Citation: Davide Sala, Andrea Giachetti, Antonio Rosato. Molecular dynamics simulations of metalloproteins: A folding study of rubredoxin from Pyrococcus furiosus[J]. AIMS Biophysics, 2018, 5(1): 77-96. doi: 10.3934/biophy.2018.1.77
[1] | Thatiane Rodrigues Mota, Dyoni Matias de Oliveira, Rogério Marchiosi, Osvaldo Ferrarese-Filho, Wanderley Dantas dos Santos . Plant cell wall composition and enzymatic deconstruction. AIMS Bioengineering, 2018, 5(1): 63-77. doi: 10.3934/bioeng.2018.1.63 |
[2] | Ronald C. Sims, Sean K. Bedingfield, Reese Thompson, Judith L. Sims . Bioenergy from wastewater-based biomass. AIMS Bioengineering, 2016, 3(1): 103-124. doi: 10.3934/bioeng.2016.1.103 |
[3] | Egidio Viola, Francesco Zimbardi, Vito Valerio, Antonio Villone . Effect of Ripeness and Drying Process on Sugar and Ethanol Production from Giant Reed (Arundo donax L.). AIMS Bioengineering, 2015, 2(2): 29-39. doi: 10.3934/bioeng.2015.2.29 |
[4] | Adam B Fisher, Stephen S Fong . Lignin biodegradation and industrial implications. AIMS Bioengineering, 2014, 1(2): 92-112. doi: 10.3934/bioeng.2014.2.92 |
[5] | Zongyuan Zhu, Rachael Simister, Susannah Bird, Simon J. McQueen-Mason, Leonardo D. Gomez, Duncan J. Macquarrie . Microwave assisted acid and alkali pretreatment of Miscanthus biomass for biorefineries. AIMS Bioengineering, 2015, 2(4): 449-468. doi: 10.3934/bioeng.2015.4.449 |
[6] | Liwei Chen, Jaslyn Lee, Wei Ning Chen . The use of metabolic engineering to produce fatty acid-derived biofuel and chemicals in Saccharomyces cerevisiae: a review. AIMS Bioengineering, 2016, 3(4): 468-492. doi: 10.3934/bioeng.2016.4.468 |
[7] | Nhuan P. Nghiem, Justin Montanti, David B. Johnston . Sorghum as a renewable feedstock for production of fuels and industrial chemicals. AIMS Bioengineering, 2016, 3(1): 75-91. doi: 10.3934/bioeng.2016.1.75 |
[8] | Chiara Schiraldi, Alberto D'Avino, Alessandro Ruggiero, Katia Della Corte, Mario De Rosa . Saccharomyces pastorianus as cell factory to improve production of fructose 1,6-diphosphate using novel fermentation strategies. AIMS Bioengineering, 2015, 2(3): 206-221. doi: 10.3934/bioeng.2015.3.206 |
[9] | Noaa Frederick, Mengxing Li, Danielle Julie Carrier, Michael D. Buser, Mark R. Wilkins . Switchgrass storage effects on the recovery of carbohydrates after liquid hot water pretreatment and enzymatic hydrolysis. AIMS Bioengineering, 2016, 3(3): 389-399. doi: 10.3934/bioeng.2016.3.389 |
[10] | Siseon Lee, Robert J. Mitchell . Perspectives on the use of transcriptomics to advance biofuels. AIMS Bioengineering, 2015, 2(4): 487-506. doi: 10.3934/bioeng.2015.4.487 |
Biomass gasification-fermentation is a two-step thermochemical-biological conversion process that can be used to produce fuels and chemicals from renewable feedstocks. In the first step, renewable feedstocks such as dedicated energy crops, crop residues, wood, and municipal solid waste are converted to primarily carbon monoxide (CO), hydrogen (H2) and carbon dioxide (CO2) through gasification using a limited amount of oxidizing agent (such as air, oxygen and/or steam) at high temperatures that typically range from 600 to 1,000 °C [1,2,3,4]. The extreme gasification environment breaks down organic compounds into a stable gaseous mixture commonly referred as synthesis gas (syngas). Syngas is composed of CO, H2, and CO2 as well as impurities, such as nitrogen (N2), methane (CH4), ethane (C2H6), ethylene (C2H4) and acetylene (C2H2), tars, ammonia (NH3) and hydrogen sulfide (H2S), and particulates. In the second step, syngas is fermented by anaerobic microbes, i.e., acetogens, such as Clostridium ljungdahlii, Clostridium autoethanolgenum, Eubacterium limosum (aka Butryibacterium methylotrophicum), Clostridium carboxidivorans, Clostridium ragsdalei, Clostridium species MAceT113 and Alkalibaculum bacchi to form acids (acetic acid, butyric acid, hexanoic acid), alcohols (ethanol, isopropanol, butanol, hexanol), and specialty chemicals (2, 3-butanediol, acetone, fatty acids, fatty alcohols) [5,6,7,8,9,10,11,12,13,14]. This process is referred to as syngas fermentation.
Biomass gasification-fermentation offers several major advantages compared to other processes that have been proposed for production of fuels and chemicals from renewable feedstocks.
1. Gasification utilizes lignin and all other organic molecules present in the process feedstock to produce syngas; hence, theoretically, all organic parts of the feedstock can be converted into fuels and chemicals during fermentation. Other processes that use renewable feedstocks, such fermentations of sugars from either sugar crops (e.g., sugar cane, sugar beets and sweet sorghum) or enzymatic hydrolysis of starch, cellulose and/or hemicellulose, can only use the carbohydrates in the feedstock for fuel and chemical production. Lignin makes up between 17 and 30% of the organic portion of biomass [15].
2. Conversion of syngas through fermentation takes place at lower temperatures and pressures than chemical catalytic conversions of syngas; thus, specially designed reactors that are resistant to high temperatures and pressures are not required to conduct syngas fermentation [12]. This reduces capital and operational costs.
3. Microbial fermentations produce fewer by-products than chemical catalyst conversion; hence, fermentation has higher product specificities [16].
4. Acetogens involved in syngas fermentation are able to operate under flexible CO:H2:CO2 ratios; thus, syngas reforming is not required to match fermentation conditions
5. Gaseous substrates, such as CO and H2, allow uncoupling of hydraulic retention time with the supply of substrate, thus offering a better control of substrate and product inhibition [12].
6. Hazardous or xenobiotic products are not formed in the fermentation process [16].
7. The mesophilic incubation temperatures, low media carbohydrate levels, low operating pH and high CO concentrations are inhibitory to many classes of contaminating microorganisms, including methanogens [4].
Despite these advantages, several challenges emerge while integrating biomass gasification with syngas fermentation. These include high temperatures (600-1000 °C) for gasification and the accompanying costs to achieve these temperatures, gas-liquid mass transfer, reactor design, low cell mass concentrations, low product yields, slow conversion rates, low stability of microbial catalysts, high cost of fermentation media, and inhibitory effects of gas impurities [17,18,19,20]. Several studies have focused on understanding the microbial physiology of syngas fermenting microbes, and optimizing and maximizing syngas fermentation [21,22,23,24]. However, few studies have focused on the integrating of the two technologies. The following review identifies and addresses the underlying research gaps and challenges, thus enabling a holistic approach for designing and commercializing biomass gasification-fermentation process.
The conversion of biomass to syngas occurs through four steps, namely: drying, devolatilization, oxidation and reduction. Drying releases moisture from biomass that further reacts to form H2. Devolatilization occurs at temperatures between 225 and 500°C and releases volatile compounds from biomass [25]. Oxidation is the next step wherein the devolatilized products (i.e., char, tar and gases) react with an oxidizing agent yielding essential heat for endothermic reactions. Reduction is the final gasification step, which reduces the hot combustion products (produced during oxidation) mainly to CO and H2.
Gasifiers are classified based on the availability of heat to drive the reactions (direct or indirect) and on the type of bed. Direct gasification involves use of air and/or oxygen (O2) as an oxidizing agent, which reacts with biomass within the gasifier and produces heat energy needed to drive endothermic gasification reactions. In indirect gasification, the gasifier is externally heated to supply essential heat for gasification reactions. Direct gasification using air (i.e., air-gasification) results in N2 diluted (50-65%) syngas with low H2 and CO concentrations while indirect gasification, which typically uses steam or oxygen as an oxidizing agent, yields syngas with higher H2 and CO concentrations. Detailed discussion on gasifier types based on the bed configuration can be found in other literature [2,3].
Biomass, such as dedicated energy crops, agricultural and forestry residues, and municipal solid waste, can be gasified to generate syngas. Biomass exhibits higher reactivity and require a lower gasification temperature than coal, which historically has been used as a gasifier feedstock to generate syngas [2]. However, biomass generally contains less carbon (39-52%) than coal (68-92%) [25,26], consequently, less carbon is available for conversion into carbon-based gaseous products in the syngas.
Feedstock properties(such as size, shape, density, chemical composition, energy content and reactivity) and gasification conditions (such as reactor type, temperature, pressure, type and flow of oxidizing agent, biomass flow, and catalyst used) have considerable influence on syngas composition and impurities [3,27]. Table 1 lists syngas composition and impurities from gasification of a variety of feedstocks as a function of gasifier reactor type, biomass feedstock and oxidizing agent.
The overall gasification reaction using air and/or steam is represented in Figure 1. The formations of char and tar are usually due to incomplete conversion of biomass into syngas [3].
Syngas contains several undesirable impurities such as tars (defined as benzene and other condensable aromatic hydrocarbons of molecular weight higher than benzene), particulate matter (pure carbon and ash), inorganic compounds (ammonia, hydrogen cyanide, hydrogen sulfide, carbonyl sulfide, and nitric oxides), and light hydrocarbons (methane, acetylene, ethylene, and ethane) [25]. Syngas impurities can profoundly affect gas utilization technology downstream of the gasifier [3,25]. Handling and removal of tars have been one of the major challenges during process scale-up, as tars tend to condense at low temperature and/or form aerosols that affect engine and turbine performance [3,25]. The condensed tars easily mix with particulates and accumulate on inside surfaces resulting in the clogging and fouling of pipes and equipment [3,25]. Cleaning and reforming raw syngas is a major downstream process for effective utilization of syngas by subsequent processes [2,3]. Tar content in syngas varies depending on gasifier type, gasification time-temperature combination, bed geometry and type of oxidizing agent used [3,25].
For successful integration of biomass gasification and syngas fermentation processes, it is important to identify the key factors in gasification that can have a potential impact on fermentation. Some of these factors are the carbon conversion efficiency, syngas impurities, H2/CO ratio and usability of CO2 in the presence of CO.
Carbon conversion efficiency (CCEG) quantifies the effectiveness of a gasification process and is defined as the ratio between the mass flow rate of carbon in the syngas to mass flow rate of carbon in the feedstock [25]. In general, the CCEG for a gasifier can vary between 60 and 99% [25], which means that 60-99% of the carbon present in the feedstock can be converted into carbon-containing gases. Carbon loss is mostly in the form of tars, char and volatile hydrocarbons. These compounds that represent carbon that is unavailable for subsequent conversion to alcohols during fermentation; therefore, conversion of these impurities to syngas CO and H2 is desirable so that they are available for fermentation. Normally, carbon in the form of volatile hydrocarbons, such as CH4 and C2H2, are accounted for in the calculation of CCEG along with CO and CO2; however, volatile hydrocarbons are not fermentable by syngas fermenting microorganisms. Hence, a modified CCEG that accounts for CO and CO2 as the only fermentable carbon is more appropriate for carbon balance calculations involving syngas fermentation. In one study, a lab scale fluidized bed air gasification system achieved a CCEG of 89% while the modified CCEG for syngas fermentation was 72% [27].
Increasing the gasifier operating temperature converts tars into CO, CO2 and H2, thereby, increasing the CCEG [3]. However, the operating costs also increase as temperature increases. An effect similar to increasing temperature can be obtained by reducing tar content with an in-bed catalyst such as dolomite and limestone used during fluidized bed gasification [3]. CCEG can also be controlled by the equivalence ratio (ER) [35], which is the ratio of actual air supplied to the stoichiometric air required for complete combustion of biomass. Higher ERs result in higher operating temperatures, which in turn increases the degree of oxidation of available carbon and thus increases CCEG [27]. However, higher ERs also result in more CO2; hence, optimum ER, which results in more CO and H2 [27], is more suitable for producing syngas for fermentation. Lv et al. [36] showed that an increase in ER from 0.19 to 0.23 in an air-steam gasification system increased CCEG by 19%. Decrease in the size of biomass particles can also result in a higher CCEG because heating rates increased with an increase in the surface area of biomass particles [36].
Cold gas efficiency (CGE) is another energy efficiency calculated for gasifiers. CGE is defined as the ratio of chemical energy of syngas to the total energy throughput to the gasifier [25]. Higher CGE results in a higher chemical energy, also called heating value, of the syngas. Higher CGE represents higher CO and H2 concentrations in the syngas, which are desirable for fermentation.
Although CCEG is generally high in gasification, the presence of impurities (compounds other than CO, CO2 and H2), such as tars, inorganic impurities and light hydrocarbons, can reduce fermentability of syngas [30,37,38]. These species can affect growth, enzyme kinetics and/or alter growth media by changing the osmolarity, pH and/or redox potential [38]. Microbial catalysts can tolerate sulfides (such as carbonyl sulfide, hydrogen sulfide and sulfur dioxide), chlorine compounds and tars [4,39]; however, the effects of these contaminants on alcohol production and cell growth have not been thoroughly studied. Understanding of gasification parameters that cause production of these contaminants and effects of these contaminants on fermentation are critical for design and development of an economically viable gasification-fermentation process. A limited number of studies have examined the effects of syngas impurities on syngas fermentation microorganisms and have shown both positive and negative effects of various impurities on growth and product distribution [30,40,41,42]. Metabolically, many syngas impurities deactivate key enzymes involved in the Wood-Ljungdahl pathway such as carbon monoxide dehydrogenase (CODH), hydrogenase, and alcohol dehydrogenase [37,38,43]. Table 2 lists the highest concentrations of various impurities reported in syngas in conjunction with microbial catalysts and their tolerance ceiling. Microbes demonstrate varied responses to impurities, so generalization of such responses is not warranted.
Several studies have compared the effects of syngas produced from biomass gasification and syngas produced from mixing pure gases on cell growth and ethanol and acetic acid production. Datar et al. [40] fermented biomass-generated syngas from a fluidized bed gasifier. Researchers observed growth inhibition of C. carboxidivorans in a bioreactor. Growth resumed when biomass-generated syngas was substituted by clean bottled gas. Another study observed that C. ragsdalei adapted to biomass-generated syngas exhibited a 133% increase in ethanol concentration and 500% decrease in acetic acid concentration in comparison to cells that were grown on pure gas mixes [41]. Cultural adaptation to impurities prior to large scale cultivations has been shown to increase ethanol to acetic acid ratios.
![]() |
Carbonaceous impurities originate from carbon, hydrogen and oxygen in the feedstock that could not be converted into CO, CO2 and H2 during gasification and include lighter hydrocarbons, such as CH4, acetylene (C2H2), ethylene (C2H4), and ethane (C2H6), and tars. CH4 is a major component of syngas, which is formed through the methanation reaction during gasification.
The formation of CH4 could be important for the overall gasification process and cannot be completely avoided as it is an exothermic process and the heat evolved by this reaction supports other endothermic reactions. The highest concentration of CH4 in syngas reported in literature is 14.9% using air-steam co-gasification of coal and plastic waste mixtures in a fluidized bed gasifier [50]. In a previous study, syngas with a CH4 concentration of 4.5% was fed to C. carboxidivorans [40]. CH4 did not affect gas utilization or growth, which confirmed the inert nature of CH4 in syngas fermentation. Ahmed et al. [30] reported that syngas containing C2H2 (0.1% v/v), C2H4 (1.4% v/v) and C2H6 (0.35% v/v) did not affect growth during syngas fermentations with C. carboxidivorans.
In general, syngas-fermenting microbes have many enzymes similar to nitrogen fixing bacteria and methanogens [51,52] and are also found in close connection to methanogens and sulfate reducing bacteria in their natural environment [53]. Therefore, some understanding of how syngas-fermenting microbes could respond to volatile hydrocarbons in syngas can be obtained by reviewing literature that investigated the effects of volatile hydrocarbons on these similar microbes. For instance, in Azotobacter vinelandii, C2H2 was found to be a slow binding, active site directed reversible inhibitor of nickel-ferrous hydrogenase involved in catalyzing the oxidation of H2 [52]. C2H2 also inhibits many methanogenic microorganisms by interfering with membrane linked proton movement and ATP synthesis [54]. However, C2H2 fermenting anaerobic microorganisms have been discovered [55]. Pelobacter acetylenicus, a C2H2 fermenting microbe has a tungsten-iron-sulfur protein called acetylene hydratase that catalyzes a net hydration reaction rather than a redox reaction producing ethanol and acetate as major fermentation products through acetaldehyde as a metabolic intermediate [55].
${C_2}{H_2} + {\rm{ }}{H_2}O \to C{H_3}CHO \Delta G{\rm{ }} = {\rm{ }} - 111.9{\rm{ }}KJ/mole$ | (1) |
Like C2H2, C2H4 also has an inhibitory effect on methanogens [54]. About 50% of CH4 formation was inhibited at 0.07% C2H4 and as C2H4 concentration increased to 0.5%, more than 94% CH4 formation was inhibited, but these inhibitions were found to be reversible [56]. In a different study, Oremland and Taylor [57] showed inhibition of CH4 formation by C2H2 and C2H4, but, C2H6 was found to be an inert gas for fermentation.
A review by Rabou et al. [58] classified tars based on physical properties into five classes: 1) GC undetectable, which are 7 carbon and higher ring compounds, 2) heterocyclic such as phenol, cresol and pyridine, 3) light aromatic such as toluene, styrene and xylene, 4) light poly-aromatic such as naphthalene, phenanthrene and anthracene, and 5) heavy poly-aromatic such as fluoroanthene, pyrene, chrysene, perylene and benzoperylene. Usually, the concentration of individual components in tars is of less practical use [58]; thus, the total tar content produced during gasification is usually determined gravimetrically by dissolving tar in acetone solution at low temperatures. Benzene and toluene have been found to be two predominant tar compounds in syngas produced from gasification of switchgrass [59]. Xu et al. [38] reported Henry’s Law constants for common syngas impurities. Henry’s Law constants give an idea about the solubility of gaseous components in aqueous media. Henry’s Law constants for naphthalene and benzene were observed to be 1,000 and 10,000 fold higher than those for CO and H2, respectively. Due to such high solubilities in aqueous media, these compounds could impart a significant impact during fermentations despite their low concentrations in syngas [38].
Ahmed et al. [30] used a 0.025 m filter to remove particulate matter and some tars from biomass-generated syngas to overcome cell dormancy. The filter removed the nitric oxide (NO) which was observed to be the potential inhibitor of hydrogenase activity responsible for reducing the amount of carbon and H2 available for product formation. The study further tested the effect of supplementing fermentation media with 1 ml of acetone solution containing tars. Initial cell growth was inhibited by the addition of tars; however, eventually the cells did grow, possibly due to microbial adaptation. This study corroborated previous observations and showed that tars inhibited cellular growth, but did not cause cell death. Also, tars promoted ethanol production and inhibited acetic acid production, thus increasing the ethanol to acetate ratio [30].
Thermal cracking of tars and light hydrocarbons into permanent gases such as CO and H2, normally occurs at temperatures above 1,000 °C, but biomass gasification generally occurs at lower temperatures [2]. Several syngas tar and light hydrocarbon removal technologies such as wet scrubbing, dry scrubbing, dry-wet scrubbing and hot gas catalytic conditioning have been commercialized [2]. Scrubbing techniques are cost effective, but often generate a stream of contaminated water that may require extensive biological or thermal treatments [3]. Catalytic conversions convert tars into CO and H2, thereby improving the overall yield of syngas in gasification [3]. Nickel, dolomite, and limestone based catalysts can also be used in fixed-bed applications (downdraft and updraft), but are not robust for fluidized bed reactor systems [25]. Commercial steam reforming and naptha cracking catalysts have been shown to successfully crack tars in a secondary reactor downstream from the gasifier; however, challenges with coking and economic viability due to their high costs still remain [60,61]. Biochar (byproduct of biomass gasification) and biochar-supported catalysts have also shown to reduce significantly syngas tars [62,63,64]. Hydrocarbons can also be removed by Rectisol wash (acid-gas removal process) that employs methanol as solvent and removes a number of species such as H2S, NH3 and HCN with concomitant separation of CO2 [26]. Removal technologies for hydrocarbons and tars have been reviewed elsewhere [3,25,58].
Nitrogenous species such as ammonia (NH3), hydrogen cyanide (HCN) and nitric oxides (NOx) are commonly present in syngas. These species are formed predominantly from the nitrogen contained in the biomass [26] and/or from nitrogen in char produced during gasification [65]. The environment in a gasifier can have both reduction zones and oxidation zones. Under reduced condition in the pyrolytic zone of the gasifier, N-H and N-C bonds in the biomass structure are converted into HCN, which is subsequently reduced to ammonia [26,65]. The proportions of these two intermediates formed depend on the type of gasifier and gasifer conditions [65]. For example, when H2 concentrations are available in surplus, nitrogen that is predominantly available as NH3 and HCN does not undergo further decompositon. Further, it has been noted that concentrations of NH3 and HCN depend on feedstock nitrogen compositon [25]. When a gasifier has excess O2 in the oxidation zone, a part of ammonia gets transformed to nitric oxide (NO) or nitrous oxide (N2O). The oxidation of ammonia to N2O is more favorable, but is not a common occurrence in gasification because N2O gets reduced to NO by the time it exits the gasifier [65].
$Nitrogeninbiomass \to HCN\mathop \to \limits^{{H_2}O} NH3\matrix{ {\mathop \to \limits^{NO} } \cr {\mathop \to \limits_{{O_2}} } \cr } \matrix{ {{N_2}} \cr {NO} \cr } $ | (2) |
The presence of chlorine compounds in gas, such as HCl, can result in the formation of ammonium chloride (NH4Cl), increase osmolarity of fermentation media and inhibit hydrogenase in enzymatic systems above certain levels [43].
NH3 concentrations ranging from 0.5 to 1% in syngas have been reported [31]. NH3 is highly soluble in water forming ammonium ions (NH4+). NH4+ can serve as a nitrogen source, but excess NH4+ is inhibitory to many syngas fermenting bacteria [38]. A very strong inhibition has been observed above 2.7 gL-1 NH4+in C. ragsdalei [38] which was due to non-competitive inhibition of hydrogenase enzymes [43].
Effect of NO on C. carboxidivorans was studied by Ahmed and Lewis [37]. They observed that concentrations of NO above 40 ppm acted as a non-competitive, reversible inhibitor to hydrogenase. At these concentrations, 5-7 times increase in ethanol production was also observed due to the increased activity of alcohol dehydrogenase. However, at NO concentrations of 140-150 ppm, H2 consumption was completely shut down. A similar metabolic shift was also observed with C. ragsdalei when biomass-generated syngas from a fluidized bed gasifier with a NO concentration of 44.5 ppm was used as a fermentation substrate [41]. In comparison to pure bottled syngas, a 133% increase in ethanol and a 200% decrease in acetic acid concentrations were observed.
NH3 and HCN have high solubility in water and can be removed by water wash systems, such as Rectisol wash that employs methanol as a solvent at low temperatures [26]. NOx and N2O gases can be removed by non-selective catalytic reduction (NSCR), selective catalytic reduction and catalytic decomposition of N2O into N2 and O2 [66,67,68].
Sulfur compounds such as hydrogen sulfide (H2S), carbonyl sulfide (COS) originate from sulfur in the feedstock. Gasification of coal generally results in higher sulfur gas concentration compared to biomass gasification due to higher sulfur content of coal (0.4-4.9%) versus biomass such as switchgrass (0.2%) [26,69]. Among all sulfur gases, H2S is the dominating species and can account for 93-96% of the total sulfur content in syngas with the balance accounted for by COS [26,70].
Chemical catalysts are less tolerant to sulfur than are microbial catalysts as certain chemical catalysts are inactivated at 0.1 ppm of sulfur in the producer gas [4]. Methanogens, hydrogenic bacteria (bacteria that perform the water gas shift reaction, which includes syngas fermentation bacteria) and acetogens have demonstrated tolerances of up to 20% sulfur in feed gases (Table 2) [42]. P. productus, E. limosum and R. rubrum were shown to degrade COS into H2S using the following reaction scheme, while C. ljungdahlii cells were unable to metabolize or degrade COS using the following reaction [71]:
$COS{\rm{ }} + {\rm{ }}{H_2}O \to {H_2}S{\rm{ }} + {\rm{ }}C{O_2}$ | (3) |
Sulfides also serve to control redox potential, remove residual O2 and act as a sulfur source for microbial metabolism [72]. Hu et al. [72] observed that mass transfer played an important role in sulfide loss from continuous bioreactors while equilibrium between the sulfide level in gaseous phase and aqueous phase played an essential role in small scale reactors. Researchers also observed that increasing sulfide concentrations from 0 to 1.9 mM in small scale batch cultivations of C. ragsdalei increased ethanol concentration from 0.9 to 2.1 g/L and decreased acetic acid concentration from 5.2 to 3.6 g/L.
Even though microbes show a higher tolerance to sulfur gases compared to chemical catalysts, it is recommended to test the effect of sulfur gases before drawing any broad conclusions from the limited studies. Removal of sulfur compounds (desulfurization or acid gas removal) can be achieved by adsorption, absorption, diffusion or chemical conversion [26]. The technique selected will depend on gas purity requirements, raw gas composition, selectivity of the process and economic considerations [26]. It is also important to note that amine absorption systems commonly used for desulfurization of gas also remove CO2, which can be used as a carbon source by syngas fermentation bacteria.
The presence of O2 and acetone in syngas has been reported in literature [9,48]. O2 is usually completely consumed during gasification; thus, its presence is usually due to downstream operations such as compressors and wash water [26]. Likewise, acetone is introduced into the syngas during wet scrubbing of gases to remove tar species [9,30,40].
O2 is one of the most toxic gases during syngas fermentation as it inhibits several enzymes such as pyruvate-ferredoxin oxidoreductase hydrogenases (including Ni- and Fe- containing) present in the Wood-Ljungdahl pathway. Its presence in fermentation media can also increase redox potential [22,73]. However, some acetogens, like A. woodii, C. magnum, C. glycolicum RD-1 and M. thermoacetica, are tolerant to O2 concentrations ranging from 0.5% to 6% [74,75]. C. glycolicum RD-1 is one of the most interesting acetogens because it can withstand up to 6% O2 and produce ethanol, lactate and H2 as end products [49]. Under anoxic conditions, acetogenic metabolism shifts towards enzyme pathways that are less sensitive to O2, resulting in production of acetic acid and ethanol [49,74]. Furthermore, a few acetogens, such as A. woodii, C. magnum, S. silvatica, M. thermoacetica andC. glycolicum RD-1, contain enzymes, such as reduced NADH-oxidase, superoxide dismutase, rubredoxin oxidoreductase (superoxide reductase), rubrerythrin (a peroxidase), catalase and cytochrome ba oxidase, that are involved in the removal of O2 or its toxic products [49,74,75].
C. carboxidivorans has been shown to tolerate O2 concentrations up to 1,900 ppm (or 0.19%) without any impact on CO and H2 utilization, growth or product formation [40]. Kundiyana et al. [48] showed that pilot scale fermentation of syngas using C. ragsdalei was unaltered when O2 concentrations between 0.04 to 2.6% were observed in syngas from a downdraft gasifier. For the integration of the gasification-syngas fermentation process, it is essential that the level of tolerance to O2 be determined for different syngas fermenting microbes.
Acetone is another compound whose presence in syngas is reported in literature. Ramachandriya et al. [9] reported that acetone concentration of 0.4% in syngas did not affect C. ragsdalei growth, but product formation was redistributed to produce isopropanol as one of the end products. Production of isopropanol is unique to the acetogenic fermentation systems and alludes further to the presence of a secondary alcohol dehydrogenase enzyme [9].
Due to the high solubility of acetone in water, any carryover of acetone from a water-acetone scrubbing system will result in its accumulation in fermentation broth; thus, the scrubbing system for tars must be carefully designed so that it does not affect the fermentation process. Any wash system must use de-aerated water to eliminate introduction of O2 in syngas [26]. O2 can be removed from syngas by installing copper scrubbers, but care has to be taken due to the presence of C2H2 in syngas as it reacts with copper to form the explosive copper acetylide [76]. Oxygen from gas streams can also be removed by using other metal based catalysts such as platinum and palladium catalysts supported on alumina [77].
From the above discussion, it is clear that that many minor gaseous components introduced during gasification and syngas processing can have both positive and negative effects on syngas fermenting bacteria. In general, the tolerance level of microbes to sulfur compounds is better than chemical catalysts and microbial adaptation to gaseous components is an important strategy to overcome any inhibitory effects. However, it is important to underscore that different microbes behave differently in the presence of these gaseous components. Understanding how each of these gaseous components interacts with the microbial system will help identify problems and better optimize a gasification-syngas fermentation process.
One of the major challenges developing syngas thermochemical conversion processes is producing syngas with fixed H2 to CO ratios for a process utilizing syngas as a substrate. For instance, catalytic synthesis of methanol, mixed alcohol synthesis and hydrocarbon fuel synthesis by the Fisher-Tropsch process require the H2 to CO ratio to be around 2, 1.1 and 0.6, respectively [29,78], which imposes additional reforming steps and increases overall process cost. On the contrary, microbial fermentations of syngas utilize different ratios of CO, H2 and CO2. Reactions for formation of ethanol and acetate with corresponding Gibbs free energies (ΔG) at 298 K can be found in Table 3.
For syngas fermentations, efficiency is measured as a ratio between experimental yields (moles or grams of product that has been produced per mole or gram of substrate consumed) and the theoretical yield (moles or grams of product that can be produced theoretically per mole or gram of substrate consumed). This ratio is commonly referred as the % of theoretical yield of gaseous substrate into products. This is different from percentage gas utilization, which is simply the ratio between the amount of gas consumed by the microbe to the amount of gas supplied. Gas utilization can be improved with better reactor designs and/or by recycling unused gases; however, fermentation efficiency is increased if the microbial catalyst is highly specific to one product.
The fermentation efficiencies of various syngas fermenting microbes for ethanol production have not been widely reported. The fermentation efficiency for ethanol production using C. ragsdalei at a H2 to CO ratio of 0.25 ranged from 12.9 to 60% [8,9], while that for C. carboxidivorans ranged from 45.1 to 54.9% [9]. Generally, direct gasification of biomass using a fluidized bed gasifier provides a H2 to CO ratio between 0.25 and 0.53 (Table 1). Rajagopalan et al. [79] reported 55% fermentation efficiency for ethanol production with C. carboxidivorans grown on 25% CO, 15% CO2 and 60% N2 mixtures. C. ljungdahlii when grown on 80% CO, 20% CO2 gas mix showed a fermentation efficiency of 71.9% for ethanol production [80]. A patent from BRI Energy (now INEOS Bio) reported production of 30 g/L ethanol with 96% conversion of CO and 60% conversion of H2 to ethanol and a productivity of 38 g/L*d using a 2 stage continuous stirred tank reactor (CSTR) [81].
Gasification canproduce different H2 to CO ratios (Table 1) depending on the feedstock, type and design of the gasifier, operating conditions, and the type of catalyst used. An equimolar concentration of CO and H2, if available, during fermentation will theoretically yield 0.67 moles of ethanol, i.e. two thirds of the carbon in CO can be converted into ethanol, which is the same as the theoretical yield of ethanol from glucose. Some Clostridium species can produce ethanol following stoichiometric equation 4, which requires a H2 to CO ratio of 2. This ratio can be produced through indirect gasification using steam. Syngas fermentation can also occur in the absence of CO with a CO2 to H2 ratio of 1:3 (equation 6, Table 3) conserving all the carbon in CO2 to ethanol. However, this molar ratio is attainable only when a reforming step such as a water gas shift reactor is added to the downstream gasification process. Such additional unit processes will increase the cost of the overall process at the expense of obtaining complete carbon conversion into fuels. Detailed process economics have to be performed to make knowledgeable recommendations for designing an economically viable gasification-fermentation process.
A major issue in syngas fermentation is the gas liquid mass transfer rate, ${\rm{d}}{N_S}^G$, which can be expressed assuming mass transfer limitation as:
${{{\rm{d}}{N_S}^G} \over {{V_L}{\rm{dt}}}}{\rm{ = }}{{{k_L}{\rm{a}}} \over {\rm{H}}}P_S^G$ | (4) |
Where, ${\rm{d}}{N_S}^G$ is the number of moles of gas transported from the gas phase
${V_L}$ is the liquid volume of the reactor
$t$ is time
${k_L}{\rm{a}}$ is the mass transfer coefficient
H is the Henry’s law constant
$P_S^G$ is the partial pressure of substrate in gas phase
From the equation, it can be observed that the rate of reaction is proportional to the partial pressure of the gaseous substrates, which depends on the total pressure of the reactor and can be improved by providing gases rich in CO, CO2 and H2. Therefore, it is highly desirable to have an undiluted syngas (with little or no nitrogen), because the resulting syngas will naturally have higher partial pressure of syngas components. This will result in improved mass transfer rates for sparingly soluble CO and H2.
Hurst and Lewis [85] showed that increasing the partial pressure of CO in 350 ml Pyrex bioreactors from 0.35 to 2.0 atm increased cell mass and ethanol concentrations and promoted the reduction of acetic acid into ethanol, thereby increasing the product specificity of the microbe towards ethanol. In a different study, increasing partial pressure of gases by using different gas compositions (Syngas I- 20% CO, 15% CO2, 5% H2 and 60% N2 versus syngas II- 40% CO, 30% CO2 and 30% H2) using newly isolated moderately alkaliphilic strains of A. bacchi resulted in increased ethanol productivity and yields from CO [7]. Increased concentrations of CO, CO2 and H2 from gasifiers are only possible through indirect gasification or by providing pure O2 as a gasification medium.
Gas-liquid mass transfer is one of the major challenges in gas-liquid fermentations. The commonly used CSTRs will have to be operated with additional impellers and at a high power density (power applied to move a unit volume of fermentation broth) to break gas bubbles and enhance gas-liquid mass transfer. Numerous studies have compared the mass transfer characteristics of different reactorconfigurations [86,87,88,89]. Munasinghe and Khanal [87] compared various reactor configurations and observed that an air lift reactor attached with a bulb diffuser would provide a better CO mass transfer coefficient in aqueous phase in comparison to a submerged composite hollow fiber membrane module and column diffuser. Orgill et al. [86] observed that a non-porous polydimethylsiloxane (PDMS) hollow fiber reactor provided the highest mass transfer coefficient (KTOTA/V) of 1,062 h−1, which was 152% and 832% higher than the total mass transfer coefficients observed in a trickle bed reactor with 6mm glass beads (421 h−1) and a stirred tank reactor (114 h−1), respectively. However, Orgill et al. [86] used O2 to calculate their mass transfer coefficients. Another study by Munasinghe and Khanal achieved a total mass transfer coefficient with a polyethylene/polypropylene/polyurethane composite hollow fiber reactor that was similar to that observed in [86] using CO (947 h−1) [88]. Shen et al. [89] achieved a CO mass transfer coefficient of 1096 h−1 with a polypropylene hollow fiber membrane. They also produced a maximum ethanol concentration of 24 g/L with a productivity of 3.44 g/L*day using C. carboxidivorans in continuous operation, demonstrating the operational feasibility of hollow fiber membrane based reactors [89]. In addition, ethanol production of 6 g/L was achieved in semi-continuous syngas fermentation in a trickle bed reactor using C. ragsdalei [90].
Another critical issue requiring attention during syngas fermentation is the toxicity of CO to hydrogenase enzymes. CO inhibition in Clostridium hydrogenases has been widely observed in acetone-butanol-ethanol (ABE) fermentations [91]. During syngas fermentation, CO and H2 serve as electron donors required for the various reduction reactions of the metabolic pathway. The production of reducing power from CO originates from the water gas-shift reaction (CO + H2O → CO2 + 2H+ + 2e−) and is catalyzed by the enzyme carbon monoxide dehydrogenase (CODH). While A. woodii, A. kivui, C. ljungdahlii, C. ragsdalei, C. carboxidivorans and A. bacchi have the ability to simultaneously consume CO and H2, other microorganisms such as P. productus, E. limosum, A. notarae, C. aceticum, C. thermoaceticum, S. sphaeroides, R. gelatinosa and R. rubrum are unable to simultaneously consume CO and H2 [7,8,9,40,92]. Ramachandriya et al. [9] showed that C. ragsdalei can consume more H2 than C. carboxidivorans under similar fermentation conditions. In the same study, it was also observed that the amount of CO2 produced by C. carboxidivorans was almost equal to the amount of CO consumed, indicating that the water-gas shift reaction was the main source of reducing power.
Hu et al. [93] indicated that CO (and not H2) was thermodynamically a preferred source of electrons at any given pH, ionic strength, gas partial pressure and electron pair carrier. The generation of reducing power from CO is undesirable because it not only makes an inefficient use of CO as a source of electrons (that could be potentially be carried out by hydrogenase enzymes using H2), but it also diverts the carbon flux of CO towards CO2 and away from ethanol or other products. Consequently, the reaction for the bioconversion of syngas into ethanol is likely occurring by the stoichiometric equation 6 (Table 3), which is an inefficient reaction because the products assimilate only 33% of carbon in CO. In such a scenario, CO2 will be a byproduct of these fermentations and most of the unconsumed H2 will come out of the bioreactor.
CO2 will always be present as an integral component of syngas irrespective of feedstock, gasifier type, gasifying medium and gasification conditions. The ratio of CO to CO2 varied between 0.80 and 2.0 for most of the gasifications performed in the presence of air in a fluidized or downdraft reactor (Table 1), except for syngas produced from an industrial fluidized bed reactor that uses O2 and steam as gasifying agents and coal as the feedstock [29]. Formation of both ethanol and acetic acid from CO (equations 6 and 11) is thermodynamically more favorable than from CO2 and H2 (equations 4 and 12), which indicates CO as the preferred substrate for carbon and energy (Table 3). Supporting experimental evidence for CO preference during syngas fermentation can be found in fermentations using C. ragsdalei, C. carboxidivorans, A. bacchi and P. productus, which preferentially consumed CO when a mixture of CO, CO2 and H2 was provided [7,8,9,79,94]. Although CO2 is not harmful to syngas fermenting microbes, unconsumed CO2 in the exhaust stream reduces the amount of usable carbon converted into products. Utilization of CO2 is important from an economic standpoint, and will require an additional fermentation step for complete utilization of CO2.
CO2 can be fermented into solvents and acids only in the presence of H2 (equations in Table 3). Ramachandriya et al. [95] demonstrated an ethanol yield of 66.5% of theoretical C. carboxidivorans with a feed gas composition of 75% H2 and 25% CO2. This study highlighted that CO2 that is typically generated from syngas fermentations can be converted into solvents and acids by additional fermentation steps utilizing H2. Such a process will increase carbon conversion from biomass to solvents. Furthermore, CO2 fermentations using syngas fermenting microbes can also become a bolt-on technology for CO2 conversion into biofuels and biochemicals if appropriate H2 sources can be identified.
In comparison to traditional yeast fermentation, the fermentation of syngas produced from biomass is a relatively new technology and has potential to produce fuels and chemicals. The merits of syngas fermentation will allow feedstock flexibility during the gasification process. Besides lignocellulosic materials, refuse such as waste tires and municipal sludge can prove to be inexpensive and renewable feedstocks for gasification. However, many issues with the integration of gasification and syngas fermentation technology must be addressed for large scale commercial deployment. Some of the challenges of this integrated platform, including conversion efficiency of the gasification and syngas fermentation processes, effect of minor gaseous species, and effect of H2:CO:CO2 ratios on syngas composition, were discussed extensively in this review. With further aggressive and concerted fundamental and applied research, gasification-syngas fermentation technology can prove to be a major source of renewable fuel and chemical production.
The authors acknowledge the support of the Oklahoma Agricultural Experiment Station in developing this review.
All authors declareno conflict of interest.
[1] |
Klepeis JL, Lindorff-Larsen K, Dror RO, et al. (2009) Long-timescale molecular dynamics simulations of protein structure and function. Curr Opin Struct Biol 19: 120–127. doi: 10.1016/j.sbi.2009.03.004
![]() |
[2] |
Stone JE, Phillips JC, Freddolino PL, et al. (2007) Accelerating molecular modeling applications with graphics processors. J Comput Chem 28: 2618–2640. doi: 10.1002/jcc.20829
![]() |
[3] |
Perez A, Morrone JA, Simmerling C, et al. (2016) Advances in free-energy-based simulations of protein folding and ligand binding. Curr Opin Struct Biol 36: 25–31. doi: 10.1016/j.sbi.2015.12.002
![]() |
[4] |
Lane TJ, Shukla D, Beauchamp KA, et al. (2013) To milliseconds and beyond: Challenges in the simulation of protein folding. Curr Opin Struct Biol 23: 58–65. doi: 10.1016/j.sbi.2012.11.002
![]() |
[5] |
Freddolino PL, Harrison CB, Liu Y, et al. (2010) Challenges in protein-folding simulations. Nat Phys 6: 751–758. doi: 10.1038/nphys1713
![]() |
[6] |
Best RB (2012) Atomistic molecular simulations of protein folding. Curr Opin Struct Biol 22: 52–61. doi: 10.1016/j.sbi.2011.12.001
![]() |
[7] | Piana S, Lindorff-Larsen K, Shaw DE (2012) Protein folding kinetics and thermodynamics from atomistic simulation. P Natl Acad Sci USA 109: 17845–17850. |
[8] |
Suárez E, Lettieri S, Zwier MC, et al. (2014) Simultaneous computation of dynamical and equilibrium information using a weighted ensemble of trajectories. J Chem Theory Comput 10: 2658–2667. doi: 10.1021/ct401065r
![]() |
[9] |
Pierce LCT, Salomon-Ferrer R, Augusto F. De Oliveira C, et al. (2012) Routine access to millisecond time scale events with accelerated molecular dynamics. J Chem Theory Comput 8: 2997–3002. doi: 10.1021/ct300284c
![]() |
[10] |
Kubelka J, Hofrichter J, Eaton WA (2004) The protein folding "speed limit." Curr Opin Struct Biol 14: 76–88. doi: 10.1016/j.sbi.2004.01.013
![]() |
[11] |
Lindorff-Larsen K, Piana S, Dror RO, et al. (2011) How fast-folding proteins fold. Science 334: 517–520. doi: 10.1126/science.1208351
![]() |
[12] | Putignano V, Rosato A, Banci L, et al. (2018) MetalPDB in 2018: a database of metal sites in biological macromolecular structures. Nucleic Acids Res 41: 459–464. |
[13] |
Li W, Wang J, Zhang J, et al. (2015) Molecular simulations of metal-coupled protein folding. Curr Opin Struct Biol 30: 25–31. doi: 10.1016/j.sbi.2014.11.006
![]() |
[14] |
Bentrop D, Bertini I, Iacoviello R, et al. (1999) Structural and dynamical properties of a partially unfolded Fe4S4 protein: Role of the cofactor in protein folding. Biochemistry 38: 4669–4680. doi: 10.1021/bi982647q
![]() |
[15] |
Blake PR, Summers MF, Park JB, et al. (1991) Determinants of protein hyperthermostability: purification and amino acid sequence of rubredoxin from the hyperthermophilic archaebacterium pyrococcus furiosus and secondary structure of the zinc adduct by NMR. Biochemistry 30: 10885–10895. doi: 10.1021/bi00109a012
![]() |
[16] | Prakash S, Sundd M, Guptasarma P (2014) The key to the extraordinary thermal stability of P. furiosus holo-rubredoxin: Iron binding-guided packing of a core aromatic cluster responsible for high kinetic stability of the native structure. PLoS One 9: e89703. |
[17] |
Hernandez G, Jenney FE, Adams MW, et al. (2000) Millisecond time scale conformational flexibility in a hyperthermophile protein at ambient temperature. Proc Natl Acad Sci USA 97: 3166–3170. doi: 10.1073/pnas.97.7.3166
![]() |
[18] | Rader AJ (2010) Thermostability in rubredoxin and its relationship to mechanical rigidity. Phys Biol 7: 016002. |
[19] |
Bonomi F, Iametti S, Ferranti P, et al. (2008) "Iron priming" guides folding of denatured aporubredoxins. J Biol Inorg Chem 13: 981–991. doi: 10.1007/s00775-008-0385-4
![]() |
[20] |
Zartler ER, Jenney FE, Terrell M, et al. (2001) Structural basis for thermostability in aporubredoxins from Pyrococcus furiosus and Clostridium pasteurianum. Biochemistry 40: 7279–7290. doi: 10.1021/bi0026831
![]() |
[21] |
Cavagnero S, Debe DA, Zhou ZH, et al. (1998) Kinetic role of electrostatic interactions in the unfolding of hyperthermophilic and mesophilic rubredoxins. Biochemistry 37: 3369–3376. doi: 10.1021/bi9721795
![]() |
[22] |
Strop P, Mayo SL (1999) Rubredoxin variant folds without iron. J Am Chem Soc 121: 2341–2345. doi: 10.1021/ja9834780
![]() |
[23] |
Hamelberg D, Mongan J, McCammon JA (2004) Accelerated molecular dynamics: A promising and efficient simulation method for biomolecules. J Chem Phys 120: 11919–11929. doi: 10.1063/1.1755656
![]() |
[24] |
Doshi U, Hamelberg D (2015) Towards fast, rigorous and efficient conformational sampling of biomolecules: Advances in accelerated molecular dynamics. BBA-Gen Subjects 1850: 878–888. doi: 10.1016/j.bbagen.2014.08.003
![]() |
[25] |
Miao Y, Feixas F, Eun C (2015) Accelerated molecular dynamics simulations of protein folding. J Comput Chem 36: 1536–1549. doi: 10.1002/jcc.23964
![]() |
[26] | Case DA, Cerutti DS, Cheatham TE, et al. (2017) Amber 2017, University of California, San Francisco. |
[27] |
Carvalho ATP, Teixeira AFS, Ramos MJ (2013) Parameters for molecular dynamics simulations of iron-sulfur proteins. J Comput Chem 34: 1540–1548. doi: 10.1002/jcc.23287
![]() |
[28] |
Bertini I, Case DA, Ferella L, et al. (2011) A grid-enabled web portal for NMR structure refinement with AMBER. Bioinformatics 27: 2384–2390. doi: 10.1093/bioinformatics/btr415
![]() |
[29] |
Wassenaar TA, van Dijk M, Loureiro-Ferreira N, et al. (2012) WeNMR: Structural biology on the grid. J Grid Comput 10: 743–767. doi: 10.1007/s10723-012-9246-z
![]() |
[30] |
Prompers JJ, Brüschweiler R, Bruschweiler R (2002) General framework for studying the dynamics of folded and nonfolded proteins by NMR relaxation spectroscopy and MD simulation. J Am Chem Soc 124: 4522–4534. doi: 10.1021/ja012750u
![]() |
[31] |
Korzhnev DM, Billeter M, Arseniev AS, et al. (2001) NMR studies of Brownian tumbling and internal motions in proteins. Prog Nucl Mag Res Sp 38: 197–266. doi: 10.1016/S0079-6565(00)00028-5
![]() |
[32] |
Kabsch W, Sander C (1983) Dictionary of protein secondary structure: Pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 22: 2577–2637. doi: 10.1002/bip.360221211
![]() |
[33] |
Rost B, Sander C (1993) Prediction of protein secondary structure at better than 70% accuracy. J Mol Biol 232: 584–599. doi: 10.1006/jmbi.1993.1413
![]() |
[34] |
Li DW, Brüschweiler R (2012) PPM: A side-chain and backbone chemical shift predictor for the assessment of protein conformational ensembles. J Biomol NMR 54: 257–265. doi: 10.1007/s10858-012-9668-8
![]() |
[35] |
Hiller R, Zhou ZH, Adams MW, et al. (1997) Stability and dynamics in a hyperthermophilic protein with melting temperature close to 200 degrees C. Proc Natl Acad Sci USA 94: 11329–11332. doi: 10.1073/pnas.94.21.11329
![]() |
[36] |
Ishima R, Torchia DA (2000) Protein dynamics from NMR. Nat Struct Biol 7: 740–743. doi: 10.1038/78963
![]() |
[37] |
Jarymowycz VA, Stone MJ (2006) Fast time scale dynamics of protein backbones: NMR relaxation methods, applications, and functional consequences. Chem Rev 106: 1624–1671. doi: 10.1021/cr040421p
![]() |
[38] |
LeMaster DM (1999) NMR relaxation order parameter analysis of the dynamics of protein side chains. J Am Chem Soc 121: 1726–1742. doi: 10.1021/ja982988r
![]() |
[39] |
Ruschak AM, Kay LE (2010) Methyl groups as probes of supra-molecular structure, dynamics and function. J Biomol NMR 46: 75–87. doi: 10.1007/s10858-009-9376-1
![]() |
[40] |
Bougault CM, Eidsness MK, Prestegard JH (2003) Hydrogen bonds in rubredoxins from mesophilic and hyperthermophilic organisms. Biochemistry 42: 4357–4372. doi: 10.1021/bi027264d
![]() |
[41] |
Prestegard JH, Bougault CM, Kishore AI (2004) Residual dipolar couplings in structure determination of biomolecules. Chem Rev 104: 3519–3540. doi: 10.1021/cr030419i
![]() |
[42] |
Cho-Chung YS, Pitot HC (1968) Regulatory effects of nicotinamide on tryptophan pyrrolase synthesis in rat liver in vivo. Eur J Biochem 3: 401–406. doi: 10.1111/j.1432-1033.1967.tb19543.x
![]() |
[43] |
Blasie CA, Berg JM (2002) Structur e-based thermodynamic analysis of a coupled metal binding-protein folding reaction involving a zinc finger peptide. Biochemistry 41: 15068–15073. doi: 10.1021/bi026621h
![]() |
[44] |
Weinkam P, Romesberg FE, Wolynes PG (2009) Chemical frustration in the protein folding landscape: Grand canonical ensemble simulations of cytochrome c. Biochemistry 48: 2394–2402. doi: 10.1021/bi802293m
![]() |
[45] |
Devereux M, Gresh N, Piquemal JP, et al. (2014) A supervised fitting approach to force field parametrization with application to the SIBFA polarizable force field. J Comput Chem 35: 1577–1591. doi: 10.1002/jcc.23661
![]() |
[46] |
Wu R, Lu Z, Cao Z, et al. (2011) A transferable nonbonded pairwise force field to model zinc interactions in metalloproteins. J Chem Theory Comput 7: 433–443. doi: 10.1021/ct100525r
![]() |
[47] |
Sakharov DV, Lim C (2005) Zn protein simulations including charge transfer and local polarization effects. J Am Chem Soc 127: 4921–4929. doi: 10.1021/ja0429115
![]() |
[48] |
Chakravorty DK, Wang B, Lee CW, et al. (2012) Simulations of allosteric motions in the zinc sensor CzrA. J Am Chem Soc 134: 3367–3376. doi: 10.1021/ja208047b
![]() |
[49] |
Chakravorty DK, Parker TM, Guerra AJ, et al. (2013) Energetics of zinc-mediated interactions in the allosteric pathways of metal sensor proteins. J Am Chem Soc 135: 30–33. doi: 10.1021/ja309170g
![]() |
[50] |
Reyes-Caballero H, Campanello GC, Giedroc DP (2011) Metalloregulatory proteins: Metal selectivity and allosteric switching. Biophys Chem 156: 103–114. doi: 10.1016/j.bpc.2011.03.010
![]() |
[51] |
Andrews CT, Elcock AH (2013) Molecular dynamics simulations of highly crowded amino acid solutions: comparisons of eight different force field combinations with experiment and with each other. J Chem Theory Comput 9: 4585–4602. doi: 10.1021/ct400371h
![]() |
[52] |
Abriata LA, Dal Peraro M (2015) Assessing the potential of atomistic molecular dynamics simulations to probe reversible protein-protein recognition and binding. Sci Rep 5: 10549. doi: 10.1038/srep10549
![]() |
![]() |
![]() |
1. | Antonio Grimalt-Alemany, Mateusz Łężyk, Lene Lange, Ioannis V. Skiadas, Hariklia N. Gavala, Enrichment of syngas-converting mixed microbial consortia for ethanol production and thermodynamics-based design of enrichment strategies, 2018, 11, 1754-6834, 10.1186/s13068-018-1189-6 | |
2. | Nulee Jang, Muhammad Yasin, Shinyoung Park, Robert W. Lovitt, In Seop Chang, Determination of volumetric gas–liquid mass transfer coefficient of carbon monoxide in a batch cultivation system using kinetic simulations, 2017, 239, 09608524, 387, 10.1016/j.biortech.2017.05.023 | |
3. | Olivera S. Stamenković, Kaliramesh Siliveru, Vlada B. Veljković, Ivana B. Banković-Ilić, Marija B. Tasić, Ignacio A. Ciampitti, Ivica G. Đalović, Petar M. Mitrović, Vladimir Š. Sikora, P.V. Vara Prasad, Production of biofuels from sorghum, 2020, 124, 13640321, 109769, 10.1016/j.rser.2020.109769 | |
4. | Eduardo Almeida Benalcázar, Baudine Gevers Deynoot, Henk Noorman, Patricia Osseweijer, John A. Posada, Production of bulk chemicals from lignocellulosic biomass via thermochemical conversion and syngas fermentation: a comparative techno‐economic and environmental assessment of different site‐specific supply chain configurations, 2017, 11, 1932-104X, 861, 10.1002/bbb.1790 | |
5. | Vineet Singh Sikarwar, Michael Pohořelý, Erik Meers, Siarhei Skoblia, Jaroslav Moško, Michal Jeremiáš, Potential of coupling anaerobic digestion with thermochemical technologies for waste valorization, 2021, 294, 00162361, 120533, 10.1016/j.fuel.2021.120533 | |
6. | Muhammad Yasin, Nulee Jang, Mungyu Lee, Hyunsoo Kang, Muhammad Aslam, Aqeel Ahmed Bazmi, In Seop Chang, Bioreactors, gas delivery systems and supporting technologies for microbial synthesis gas conversion process, 2019, 7, 2589014X, 100207, 10.1016/j.biteb.2019.100207 | |
7. | Antonio Grimalt‐Alemany, Ioannis V. Skiadas, Hariklia N. Gavala, Syngas biomethanation: state‐of‐the‐art review and perspectives, 2018, 12, 1932-104X, 139, 10.1002/bbb.1826 | |
8. | Konstantinos Chandolias, Tobias Richards, Mohammad J. Taherzadeh, 2018, 9780444639929, 157, 10.1016/B978-0-444-63992-9.00005-7 | |
9. | Peter Riegler, Thomas Chrusciel, Alexander Mayer, Kathrin Doll, Dirk Weuster-Botz, Reversible retrofitting of a stirred-tank bioreactor for gas-lift operation to perform synthesis gas fermentation studies, 2019, 141, 1369703X, 89, 10.1016/j.bej.2018.09.021 | |
10. | E.T. Liakakou, B.J. Vreugdenhil, N. Cerone, F. Zimbardi, F. Pinto, R. André, P. Marques, R. Mata, F. Girio, Gasification of lignin-rich residues for the production of biofuels via syngas fermentation: Comparison of gasification technologies, 2019, 251, 00162361, 580, 10.1016/j.fuel.2019.04.081 | |
11. | 2020, 9780444643216, 289, 10.1016/B978-0-444-64321-6.00015-X | |
12. | Kan Liu, John R. Phillips, Xiao Sun, Sayeed Mohammad, Raymond L. Huhnke, Hasan K. Atiyeh, Investigation and Modeling of Gas-Liquid Mass Transfer in a Sparged and Non-Sparged Continuous Stirred Tank Reactor with Potential Application in Syngas Fermentation, 2019, 5, 2311-5637, 75, 10.3390/fermentation5030075 | |
13. | E.T. Liakakou, A. Infantes, A. Neumann, B.J. Vreugdenhil, Connecting gasification with syngas fermentation: Comparison of the performance of lignin and beech wood, 2021, 290, 00162361, 120054, 10.1016/j.fuel.2020.120054 | |
14. | Nulee Jang, Muhammad Yasin, Mungyu Lee, Hyunsoo Kang, In Seop Chang, Gas circulation rate and medium exchange ratio as influential factors affecting ethanol production in carbon monoxide fermentation using a packed-bed reactor, 2020, 4, 2398-4902, 1963, 10.1039/C9SE00943D | |
15. | Haris Nalakath Abubackar, María C. Veiga, Christian Kennes, 2019, 9780444642004, 207, 10.1016/B978-0-444-64200-4.00015-3 | |
16. | Sunita Varjani, Ashok Pandey, Edgard Gnansounou, Samir Khanal, Raveendran Sindhu, 2020, 9780444643216, 3, 10.1016/B978-0-444-64321-6.00001-X | |
17. | Steven Wainaina, Ilona Sárvári Horváth, Mohammad J. Taherzadeh, Biochemicals from food waste and recalcitrant biomass via syngas fermentation: A review, 2018, 248, 09608524, 113, 10.1016/j.biortech.2017.06.075 | |
18. | Alba Infantes, Michaela Kugel, Klaus Raffelt, Anke Neumann, Side-by-Side Comparison of Clean and Biomass-Derived, Impurity-Containing Syngas as Substrate for Acetogenic Fermentation with Clostridium ljungdahlii, 2020, 6, 2311-5637, 84, 10.3390/fermentation6030084 | |
19. | Ho Ting Luk, Cecilia Mondelli, Daniel Curulla Ferré, Joseph A. Stewart, Javier Pérez-Ramírez, Status and prospects in higher alcohols synthesis from syngas, 2017, 46, 0306-0012, 1358, 10.1039/C6CS00324A | |
20. | Cosetta Ciliberti, Antonino Biundo, Roberto Albergo, Gennaro Agrimi, Giacobbe Braccio, Isabella de Bari, Isabella Pisano, Syngas Derived from Lignocellulosic Biomass Gasification as an Alternative Resource for Innovative Bioprocesses, 2020, 8, 2227-9717, 1567, 10.3390/pr8121567 | |
21. | Syed Haseeb Sultan, Arkom Palamanit, Kua-Anan Techato, Muhammad Amin, Khurshid Ahmed Baloch, Physiochemical Characterization and Potential of Synthesis Gas Production from Rubber Wood Biomass by Using Downdraft Gasifier, 2021, 40, 02547821, 1, 10.22581/muet1982.2101.01 | |
22. | Carolina Benevenuti, Priscilla Amaral, Tatiana Ferreira, Peter Seidl, Impacts of Syngas Composition on Anaerobic Fermentation, 2021, 2, 2624-781X, 391, 10.3390/reactions2040025 | |
23. | Yunheng Tang, Yun Huang, Wentian Gan, Ao Xia, Qiang Liao, Xianqing Zhu, Ethanol production from gas fermentation: Rapid enrichment and domestication of bacterial community with continuous CO/CO2 gas, 2021, 175, 09601481, 337, 10.1016/j.renene.2021.04.134 | |
24. | Tharaka Rama Krishna C. Doddapaneni, Fiaz Ahmad, Kaspar Valgepea, Timo Kikas, 2022, 9780323898553, 67, 10.1016/B978-0-323-89855-3.00025-X | |
25. | Venkatesa Prabhu Sundramurthy, Thirumullaivoyal G. Nithya, Chandran Masi, Chinnasamy Gomadurai, Ebrahim M. Abda, Recent advances and prospects for industrial waste management and product recovery for environmental appliances, 2021, 0, 2365-6581, 10.1515/psr-2021-0063 | |
26. | Flávio C. F. Baleeiro, Sabine Kleinsteuber, Heike Sträuber, Recirculation of H2, CO2, and Ethylene Improves Carbon Fixation and Carboxylate Yields in Anaerobic Fermentation, 2022, 10, 2168-0485, 4073, 10.1021/acssuschemeng.1c05133 | |
27. | George M. Teke, Siew L. Tai, Robert W. M. Pott, Extractive Fermentation Processes: Modes of Operation and Application, 2022, 9, 2196-9744, 146, 10.1002/cben.202100028 | |
28. | Elisa M. de Medeiros, Henk Noorman, Rubens Maciel Filho, John A. Posada, Multi-Objective Sustainability Optimization of Biomass Residues to Ethanol via Gasification and Syngas Fermentation: Trade-Offs between Profitability, Energy Efficiency, and Carbon Emissions, 2021, 7, 2311-5637, 201, 10.3390/fermentation7040201 | |
29. | Francesco Regis, Alessandro Hugo Antonio Monteverde, Debora Fino, A techno-economic assessment of bioethanol production from switchgrass through biomass gasification and syngas fermentation, 2023, 274, 03605442, 127318, 10.1016/j.energy.2023.127318 | |
30. | Rendy Mukti, Made Tri Ari Penia Kresnowati, Tjandra Setiadi, Challenges in Syngas Fermentation for Bioethanol Production: Syngas Composition, 2023, 1, 2813-6535, 9, 10.4028/p-9g14o1 | |
31. | Budi Mandra Harahap, Birgitte K. Ahring, Acetate Production from Syngas Produced from Lignocellulosic Biomass Materials along with Gaseous Fermentation of the Syngas: A Review, 2023, 11, 2076-2607, 995, 10.3390/microorganisms11040995 | |
32. | J. Figueras, H. Benbelkacem, C. Dumas, P. Buffiere, Inhibitions from Syngas Impurities: Impact of Common Tar Components on a Consortium Adapted for Syngas Biomethanation, 2024, 15, 1877-2641, 2427, 10.1007/s12649-023-02237-x | |
33. | Michael Mol, Magda Stephania Ardila, Bronwyn Ashleigh Mol, Habibu Aliyu, Anke Neumann, Pieter de Maayer, The effects of synthesis gas feedstocks and oxygen perturbation on hydrogen production by Parageobacillus thermoglucosidasius, 2024, 23, 1475-2859, 10.1186/s12934-024-02391-4 | |
34. | A.R. Akinwumi, O.C. Nwinyi, A.O. Ayeni, S. Venkata Mohan, Influence of synthetic carbon grade on the metabolic flux of polyhydroxyalkanoate monomeric constitution synthesized by Bacillus cereus AAR-1, 2024, 27, 2589014X, 101958, 10.1016/j.biteb.2024.101958 | |
35. | Obid Tursunov, Katarzyna Śpiewak, Nurislom Abduganiev, Yang Yang, Alexander Kustov, Islom Karimov, Thermogravimetric and thermovolumetric study of municipal solid waste (MSW) and wood biomass for hydrogen-rich gas production: a case study of Tashkent region, 2023, 30, 1614-7499, 112631, 10.1007/s11356-023-30368-0 | |
36. | Shaik Muntasir Shovon, Faysal Ahamed Akash, Wahida Rahman, Md Abdur Rahman, Prosenjeet Chakraborty, H.M. Zakir Hossain, Minhaj Uddin Monir, Strategies of managing solid waste and energy recovery for a developing country – A review, 2024, 10, 24058440, e24736, 10.1016/j.heliyon.2024.e24736 | |
37. | Philipp Leuter, Sebastian Fendt, Hartmut Spliethoff, Requirements on synthesis gas from gasification for material and energy utilization: a mini review, 2024, 12, 2296-598X, 10.3389/fenrg.2024.1382377 | |
38. | Pedro Acuña López, Stefano Rebecchi, Elodie Vlaeminck, Koen Quataert, Christian Frilund, Jaana Laatikainen-Luntama, Ilkka Hiltunen, Karel De Winter, Wim K. Soetaert, Demonstrating Pilot-Scale Gas Fermentation for Acetate Production from Biomass-Derived Syngas Streams, 2024, 10, 2311-5637, 285, 10.3390/fermentation10060285 | |
39. | Berivan Tunca, Anastasia Rovithi, Sambit Dutta, Cesar Quintela, Manuel Pinelo, Ioannis V. Skiadas, Hariklia N. Gavala, Enhanced Carbon Monoxide Bioconversion Through Adaptation of Mixed Microbial Consortia in Trickle Bed Reactors, 2025, 1877-2641, 10.1007/s12649-025-02945-6 | |
40. | Nathawat Unsomsri, Pathipan Manchit, Puthanate Tabrak, Songkran Wiriyasart, Sommas Kaewluan, Effect of Oxygen-Enriched Air on Flame Temperature and Emission Characteristics During Hot Producer Gas Combustion in a Modified Cross-draft Gasifier-Burner, 2024, 0010-2202, 1, 10.1080/00102202.2024.2438743 |
![]() |