyrococcus furiosus. Appl Environ Microbiol. 1997;63:1160–1164. doi: 10.1128/aem.63.3.1160-1164.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Farkas J, et al. Recombinogenic properties of Pyrococcus furiosus strain COM1 enable rapid selection of targeted mutants. Appl Environ Microbiol. 2012;78:4669–4676. doi: 10.1128/AEM.00936-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Lipscomb GL, et al. Natural competence in the hyperthermophilic archaeon Pyrococcus furiosus facilitates genetic manipulation: construction of markerless deletions of genes encoding the two cytoplasmic hydrogenases. Appl Environ Microbiol. 2011;77:2232–2238. doi: 10.1128/AEM.02624-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Schut GJ, Lipscomb GL, Nguyen DMN, Kelly RM, Adams MWW. Heterologous production of an energy-conserving carbon monoxide dehydrogenase complex in the hyperthermophile Pyrococcus furiosus. Front Microbiol. 2016;7:1–9. doi: 10.3389/fmicb.2016.00029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Hawkins AB, et al. Bioprocessing analysis of Pyrococcus furiosus strains engineered for CO2-based 3-hydroxypropionate production. Biotechnol Bioeng. 2015;112:1533–1543. doi: 10.1002/bit.25584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Nguyen DMN, et al. Temperature-dependent acetoin production by Pyrococcus furiosus is catalyzed by a biosynthetic acetolactate synthase and its deletion improves ethanol production. Metab Eng. 2016;34:71–79. doi: 10.1016/j.ymben.2015.12.006. [DOI] [PubMed] [Google Scholar]
- 104.Basen M, Sun J, Adams MWW. Engineering a hyperthermophilic archaeon for temperature-dependent product formation. MBio. 2012;3:1–8. doi: 10.1128/mBio.00053-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Keller MW, et al. A hybrid synthetic pathway for butanol production by a hyperthermophilic microbe. Metab Eng. 2015;27:101–106. doi: 10.1016/j.ymben.2014.11.004. [DOI] [PubMed] [Google Scholar]
- 106.Blumer-Schuette SE, et al. Thermophilic lignocellulose deconstruction. FEMS Microbiol Rev. 2014;38:393–448. doi: 10.1111/1574-6976.12044. [DOI] [PubMed] [Google Scholar]
- 107.VanFossen AL, Ozdemir I, Zelin SL, Kelly RM. Glycoside hydrolase inventory drives plant polysaccharide deconstruction by the extremely thermophilic bacterium Caldicellulosiruptor saccharolyticus. Biotechnol Bioeng. 2011;108:1559–1569. doi: 10.1002/bit.23093. [DOI] [PubMed] [Google Scholar]
- 108.Zurawski JV, et al. Comparative analysis of extremely thermophilic Caldicellulosiruptor species reveals common and unique cellular strategies for plant biomass utilization. Appl Environ Microbiol. 2015;81:7159–7170. doi: 10.1128/AEM.01622-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Blumer-Schuette SE, et al. Caldicellulosiruptor core and pangenomes reveal determinants for noncellulosomal thermophilic deconstruction of plant biomass. J Bacteriol. 2012;194:4015–28. doi: 10.1128/JB.00266-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Blumer-Schuette SE, Lewis DL, Kelly RM. Phylogenetic, microbiological, and glycoside hydrolase diversities within the extremely thermophilic, plant biomass-degrading genus Caldicellulosiruptor. Appl Environ Microbiol. 2010;76:8084–8092. doi: 10.1128/AEM.01400-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Vanfossen AL, Verhaart MRA, Kengen SMW, Kelly RM. Carbohydrate utilization patterns for the extremely thermophilic bacterium Caldicellulosiruptor saccharolyticus reveal broad growth substrate preferences. Appl Environ Microbiol. 2009;75:7718–24. doi: 10.1128/AEM.01959-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Vanfossen AL, Lewis DL, Nichols JD, Kelly RM. Polysaccharide degradation and synthesis by extremely thermophilic anaerobes. Ann N Y Acad Sci. 2008;1125:322–337. doi: 10.1196/annals.1419.017. [DOI] [PubMed] [Google Scholar]
- 113.Bredholt S, Sonne-Hansen J, Nielsen P, Mathrani IM, Ahring BK. Caldicellulosiruptor kristjanssonii sp nov., a cellulolytic, extremely thermophilic, anaerobic bacterium. Int J Syst Bacteriol. 1999;49:991–996. doi: 10.1099/00207713-49-3-991. [DOI] [PubMed] [Google Scholar]
- 114.Bielen AAM, Verhaart MRA, van der Oost J, Kengen SWM. Biohydrogen production by the thermophilic bacterium Caldicellulosiruptor saccharolyticus: current status and perspectives. Life (Basel, Switzerland) 2013;3:52–85. doi: 10.3390/life3010052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.van de Werken HJG, et al. Hydrogenomics of the extremely thermophilic bacterium Caldicellulosiruptor saccharolyticus. Appl Environ Microbiol. 2008;74:6720–6729. doi: 10.1128/AEM.00968-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Zurawski Jeffrey V, Blumer-Schuette Sara E, Conway Jonathan M, Kelly RM. In: Adv In Photosynthesis and Respiration, Volume on Microbial Bioenergy: Hydrogen Production. Zannoni D, De Philippis R, editors. Springer Science+Business Media; 2014. pp. 177–195. [DOI] [Google Scholar]
- 117.Mladenovska Z, Mathrani IM, Ahring BK. Isolation and characterization of Caldicellulosiruptor lactoaceticus sp nov., an extremely thermophilic, cellulolytic, anaerobic bacterium. Arch Microbiol. 1995;163:223–230. [Google Scholar]
- 118.Hamilton-Brehm SD, et al. Caldicellulosiruptor obsidiansis sp nov., an anaerobic, extremely thermophilic, cellulolytic bacterium isolated from Obsidian Pool, Yellowstone National Park. Appl Environ Microbiol. 2010;76:1014–1020. doi: 10.1128/AEM.01903-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.de Vrije T, et al. Glycolytic pathway and hydrogen yield studies of the extreme thermophile Caldicellulosiruptor saccharolyticus. Appl Microbiol Biotechnol. 2007;74:1358–1367. doi: 10.1007/s00253-006-0783-x. [DOI] [PubMed] [Google Scholar]
- 120.van de Werken HJG, et al. Hydrogenomics of the extremely thermophilic bacterium Caldicellulosiruptor saccharolyticus. Appl Environ Microbiol. 2008;74:6720–9. doi: 10.1128/AEM.00968-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Brunecky R, et al. Revealing nature’s cellulase diversity: the digestion mechanism of Caldicellulosiruptor bescii CelA. Science. 2013;342:1513–6. doi: 10.1126/science.1244273. [DOI] [PubMed] [Google Scholar]
- 122.Chung DDD, et al. Homologous expression of the Caldicellulosiruptor bescii CelA reveals that the extracellular protein Is glycosylated. PLoS One. 2015;10:e0119508. doi: 10.1371/journal.pone.0119508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Young J, Chung D, Bomble YJ, Himmel ME, Westpheling J. Deletion of Caldicellulosiruptor bescii CelA reveals its crucial role in the deconstruction of lignocellulosic biomass. Biotechnol Biofuels. 2014;7:142–149. doi: 10.1186/s13068-014-0142-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Ozdemir I, Blumer-Schuette SE, Kelly RM. S-layer homology domain proteins Csac_0678 and Csac_2722 are implicated in plant polysaccharide deconstruction by the extremely thermophilic bacterium Caldicellulosiruptor saccharolyticus. Appl Environ Microbiol. 2012;78:768–777. doi: 10.1128/AEM.07031-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Conway JM, et al. Multidomain, surface layer-associated glycoside hydrolases contribute to plant polysaccharide degradation by Caldicellulosiruptor species. J Biol Chem. 2016;291:6732–6747. doi: 10.1074/jbc.M115.707810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Blumer-Schuette SE, et al. Discrete and structurally unique proteins (tapirins) mediate attachment of extremely thermophilic Caldicellulosiruptor species to cellulose. J Biol Chem. 2015;290:10645–10656. doi: 10.1074/jbc.M115.641480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Basen M, et al. Degradation of high loads of crystalline cellulose and of unpretreated plant biomass by the thermophilic bacterium Caldicellulosiruptor bescii. Bioresour Technol. 2014;152:384–392. doi: 10.1016/j.biortech.2013.11.024. [DOI] [PubMed] [Google Scholar]
- 128.Bing W, et al. Caldicellulosiruptor changbaiensis sp nov., a cellulolytic and hydrogen-producing bacterium from a hot spring. Int J Syst Evol Microbiol. 2015;65:293–7. doi: 10.1099/ijs.0.065441-0. [DOI] [PubMed] [Google Scholar]
- 129.Dam P, et al. Insights into plant biomass conversion from the genome of the anaerobic thermophilic bacterium Caldicellulosiruptor bescii DSM 6725. Nucleic Acids Res. 2011;39:3240–3254. doi: 10.1093/nar/gkq1281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Isern NG, Xue J, Rao JV, Cort JR, Ahring BK. Novel monosaccharide fermentation products in Caldicellulosiruptor saccharolyticus identified using NMR spectroscopy. Biotechnol Biofuels. 2013;6:47. doi: 10.1186/1754-6834-6-47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Miroshnichenko ML, et al. Caldicellulosiruptor kronotskyensis sp nov and Caldicellulosiruptor hydrothermalis sp nov., two extremely thermophilic, cellulolytic, anaerobic bacteria from Kamchatka thermal springs. Int J Syst Evol Microbiol. 2008;58:1492–6. doi: 10.1099/ijs.0.65236-0. [DOI] [PubMed] [Google Scholar]
- 132.Bielen AAM, et al. A thermophile under pressure: transcriptional analysis of the response of Caldicellulosiruptor saccharolyticus to different H2 partial pressures. Int J Hydrogen Energy. 2013;38:1837–1849. [Google Scholar]
- 133.Zeidan AA, Van Niel EWJ. Developing a thermophilic hydrogen-producing co-culture for efficient utilization of mixed sugars. Int J Hydrogen Energy. 2009;34:4524–4528. [Google Scholar]
- 134.van Niel EWJ, et al. Distinctive properties of high hydrogen producing extreme thermophiles, Caldicellulosiruptor saccharolyticus and Thermotoga elfii. Int J Hydrogen Energy. 2002;27:1391–1398. [Google Scholar]
- 135.Willquist K, van Niel EWJ. Lactate formation in Caldicellulosiruptor saccharolyticus is regulated by the energy carriers pyrophosphate and ATP. Metab Eng. 2010;12:282–290. doi: 10.1016/j.ymben.2010.01.001. [DOI] [PubMed] [Google Scholar]
- 136.de Vrije T, et al. Hydrogen production from carrot pulp by the extreme thermophiles Caldicellulosiruptor saccharolyticus and Thermotoga neapolitana. Int J Hydrogen Energy. 2010;35:13206–13213. [Google Scholar]
- 137.Herbel Z, et al. Exploitation of the extremely thermophilic Caldicellulosiruptor saccharolyticus in hydrogen and biogas production from biomasses. Environ Technol. 2010;31:1017–1024. doi: 10.1080/09593330.2010.484075. [DOI] [PubMed] [Google Scholar]
- 138.Ivanova G, Rakhely G, KOVACS K. Hydrogen production from biopolymers by Caldicellulosiruptor saccharolyticus and stabilization of the system by immobilization. Int J Hydrogen Energy. 2008;33:6953–6961. [Google Scholar]
- 139.Ivanova G, Rákhely G, Kovács KL. Thermophilic biohydrogen production from energy plants by Caldicellulosiruptor saccharolyticus and comparison with related studies. Int J Hydrogen Energy. 2009;34:3659–3670. [Google Scholar]
- 140.Kádár Z, et al. Yields from glucose, xylose, and paper sludge hydrolysate during hydrogen production by the extreme thermophile Caldicellulosiruptor saccharolyticus. Appl Biochem Biotechnol. 2004;113–116:497–508. doi: 10.1385/abab:114:1-3:497. [DOI] [PubMed] [Google Scholar]
- 141.Ljunggren M, Willquist K, Zacchi G, van Niel EW. A kinetic model for quantitative evaluation of the effect of hydrogen and osmolarity on hydrogen production by Caldicellulosiruptor saccharolyticus. Biotechnol Biofuels. 2011;4:31. doi: 10.1186/1754-6834-4-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Martinez-Porqueras E, Wechselberger P, Herwig C. Effect of medium composition on biohydrogen production by the extreme thermophilic bacterium Caldicellulosiruptor saccharolyticus. Int J Hydrogen Energy. 2013;38:11756–11764. [Google Scholar]
- 143.Panagiotopoulos IA, Bakker RR, de Vrije T, Koukios EG, Claassen PAM. Pretreatment of sweet sorghum bagasse for hydrogen production by Caldicellulosiruptor saccharolyticus. Int J Hydrogen Energy. 2010;35:7738–7747. [Google Scholar]
- 144.Pawar SS, Vongkumpeang T, Grey C, van Niel EW. Biofilm formation by designed co-cultures of Caldicellulosiruptor species as a means to improve hydrogen productivity. Biotechnol Biofuels. 2015;8:19. doi: 10.1186/s13068-015-0201-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Pawar SS, Nkemka VN, Zeidan AA, Murto M, van Niel EWJ. Biohydrogen production from wheat straw hydrolysate using Caldicellulosiruptor saccharolyticus followed by biogas production in a two-step uncoupled process. Int J Hydrogen Energy. 2013;38:9121–9130. [Google Scholar]
- 146.Talluri S, Raj SM, Christopher LP. Consolidated bioprocessing of untreated switchgrass to hydrogen by the extreme thermophile Caldicellulosiruptor saccharolyticus DSM 8903. Bioresour Technol. 2013;139:272–279. doi: 10.1016/j.biortech.2013.04.005. [DOI] [PubMed] [Google Scholar]
- 147.Van Niel EWJ, Claassen PAM, Stams AJM. Substrate and product inhibition of hydrogen production by the extreme thermophile, Caldicellulosiruptor saccharolyticus. Biotechnol Bioeng. 2003;81:255–262. doi: 10.1002/bit.10463. [DOI] [PubMed] [Google Scholar]
- 148.Van Groenestijn JW, et al. Performance and population analysis of a non-sterile trickle bed reactor inoculated with Caldicellulosiruptor saccharolyticus, a thermophilic hydrogen producer. Biotechnol Bioeng. 2009;102:1361–1367. doi: 10.1002/bit.22185. [DOI] [PubMed] [Google Scholar]
- 149.Zeidan AA, van Niel EWJ. A quantitative analysis of hydrogen production efficiency of the extreme thermophile Caldicellulosiruptor owensensis OLT. Int J Hydrogen Energy. 2010;35:1128–1137. [Google Scholar]
- 150.Thauer RK, Jungermann K, Decker K. Energy conservation in chemotrophic anaerobic bacteria. Bacteriol Rev. 1977;41:100–180. doi: 10.1128/br.41.1.100-180.1977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Willquist K, Pawar SS, Van Niel EWJ. Reassessment of hydrogen tolerance in Caldicellulosiruptor saccharolyticus. Microb Cell Fact. 2011;10:111. doi: 10.1186/1475-2859-10-111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Peters JW, Miller AF, Jones AK, King PW, Adams MWW. Electron bifurcation. Curr Opin Chem Biol. 2016;31:146–152. doi: 10.1016/j.cbpa.2016.03.007. [DOI] [PubMed] [Google Scholar]
- 153.Farkas J, et al. Improved growth media and culture techniques for genetic analysis and assessment of biomass utilization by Caldicellulosiruptor bescii. J Ind Microbiol Biotechnol. 2013;40:41–9. doi: 10.1007/s10295-012-1202-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Chung DD, et al. Construction of a stable replicating shuttle vector for Caldicellulosiruptor species: use for extending genetic methodologies to other members of this genus. PLoS One. 2013;8:e62881. doi: 10.1371/journal.pone.0062881. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Chung D, Farkas J, Huddleston JR, Olivar E, Westpheling J. methylation by a unique α-class N4-cytosine methyltransferase is required for DNA transformation of Caldicellulosiruptor bescii DSM6725. PLoS One. 2012;7:e43844. doi: 10.1371/journal.pone.0043844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Chung D, et al. Overcoming restriction as a barrier to DNA transformation in Caldicellulosiruptor species results in efficient marker replacement. Biotechnol Biofuels. 2013;6:82. doi: 10.1186/1754-6834-6-82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Cha M, Chung D, Elkins JG, Guss AM, Westpheling J. Metabolic engineering of Caldicellulosiruptor bescii yields increased hydrogen production from lignocellulosic biomass. Biotechnol Biofuels. 2013;6:85. doi: 10.1186/1754-6834-6-85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Chung D, Cha M, Guss AM, Westpheling J. Direct conversion of plant biomass to ethanol by engineered Caldicellulosiruptor bescii. Proc Natl Acad Sci U S A. 2014;111:8931–8936. doi: 10.1073/pnas.1402210111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Chung D, et al. Cellulosic ethanol production via consolidated bioprocessing at 75 °C by engineered Caldicellulosiruptor bescii. Biotechnol Biofuels. 2015;8:163. doi: 10.1186/s13068-015-0346-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Cha M, Chung D, Westpheling J. Deletion of a gene cluster for [Ni-Fe] hydrogenase maturation in the anaerobic hyperthermophilic bacterium Caldicellulosiruptor bescii identifies its role in hydrogen metabolism. Appl Microbiol Biotechnol. 2016;100:1823–1831. doi: 10.1007/s00253-015-7025-z. [DOI] [PubMed] [Google Scholar]
- 161.Chung D, et al. Deletion of a gene cluster encoding pectin degrading enzymes in Caldicellulosiruptor bescii reveals an important role for pectin in plant biomass recalcitrance. Biotechnol Biofuels. 2014;7:147. doi: 10.1186/s13068-014-0147-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Bhandari V, Gupta RS. In: The Prokaryotes – Other Major Lineages of Bacteria and the Archaea. Rosenberg E, editor. 2014. pp. 989–1010. [DOI] [Google Scholar]
- 163.Belkin S, Wirsen CO, Jannasch HW. A new sulfur-reducing, extremely thermophilic eubacterium from a submarine thermal vent. Appl Environ Microbiol. 1986;51:1180–1185. doi: 10.1128/aem.51.6.1180-1185.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Huber R, et al. Thermotoga maritima sp nov represents a new genus of unique extremely thermophilic eubacteria growing up to 90C. Arch Microbiol. 1986;144:324–333. [Google Scholar]
- 165.Nelson KE, et al. Evidence for lateral gene transfer between archaea and bacteria from genome sequence of Thermotoga maritima. Nature. 1999;399:323–329. doi: 10.1038/20601. [DOI] [PubMed] [Google Scholar]
- 166.Rodionov Da, et al. Genomic identification and in vitro reconstitution of a complete biosynthetic pathway for the osmolyte di-myo-inositol-phosphate. Proc Natl Acad Sci U S A. 2007;104:4279–4284. doi: 10.1073/pnas.0609279104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Butzin NC, et al. Reconstructed ancestral myo-inositol-3-phosphate synthases indicate that ancestors of the Thermococcales and Thermotoga species were more thermophilic than their descendants. PLoS One. 2013;8:e84300. doi: 10.1371/journal.pone.0084300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Ben Hania W, et al. Cultivation of the first mesophilic representative (‘mesotoga’) within the order Thermotogales. Syst Appl Microbiol. 2011;34:581–585. doi: 10.1016/j.syapm.2011.04.001. [DOI] [PubMed] [Google Scholar]
- 169.Rodionova IA, et al. Novel inositol catabolic pathway in Thermotoga maritima. Environ Microbiol. 2013;15:2254–2266. doi: 10.1111/1462-2920.12096. [DOI] [PubMed] [Google Scholar]
- 170.Conners SB, et al. Microbial biochemistry, physiology, and biotechnology of hyperthermophilic Thermotoga species. FEMS Microbiol Rev. 2006;30:872–905. doi: 10.1111/j.1574-6976.2006.00039.x. [DOI] [PubMed] [Google Scholar]
- 171.Takahata Y, Nishijima M, Hoaki T, Maruyama T. Thermotoga petrophila sp nov and Thermotoga naphthophila sp nov., two hyperthermophilic bacteria from the Kubiki oil reservoir in Niigata, Japan. Int J Syst Evol Microbiol. 2001;51:1901–1909. doi: 10.1099/00207713-51-5-1901. [DOI] [PubMed] [Google Scholar]
- 172.Balk MM, Weijma J, Stams AJM. Thermotoga lettingae sp nov., a novel thermophilic, methanol-degrading bacterium isolated from a thermophilic anaerobic reactor. Int J Syst Evol Microbiol. 2002;52:1361–1368. doi: 10.1099/00207713-52-4-1361. [DOI] [PubMed] [Google Scholar]
- 173.Ravot G, et al. Thermotoga elfii sp nov., a novel thermophilic bacterium from an African oil-producing well. Int J Syst Bacteriol. 1995;45:308–314. doi: 10.1099/00207713-45-2-308. [DOI] [PubMed] [Google Scholar]
- 174.Jeanthon C, et al. Thermotoga subterranea sp nov., a new thermophilic bacterium isolated from a continental oil reservoir. Arch Microbiol. 1995;164:91–97. [PubMed] [Google Scholar]
- 175.Windberger E, Huber R, Trincone A, Fricke H, Stetter KO. Thermotoga thermarum sp nov and Thermotoga neapolitana occurring in African continental solfataric springs. Arch Microbiol. 1989;151:506–512. [Google Scholar]
- 176.Conners SB, et al. An expression-driven approach to the prediction of carbohydrate transport and utilization regulons in the hyperthermophilic bacterium Thermotoga maritima. J Bacteriol. 2005;187:7267–7282. doi: 10.1128/JB.187.21.7267-7282.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Chhabra SR, et al. Carbohydrate-induced differential gene expression patterns in the hyperthermophilic bacterium Thermotoga maritima. J Biol Chem. 2003;278:7540–7552. doi: 10.1074/jbc.M211748200. [DOI] [PubMed] [Google Scholar]
- 178.Frock AD, Notey JS, Kelly RM. The genus Thermotoga: recent developments. Environ Technol. 2010;31:1169–1181. doi: 10.1080/09593330.2010.484076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Schröder C, Selig M, Schönheit P. Glucose fermentation to acetate, CO2 and H2 in the anaerobic hyperthermophilic eubacterium Thermotoga maritima: involvement of the Embden-Meyerhof pathway. Arch Microbiol. 1994;161:460–470. [Google Scholar]
- 180.Selig M, Xavier KB, Santos H, Schönheit P. Comparative analysis of Embden-Meyerhof and Entner-Doudoroff glycolytic pathways in hyperthermophilic archaea and the bacterium Thermotoga. Arch Microbiol. 1997;167:217–32. doi: 10.1007/BF03356097. [DOI] [PubMed] [Google Scholar]
- 181.Frock AD, Gray SR, Kelly RM. Hyperthermophilic Thermotoga species differ with respect to specific carbohydrate transporters and glycoside hydrolases. Appl Environ Microbiol. 2012;78:1978–1986. doi: 10.1128/AEM.07069-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.McCutchen CM, et al. Characterization of extremely thermostable enzymatic breakers (a-1,6-galactosidase and B-1,4-mannanase) from the hyperthermophilic bacterium Thermotoga neapolitana 5068 for hydrolysis of guar gum. Biotechnol Bioeng. 1996;52:332–339. doi: 10.1002/(SICI)1097-0290(19961020)52:2<332::AID-BIT13>3.0.CO;2-L. [DOI] [PubMed] [Google Scholar]
- 183.Blumer-Schuette SE, Kataeva I, Westpheling J, Adams MW, Kelly RM. Extremely thermophilic microorganisms for biomass conversion: status and prospects. Curr Opin Biotechnol. 2008;19:210–217. doi: 10.1016/j.copbio.2008.04.007. [DOI] [PubMed] [Google Scholar]
- 184.Xu H, Han D, Xu Z. Expression of heterologous cellulases in Thermotoga sp Strain RQ2. Biomed Res Int. 2015;2015:304–523. doi: 10.1155/2015/304523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Han D, Norris SM, Xu Z. Construction and transformation of a Thermotoga-E. coli shuttle vector. BMC Biotechnol. 2012;12:2. doi: 10.1186/1472-6750-12-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Chong ML, et al. Biohydrogen production from glycerol using Thermotoga spp. Proc 2001 DOE Hydrog Progr Rev. 2001 doi: 10.1016/j.egypro.2012.09.036. [DOI] [Google Scholar]
- 187.Van Ooteghem SA, Jones A, Van Der Lelie D, Dong B, Mahajan D. H2 production and carbon utilization by Thermotoga neapolitana under anaerobic and microaerobic growth conditions. Biotechnol Lett. 2004;26:1223–1232. doi: 10.1023/B:BILE.0000036602.75427.88. [DOI] [PubMed] [Google Scholar]
- 188.Munro SA, Zinder SH, Walker LP. The fermentation stoichiometry of Thermotoga neapolitana and influence of temperature, oxygen, and pH on hydrogen production. Biotechnol Prog. 2009;25:1035–1042. doi: 10.1002/btpr.201. [DOI] [PubMed] [Google Scholar]
- 189.Schut GJ, Adams MWW. The iron-hydrogenase of Thermotoga maritima utilizes ferredoxin and NADH synergistically: A new perspective on anaerobic hydrogen production. J Bacteriol. 2009;191:4451–4457. doi: 10.1128/JB.01582-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Muralidharan V, Rinker KD, Hirsh IS, Bouwer EJ, Kelly RM. Hydrogen transfer between methanogens and fermentative heterotrophs in hyperthermophilic cocultures. Biotechnol Bioeng. 1997;56:268–278. doi: 10.1002/(SICI)1097-0290(19971105)56:3<268::AID-BIT4>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
- 191.Johnson MR, et al. The Thermotoga maritima phenotype is impacted by syntrophic interaction with Methanococcus jannaschii in hyperthermophilic coculture. Appl Environ Microbiol. 2006;72:811–818. doi: 10.1128/AEM.72.1.811-818.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Fardeau ML, et al. Thennotoga hypogea sp nov., a xylanolytic, thermophilic bacterium from an oil-producing well. Int J Syst Bacteriol. 1997;47:1013–1019. doi: 10.1099/00207713-47-4-1013. [DOI] [PubMed] [Google Scholar]
- 193.Ying X, Wang Y, Badiei HR, Karanassios V, Ma K. Purification and characterization of an iron-containing alcohol dehydrogenase in extremely thermophilic bacterium Thermotoga hypogea. Arch Microbiol. 2007;187:499–510. doi: 10.1007/s00203-007-0217-x. [DOI] [PubMed] [Google Scholar]
- 194.Eram MS, Wong A, Oduaran E, Ma K. Molecular and biochemical characterization of bifunctional pyruvate decarboxylases and pyruvate ferredoxin oxidoreductases from Thermotoga maritima and Thermotoga hypogea. J Biochem. 2015;158:459–466. doi: 10.1093/jb/mvv058. [DOI] [PubMed] [Google Scholar]
- 195.Nguyen TAD, Han SJ, Kim JP, Kim MS, Sim SJ. Hydrogen production of the hyperthermophilic eubacterium, Thermotoga neapolitana under N2 sparging condition. Bioresour Technol. 2010;101:S38–S41. doi: 10.1016/j.biortech.2009.03.041. [DOI] [PubMed] [Google Scholar]
- 196.Menendez C, et al. Presence of acetyl coenzyme A (CoA) carboxylase and propionyl-CoA carboxylase in autotrophic crenarchaeota and indication for operation of a 3-hydroxypropionate cycle in autotrophic carbon fixation. J Bacteriol. 1999;181:1088–1098. doi: 10.1128/jb.181.4.1088-1098.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Teufel R, Kung JW, Kockelkorn D, Alber BE, Fuchs G. 3-Hydroxypropionyl-coenzyme A dehydratase and acryloyl-coenzyme A reductase, enzymes of the autotrophic 3-hydroxypropionate/4-hydroxybutyrate cycle in the Sulfolobales. J Bacteriol. 2009;191:4572–4581. doi: 10.1128/JB.00068-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Ramos-Vera WH, Weiss M, Strittmatter E, Kockelkorn D, Fuchs G. Identification of missing genes and enzymes for autotrophic carbon fixation in Crenarchaeota. J Bacteriol. 2011;193:1201–1211. doi: 10.1128/JB.01156-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Hawkins AS, Han Y, Bennett RK, Adams MWW, Kelly RM. Role of 4-hydroxybutyrate-co a synthetase in the CO2 fixation cycle in thermoacidophilic archaea. J Biol Chem. 2013;288:4012–4022. doi: 10.1074/jbc.M112.413195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Alber BE, Kung JW, Fuchs G. 3-Hydroxypropionyl-coenzyme A synthetase from Metallosphaera sedula, an enzyme involved in autotrophic CO2 fixation. J Bacteriol. 2008;190:1383–1389. doi: 10.1128/JB.01593-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Han Y, Hawkins AS, Adams MWW, Kelly RM. Epimerase (Msed_0639) and mutase (Msed_0638 and Msed_2055) convert (S)-methylmalonyl-coenzyme a (CoA) to succinyl-coA in the Metallosphaera sedula 3-hydroxypropionate/4-hydroxybutyrate cycle. Appl Environ Microbiol. 2012;78:6194–6202. doi: 10.1128/AEM.01312-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Kockelkorn D, Fuchs G. Malonic semialdehyde reductase, succinic semialdehyde reductase, and succinyl-coenzyme A reductase from Metallosphaera sedula: enzymes of the autotrophic 3-hydroxypropionate/4-hydroxybutyrate cycle in Sulfolobales. J Bacteriol. 2009;191:6352–6362. doi: 10.1128/JB.00794-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Oshima T, Imahori K. Description of Thermus thermophilus (Yoshida and Oshima) comb. nov., a nonsporulating thermophilic bacterium from a Japanese thermal spa. Int J Syst Bacteriol. 1974;24:102–112. [Google Scholar]
- 204.Lee NRR, et al. Genome-scale metabolic network reconstruction and in silico flux analysis of the thermophilic bacterium Thermus thermophilus HB27. Microb Cell Fact. 2014;13:61–73. doi: 10.1186/1475-2859-13-61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Swarup A, Lu J, DeWoody KC, Antoniewicz MR. Metabolic network reconstruction, growth characterization and 13C-metabolic flux analysis of the extremophile Thermus thermophilus HB8. Metabolic Eng. 2014;24:173–180. doi: 10.1016/j.ymben.2014.05.013. [DOI] [PubMed] [Google Scholar]
- 206.Koyama Y, Hoshino T, Tomizuka N, Furukawa K. Genetic Transformation of the Extreme Thermophile Thermus thermophilus and of Other Thermus spp. J Bacteriol. 1986;166:338–340. doi: 10.1128/jb.166.1.338-340.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Sazanov La, Hinchliffe P. Structure of the hydrophilic domain of respiratory complex I from Thermus thermophilus. Science. 2006;311:1430–1436. doi: 10.1126/science.1123809. [DOI] [PubMed] [Google Scholar]
- 208.Yokoyama K, et al. V-Type H-ATPase/Synthase from a Thermophilic Eubacterium, Thermus thermophilus. Biochemistry. 2000;275:13955–13961. doi: 10.1074/jbc.275.18.13955. [DOI] [PubMed] [Google Scholar]
- 209.Yusupov MM, et al. Crystal structure of the ribosome at 5.5 A resolution. Science. 2001;292:883–896. doi: 10.1126/science.1060089. [DOI] [PubMed] [Google Scholar]
- 210.Selmer M, et al. Structure of the 70S ribosome complexed with mRNA and tRNA. 2006;313:1935–1942. doi: 10.1126/science.1131127. [DOI] [PubMed] [Google Scholar]
- 211.Krutsakorn B, et al. In vitro production of n-butanol from glucose. Metab Eng. 2013;20:84–91. doi: 10.1016/j.ymben.2013.09.006. [DOI] [PubMed] [Google Scholar]
- 212.Karhumaa K, Hahn-Hägerdal B, Gorwa-Grauslund MF. Investigation of limiting metabolic steps in the utilization of xylose by recombinant Saccharomyces cerevisiae using metabolic engineering. Yeast. 2005;22:359–368. doi: 10.1002/yea.1216. [DOI] [PubMed] [Google Scholar]
- 213.Ramírez-Arcos S, Fernández-Herrero LA, Marín I, Berenguer J. Anaerobic growth, a property horizontally transferred by an Hfr-like mechanism among extreme thermophiles. J Bacteriol. 1998;180:3137–3143. doi: 10.1128/jb.180.12.3137-3143.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Moreno R, et al. High-Level Overproduction of His-Tagged Tth DNA Polymerase in Thermus thermophilus High-Level Overproduction of His-Tagged Tth DNA Polymerase in Thermus thermophilus. 2005;71:1–4. doi: 10.1128/AEM.71.1.591-593.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Hidalgo A, et al. Thermus thermophilus as a cell factory for the production of a thermophilic Mn-dependent catalase which fails to be synthesized in an active form in Escherichia coli. Appl Environ Microbiol. 2004;70:3839–3844. doi: 10.1128/AEM.70.7.3839-3844.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]





