The denitrifying betaproteobacterium Aromatoleum aromaticum EbN1T is a facultative anaerobic degradation specialist and belongs to the environmental bacteria studied best on the proteogenomic level. This review summarizes the current state of knowledge about the anaerobic and aerobic degradation (to CO2) of 47 organic growth substrates (23 aromatic, 21 aliphatic, and 3 amino acids) as well as the modes of respiratory energy conservation (denitrification vs. O2-respiration). The constructed catabolic network is comprised of 256 genes, which occupy ∼7.5% of the coding regions of the genome. In total, 219 encoded proteins have been identified by differential proteomics, yielding a proteome coverage of ∼74% of the network. Its degradation section is composed of 31 peripheral and 4 central pathways, with several peripheral modules (e.g., for 4-ethylphenol, 2-phenylethylamine, indoleacetate, and phenylpropanoids) discovered only after the complete genome [Arch Microbiol. 2005 Jan;183(1):27–36] and a first proteomic survey [Proteomics. 2007 Jun;7(13):2222–39] of A. aromaticum EbN1T were reported. The activation of recalcitrant aromatic compounds involves a suite of biochemically intriguing reactions ranging from C−H-bond activation (e.g., ethylbenzene dehydrogenase) via carboxylation (e.g., acetophenone carboxylase) to oxidative deamination (e.g., benzylamine), reductive dearomatization (benzoyl-CoA), and epoxide-forming oxygenases (e.g., phenylacetyl-CoA). The peripheral reaction sequences are substrate-specifically induced, mediated by specific transcriptional regulators with in vivo response thresholds in the nanomolar range. While lipophilic substrates (e.g., phenolics) enter the cells via passive diffusion, polar ones require active uptake that is driven by specific transporters. Next to the protein repertoire for canonical complexes I−III, denitrification, and O2-respiration (low- and high-affinity oxidases), the genome encodes an Ndh-II, a tetrathionate reductase, two ETF:quinone oxidoreductases, and two Rnf-type complexes, broadening the electron transfer flexibility of the strain. Taken together, the detailed catabolic network presented here forms a solid basis for future systems biology-level studies with A. aromaticum EbN1T.

Aromatic compounds represent, next to carbohydrates, the second most abundant organic molecules on the planet. In the biosphere they are encountered as products of anaerobic biodegradation [Fischer-Romero et al., 1996], as building blocks of major biomacromolecules [de Leeuw et al., 2006], or as phytohormones [Pichersky and Raguso, 2018]. In the geosphere, aromatic compounds are abundant constituents of kerogen [Vandenbroucke and Largeau, 2007] and crude oil [Marshall and Rodgers, 2008]. Anthropogenic activities including industrial synthesis and fossil energy production provide further sources of aromatic compounds in the environment [ATSDR, 2008]. While aromatic compounds are energy-rich substrates for supporting microbial growth, they are at the same time biochemically challenging due to their chemical stability [Wilkes and Schwarzbauer, 2010] and physiologically adverse on account of their toxicology [Sikkema et al., 1995]. To activate and cleave the aromatic ring, O2-dependent oxygenase-based reactions and a large variety of intriguing O2-independent reactions are employed by aerobic and anaerobic microorganisms, respectively [Rabus et al., 2016; Boll et al., 2020; Cheng et al., 2022].

Since transitions between oxia and anoxia occur in many different environments, facultative degradation specialists for aromatic compounds represent keystone members of the microbial communities in these habitats, e.g., the facultative denitrifying betaproteobacterial genera Aromatoleum [Rabus et al., 2019] and Thauera [Anders et al., 1995]. The genus Aromatoleum comprises currently 10 validly described species, with Aromatoleum aromaticum EbN1T (DSM 19018T) as the type species and versatile representative [Rabus and Widdel, 1995]. This strain is the first anaerobic hydrocarbon degrader with a completely sequenced genome [Rabus et al., 2005] and is particularly well studied on the physiological, genetic, and OMICS levels [e.g., Büsing et al., 2015a, b; Rabus et al., 2014; Trautwein et al., 2012a, 2012b; Vagts et al., 2020; Wöhlbrand et al., 2007, 2008]. In the case of the catabolic signature property of A. aromaticum EbN1T, namely the anaerobic degradation of ethylbenzene, the crystal structures of the pathway’s first three enzymes have been determined: ethylbenzene dehydrogenase [Kloer et al., 2006], (S)-1-phenylethanol dehydrogenase [Höffken et al., 2006], and acetophenone carboxylase [Weidenweber et al., 2017]. Most recently, A. aromaticum EbN1T was established as a model for systems biology-level investigations, including a manual re-annotation of the complete genome, multi-OMICS integration, and a genome-wide metabolic model [Becker et al., 2022a].

Here, we summarize the state of knowledge on the catabolism of A. aromaticum EbN1T. For this purpose, we integrated an up-to-date manual annotation of its genome with all currently available data from differential proteomics and genetic/biochemical analyses conducted over the past more than 2 decades with A. aromaticum EbN1T. The functional predictions were supported by pathway-specific literature on related, biochemically well-studied Thauera aromatica K172T, Aromatoleum evansii KB740T (formerly Azoarcus evansii KB740T [Rabus et al., 2019]), Rhodopseudomonas palustris, and other strains, as well as by literature on general catabolic aspects using standard bacteria such as Escherichia coli, Bacillus subtilis, and Pseudomonas species.

Genomic Localization of “Catabolic” Genes and Proteomic Coverage

Circular representations of the chromosome and plasmid 2 of A. aromaticum EbN1T (Fig. 1A) illustrate the localization of genes attributed to the substrate-specific degradation pathways for aromatic compounds, aliphatic compounds, and amino acids, to the central pathways for terminal oxidation to CO2, as well as to the modules for respiratory energy conservation employed under anoxic and oxic conditions. In most cases, the genes constituting the various degradation and respiration modules are arranged in operon-like structures, which for the most part are scattered across the chromosome and the plasmid (2nd plasmid not shown in Fig. 1A). This reflects the previously reported high plasticity of the genome of A. aromaticum EbN1T [Rabus et al., 2005].

Fig. 1.

Genomic and proteomic overview of the catabolic network of A. aromaticumEbN1T. A Genomic localization of “degradation” and “respiration” genes. B Protein inventory, genome occupancy, and proteomic coverage of the catabolic network.

Fig. 1.

Genomic and proteomic overview of the catabolic network of A. aromaticumEbN1T. A Genomic localization of “degradation” and “respiration” genes. B Protein inventory, genome occupancy, and proteomic coverage of the catabolic network.

Close modal

In total, the degradation and respiration modules are built on 279 and 91 genes, respectively, occupying 7.4% and 2.0% of the genome’s coding space, with 178 and 65 encoded proteins identified by differential proteomics (Fig. 1B). Overall, A. aromaticum EbN1T realizes its catabolic network with 370 predicted proteins (∼67% identified, ∼9.5% of coding region of its 4.7 Mbp genome). This is in a similar range as recently determined for closely related A. aromaticum pCyN1 (215 predicted proteins, 89% proteomic coverage, ∼6% of coding region of the 4.4 Mbp genome) [Becker et al., 2022b] and deltaproteobacterial, marine, versatile, sulfate-reducing Desulfonema magnum (145 predicted proteins, 90% identified, ∼2.16% of coding region of the 8 Mbp genome) [Schnaars et al., 2021].

Overall Structure of Degradation Network

An integrative scheme of the degradation network of A. aromaticum EbN1T is presented in Figure 2, illustrating the complex structure of the network, interlinkages, and central modules. This scheme provides a roadmap to the diverse reaction sequences employed for the degradation of 23/21 aromatic/aliphatic compounds and 3 amino acids to CO2 (via the TCA cycle; in few cases, requiring preparatory passage through the methylmalonyl-CoA pathway). Summed up, about a dozen anaerobic peripheral degradation routes channel a total of 15 different aromatic compounds into the anaerobic benzoyl-CoA pathway, underpinning its central role in the degradation network. Aerobic degradation of 4 aromatic compounds proceeds via the “box” pathway. Peripheral degradation of aliphatic compounds enters the TCA cycle at the levels of citrate via acetyl-CoA (10 compounds), of 2-oxoglutarate (3 compounds) or of succinate (6 compounds).

Fig. 2.

Overall structure of the degradation network of A. aromaticumEbN1T. Indicated figures provide detailed information on the reaction sequence, coding genes, and involved proteins of the respective pathways. Coloring of arrows: red, anaerobic; blue, aerobic; brown, anaerobic and aerobic. Thick arrows highlight the four central modules of the degradation network: anaerobic benzoyl-CoA pathway, aerobic box pathway, TCA cycle and methylmalonyl-CoA pathway. Green dots indicate substrates of A. aromaticumEbN1T recognized only after initial publication of the isolation of the strain [Rabus and Widdel, 1995] and determination of the genome sequence [Rabus et al., 2005].

Fig. 2.

Overall structure of the degradation network of A. aromaticumEbN1T. Indicated figures provide detailed information on the reaction sequence, coding genes, and involved proteins of the respective pathways. Coloring of arrows: red, anaerobic; blue, aerobic; brown, anaerobic and aerobic. Thick arrows highlight the four central modules of the degradation network: anaerobic benzoyl-CoA pathway, aerobic box pathway, TCA cycle and methylmalonyl-CoA pathway. Green dots indicate substrates of A. aromaticumEbN1T recognized only after initial publication of the isolation of the strain [Rabus and Widdel, 1995] and determination of the genome sequence [Rabus et al., 2005].

Close modal

The network architecture of its degradation and respiration parts is based on the recent manual re-annotation of the complete genome of A. aromaticum EbN1T [Becker et al., 2022a]. To facilitate comparison of the catabolic network with previous proteogenomic publications on this strain, the tabulated protein constituents of each individual catabolic module (Fig. 3−42) are presented with new [Becker et al., 2022a] and original locus tags [Rabus et al., 2005]; in addition, references for experimental evidence (proteomic, molecular genetic, and/or biochemical) are indicated.

Toluene

Degradation Pathway

Toluene represents the best-studied model compound in the context of the anaerobic degradation of aromatic hydrocarbons (alkylbenzenes). The involved reactions and enzymes have been mainly elucidated with T. aro-matica K172T and Azoarcus sp. strain T [for overview, see Heider et al. [2016a], and the coding genes have been identified in T. aromatica K172T and A. aromaticum EbN1T [Leuthner et al., 1998; Leuthner and Heider, 2000; Hermuth et al., 2002; Kube et al., 2004]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 3A, B), which proceeds via six steps. (i) In the initial reaction step, the glycyl-radical enzyme benzylsuccinate synthase (BssABC) adds toluene to the double bond of the fumarate cosubstrate, forming (R)-benzylsuccinate, with the stereospecificity elucidated among others by modeling and using [d3-methyl]toluene combined with NMR-analysis [Leuthner et al., 1998; Krieger et al., 2001; Qiao and Marsh, 2005; Seyhan et al., 2016; Szaleniec and Heider, 2016]. Benzylsuccinate synthase belongs to the growing family of glycyl-radical-containing enzymes. This family includes related enzymes activating other aromatic and aliphatic hydrocarbons as well as well-studied pyruvate formate-lyase and anaerobic ribonucleotide reductase, with the latter two, however, catalyzing very different reactions [Frey, 2001; Strijkstra et al., 2014; Heider et al., 2016a]. EPR and Mössbauer spectroscopic, crystal structure, and biochemical analyses revealed that the heterotrimeric benzylsuccinate synthase harbors the glycyl (storage)/thiyl (catalysis)-radical sites in the catalytic α-subunit (BssA) and FeS clusters in the accessory β- and γ-subunits (BssBC) [Duboc-Toia et al., 2003; Li et al., 2009; Hillberg et al., 2012; Funk et al., 2014]. The glycyl storage radical is assumed to be generated by activase BssD according to the known route involving S-adenosylmethionine (SAM). During catalysis toluene is suggested to be sandwiched between the thiyl-radical and the fumarate cosubstrate resulting in syn-addition [Li and Marsh, 2006; Funk et al., 2015; Szaleniec and Heider, 2016], i.e., cleavage of the benzylic C−H-bond concurrent with C−C-bond formation between the benzyl radical and fumarate. This is analogous to the previously proposed concerted mechanism of anaerobic conversion of n-hexane to (R)-(1-methylpentyl)succinate in Aromatoleum sp. strain HxN1. Concurrent to the homolytic abstraction of the pro-S hydrogen from carbon atom 2 of n-hexane, the new bond between the latter and one of the two carbon atoms of the double bond of fumarate is already formed on the other side of the n-hexane molecule, thus avoiding a free alk-2-yl radical intermediate [Jarling et al., 2012; Wilkes et al., 2016]. (ii) (R)-Benzylsuccinate is then prepared for subsequent β-oxidation by activation to (R)-benzylsuccinyl-CoA catalyzed by a specific succinyl-CoA:(R)-benzylsuccinate CoA transferase (BbsEF) having an α2β2 conformation [Leutwein and Heider, 2001]. (iii) (R)-Benzylsuccinyl-CoA is then oxidized by substrate-specific, homotetrameric, and FAD-containing (R)-benzylsuccinyl-CoA dehydrogenase (BbsG) forming (E)-phenylitaconyl-CoA (synonym: 2-(E)-benzylidenesuccinyl-CoA) [Leutwein and Heider, 1999, 2002]. Electrons from this oxidation reaction are channeled via electron transfer flavoproteins (ETF) in conjunction with ETF:quinone oxidoreductase directly into the respiratory quinone pool [Vogt et al., 2019]. (iv) Then, homotrimeric enoyl-CoA hydratase (BbsH) adds water to the double bond of (E)-phenylitaconyl-CoA forming 2-(α-hydroxyphenyl) methylsuccinyl-CoA (synonym: 2-(α-hydroxybenzyl)succinyl-CoA) [Leuthner and Heider, 2000; von Horsten et al., 2022]. (v) The latter is then oxidized by the NAD+-dependent, heterotetrameric (S,R)-2-(α-hydroxybenzyl)succinyl-CoA dehydrogenase (BbsCD) yielding (S)-2-benzoylsuccinyl-CoA, according to enzymatic and crystal structure analyses [Leuthner and Heider, 2000; von Horsten et al., 2022]. (vi) Finally, 3-oxoacyl-CoA thiolase (BbsAB), catalyzes the cleavage of (S)-2-benzoylsuccinyl-CoA into the central intermediate benzoyl-CoA and succinyl-CoA, which can be reused for CoA transfer to benzylsuccinate and formation of the fumarate cosubstrate [Leuthner and Heider, 2000]. The activity of this novel thiolase requires both subunits, the thio­-lase family member BbsB and the Zn-finger protein BbsA, jointly positioning the CoA moiety of the substrate into the catalytically productive conformation [Weidenweber et al., 2022].

Fig. 3.

Anaerobic degradation of toluene by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Kühner et al. [2005]); for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Kube et al. [2004]) and T. aromaticaK172T (AJ 278289). Genes encoding the predicted two-component sensory/regulatory system TdiSR are marked in black and those for electron transfer proteins in brown. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 3.

Anaerobic degradation of toluene by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Kühner et al. [2005]); for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Kube et al. [2004]) and T. aromaticaK172T (AJ 278289). Genes encoding the predicted two-component sensory/regulatory system TdiSR are marked in black and those for electron transfer proteins in brown. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake and Efflux

Due to its highly lipophilic property, toluene does not require a specific uptake system, but rather enters the cell by passive diffusion. Notably, a novel type of efflux system appears to be encoded in proximity to the bbs gene cluster (EbN1_C01076-80), which was specifically detected in toluene-metabolizing cells of A. aromaticum EbN1T and probably contributes to solvent tolerance [Wöhlbrand et al., 2008].

Transcriptional Regulation

The genes encoding the protein constituents of the anaerobic degradation of toluene are organized in the two neighboring, albeit opposingly oriented operons bss (benzylsuccinate synthase) and bbs (β-oxidation of benzylsuccinate), which are preceded by the tdiSR genes coding for a presumptively toluene-specific two-component sensory/regulatory system (Fig. 3B) [Leuthner and Heider, 1998]. Differential global profiling of substrate-adapted cells of A. aromaticum EbN1T revealed specific formation of transcripts/proteins of the bss and bbs operons [Champion et al., 1999; Kühner et al., 2005]. In conjunction with the presence of conserved sequences in the promoter regions of both operons, the latter are proposed to be regulated by TdiSR in a concerted manner [Rabus et al., 2014].

Ethylbenzene, 1-Phenylethanol, and Acetophenone

Degradation Pathway

The anaerobic growth with ethylbenzene is the physio­logical signature property of A. aromaticum EbN1T. (S)-1-Phenylethanol and acetophenone represent intermediates of the degradation pathway and additionally serve as growth-supporting substrates [Rabus and Widdel, 1995; Kühner et al., 2005]. The key enzyme of this degradation pathway, ethylbenzene dehydrogenase, is archetypic for oxygen-independent alkyl chain hydroxylation of various hydrocarbons [Heider et al., 2016b]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 4A, B), which proceeds via five steps. (i) Ethylbenzene is stereospecifically hydroxy­lated at its methylene carbon forming (S)-1-phenylethanol by the novel molybdenum/iron-sulfur/heme-containing ethylbenzene dehydrogenase (EbdA1B1C1), which is localized in the periplasm [Rabus and Heider, 1998; Johnson et al., 2001; Kniemeyer and Heider, 2001a]. The crystal structure and characterization of redox centers revealed that the α-subunit (EbdA) harbors the Mo-bisMGD (molybdopterin-guanine-dinucleotide) cofactor and one [4Fe-4S] cluster (FS0), that the β-subunit (EbdB) contains three [4Fe-4S] clusters (FS1−3) and one [3Fe-3S] cluster (FS4), and that the γ-subunit (EbdC) carries the heme b cofactor [Kloer et al., 2006; Hagel et al., 2022]. Hydroxylation of ethylbenzene proceeds via two successive one-electron steps reducing the Mo-bisMGD cofactor (from MoVI to MoIV), with the electrons sequentially transferred via the FeS clusters to the heme b cofactor. Modeling approaches of the reaction indicated the involvement of two transition states, i.e., radical-type followed by carbocation-type, and that stereospecificity results from markedly faster activation of the pro(S) hydrogen compared to its pro(R) counterpart [Szaleniec et al., 2007, 2010, 2014]. Electrocatalytic activity studies with ethylbenzene dehydrogenase using hydroxymethylferrocenium as artificial electron acceptor combined with digital modeling provided new insights into the kinetics of the catalytic mechanism [Kalimuthu et al., 2015]. With a more applied perspective, comprehensive substrate and inhibitor spectra as well as production of industrial chiral alcohols by ethylbenzene dehydrogenase were investigated [Knack et al., 2012, 2013; Tataruch et al., 2014]. (ii) (S)-1-Phenylethanol is then oxidized by stereoselective (S)-1-phenylethanol dehydrogenase (Ped1) to acetophenone [Kniemeyer and Heider, 2001b]; (R)-1-phenylethanol, which also supports growth of A. aromaticum EbN1T, is oxidized by a different dehydrogenase. The crystal structure of Ped1 revealed the structural features determining enantioselectivity and steady-state kinetic analysis showed high cooperativity of the enzyme with the (S)-enantiomer [Höffken et al., 2006]. From an applied perspective, heterologously produced Ped1 was shown to reduce a wide range of prochiral ketones and 3-oxo esters to the respective enantiopure secondary alcohols [Dudzik et al., 2015]. Notably, the genome of A. aromaticum EbN1T harbors a 2ndebd-ped gene cluster, the function of which remains, however, elusive since the initial report [Rabus et al., 2005]; an unmarked in-frame deletion of ebdC2, which encodes the γ-subunit of this paralogous ethylbenzene dehydrogenase, did not show a phenotype [Wöhlbrand and Rabus, 2009]. (iii) Subsequently, acetophenone is carboxylated by a novel type of carboxylase, acetophonene carboxylase (encoded by the apc1−5 genes), to benzoylacetate, requiring 2 mol ATP per mol acetophonene [Jobst et al., 2010]. During purification, the enzyme dissociates into a heterooctameric Apc(αα′βγ)2 core complex and Apcε. The crystal structure of Apc(αα′βγ)2 allowed proposal of the following four-step reaction mechanism [Weidenweber et al., 2017]: First, Apcα binds ATP/acetophenone and Apcα′ binds ATP/bicarbonate leading to a tightened Apcαα′βγ core complex. Second, the kinase activities of Apcαα′ are triggered by Apcε attaching to the core complex, which lead to the formation of phosphoenol acetophenone and carboxyphosphate, respectively. Third, formation of electrophilic CO2 and its approximation to phosphoenol acetophenone by further tightening of the Apc complex facilitates the carboxylation reaction followed by dephosphorylation to benzoylacetate. Fourth, Apcε detaches, the pro­ducts are released, and the relaxed Apc core complex is restored. (iv) Based on proteogenomic and transcript analyses, benzoylacetate was suggested to be activated to its CoA ester by a predicted benzoylacetate-CoA ligase (Bal) [Rabus et al., 2002; Kühner et al., 2005], which was corroborated by enzyme activity measurements [Muhr et al., 2015]. (v) Finally, benzoylacetyl-CoA is suggested to be thiolytically cleaved by a 3-oxo acyl-CoA thiolase yielding the central intermediate benzoyl-CoA and acetyl-CoA [Rabus et al., 2002]; the identity of the thiolase is unclear at present.

Fig. 4.

Anaerobic degradation of ethylbenzene, (S)-1-phenylethanol, and acetophenone by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Kühner et al. [2005]); for further degradation see Fig. 15. B Assigned gene cluster (modified from Rabus et al. [2002]). Genes encoding the predicted two-component sensory/regulatory systems EdiSR and AdiSR are marked in black. C Protein inventory and proteomic coverage of the pathway.

Fig. 4.

Anaerobic degradation of ethylbenzene, (S)-1-phenylethanol, and acetophenone by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Kühner et al. [2005]); for further degradation see Fig. 15. B Assigned gene cluster (modified from Rabus et al. [2002]). Genes encoding the predicted two-component sensory/regulatory systems EdiSR and AdiSR are marked in black. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Ethylbenzene does not require a specific uptake system, since it is oxidized in the periplasm. Due to their highly lipophilic properties, neither (S)-1-phenylethanol nor acetophenone requires a specific uptake system, but rather enter the cell by passive diffusion.

Transcriptional Regulation

The genes encoding the protein constituents of the anaerobic degradation of ethylbenzene are organized in two neighboring operons in A. aromaticum EbN1T (Fig. 4B): ebd-ped (upper pathway section from ethylbenzene to acetophenone) and apc-bal (lower pathway section from acetophenone to benzoylacetyl-CoA). These two operons flank genes encoding 2 two-component sensory/regulatory systems proposed to be specific for ethylbenzene (ediSR, formerly tcs1/tcr1) and acetophenone (adiSR, formerly tcs2/tcr2), respectively [Rabus et al., 2002]. Differential profiling of transcript and protein patterns [Champion et al., 1999; Kühner et al., 2005] and of enzymatic activities [Muhr et al., 2015] indicated substrate-specific, sequential regulation of the two operons.

Phenol

Degradation Pathway

The anaerobic degradation of phenol was mainly elucidated with T. aromatica K172T and found to proceed in the same way in A. aromaticum EbN1T, based on homology and organization of genes [Breinig et al., 2000; Rabus et al., 2005] and on phenol-specific formation of involved enzyme subunits [Wöhlbrand et al., 2007]. The following metabolites, reactions, and enzymes are involved in the degradation pathway (Fig. 5A), which proceeds via three steps. (i) The initial reaction step is catalyzed by phenylphosphate synthase (PpsA1BC) that transfers the β-phosphate group of MgATP to phenol furnishing phenylphosphate. Similar to phosphoenolpyruvate synthase, a ping-pong mechanism of catalysis is proposed involving binding of diphosphate (PPi) to the catalytic His-residue in PpsA followed by release of an orthophosphate (Pi) (rendering the reaction unidirectional) and transfer of the retained Pi from the His-residue to phenol [Schmeling et al., 2004; Narmandakh et al., 2006]. (ii) Then, phenylphosphate is converted to 4-hydroxybenzoate by phenylphosphate carboxylase (PpcBDAC) via a reaction resembling a Kolbe-Schmitt carboxylation. The enzyme requires K+ and Mg2+, is highly oxygen sensitive, and requires tight subunit interactions for achieving phenylphosphate dephosphorylation (by the δ-subunit) with concomitant carboxylation (by αβγ-subunits, constituting the core carboxylase) of the intermediary enzyme-bound phenolate [Lack and Fuchs, 1992; Schühle and Fuchs, 2004; Schmeling and Fuchs, 2009]. Anaerobic conversion of catechol via catechylphosphate to protocatechuate (3,4-dihydroxybenzoate) can also be catalyzed by Pps and Ppc, indicating promiscuity of both enzymes [Ding et al., 2008]. (iii) Further degradation of 4-hydroxybenzoate to the central intermediate benzoyl-CoA is as described in section “4-Hydroxybenzoate” below.

Fig. 5.

Anaerobic degradation of phenol by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Breinig et al. [2000]); for further degradation see Fig. 11 and 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2002]) and T. aromaticaK172T (modified from Breinig et al. [2000]). The gene encoding the predicted regulator PheR is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 5.

Anaerobic degradation of phenol by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Breinig et al. [2000]); for further degradation see Fig. 11 and 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2002]) and T. aromaticaK172T (modified from Breinig et al. [2000]). The gene encoding the predicted regulator PheR is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Due to its highly lipophilic property, phenol does not require a specific uptake system, but rather enters the cell by passive diffusion.

Transcriptional Regulation

The genes coding for Pps and Ppc are organized in an operon-like structure (Fig. 5B), directly preceded by an opposingly oriented gene encoding the predicted, σ54-dependent transcriptional regulator PheR demonstrated to control expression of the pps and ppc genes [Breinig et al., 2000; Rabus et al., 2005; Buschen et al. 2023].

4-Methylphenol (p-Cresol)

Degradation Pathway

First evidence for O2-independent oxidation of 4-methylphenol at its methyl group arose from studies with Pseudomonas putida [Hopper, 1976; Hopper and Taylor, 1977], which was subsequently confirmed for several species of the genera Thauera and Aromatoleum (most species affiliating with the latter were formerly classified as Azoarcus) [Rudolphi et al., 1991]. The following metabolites, reactions, and enzymes are involved in the degradation pathway (Fig. 6A), which proceeds via three steps. (i) The initial reaction has been intensively studied in Ps. putida. It is an anaerobic hydroxylation of the methyl-group of 4-methylphenol forming 4-hydroxybenzyl alcohol followed by oxidation to 4-hydroxybenzaldehyde. These consecutive reactions are both catalyzed by the flavocytochrome enzyme termed p-cresol methylhydroxylase (PCMH), which is localized in the periplasm [Hopper et al., 1985]. Several structural studies on heterotetrameric α2β2 PCHM of Ps. putida indicated that a structural change tunes the spatial flavoprotein-cytochrome interactions for optimized intermolecular electron transfer and supported the following proposed reaction mechanism [Shamala et al., 1986; Mathews et al., 1991; McLendon et al., 1991; Cunane et al., 2000, 2005]: 4-Methylphenol is first oxidized to a quinone methide intermediate, with the released two electrons received by the flavin yielding a fully reduced FAD, wherefrom the electrons are sequentially passed on to the heme cofactor. Water is then assumed to attack (nucleophilically) the methide carbon of the intermediate, resulting in formation of 4-hydroxybenzyl alcohol, which is then oxidized by PCMH to 4-hydroxybenzaldehyde. In A. aromaticum EbN1T the conversion of 4-methylphenol to 4-hydroxybenzaldehyde is apparently catalyzed by a different albeit also FAD-dependent p-cresol methylhydroxylase (Cmh), which is apparently localized in the cytoplasm (no predicted signal peptide). (ii) 4-Hydroxybenzaldehyde is then oxidized to 4-hydroxybenzoate by 4-hydroxybenzaldehyde dehydrogenase (PchA) [Cronin et al., 1999]; this enzyme is termed Hbd in A. aromaticum EbN1T. (iii) Further degradation of 4-hydroxybenzoate to the central intermediate benzoyl-CoA is as described in section “4-Hydroxybenzoate” below.

Fig. 6.

Anaerobic degradation of 4-methylphenol (p-cresol) by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Vagts et al. [2020]); for further degradation see Fig. 11 and 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Vagts et al. [2020]) and P. denitrificansNCIMB 9866 (modified from Kim et al. [1994]). Genes encoding the two-component sensory/regulatory system PcrSR are marked in black. seq %id., amino acid sequence identities. C Protein inventory and proteomic coverage of the pathway.

Fig. 6.

Anaerobic degradation of 4-methylphenol (p-cresol) by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Vagts et al. [2020]); for further degradation see Fig. 11 and 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Vagts et al. [2020]) and P. denitrificansNCIMB 9866 (modified from Kim et al. [1994]). Genes encoding the two-component sensory/regulatory system PcrSR are marked in black. seq %id., amino acid sequence identities. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Due to its highly lipophilic property, 4-methylphenol does not require a specific uptake system, but rather enters the cell by passive diffusion [Vagts et al., 2020].

Transcriptional Regulation

The structural genes pchC and pchF encoding the cytochrome (α) and flavoprotein (β) subunits of PCMH, respectively, and pchA encoding 4-hydroxybenzaldehyde dehydrogenase were first determined in Ps. putida [Kim et al., 1994; Cronin et al., 1999]; the function of the pchX gene additionally present in the pch gene cluster is still unknown. Homologs of the pchCF genes were also found in the genome of A. aromaticum EbN1T, but assigned to the anaerobic degradation of 4-ethylphenol and renamed as emhCF genes (see below). Transcriptional regulation of the pch operon in Pseudomonas spp. was suggested to be mediated by a one-component regulatory protein (DmpR/PchR) [Pavel et al., 1994; Jõesaar et al., 2010], belonging to the σ54-dependent regulators of the XylR/DmpR-family [Tropel and van der Meer, 2004]. By contrast, in A. aromaticum EbN1T transcription of the cmh-hbd genes (Fig. 6B) is controlled by the σ54-dependent two-component sensory/regulatory system PcrSR, as evidenced by unmarked, in-frame deletion mutagenesis [Vagts et al., 2020]. The in vivo responsiveness of A. aromaticum EbN1T toward p-cresol falls in the nanomolar range (<100 nM), as determined by targeted, quantitative transcript analyses [Vagts et al., 2020].

4-Ethylphenol and 4-Hydroxyacetophenone

Degradation Pathway

The anaerobic growth of A. aromaticum EbN1T with 4-ethylphenol represents a physiological trait, which was not predicted during initial genome annotation and only subsequently discovered by differential proteomics. Likewise, 1-(4-hydroxyphenyl)ethanol and 4-hydroxyacetophenone were recognized as degradation intermediates as well as growth-supporting substrates [Wöhlbrand et al., 2008]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 7A, B), which proceeds via five steps as previously proposed based on differential proteomic and metabolite analyses [Wöhlbrand et al., 2008]. (i) In the initial reaction, 4-ethylphenol is hydroxylated at the methylene carbon forming (R)-1-(4-hydroxyphenyl)ethanol catalyzed by 4-ethylphenol methylenehydroxylase (EmhCF, revised from original PchCF annotation) of A. aromaticum EbN1T. EmhCF is similar to PCMH of Ps. putida NCIMB 9866, which was previously demonstrated to also convert 4-ethylphenol [McIntire et al., 1985]. Thus, intermediary formation of a quinone methide appears plausible, too, as inferred from the redox potential of the flavin cofactor of ethylphenol methylhydroxylase from Ps. putida JD1 [Reeve et al., 1989]. EmhCF is probably localized in the cytoplasm, since the large, flavin cofactor-containing F-subunit is apparently devoid of a signal peptide and only for the small C subunit a Sec-type signal peptide can be predicted. (ii) (R)-1-(4-Hydroxyphenyl)ethanol is then further oxidized to 4-hydroxyacetophenone by (R)-specific 1-(4-hydroxyphenyl)ethanol dehydrogenase (Hped1, formerly ChnA), as demonstrated by unmarked, in-frame mutagenesis and analysis of the crystal structure [Büsing et al., 2015a]. Molecular genetic analysis further demonstrated that the dehydrogenase encoded by the hped2 gene (directly neighboring the hped1 gene) does not serve this function [Büsing et al., 2015a]. Prompted by the biotechnological potential of Hped1, the pH-dependent stability of this enzyme was recently investigated under storage and in-process conditions [Tataruch et al., 2023]. (iii) p-Hydroxyacetophenone is then proposed to be converted to 4-hydroxybenzoylacetate by the carboxylase HacABC (formerly XccABC) [Wöhlbrand et al., 2008], which should first carboxylate its biotin cofactor followed by transcarboxylation of the actual substrate as known from other well-studied biotin-dependent carboxylases [Attwood and Wallace, 2002]. (iv) Further degradation is proposed to proceed via activation of 4-hydroxybenzoylacetate to its CoA ester catalyzed by the acetoacetyl-CoA ligase-like protein AcsA1, followed by thiolytic removal of an acetyl-CoA moiety by the predicted thiolase TioL1 yielding 4-hydroxybenzoyl-CoA. (v) Further degradation of 4-hydroxybenzoyl-CoA to the central intermediate benzoyl-CoA is as described in section “4-Hydroxybenzoate” below. Notably, the proposed degradation pathway from 4-ethylphenol to 4-hydroxy-benzoyl-CoA is analogous to the one described above for ethylbenzene, albeit involving a completely different set of enzymes.

Fig. 7.

Anaerobic degradation of 4-ethylphenol and 4-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Vagts et al. [2020]); for further degradation see Fig. 11 and 15. B Assigned gene cluster (modified from Wöhlbrand et al. [2008]). The gene encoding the EtpR sensor/regulator is marked in black. C Protein inventory and proteomic coverage of the pathway.

Fig. 7.

Anaerobic degradation of 4-ethylphenol and 4-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Vagts et al. [2020]); for further degradation see Fig. 11 and 15. B Assigned gene cluster (modified from Wöhlbrand et al. [2008]). The gene encoding the EtpR sensor/regulator is marked in black. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake and Efflux

Due to its highly lipophilic property, 4-ethylphenol does not require a specific uptake system, but rather enters the cell by passive diffusion [Vagts et al., 2020]. This can also be envisioned for 1-(4-hydroxyphenyl)ethanol and 4-hydroxyacetophenone. Genes encoding a novel type of efflux system (EbN1_C001730-60) are located upstream of the acsA1 gene. This efflux system is homologous to the one for toluene, but specifically formed in 4-ethylphenol or 4-ethylacetophenone-metabolizing cells of A. aromaticum EbN1T, and probably contributes to the strain’s solvent tolerance [Wöhlbrand et al., 2008].

Transcriptional Regulation

The genes encoding all above described enzymes for the anaerobic degradation of 4-ethylphenol are localized in a single large operon-like structure (Fig. 7B) [Wöhlbrand et al., 2008]. Differential abundance and activity profiles demonstrated the substrate-specific formation of these enzymes [Wöhlbrand et al., 2008; Muhr et al., 2015]. The direct genomic neighborhood of the etpR gene (formerly ebA324) in conjunction with the presence of the 12/24 consensus sequence of σ54-dependent promotors [Helmann and Chamberlin, 1988] upstream of acsA1 as the 1st gene of the operon, suggests EtpR as its transcriptional regulator [Wöhlbrand et al., 2008]. In accord, unmarked, in-frame deletion of the etpR gene resulted in loss of the capacity for anaerobic growth with 4-ethylphenol or 4-hydroxyacetophenone, additionally underpinning its capacity to recognize both aromatic compounds [Büsing et al., 2015b]. The in vivo responsiveness of A. aromaticum EbN1T toward them falls in the lower nanomolar range, as determined by targeted, quantitative transcript analyses [Vagts et al., 2018, 2020].

2-Phenylethylamine and Benzylamine

Degradation Pathway

The aromatic primary amines 2-phenylethylamine and benzylamine were recently recognized to support robust anaerobic growth of A. aromaticum EbN1T [Schmitt et al., 2019]. While amine dehydrogenases are long known for the degradation of aliphatic and aromatic amines in Pseudomonas spp. [e.g., Durham and Perry, 1978; Iwaki et al., 1983] the findings with A. aromaticum EbN1T represent the first example of an anaerobic metabolism. Here, the oxidative deamination of 2-phenylethylamine and benzylamine (both Fig. 8A) is catalyzed by two specific periplasmic quinohemoprotein amine dehydrogenases [Schmitt et al., 2019] characterized by an encaged quinoid cofactor as known from Paracoccus denitrificans [Datta et al., 2001; Fujieda et al., 2003] and Ps. putida [Satoh et al., 2002]. 2-Phenylamine is converted to phenylacetaldehyde by 2-phenylethylamine dehydrogenase (2-PEADH) (QhpA1B1C1), which has an α2β2γ2-composition and a rather broad substrate spectrum including also aliphatic amines. Further catabolism branches into the pathway for degradation of phenylalanine (see below). Benzylamine is converted to benzaldehyde by benzylamine dehydrogenase (BAmDH) (QhpA2B1C2), which has an αβγ-composition and a restricted substrate spectrum with only 2-phenylethylamine as additional substrate. Further catabolism branches into the pathway for benzaldehyde degradation (see below). 2-PEADH and BAmDH contain 2 heme c cofactors in their α-subunits and the cysteine tryptophylquinone (CTQ) cofactor. The QhpDEF-paralogs encoded in both qhp operons, are assumed to function as maturation factors. QhpD1 and QhpD2 are S-adenosylmethionine (SAM)-radical enzymes presumably forming the thioether bonds in the γ-subunits [Nakai et al., 2015]. QhpF1 and QhpF2 are membrane-bound proteins possibly involved in exporting the modified γ-subunits to the periplasm [Nakai et al., 2014]. QhpE1 and QhpE2 are predicted serine proteases assumed to cleave the nonstandard signal peptides of the γ-subunits [Nakai et al., 2012]. QhpG is only encoded in the qhp2 operon (BAmDH) and represents a putative FAD-dependent monooxygenase possibly involved in forming the quinoid cofactor in the γ-subunit [Nakai et al., 2014]. The CcmF protein, which is also only encoded in the qhp2 operon (BAmDH), is a predicted membrane-bound cytochrome c maturation protein possibly delivering heme c to the α-subunit in the periplasmic space [Pearce et al., 1998]. Each operon (qhp1 and qhp2) harbors a gene (qhpF1 and qhpF2) encoding the permease of an ABC-type transporter (hybrid with ATPase domain); each permease is assumed to translocate its cognate γ-subunit (qhpC1 and qhpC2) to the periplasm.

Fig. 8.

Oxidative deamination of 2-phenylethylamine and benzylamine by A. aromaticumEbN1T. A Proposed reaction for 1-phenylethylamine (left) and benzylamine (right) degradation; for further degradation see Fig. 15 and 18. B Assigned gene clusters of A. aromaticumEbN1T and Pc. denitrificans. The gene encoding the predicted QhpR regulator is marked in black and the genes for the two permeases in gray. seq %id., amino acid sequence identities. Parts A and B were modified from Schmitt et al. [2019]. C Protein inventory and proteomic coverage of the pathway.

Fig. 8.

Oxidative deamination of 2-phenylethylamine and benzylamine by A. aromaticumEbN1T. A Proposed reaction for 1-phenylethylamine (left) and benzylamine (right) degradation; for further degradation see Fig. 15 and 18. B Assigned gene clusters of A. aromaticumEbN1T and Pc. denitrificans. The gene encoding the predicted QhpR regulator is marked in black and the genes for the two permeases in gray. seq %id., amino acid sequence identities. Parts A and B were modified from Schmitt et al. [2019]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Neither of the two aromatic amines requires a dedictated uptake system since their initial oxidative deamination occurs in the periplasm. The unpolar products of this reaction (phenylacetaldehyde and benzaldehyde) can be assumed to enter the cell by passive diffusion.

Transcriptional Regulation

The genes coding for 2-PEADH and BAmDH and their biogenesis are located in two distinct operon-like structures in the genome of A. aromaticum EbN1T (Fig. 8B) [Schmitt et al., 2019], resembling the gene organization identified in Pc. denitrificans [Nakai et al., 2014]. Only the qhp2 operon (BAmDH) harbors the qhpR gene encoding an AraC-type regulator, which activates transcription in response to the aromatic amine substrate; notably BAmDH was most strongly induced and active with benzylamine. Whether the qhp1 operon is also under control of QhpR is unclear at present.

o-Phthalate

Degradation Pathway

The anaerobic degradation of o-phthalate (monomer of industrially produced phthalate esters) was discovered by differential proteomic analyses in members of the genera Aromatoleum and Thauera, and shown to be plasmid-born in A. aromaticum EbN1T [Ebenau-Jehle et al., 2017]. The following metabolites, reactions, and enzymes are involved in the degradation pathway (Fig. 9A), which proceeds via two steps. (i) o-Phthalate is initially activated to phthaloyl-CoA by the highly specific succinyl-CoA:phthalate CoA transferase (SptAB), with the formed CoA ester intermediate showing an extreme lability [Mergelsberg et al., 2018]. (ii) Phthaloyl-CoA is then decarboxylated by phthaloyl-CoA decarboxylase (Pcd1/2) generating the central intermediate benzoyl-CoA. Pcd of T. chlorobenzoica was demonstrated to have a hexameric structure and to contain prenylated FMN, K+ and Fe2+ as cofactors. High intracellular concentrations of Pcd in conjunction with the irreversibility of the catalyzed decarboxylation reaction allow for efficient o-phthalate conversion despite the lability of the initial phthaloyl-CoA intermediate [Mergelsberg et al., 2017, 2018]. The sptAB and pcd1 genes form an operon-like structure in the genome of A. aromaticum EbN1T (Fig. 9B), with the pcd2 gene localized 11.7 kbp upstream of the pcd1 gene.

Fig. 9.

Anaerobic degradation of o-phthalate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T and T. chlorobenzoica. The gene encoding a predicted regulator is marked in black and those coding for two uptake systems in gray. seq %id., amino acid sequence identities. Parts A and B were modified from Ebenau-Jehle et al. [2017] and Mergelsberg et al. [2017, 2018]. C Protein inventory and proteomic coverage of the pathway.

Fig. 9.

Anaerobic degradation of o-phthalate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T and T. chlorobenzoica. The gene encoding a predicted regulator is marked in black and those coding for two uptake systems in gray. seq %id., amino acid sequence identities. Parts A and B were modified from Ebenau-Jehle et al. [2017] and Mergelsberg et al. [2017, 2018]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The sptAB and pcd1 genes are flanked by genes encoding a TRAP transporter (DctMQ and DctP4) and an ABC transporter (EbN1_PII00660/70/80) (Fig. 9B) [Ebenau-Jehle et al., 2017], which could both be involved in the uptake of o-phthalate and possibly further substrates of the Spt-Pcd degradation pathway. Since, however, only the TRAP-transporter is encoded next to the sptAB and pcd genes in T. chlorobenzoica, it may be the obvious candidate for the uptake of o-phthalate.

Transcriptional Regulation

The spt-pcd genetic neighborhood also harbors a gene coding for an IclR-type one-component sensory/regulatory system (EbN1_PII00710) (Fig. 9B) that could control spt-pcd gene expression in response to o-phthalate, in particular as this gene is also conserved and shows a similar genetic neighborhood in T. chlorobenzoica [Ebenau-Jehle et al., 2017].

Indoleacetate (Auxin)

Degradation Pathway

The anaerobic degradation of indoleacetate (IAA; synonym: indole-3-acetate), which is also known as the plant hormone auxin, was elucidated in Aromatoleum evansii KB740T (formerly A. evansii KB740T) and A. aromaticum EbN1T. The following metabolites, reactions, and enzymes are involved in the degradation pathway (Fig. 10A), which proceeds via nine steps [Ebenau-Jehle et al., 2012; Schühle et al., 2016, 2021]. (i) In the initial reaction, indoleacetate is activated to its CoA ester by a highly specific indoleacetate-CoA ligase (IaaB). (ii) Indoleacetyl-CoA is then anaerobically hydroxylated to the enol form of 2-oxo-indoleacetyl-CoA by a predicted molybdenum cofactor-containing, heterotrimeric dehydrogenase belonging to the xanthine dehydrogenase family (IaaIJK). (iii) The keto-form of 2-oxo-indoleacetyl-CoA is assumed to be hydrolytically cleaved by an ATP-dependent hydrolase (IaaCE) forming 3′-(2-aminophenyl)succinyl-CoA. (iv) The latter is then isomerized by a phenylsuccinyl-CoA transferase (IaaL) to its regioisomer 2′-(2-aminophenyl)succinyl-CoA, the C-skeleton of which is (v) then rearranged by a methylmalonyl-CoA mutase-like enzyme (IaaGH) to 2-aminobenzylmalonyl-CoA. (vi) The carboxyl group in α-position of 2-aminobenzylmalonyl-CoA is then proposed to be oxidatively removed by a benzylmalonyl-CoA dehydrogenase (IaaF) forming (2-aminophenyl)propenoyl-CoA. The latter is then further catabolized according to a β-oxidation-like reaction sequence via (vii) hydration to 3-hydroxy-3-(2-aminophenyl)-propanoyl-CoA followed by (viii) oxidation to 3-oxo-3-(2-aminophenyl)-propanoyl-CoA catalyzed by a fused enoyl-CoA hydratase/3-hydroxyacyl-CoA dehydrogenase (IaaP). (ix) Finally, cleaving off acetyl-CoA by a 3-oxoacyl-CoA thiolase (IaaA) yields 2-aminobenzoyl-CoA. Further degradation of 2-aminobenzoyl-CoA to the central intermediate benzoyl-CoA is as described in section “2-Aminobenzoate” below.

Fig. 10.

Anaerobic degradation of indoleacetate (auxin) by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene cluster. Genes encoding predicted regulators and an uptake system are marked in black and gray, respectively. Parts A and B were modified from Ebenau-Jehle et al. [2012] and Schühle et al. [2016]. C Protein inventory and proteomic coverage of the pathway.

Fig. 10.

Anaerobic degradation of indoleacetate (auxin) by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene cluster. Genes encoding predicted regulators and an uptake system are marked in black and gray, respectively. Parts A and B were modified from Ebenau-Jehle et al. [2012] and Schühle et al. [2016]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The iaa gene cluster (Fig. 10B) harbors the iaaM gene coding for the periplasmic solute-binding protein of an ABC transporter, potentially being involved in high-affinity uptake of indoleacetate.

Transcriptional Regulation

All genes encoding proteins for the anaerobic degradation of indoleacetate are organized in a coherent cluster in the genome of A. aromaticum EbN1T (Fig. 10B). The cluster harbors the iaaQ and iaaR genes encoding transcriptional regulators belonging to the TetR [Cuthbertson and Nodwell, 2013] and GntR families [Rigali et al., 2002], respectively.

4-Hydroxybenzoate

Degradation Pathway

The anaerobic degradation of 4-hydroxybenzoate, was mainly elucidated in T. aromatica K172T and R. palustris [e.g., Breese and Fuchs, 1998; Egland and Harwood, 2000]. The following metabolites, reactions, and enzymes are involved in the degradation pathway (Fig. 11A), which proceeds via two steps. (i) In the initial reaction, 4-hydroxybenzoate is activated to its CoA ester by predicted Mg2+/ATP-dependent 4-hydroxybenzoate-CoA ligase (Hcl2), as previously reported for characterized homologs in T. aromatica K172T and R. palustris [Biegert et al., 1993; Gibson et al., 1994]. Note, that 4-hydroxybenzoyl-CoA is also generated during the anaerobic degradation of 4-methylphenol, 4-ethylphenol, and para-hydroxylated phenylpropanoids. (ii) 4-Hydroxybenzoyl-CoA is then reductively dehydroxylated to the central intermediate benzoyl-CoA by the molybdenum-containing 4-hydroxybenzoyl-CoA reductase (HcrABC) [Gibson et al., 1997; Breese and Fuchs, 1998], which was to date, however, not detected in anaerobic cultures of A. aromaticum EbN1T by differential proteomics [e.g., Wöhlbrand et al., 2007]. Heterodimeric HcrABC belongs to the xanthine oxidase (XO) family of molybdenum enzymes, possesses an (αβγ)2 structure, harbors two molybdenum cofactors, four [2Fe-2S] and two [4Fe-4S] clusters as well as two FADs. Ferredoxin serves as electron donor and the active site of Hcr has a sidechain lining that is assumed to stabilize the postulated radical intermediate. Two successive one-electron steps via the molybdenum cofactor (MoIV to MoVI) to 4-hydroxybenzoyl-CoA according to a Birch-like mechanism are proposed to be involved. MoIV is probably regenerated by a two electron and two proton transfer after release of the re-aromatized benzoyl-CoA [Breese and Fuchs, 1998; Boll et al., 2001; Unciuleac et al., 2004a, b; Johannes et al., 2008].

Fig. 11.

Anaerobic degradation of 4-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Boll et al. [2001]); for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2005]) and T. aromaticaK172T (AJ 278289). The gene for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 11.

Anaerobic degradation of 4-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Boll et al. [2001]); for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2005]) and T. aromaticaK172T (AJ 278289). The gene for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

In Ps. putida and Acinetobacter sp. ADP1, PcaK, which belongs to the aromatic acid:H+ symporters (AAHS) of the major facilitator superfamily (MFS) [Pao et al., 1998], has been reported to function in uptake of 4-hydroxy-benzoate [Harwood et al., 1994; Pernstich et al., 2014]. The genome of A. aromaticum EbN1T encodes a (di)hydroxybenzoate transporter (PcaK); however, the pcaK gene is not adjacent to the genes encoding enzymes for anaerobic degradation of 4-hydroxybenzoate (Fig. 11B).

Transcriptional Regulation

In R. palustris, the HbaR protein, which belongs to the Crp-Fnr superfamily [Körner et al., 2003], activates the transcription of the genes for anaerobic degradation of 4-hydroxybenzoate [Egland and Harwood, 2000]. While the genome of A. aromaticum EbN1T does not contain a gene unambiguously encoding a HbaR homolog in direct neighborhood to the “catabolic” genes, a TaoR-like transcriptional regulator belonging to the LysR-family is encoded (Fig. 11B) [Maddocks and Oyston, 2008; Rajeev et al., 2019].

3-Hydroxybenzoate

Degradation Pathway

Anaerobic degradation of 3-hydroxybenzoate was mainly investigated with T. aromatica K172T [Laempe et al., 2001] and A. aromaticum EbN1T [Rabus et al., 2005; Wöhlbrand et al., 2007]. The current differential metabolite, transcript, and proteomic data suggest a pathway deviating from the one hypothesized earlier [Wöhlbrand et al., 2007]. Next to the enzymes encoded in the “3-hydroxybenzoate” gene cluster also Oah from the central benzoyl-CoA pathway is involved in feeding into 3-hydroxypimeloyl-CoA. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 12A, B), which proceeds via eight steps. (i) Early studies with T. aromatica K172T revealed inducible formation of a 3-hydroxybenzoate-CoA ligase (Hcl1) activating 3-hydroxybenzoate to 3-hydroxybenzoyl-CoA [Laempe et al., 2001], which was subsequently corroborated by means of differential proteomics with A. aromaticum EbN1T [Wöhlbrand et al., 2007]. (ii) Then, 3-hydroxybenzoyl-CoA is reductively dearomatized by a paralog of class I BCR (termed HbrADBC) forming 3-hydroxydienoyl-CoA, i.e., the meta-hydroxy group is retained. Please note that the paralogous subunit pairs BcrA/HbrA and BcrD/HbrD share 100% sequence identity. Therefore, abundance profiles reflect the pool size in each case, hindering the assignment of the specific abundance value to each individual paralog. (iii) The 3-hydroxy-dienoyl-CoA is then reduced to 5-hydroxycyclohex-2-ene-1-carbonyl-CoA by a predicted (and experimentally detected) dehydrogenase (EbN1_C03840), which is probably FAD-dependent. (iv) The 5-hydroxy-enoyl-CoA is then hydrated to 2,5-dihydroxycyclohex-1-carbonyl-CoA by a predicted (and experimentally detected) enoyl-CoA hydratase/isomerase (EbN1_C03750 or EbN1_C03770). (v) 2,5-Dihydroxycyclohex-1-carbonyl-CoA is then assumed to be dehydrated to 6-hydroxy-cyclohex-2-ene-1-carbonyl-CoA again by either of the two enoyl-CoA hydratases/isomerases (EbN1_C03750 and EbN1_C03770). (vi) Then water is added to the double bond of 6-hydroxycyclohex-2-ene-1-carbonyl-CoA yielding 2,6-dihydroxycyclohex-1-carbonyl-CoA, a reaction presumably catalyzed again by either of the two enoyl-CoA hydratases/isomerases (EbN1_C03750 and EbN1_C03770) or possibly by another enoyl-CoA hydratase (Dch); however, Dch is a constituent of the anaerobic benzoyl-CoA pathway. Note that the speci­ficity of these two enoyl-CoA hydratases/isomerases (EbN1_C03750 and EbN1_C03770) for substrate positions 2, 5, and 6 subsequently targeted during the three aforementioned reactions cannot be differentiated at present [Becker et al., 2022a]. The hydroxy-group shift from C5 to C6 is reminiscent of the hydroxy-group shift occurring during catabolism of levulinic acid in Ps. putida KT2440, where 4-hydroxyvaleryl-CoA is converted to 3-hydroxyvaleryl-CoA involving 4-phosphovaleryl-CoA as key intermediate [Rand et al., 2017]. Such a mechanism is most likely not involved in A. aromaticum EbN1T since the key enzyme phosphoryl transferase (LvaAB) is not encoded in its genome. (vii) 2,6-Dihydroxycyclohex-1-carbonyl-CoA is then oxidized to 2-oxo-6-hydroxycyclohex-1-carbonyl-CoA by 6-hydroxycyclohex-1-ene-1-carbonyl-CoA dehydrogenase (EbN1_C03780), which is specifically detected in cells adapted to anaerobic growth with 3-hydroxybenzoate. Alternatively, a similarily specifically detected oxoacyl-CoA reductase (EbN1_C03850) could be involved in this reaction step, but also functions in a possible side route of anaerobic degradation of 3-hydroxybenzoate [Becker et al., 2022a]. (viii) Finally, the ring structure of 2-oxo-6-hydroxycyclohex-1-carbonyl-CoA is hydrolytically cleaved to 3-hydroxy-pimelyl-CoA by 6-oxocyclohex-1-ene-1-carbonyl-CoA hydratase (Oah), recruited from the central benzoyl-CoA pathway.

Fig. 12.

Anaerobic degradation of 3-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Laempe et al. [2001]). B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2005]) and T. aromaticaK172T (AJ 278289). Genes coding for a predicted regulator and two uptake systems are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 12.

Anaerobic degradation of 3-hydroxybenzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Laempe et al. [2001]). B Assigned gene clusters of A. aromaticumEbN1T (modified from Rabus et al. [2005]) and T. aromaticaK172T (AJ 278289). Genes coding for a predicted regulator and two uptake systems are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The chromosomal locus for anaerobic degradation of 3-hydroxybenzoate encodes, next to the catabolic enzymes described above, also two uptake systems. First, in-between EbN1_C03780 and hcl1 the EbN1_C03790 gene is localized, which codes for a predicted MFS transporter. However, only low transcript and protein abundances could be observed [Becker et al., 2022a]. Second, in-between the hcl1 and EbN1_C03840 genes, the tarPMQ genes are localized, which code for a TRAP transporter (lignin-derived aromatic compounds). Notably, transcripts and protein components of this transporter are abundant during anaerobic as well as aerobic growth with 3-hydroxybenzoate [Becker et al., 2022a]. This implies that the TarPMQ system could be responsible for the uptake of this aromatic growth substrate.

Transcriptional Regulation

Downstream of the EbN1_C03750 gene, the EbN1_C03740 gene is localized, which encodes an TetR-like transcriptional regulator. The latter could thus control transcription of the genes for the uptake and anaerobic degradation of 3-hydroxybenzoate.

Phenylacetate

Degradation Pathway

The anaerobic degradation of phenylacetate (via phenyl­acetyl-CoA) was elucidated by means of biochemical studies with A. evansii KB740T and T. aro-matica K172T. The genes encoding the protein constituents of this pathway cluster together in an operon-like structure and have been identified in the genomes of A. evansii KB740T and A. aromaticum EbN1T [Rabus et al., 2005; Wöhlbrand et al., 2007]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 13A, B), which proceeds via three steps. (i) The initial step of anaerobic phenylacetate degradation is the activation to phenylacetyl-CoA by phenylacetate-CoA ligase (PadJ), as first demonstrated with T. aromatica K172T [Mohamed and Fuchs, 1993]. Note that anaerobic degradation of phenylalanine branches in at this point. (ii) Phenylacetyl-CoA is then subject to a four-electron α-oxidation yielding phenylglyoxylate. This reaction is catalyzed by heterotrimeric phenylacetyl-CoA:acceptor oxidoreductase (synonym: phenylacetyl-CoA dehydrogenase) (PadBCD), which was discovered and characterized in T. aromatica K172T. The enzyme is membrane-bound, presumably uses ubiquinone as an electron acceptor, contains molybdenum as well as iron-sulfur clusters, and has an apparent (αβγ)2 structure. While the enzyme is remarkably stable against exposure to oxygen, it is not formed during aerobic growth with phenylalanine. Intermediary released phenylglyoxylyl-CoA is proposed to be cleaved by the same enzyme to phenylglyoxylate and CoA. Formation of phenylglyoxylyl-CoA and phenylglyoxylate was demonstrated with enzyme assays provided with [1-14C] phenylacetyl-CoA. Based on homology analyses, the large subunit (PadB) is suggested to carry the molybdenum cofactor [Schneider and Fuchs, 1998; Rhee and Fuchs, 1999]. Specific formation of PadA by A. aromaticum EbN1T [Becker et al., 2022a] may hint to a functional role, which is most likely that of a chaperone during maturation of the bis-MGD cofactor of PadBCD, as known from homologous FdhD chaperone of formate dehydrogenase in E. coli [Arnoux et al., 2015; Schwanhold et al., 2018]. In A. aromaticum EbN1T, the predicted thioesterase EbN1_C31020 (here named PadK) was detected with highest abundances (transcript and protein level) during anaerobic growth with phenylalanine [Becker et al., 2022a]. Moreover, the padK gene is part of the pad gene locus. Taken together, the PadK protein is the more suitable candidate for catalyzing the removal of CoA from phenylglyoxylyl-CoA compared to PadBCD, also considering that the latter does not possess a recognizable domain for such an enzymatic activity. A further alternative thioesterase, based on transcript/protein abundance profiles, could be EbN1_C20160 [Becker et al., 2022a]. (iii) Finally, oxidative decarboxylation of phenylglyoxylate to the central intermediate benzoyl-CoA is achieved by multi-subunit, FeS-containing phenylglyoxylate:NAD+ oxidoreductase (PadEFGHI), purified and biochemically characterized in A. evansii KB740T. The core enzyme is suggested to have an α2β2γ2δ2 composition with the additional ε2 units transferring electrons from a small ferredoxin-like subunit of the core enzyme to NAD+. The enzyme was found to be specifically formed in cells growing anaerobically with phenylalanine, phenylacetate or phenylglyoxylate. The enzyme was found to be highly oxy­gen-sensitive and not formed in cells grown aerobically with phenylalanine [Hirsch et al., 1998].

Fig. 13.

Anaerobic degradation of phenylacetate to the level of benzoyl-CoA by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Schneider and Fuchs [1998]); for further degradation see Fig. 15. B Assigned gene cluster of A. aromaticumEbN1T [Rabus et al., 2005] compared to that of A. evansiiKB740T (AJ428571). Note, that the second padgene cluster (EbN1_C32400-20) present in the genome of A. aromaticumEbN1T is not depicted here, since neither respective transcripts nor proteins were detected to date. The gene coding for the presumptive transcriptional repressor PadR is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 13.

Anaerobic degradation of phenylacetate to the level of benzoyl-CoA by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Schneider and Fuchs [1998]); for further degradation see Fig. 15. B Assigned gene cluster of A. aromaticumEbN1T [Rabus et al., 2005] compared to that of A. evansiiKB740T (AJ428571). Note, that the second padgene cluster (EbN1_C32400-20) present in the genome of A. aromaticumEbN1T is not depicted here, since neither respective transcripts nor proteins were detected to date. The gene coding for the presumptive transcriptional repressor PadR is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The genetic neighborhood of the pad gene cluster provides no hints for a possible uptake system for phenylace­tate.

Transcriptional Regulation

The padR gene is localized less than 1 kbp upstream of the pad gene clusters (opposite strand). The encoded PadR protein probably functions as transcriptional repressor, since it shares a domain with the PaaX protein, which was previously reported to repress transcription of the paa gene cluster for aerobic degradation of phenylacetate in E. coli [Ferrández et al., 2000].

2-Aminobenzoate (Anthranilate)

Degradation Pathway

Anaerobic degradation of 2-aminobenzoate has been investigated in T. aromatica K172T and A. aromaticum EbN1T and was suggested to proceed via two steps (Fig. 14A). (i) The degradation is initiated by benzoate-CoA ligase, activating 2-aminobenzoate to its CoA ester (see above), albeit at a substantially lower rate compared to benzoate; the CoA ligase apparently functions anaerobically as well as aerobically [Schühle et al., 2003]. Further degradation then separates into an anaerobic and an aerobic branch. (ii) Under anoxic conditions, 2-aminobenzoyl-CoA is then assumed to be reductively dearomatized by HbrCABD yielding 6-aminocyclohex-1-ene-1-carbonyl-CoA [Wöhlbrand et al., 2007]. The amino group is then spontaneously removed in the presence of water yielding 6-oxocyclohex-1-ene-1-carbonyl-CoA, which represents a downstream intermediate of the central anaerobic benzoyl-CoA pathway.

Fig. 14.

Anaerobic degradation of 2-aminobenzoate (anthranilate) by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T. The genes coding for the DctPM uptake system are marked in gray. Parts A and B were modified from Wöhlbrand et al. [2007]. C Protein inventory and proteomic coverage of the pathway.

Fig. 14.

Anaerobic degradation of 2-aminobenzoate (anthranilate) by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15. B Assigned gene clusters of A. aromaticumEbN1T. The genes coding for the DctPM uptake system are marked in gray. Parts A and B were modified from Wöhlbrand et al. [2007]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The gene cluster framed by the hcl1 and hbrCABD genes also encodes a TRAP-type uptake system (dctPM), which could function in import of 2-aminobenzoate (Fig. 14B).

Transcriptional Regulation

No hints for a possible sensory/regulatory system responsive to 2-aminobenzoate can be inferred from the genetic neighborhood.

Benzoate

Degradation Pathway

Insights into the challenging anaerobic degradation of benzoate by denitrifying bacteria mainly emerged from biochemical investigations using T. aromatica K172T. Genes coding for proteins involved in this pathway were determined for several bacteria, e.g., R. palustris [Egland et al., 1997], T. aromatica K172T [Breese et al., 1998; Schühle et al., 2003], Azoarcus sp. strain CIB (needs reclassification as Aromatoleum sp.) [López Barragán et al., 2004], and A. aromaticum EbN1T [Rabus et al., 2005]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 15A, B), which proceeds via five steps. (i) Benzoate is activated to benzoyl-CoA by ATP-dependent benzoate-CoA ligase (Bcl) [Breese et al., 1998]. (ii) The reductive dearomatization of benzoyl-CoA via a Birch-like reaction yielding trans-cyclohex-1,5-diene-1-carbonyl-CoA (dienoyl-CoA) is catalyzed by Mg2+/ATP-dependent, FeS cluster-containing, oxygen-sensitive, four-subunit benzoyl-CoA reductase (BcrCBAD). The dienoyl-CoA product (from [14C]benzoyl-CoA) was identified by TLC. The electrons for ring reduction are donated by reduced ferredoxin (Fxd), which is delivered by a 2-oxoglutarate:ferredoxin oxidoreductase (KorAB) [Boll and Fuchs, 1995; Boll et al., 1997; Dörner and Boll, 2002; Möbitz and Boll, 2002; Thiele et al., 2008]. While fdx and korAB genes are localized in proximity to the bcr genes in T. aromatica K172T, this is only the case for fdx with A. aromaticum EbN1T. However, a 2-oxoglutarate:ferredoxin oxidoreductase is also encoded in the genome of A. aromaticum EbN1T (EbN1_C26200-20); while no differential protein abundance profiles were observed, transcript profiles were 5‒8-fold higher during anaerobic growth with phenylalanine, benzoate, 3-hydroxybenzoate, and 3-(4-hydroxyphenyl)propanoate as compared to the other tested substrate adaptation conditions [Becker et al., 2022a]. Notably, members of the genera Thauera and Aromatoleum possess different subtypes of class I BCR [e.g., Buckel et al., 2014]. (iii) The dienoyl-CoA is then hydrated to 6-hydroxycyclohex-1-ene-1-carbonyl-CoA catalyzed by cyclo­hex-1,5-diene-1-carbonyl-CoA hydratase (Dch). The metabolite identification was based on reactions with [ring-13C6]benzoyl-CoA or [ring-14C]benzoyl-CoA and the purified enzyme, followed by HPLC analysis and NMR spectroscopy [Laempe et al., 1998; Boll et al., 2000]. (iv) 6-Hydroxycyclohex-1-ene-1-carbonyl-CoA is oxidized to 6-oxocyclohex-1-ene-1-carbonyl-CoA by NAD+-dependent 6-hydroxycyclohex-1-ene-1-carbonyl-CoA dehydrogenase (Had). Metabolite identification was based on enzymatic assays with [phenyl-14C]6-hydroxycyclohex-1-ene-1-carbonyl-CoA coupled to HPLC separation, radioactivity detection, and UV spectra [Laempe et al., 1999]. (v) 6-Oxocyclohex-1-ene-1-carbonyl-CoA is then subject to hydrolytic ring cleavage yielding 3-hydroxypimelyl-CoA catalyzed by 6-oxocyclo­hex-1-ene-1-carbonyl-CoA hydratase (Oah). 3-Hydroxypimeloyl-CoA detection was as described above for 6-oxocyclohex-1-ene-1-carbonyl-CoA [Laempe et al., 1999]. Further degradation proceeds via conventional β-oxidation.

Fig. 15.

Anaerobic degradation of benzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Fuchs et al. [2011]). B Assigned gene clusters of A. aromaticumEbN1T compared to that of A. evansiiK740T. Genes coding for predicted regulators and uptake systems are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 15.

Anaerobic degradation of benzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Fuchs et al. [2011]). B Assigned gene clusters of A. aromaticumEbN1T compared to that of A. evansiiK740T. Genes coding for predicted regulators and uptake systems are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Genes coding for a predicted ABC transporter (EbN1_C30450-90) and for a symporter (benK) are localized directly upstream of the bcl1 gene. The transcripts and protein components of both of these transporters were essentially observed at high abundances during anaerobic and aerobic growth with all studied aromatic substrates. In the case of acetate, the abundance profiles were for the most part markedly lower [Trautwein et al., 2012a; Becker et al., 2022a].

Transcriptional Regulation

The gene cluster for anaerobic degradation of benzoate is preceded by the bzdR gene. BzdR was previously shown in related Aromatoleum sp. strain CIB to act as transcriptional repressor being liberated from the PN promoter in the presence of benzoyl-CoA [Durante-Rodríguez et al., 2010]. Notably, BzdR displays high sequence similarity with BoxR of the aerobic box pathway (see section “Aerobic degradation of aromatic compounds, benzoate” below). In the case of Aromatoleum sp. strain CIB, both proteins were shown to synergistically control, in a benzoyl-CoA-dependent manner, the transcription of the genes encoding the enzyme constituents of the anaerobic as well as aerobic benzoyl-CoA pathways [Valderrama et al., 2012].

Phenylalanine

Degradation Pathway

Anaerobic and aerobic degradation pathways for phenylalanine share the initial deamination followed by shared and specific reactions yielding the joint intermediate phenylacetyl-CoA [Rabus et al., 2005; Wöhlbrand et al., 2007]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 16A, B), which proceeds via up to five steps. (i) Initially, the NAD+-dependent phenylalanine:2-oxo-glutarate transaminase (TyrB; broad substrate spectrum aromatic-amino-acid transaminase) oxidatively deaminates phenylalanine to phenylpyruvate. This reaction was previously described for A. aromaticum EbN1T using differential proteomics [Wöhlbrand et al., 2007] and for T. aro­matica K172T using enzyme activity measurements [Schneider et al., 1997]. The reaction/enzyme functions under anoxic as well as oxic conditions. (ii) Phenylpyruvate is then converted to phenylacetaldehyde via decarboxylation, catalyzed by indolepyruvate decarboxylase (IpdC) [Schneider et al., 1997; Rabus et al., 2005; Wöhlbrand et al., 2007], predicted to have a broad substrate spectrum (e.g., pyruvate, indolepyruvate, and phenylpyruvate). The reaction/enzyme functions under anoxic as well as oxic conditions. (iii) In principle, a direct conversion of phenylpyruvate to phenylacetyl-CoA could be catalyzed by an indolepyruvate:ferredoxin oxidoreductase (IorBA). Previous proteomic and transcriptomic data [Wöhlbrand et al., 2007; Becker et al., 2022a] provide circumstantial evidence for this possibility. However, this enzyme was only detected at low scores, implicating an only minor role, if at all, in anaerobic degradation of phenylalanine in A. aromaticum EbN1T. (iv) Oxidation of phenylacetaldehyde to phenylacetate can be achieved by two different enzymes: First, NAD(P)+-dependent phenyl­acetaldehyde dehydrogenase (Pdh; synonym: StyD according to Beltrametti et al. [1997] and Becker et al. [2022a]), which is, however, detected only with small score values by proteomics. Second, a new tungsten-containing, aldehyde oxidoreductase (AOR; synonym: aldehyde dehydrogenase [NAD+]) was suggested to catalyze this reaction and denoted AORAa, in order to differentiate it from the various other prokaryotic AORs, but is apparently rather unspecific as it serves a broad variety of aromatic and aliphatic aldehydes [Wöhlbrand et al., 2007; Debnar-Daumler et al., 2014; Schmitt et al., 2017; Arndt et al., 2019]. AORAa (like class II BCR) belongs to the obligate tungsten-dependent members of the AOR family [Seelmann et al., 2020]. Notably, AORAa can also reduce carboxylic acids and NAD+ employing hydrogen as electron donor [Winiarska et al., 2022] and is markedly resistant against exposure to air [Arndt et al., 2019]. The AroA subunit harbors 4 FeS clusters (FS1‒FS4), AorB carries a single FeS cluster (FS0) and the enzymatically active tungsten(W-bis-MPT) cofactor (Wco), AorC contains the FAD cofactor, where NAD+ is reduced, and AorD and AorE are assumed to function as maturation factors for the tungsten cofactor. Most recent structural analysis of the heterologously overproduced, purified enzyme by means of cryo-electron microscopy revealed that AORAa forms a heteromultimeric filament (Aor[AB]nC) from multiples of AorAB proteomers, which are oligomerized starting from a single AorC. While the Wco-containing AorBs locate to the outer face of the filament, the AorAs are centrally positioned, allowing their numerous FeS clusters to be aligned into a “nanowire”-like structure. The latter can be highly charged with electrons and forms a connecting conduit between the multiple Wcos and the single FAD cofactor. Furthermore, these FeS-filaments are shielded from the enzyme's surface, whereby the stability against oxygen exposure is achieved [Winiarska et al., 2023]. Based on these structural insights and QM:MM modeling of the W-bis-MPT cofactor, the following intriguing mechanism was hypothesized: Initially, the aldehyde is activated to the 1,1-geminal alcohol by spontaneous addition of water. Then, the α-hydrogen is abstracted from the carbonyl carbon and bound by one of the oxo ligands of the tungsten cofactor, which is thereby reduced. Finally, a proton is abstracted from the carbocation intermediate of the aldehyde, leading to the formation of the carboxylic acid product. Two one-electron transfers from the reduced tungsten cofactor, via the FS0‒FS4 nanowire to the FAD cofactor, restore its catalytic active, oxidized form [Winiarska et al., 2023]. Furthermore, efficient electrocatalytic oxidation of aldehydes could be demonstrated with AORAa immobilized on glassy carbon working electrodes and applying artificial electron acceptors [Kalimuthu et al., 2023]. Most recently, AORAa was employed as key enzyme for ATP production directly from electricity via the synthetic acid/aldehyde ATP (AAA) cycle [Luo et al., 2023]. (v) Activation of phenylacetate to its CoA ester is performed by two different phenylacetate-CoA ligases employed in the anaerobic (PadJ) and aerobic pathway sections (PaaK1/2), respectively. Noteworthy, PadJ and PaaK1 displayed highest transcript and protein abundance when phenylalanine was provided as substrate, but in both cases during anaerobic as well as aerobic growth. Furthermore, gene expression and protein formation were found for these two CoA ligases also with other tested substrates [Becker et al., 2022a]. The reaction/enzyme functions under anoxic as well as oxic conditions.

Fig. 16.

Anaerobic and aerobic degradation of phenylalanine to the level of phenylacetate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Wöhlbrand et al. [2007]); for further degradation see Figs. 13, 15, and 20. B Assigned gene clusters of A. aromaticumEbN1T in comparison to T. aromaticaK172T. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 16.

Anaerobic and aerobic degradation of phenylalanine to the level of phenylacetate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Wöhlbrand et al. [2007]); for further degradation see Figs. 13, 15, and 20. B Assigned gene clusters of A. aromaticumEbN1T in comparison to T. aromaticaK172T. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Uptake of phenylalanine during anaerobic and aerobic growth is probably mediated by an ABC transporter encoded in proximity to the 1stpaa gene cluster. Further details are provided in section “Phenylacetate (aerobic)” below.

Transcriptional Regulation

No genes for potential transcriptional regulators were found in the vicinity of the pat and pdc genes. A LysR-type regulator (EbN1_C28810) is encoded about 2.6 kbp downstream of the aor genes and a two-component sensory/regulatory system (EbN1_C00670-80) directly upstream of the iorBA genes. However, since the “catabolic” genes of the anaerobic and aerobic degradation of phenylalanine are scattered across the genome, it remains unclear at present whether these sensory/regulatory systems are indeed involved in their transcriptional regulation.

Phenylpropanoids and Protocatechuate

Degradation Pathway

Plant-derived 3-phenylpropanoids support anaerobic as well as aerobic growth of A. aromaticum EbN1T. The respective gene cluster and β-oxidation-type degradation pathway were previously elucidated by combining differential proteogenomics and targeted metabolite analyses [Trautwein et al., 2012b]. This gene cluster is adjacent to the one for anaerobic degradation of benzoate, hinting at a prominent role of 3-phenylpropanoids for the nutrition of strain EbN1T under anoxic conditions. In the case of 3-(4-hydroxyphenyl)propanoate, the β-oxidation of the acyl-sidechain furnishing the central benzoyl-CoA involves the following metabolites, reactions, enzymes, and genes (Fig. 17A, B), and proceeds via six steps. (i) 3-(4-Hydroxyphenyl)propanoate is first activated by a 3-phenylpropanoate-CoA ligase (synonym: trans-feruloyl-CoA synthase) (PprA) forming 3-(4-hydroxyphenyl)propanoyl-CoA. (ii) 3-(4-Hydroxyphenyl)propanoyl-CoA is then oxidized by a 3-phenylpropanoyl-CoA dehydrogenase (synonym: acyl-CoA dehydrogenase) (PprB) yielding 4-coumaroyl-CoA. (iii) The latter is then hydrated at the double bond of its acyl-CoA sidechain by a 3-phenylpropenoyl-CoA 2,3-hydratase (synonym: enoyl-CoA hydratase/isomerase) (PprC) to 3-hydroxy-3-(4-hydroxyphenyl)propanoyl-CoA. (iv) The latter is then oxidized by a 3-aryl-3-hydroxypropanoyl-CoA dehydrogenase (synonym: 3-hydroxyacyl-CoA dehydrogenase) (PprD) forming 4-hydroxy-benzoylacetyl-CoA. (v) The latter is then thiolytically cleaved by a proposed 4-hydroxy-benzoylacetyl-CoA thiolase (synonym: acetyl-CoA C-acetyltransferase) (PprE) into 4-hydroxybenzoyl-CoA and acetyl-CoA. (vi) Finally, 4-hydroxybenzoyl-CoA is proposed to be reductively dehydroxylated by 4-hydroxybenzoyl-CoA reductase (HcrBAC) to benzoyl-CoA [Brackmann and Fuchs, 1993; Unciuleac et al., 2004b]. The complete enzyme set for conversion of 3-(4-hydroxyphenyl)propanoate to 4-hydroxybenzoyl-CoA (PprACEDB) is specifically formed during aerobic and anaerobic growth with this phenylpropanoid, as was also observed for the respective mRNAs [Trautwein et al., 2012b; Becker et al., 2022a]. On the contrary, specific formation of neither transcripts nor protein components of Hcr was detectable [Wöhlbrand et al., 2007; Becker et al., 2022a]. This pathway has a broad substrate range, since it serves the degradation of seven further growth-supporting phenylpropanoids (cinnamyl alcohol, 3-phenylpropanoate, cinnamate, m- and p-coumarate, caffeate, and 3-(3,4-dihydroxyphenyl)propanoate) and performs co-metabolic transformation of non growth-supporting phenylpropanoids (e.g., o-coumarate, ferulate, and sinapate) [Trautwein et al., 2012b].

Fig. 17.

Anaerobic and aerobic degradation of 3-(4-hydroxyphenyl)propanoate and seven further phenylpropanoids by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Trautwein et al. [2012b]); for further degradation see Figs. 11, 12, 15, 19, and 22. Names for the variants of compounds 1–5 are given in Trautwein et al. [2012b]. B Assigned gene clusters of A. aromaticumEbN1T and T. aromaticaK172T. Abbreviations: AB, ATP-binding subunit; SB, solute-binding protein serving the ABC transporter; TM, transmembrane permease subunit; seq %id., amino acid sequence identities (%). Genes coding for a predicted regulator and two uptake systems are marked in black and gray, respectively; the single gene belonging to the anaerobic degradation of benzoate is marked in red. C Protein inventory and proteomic coverage of the pathway.

Fig. 17.

Anaerobic and aerobic degradation of 3-(4-hydroxyphenyl)propanoate and seven further phenylpropanoids by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Trautwein et al. [2012b]); for further degradation see Figs. 11, 12, 15, 19, and 22. Names for the variants of compounds 1–5 are given in Trautwein et al. [2012b]. B Assigned gene clusters of A. aromaticumEbN1T and T. aromaticaK172T. Abbreviations: AB, ATP-binding subunit; SB, solute-binding protein serving the ABC transporter; TM, transmembrane permease subunit; seq %id., amino acid sequence identities (%). Genes coding for a predicted regulator and two uptake systems are marked in black and gray, respectively; the single gene belonging to the anaerobic degradation of benzoate is marked in red. C Protein inventory and proteomic coverage of the pathway.

Close modal

The degradation of protocatechuate (synonym: 3,4-dihydroxybenzoate) is assumed to be channeled into the 3-hydroxybenzoate degradation pathway following a hypothesized activation to its CoA ester and reductive dehydroxylation [Trautwein et al., 2012b].

Uptake

The genes encoding an ABC transporter and a symporter are directly flanked by the gene clusters for degradation of 3-(4-hydroxyphenyl)propanoate (anaerobic and aerobic) and benzoate (anaerobic), suggesting their shared utilization for the uptake of both types of substrates. Notably, a 3-phenylpropanoid-specific periplasmic solute-binding protein (PprF) is encoded in the “3-phenylpropanoid” gene cluster. Furthermore, the distantly encoded TRAP transporter TarPMQ is predicted to serve lignin-derived aromatic compounds, however, no transcript and protein abundance increases were observed in response to the presence of the pathway substrate 3-(4-hydroxyphenyl)propanoate [Becker et al., 2022a].

Transcriptional Regulation

The pprR gene codes for a predicted transcriptional repressor and has previously been indicated in transcriptional control of the “3-phenylpropanoid” gene cluster [Trautwein et al., 2012b]. Most recent studies in our group proved this function based on unmarked, in-frame deletion mutagenesis and further revealed that catabolite repression and additional transcriptional activation should be involved in regulation of gene expression [Vagts et al., 2021]. The transcriptional response threshold toward 3-phenylpropanoids was recently found to also be in the nanomolar range [Vagts et al., 2021].

Benzyl Alcohol and Benzaldehyde

Degradation Pathway

Benzyl alcohol is degraded via benzaldehyde to benzoate (Fig. 18A) involving sequential activity of benzyl alcohol dehydrogenase (AdhB) and benzaldehyde dehydrogenase (Ald3) [Wöhlbrand et al., 2007], as also reported for T. aromatica K172T [Biegert et al., 1995] and Ps. putida [Shaw and Harayama, 1990]. The coding genes in A. aromaticum EbN1T are distinctly located on the chromosome (Fig. 18B). Further catabolism proceeds via the respective anaerobic versus aerobic pathways for benzoate.

Fig. 18.

Anaerobic and aerobic degradation of benzyl alcohol and benzaldehyde by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15 and 20. B Assigned genes. Parts A and B were modified from Wöhlbrand et al. [2007]. C Protein inventory and proteomic coverage of the pathway.

Fig. 18.

Anaerobic and aerobic degradation of benzyl alcohol and benzaldehyde by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 15 and 20. B Assigned genes. Parts A and B were modified from Wöhlbrand et al. [2007]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Due to their rather lipophilic properties, benzyl alcohol and benzaldehyde are not expected to require a specific uptake system, but rather enter the cell by passive diffusion.

Transcriptional Regulation

No hints for specific sensory/regulatory systems can be found in the genetic neighborhood of the “catabolic” genes.

3-Hydroxybenzoate and Gentisate

Degradation Pathway

The aerobic degradation of 3-hydroxybenzoate has long been studied in various bacterial genera, such as Klebsiella, Rhodococcus, and Pseudomonas. The gene nomenclature of this pathway is rather diverse. For example, in Ralstonia sp. strain U2 the nag gene designation is used, reflecting the degradation of naphthalene via gentisate [Zhou et al., 2001], contrasted by the gen and nar gene designations in Rhodococcus opacus R7 [Di Gennaro et al., 2010] and Rhodococcus sp. strain NCIMB 12038 [Liu et al., 2011], respectively. We adopt the one used by Xu et al. [2012], which is more intuitive with respect to the degradation pathway presented here: mhb for meta-hydroxybenzoate. The mhb gene cluster from Klebsiella pneumoniae M5a1 is depicted in Figure 19B and compared to its counterpart in A. aromaticum EbN1T (EbN1_C07550-630). The gene order is rather similar and the encoded enzymes of the pathway share amino acid sequence identities in the range of 35–60%. Notably, a paralogous partial gene cluster exists in the genome of A. aromaticum EbN1T (EbN1_C32310-90), albeit with significantly lower sequence identities of its encoded proteins (30%) and lack of expression [Becker et al., 2022a]. According to the current understanding, aerobic degradation of 3-hydroxbenzoate involves the following metabolites, reactions, enzymes, and genes (Fig. 19A, B), and proceeds via four steps. (i) Initially, 3-hydroxy-benzoate is converted by an NADH-dependent monooxygenase (MhbM), (synonym: 3-hydroxybenzoate 6-hydroxylase), to gentisate (2,5-dihydroxybenzoate) [e.g., Gao et al., 2005]. Structural and modeling approaches with the homolog NagI from E. coli and Silicibacter pomeroyi implicated gentisate and O2 to bind in vicinity of the Fe2+ atom in the active site [Adams et al., 2006; Chen et al., 2008]. (ii) Ring cleavage of gentisate is then achieved by gentisate 1,2-dioxygenase (MhbD) yielding maleylpyruvate [e.g., Harpel and Lipscomb, 1990]. The homolog NagL was shown to have a dimeric structure, to possess the glutathione S-transferase fold and to attack the carbon atom 2 of the substrate with the glutathione thiolate [Marsh et al., 2008; Fang et al., 2011; Hong et al., 2019]. (iii) Maleyl­pyruvate is then transformed to its isomer fumarylpyruvate by maleylpyruvate isomerase (MhbI) [e.g., Jones and Cooper, 1990]. (iv) Finally, fumarylpyruvate is hydrolytically cleaved into pyruvate and fumarate by the fumarylpyruvate hydrolase (MhbH) [e.g., Bayly et al., 1980], which belongs to the large fumarylacetoacetate hydrolase (FAH) superfamily [Hong et al., 2020].

Fig. 19.

Aerobic degradation of 3-hydroxybenzoate and gentisate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Wöhlbrand et al. [2007]). B Assigned gene clusters of K. pneumoniaM5a1 and A. aromaticumEbN1T. Genes coding for the predicted regulators and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 19.

Aerobic degradation of 3-hydroxybenzoate and gentisate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Wöhlbrand et al. [2007]). B Assigned gene clusters of K. pneumoniaM5a1 and A. aromaticumEbN1T. Genes coding for the predicted regulators and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

A putative MFS-type transporter (EbN1_C07630) of unknown substrate specificity is encoded one gene further downstream of the mhb gene cluster on the same strand. Since expression of EbN1_C07630 could not be observed across the tested growth conditions [Becker et al., 2022a], its possible role in the uptake of 3-hydroxybenzoate remains unclear at present.

Transcriptional Regulation

A LysR-type transcriptional regulator (MhbR) is encoded directly upstream of the mhb gene cluster on the opposite strand. However, no transcripts of mhbR could be detected so far. Two further genes upstream on the same strand, another LysR-type transcriptional regulator (EbN1_C07550) is encoded, which displays low transcript abundances during aerobic growth [Becker et al., 2022a]. At present it is unclear which of the two LysR-type regulator (if at all) is responsible for transcriptional control of the mhb gene cluster.

Phenylacetate

Degradation Pathway

First insights into the aerobic degradation of phenylacetate (via phenylacetyl-CoA) emerged from molecular genetic studies with Ps. putida and E. coli [Ferrandez et al., 1998; Olivera et al., 1998]. Subsequently, the involved reaction sequence was elucidated using purified enzymes and mass spectrometric- as well as NMR-based analyses of intermediates formed from [U-13C8]phenylacetate [Ismail et al., 2003; Teufel et al., 2010]. This pathway is widespread in bacteria [Teufel et al., 2010] and has also been demonstrated in A. evansii KB740T [Mohamed et al., 2002]. The unconventional conversion of phenylacetate into two acetyl-CoA and one succinyl-CoA involves the following metabolites, reactions, enzymes, and genes (Fig. 20A, B), and proceeds via eight steps. (i) Unusually for aerobic degradation, phenylacetate is initially activated to phenylacetyl-CoA by AMP-forming phenylacetate-CoA ligase (PaaK1). (ii) Subsequent epoxidation of phenyl­acetyl-CoA forming ring 1,2-epoxy-phenylacetyl-CoA is then catalyzed by a phenylacetyl-CoA 1,2-epoxidase (PaaA1B1C1E1). The PaaA subunit harbors the catalytically relevant di-iron center. The enzyme complex has a composition of PaaA2B3–4C2E1 or multiples of this. Notably, PaaABCE also performs the reverse reaction, i.e., the reductive removal of the epoxide oxygen forming phenyl­acetyl-CoA, which was suggested to represent a relevant detoxification route for toxic epoxides. Taken together, the PaaA1B1C1E1 enzyme represents a bifunctional monooxygenase/deoxygenase [Teufel et al., 2012]. The role of subunit D is at present unclear, it may be involved in protein maturation. (iii) The epoxide is then isomerized to 2-oxepin-2[3H]-ylideneacetyl-CoA by oxepin-CoA forming ring 1,2-epoxy-phenylacetyl-CoA isomerase (PaaG) [Teufel et al., 2010]. PaaG is a bifunctional enzyme also isomerizing 2,3-dehydroadipyl-CoA (cis to trans) downstream in the degradation pathway. (iv) Ring opening of 2-oxepin-2[3H]-ylideneacetyl-CoA proceeds via two steps catalyzed by the bifunctional fusion protein oxepin-CoA hydrolase/3-oxo-5,6-dehydrosuberyl-CoA semialdehyde dehydrogenase (PaaZ1). First, the C-terminal (R)-specific enoyl-CoA hydrolase domain (PaaZ1-ECH) cleaves the ring forming the highly reactive 3-oxo-5,6-dehydrosuberyl-CoA semialdehyde. Second, the latter is then oxidized by the N-terminal NADP+-dependent aldehyde dehydrogenase domain (PaaZ1-ALDH) yielding 3-oxo-5,6-dehydrosuberyl-CoA [Teufel et al., 2011]. (v) The subsequent conversion of 3-oxo-5,6-dehydrosuberyl-CoA corresponds to a conventional β-oxidation reaction sequence as outlined in the following. A 3-oxoacyl-CoA thiolase (PaaJ1, encoded in ∼1 Mbp distance to the 1stpaa gene cluster and formed during aerobic growth with phenylalanine) then thiolytically removes an acetyl-CoA moiety from 3-oxo-5,6-dehydrosuberyl-CoA yielding in cooperation with PaaG trans-2,3-dehydroadipyl-CoA. PaaJ1 is a bifunctional thiolase also catalyzing the last step of the pathway, see below. (vi) trans-2,3-Dehydroadipyl-CoA is then hydrated by 2,3-dehydroadipyl-CoA hydratase (PaaF) forming 3-hydroxyadipyl-CoA. (vii) The latter is oxidized by NAD+-dependent 3-hydroxyadipyl-CoA dehydrogenase (PaaH), which is probably (S)-specific, forming 3-oxoadipyl-CoA. (viii) Finally, 3-oxoadipyl-CoA is thiolytically cleaved to succinyl-CoA and acetyl-CoA by the same 3-oxoacyl-CoA thiolase (PaaJ1), employed before (step v) for 3-oxo-5,6-dehydrosuberyl-CoA processing.

Fig. 20.

Aerobic degradation of phenylacetate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Olivera et al. [1998]); for further degradation see Fig. 20. B Assigned 1st paagene cluster of A. aromaticumEbN1T compared to that of E. coliK12T. Genes coding for predicted regulators and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 20.

Aerobic degradation of phenylacetate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Olivera et al. [1998]); for further degradation see Fig. 20. B Assigned 1st paagene cluster of A. aromaticumEbN1T compared to that of E. coliK12T. Genes coding for predicted regulators and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The EbN1_C20250-80 genes are organized in an operon-like structure, located in proximity (downstream) to the 1stpaa gene cluster, and encoding the components of an ABC-type transporter, i.e., 1 periplasmic solute-binding protein, 2 transmembrane permeases, and 1 cytoplasmic ATP-binding protein. All EbN1_C20250-80 genes were expressed at high level during anaerobic and aerobic growth with phenylalanine and the respective protein products were detected with high meta scores in particular in the membrane protein-enriched fraction. Transcript and protein levels were markedly lower under the other tested substrate conditions, pointing to a specificity of this ABC transporter for phenylalanine [Becker et al., 2022a].

Transcriptional Regulation

The genome of A. aromaticum EbN1T harbors two paa gene clusters (EbN1_C20120-230 and EbN1_C32870-950). The 1stpaa gene cluster was found to be expressed [Becker et al., 2022a] and ends with the paaR gene coding for a predicted transcriptional repressor, as previously demonstrated for E. coli [Ferrández et al., 2000]. The 2ndpaa gene cluster is preceded by genes encoding a two-component sensory/regulatory system (EbN1_C32850/60). Both regulatory systems were found to be expressed. However, in accord with the 1stpaa gene cluster being responsible for aerobic degradation of phenylacetyl-CoA, the paaR gene showed highest transcript abundance during growth with phenylalanine (from which phenylacetyl-CoA is derived) [Becker et al., 2022a].

2-Aminobenzoate (Anthranilate)

Degradation Pathway

Aerobic degradation of N-heterocyclic compounds such as tryptophan and indole is long known to proceed via 2-aminobenzoate and catechol or gentisate [Cain, 1968; Anderson and Dagley, 1981]. An unconventional pathway for aerobic degradation of 2-aminobenzoate has been discovered to be encoded in the genome of A. evansii KB740T and was found to proceed via two steps (Fig. 21A). (i) As described above for the anaerobic degradation of 2-aminobenzoate, benzoate-CoA ligase is assumed to activate 2-aminobenzoate to its CoA ester [Schühle et al., 2003]. (ii) Then, 2-aminobenzoyl-CoA is converted by homodimeric, bifunctional 2-aminobenzoyl-CoA monooxygenase/reductase (AbmA) to 2-amino-5-oxocyclohex-1-ene-1-carboxyl-CoA [Langkau et al., 1990; 1995]. The coding genes (amb) were identified in A. evansii KB740T (Fig. 21B). In A. aromaticum EbN1T there is fair evidence for the presence of an ambA gene, which is isolated on the genome contrasting the operon-like structure known from of A. evansii KB740T [Schühle et al., 2001]. Further degradation of 2-amino-5-oxo-cyclohex-1-ene-1-carboxyl-CoA then proceeds via β-oxidation, which in the case of A. aromaticum EbN1T could be realized by sharing the respective enzymes of the anaerobic degradation of 3-hydroxybenzoate:3-oxoacyl-CoA dehydrogenase (EbN1_C03850, AbmB-like), enoyl-CoA hydratase (EbN1_C03770, AbmC-like), and acyl-CoA dehydrogenase (EbN1_C03840, AbmD-like).

Fig. 21.

Aerobic degradation of 2-aminobenzoate (anthranilate) by A. aromaticumEbN1T. A Proposed reaction sequence. B Assigned gene clusters of A. evansiiKB740T and A. aromaticumEbN1T. Genes coding for components of uptake systems are marked in gray. seq %id., amino acid sequence identities (%). Parts A and B were modified from Wöhlbrand et al. [2007] and Schühle et al. [2001]. C Protein inventory and proteomic coverage of the pathway.

Fig. 21.

Aerobic degradation of 2-aminobenzoate (anthranilate) by A. aromaticumEbN1T. A Proposed reaction sequence. B Assigned gene clusters of A. evansiiKB740T and A. aromaticumEbN1T. Genes coding for components of uptake systems are marked in gray. seq %id., amino acid sequence identities (%). Parts A and B were modified from Wöhlbrand et al. [2007] and Schühle et al. [2001]. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Identification of a gene (abmH) coding for a solute-binding protein of an ABC transporter in A. evansii KB740T [Schühle et al., 2001] implicated uptake of 2-aminobenzoate via such an active transporter. In A. aromaticum EbN1T no equivalent gene is present in the vicinity of the abmA gene.

Transcriptional Regulation

Studies with A. evansii KB740T revealed coordinated expression of the abm genes in response to 2-aminobenzoate and suggested the transcriptional control to be mediated by the MarR-like regulator AbmF [Schühle et al., 2001]. In A. aromaticum EbN1T no equivalent gene is present in the vicinity of the abmA gene.

Benzoate

Degradation Pathway

A special pathway for the aerobic degradation of benzoate (Box, benzoate oxidation) was discovered in A. evansii KB740T by means of differential proteome analysis, revealing the involved proteins to be encoded in an operon-like structure [Gescher et al., 2002]. This box pathway is rather widespread among proteobacteria and has also been studied in Burkholderia xenovorans LB400 [Denef et al., 2005; Bains et al., 2009]. According to the current understanding from mechanistic studies with A. evansii KB740T, this aerobic degradation pathway involves the following metabolites, reactions, enzymes, and genes (Fig. 22A, B), and proceeds via four steps. (i) Initially, benzoate is activated by an ATP-dependent benzoate-CoA ligase (Bcl2) forming benzoyl-CoA, with the CoA ligase functioning in A. evansii KB740T under oxic as well as anoxic conditions [Schühle et al., 2003]. It is noteworthy that A. aromaticum EbN1T apparently employs distinct CoA ligases under these two conditions [Becker et al., 2022a]. (ii) Then, benzoyl-CoA 2,3-epoxidase (BoxBA; synonym: benzoyl-CoA,NADPH:oxygen oxidoreductase (2,3-epoxydizing)) converts benzoyl-CoA into 2,3-epoxybenzoyl-CoA (in tautomeric equilibrium with oxepin, which features a seven-membered ring). The structure of this 2,3-epoxide was inferred in assays with the purified enzyme applying CID-induced MS fragmentation, isotopic resolved MS in the presence of 18O2, and 13C NMR analysis [Rather et al., 2010]. The [4Fe-4S] cluster-containing BoxA protein is the reducing component, which transfers 2 electrons from NADPH to the epoxidase component BoxB [Rather et al., 2011a]. The di-iron enzyme BoxB upon reduction activates O2, from which one oxygen atom is then introduced into bound benzoyl-CoA while the other is released as H2O [Rather et al., 2011b]. The BoxAB enzyme belongs to the enoyl-CoA hydratase/isomerase (crotonase) protein family. (iii) Non-oxygenolytic ring cleavage of 2,3-epoxybenzoyl-CoA is then catalyzed by homodimeric 2,3-epoxy-benzoyl-CoA dihydrolase (BoxC) [Rather et al., 2010]. Regiospecific addition of OH to ring C2 of the epoxide (or its oxepin tautomer) leads to ring opening. A subsequent second OH addition at the same C-atom hydrolyzes the ring C2 to release formic acid, which results in the formation of 3,4-dehydroadipyl-CoA semialdehyde. The structure of this product was resolved by 13NMR analysis of the respective semicarbazone derivative [Gescher et al., 2005]. It is assumed that the BoxAB and BoxC enzymes are associated in vivo. (iv) 3,4-Dehydroadipyl-CoA semialdehyde is then oxidized by homodimeric, NADP+-dependent 3,4-dehydroadipyl-CoA semialdehyde dehydrogenase (BoxD) to cis-3,4-dehydroadipyl-CoA, the structure of which was resolved by interpretation of mass spectrometric fragmentation patterns in conjunction with 13NMR analysis [Gescher et al., 2006].

Fig. 22.

Aerobic degradation of benzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Rather et al. [2010]). B Assigned gene clusters of A. evansiiKB740T and A. aromaticumEbN1T. Proposed reactions that cannot be supported at present by predicted/identified proteins in A. aromaticumEbN1T are marked with dashed arrows. Genes coding for a predicted regulator and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 22.

Aerobic degradation of benzoate by A. aromaticumEbN1T. A Proposed reaction sequence (modified from Rather et al. [2010]). B Assigned gene clusters of A. evansiiKB740T and A. aromaticumEbN1T. Proposed reactions that cannot be supported at present by predicted/identified proteins in A. aromaticumEbN1T are marked with dashed arrows. Genes coding for a predicted regulator and an uptake system are marked in black and gray, respectively. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Evidence for the involvement of the 3-oxoadipyl-CoA pathway in the aerobic degradation of benzoate has previously been obtained with A. evansii KB740T and Bacillus stearothermophilus PK1. Using [ring-13C6]benzoyl-CoA as substrate in cell extracts of both bacterial strains and NMR spectroscopy, Zaar et al. [2001; 2004] were able to identify the following 13C-labeled conversion products: cis-3,4-[2,3,4,5,6-13C5]dehydroadipyl-CoA, trans-2,3-[2,3,4,5,6-13C5]dehydroadipyl-CoA, the 3,6-lactone of 3-[2,3,4,5,6-13C5]hydroxyadipyl-CoA, and 3-[2,3,4,5,6-13C5]hydroxyadipyl-CoA. These metabolites were suggested to be arranged in a pathway involving cis-trans isomerization and leading to 3-oxoadipyl-CoA. In the recent study by Becker et al. [2022a], three potential CoA esters forming distinct chromatographic peaks and having the sum formula C27H42N7O19P3S were detected in extracts of A. aromaticum EbN1T grown aerobically with benzoate or 3-(4-hydroxyphenyl)propanoate. The following plausible structural interpretations suggest the involvement of the 3-oxoadipyl-CoA pathway mentioned above: (i) The resolved sum formula (C27H42N7O19P3S) corresponds to that of an unsaturated adipyl-CoA. An unambiguous HPLC-MS-based differentiation between cis-3,4-dehydroadipyl-CoA and trans-2,3-dehydroadipyl-CoA is prevented by two opposing chromatographic effects. On the one hand, a double bond in 3,4-position to the CoA ester should lead to a shorter elution time compared to when the double bond is at the 2,3-position. On the other hand, the V-shape of the cis-isomer should increase the retention time compared to the linear shape of the trans isomer. Isomerization of cis-3,4-dehydroadipyl-CoA to trans-2,3-dehydroadipyl-CoA should occur prior to β-oxidation [Ren et al., 2004], since the latter mostly acts on trans-2,3-alkanoyl-CoAs. However, a possible enzyme catalyzing such a cis-trans isomerization, e.g., a Δ32-enoyl-CoA isomerase [Yang et al., 1995; Teufel et al., 2010], has not yet been identified in A. aromaticum EbN1T, or A. evansii KB740T and B. stearothermophilus PK1. cis-3,4-Didehydroadipyl-CoA was specifically detected in aerobic cultures of A. aromaticum EbN1T provided with benzoate or 3-(4-hydroxyphenyl)propanoate. trans-2,3-Dehydroadipyl-CoA was also found during growth with phenylalanine [Becker et al., 2022a], where this substance is known to be a degradation intermediate [Teufel et al., 2010]. (ii) The role and origin of 3-hydroxyadipyl-CoA 3,6-lactone is not clear, yet; it might originate from one of the unsaturated adipyl-CoA intermediates [Becker et al., 2022a]. (iii) The 3,6-lactone intermediate was previously assumed to be opened by a lactonase to 3-hydroxyadipyl-CoA [Zaar et al., 2001, 2004; Ren et al., 2004]. Such an enzyme could not be predicted from the genome of A. aromaticum EbN1T. The EbN1_C15610 gene, which is localized in proximity to the box gene cluster of strain EbN1T and has an ortholog in the box gene cluster of A. evansii KB740T, belongs to the α/β-hydrolases, but is predicted to be involved in vitamin K2 (menaquinone) biosynthesis. Alternatively, 3-oxoadipyl-CoA could be formed from 3-hydroxyadipyl-CoA, the hydration product of trans-2,3-dehydroadipyl-CoA. No speci­fic β-oxidation enzymes could be assigned to these two reactions. (iv) As final reaction of the presumptive 3-oxo­adipyl-CoA pathway, a thiolase (BoxE) has been proposed for A. evansii KB740T [Fuchs et al., 2011] and characterized from Pseudomonas sp. strain B13 [Kaschabek et al., 2002]. Accordingly, a boxE gene is present in the box gene cluster of both A. aromaticum EbN1T and A. evansii KB740T. The BoxE protein was identified in A. aromaticum EbN1T with particularly high score in cells adapted to aerobic growth with benzoate, 3-(4-hydroxyphenyl)propanoate, and phenylalanine [Becker et al., 2022a]. Taken together, the findings described above indicate the aerobic benzoate degradation in A. aromaticum EbN1T to involve the 3-oxo adipyl-CoA pathway, albeit the reaction sequence from cis-3,4-dehydroadipyl-CoA to 3-oxoadipyl-CoA has to await further enzymatic support.

Uptake

A solute-binding protein (EbN1_C15620) is encoded directly upstream of the bcl2 gene and is detected under all tested conditions [Becker et al., 2022a]. Since, however, the chromosomal neighborhood of the box gene cluster does not contain genes for an ABC transporter, it remains unclear what system might recruit EbN1_C15620 for the uptake of benzoate. Nevertheless, one may speculate that the ABC transporter encoded next to the genes of the anaerobic benzoyl-CoA pathway may serve this function; it is likewise formed with aromatic substrates during anaerobic as well as aerobic growth [Becker et al., 2022a].

Transcriptional Regulation

The box gene cluster contains the boxR gene, which is transcribed and translated during aerobic growth with benzoate. Furthermore, boxR transcripts were observed at high abundances also during anaerobic growth with each of the tested aromatic substrates [Becker et al., 2022a]. Note that the homologous BoxR and BzdR proteins are functionally redundant transcriptional repressors, which control the benzoate-dependent transcriptional regulation of the anaerobic and the aerobic benzoyl-CoA pathways, as demonstrated for Aromatoleum sp. strain CIB [Valderrama et al., 2012].

Cyclohexane Carboxylate

Degradation Pathway

The degradation of cyclohexane carboxylate has been mainly investigated in R. palustris [Küver et al., 1995; Pelletier and Harwood, 2000]. The following metabolites, reactions, enzymes, and genes are involved in the degradation pathway (Fig. 23A, B), which proceeds via five steps. (i) Initially, cyclohexane carboxylate is activated to its CoA ester by an ATP-dependent cyclohexane carboxylate-CoA ligase (AliA). (ii) The cyclohexane carbonyl-CoA formed is then oxidized to cyclohex-1-ene-1-carbonyl-CoA, catalyzed by a FAD-dependent dehydrogenase (AliB). At this point, the degradation pathway continues according to that for anaerobic benzoyl-CoA degradation in R. palustris. (iii) Then, cyclohex-1-ene-1-carbonyl-CoA hydratase (BadK) adds water to the C−C-double bond yielding 2-hydroxy-cyclohex-1-carbonyl-CoA, which should be S-configured according to a recent study on anaerobic degradation of p-cymene in A. aromaticum pCyN1 [Küppers et al., 2019]. (iv) Subsequently, NAD+-dependent 2-hydroxy-cyclohexane-1-carbonyl-CoA dehydrogenase (BadH) oxidizes the secondary alcohol forming 2-oxo-cyclohexane-1-carbonyl-CoA. (v) Finally, 2-oxo-cyclohexane-1-carbonyl-CoA hydrolase (BadI) hydrolytically cleaves the alicyclic ring forming pimeloyl-CoA, which is further degraded via conventional β-oxidation.

Fig. 23.

Anaerobic and aerobic degradation of cyclohexane carboxylate by A. aromaticumEbN1T. A Proposed reaction sequence. B Assigned gene clusters of R. palustrisand A. aromaticumEbN1T [Pelletier and Harwood, 2000]. The gene coding for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 23.

Anaerobic and aerobic degradation of cyclohexane carboxylate by A. aromaticumEbN1T. A Proposed reaction sequence. B Assigned gene clusters of R. palustrisand A. aromaticumEbN1T [Pelletier and Harwood, 2000]. The gene coding for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

The genetic context of the ali-bad genes on the chromosome of A. aromaticum EbN1T does not provide any clue about a possible uptake system for cyclohexane carboxylate. However, given the charged carboxyl group of the compound, it must be assumed that import of cyclohexane carboxylate requires an active uptake system.

Transcriptional Regulation

The ali-bad genes encoding the enzymes for degradation of cyclohexane carboxylate (chc) are organized in an operon-like structure on the genome of A. aromaticum EbN1T. This is similar to the gene arrangement known from R. palustris (Fig. 23B). The ali-bad gene cluster also harbors the badR gene coding for a transcriptional repressor of the MarR family [Gupta et al., 2019]. BadR was demonstrated in R. palustris to regulate transcription of the ali-bad gene cluster [Egland and Harwood, 1999] and to abrogate repression upon binding of the degradation intermediate 2-oxo-cyclohexane-1-carbonyl-CoA [Hirakawa et al., 2015].

3-Hydroxypropanoate and Acrylate

Degradation Pathway

The degradation of 3-hydroxypropanoate via a reductive route has been elucidated in the anoxygenic phototroph Rhodobacter sphaeroides [Schneider et al., 2012; Asao and Alber, 2013]. The following metabolites, reactions, enzymes, and genes (all present in the genome of A. aromaticum EbN1T) are involved in the degradation pathway (Fig. 24), which proceeds via three steps. (i) Initially, 3-hydroxypropanoate is activated to its CoA ester by a predicted CoA-ligase (PrpE) homologous to the enzyme from Salmonella typhimurium LT2 [Horswill and Escalante-Semerena, 1999]. (ii) The 3-hydroxypropanoyl-CoA formed is then assumed to be oxidized by elimination of water to acrylyl-CoA; the catalyzing dehydratase in A. aromaticum EbN1T is assumed to be EbN1_C12250. (iii) Finally, acrylyl-CoA is reduced to propanoyl-CoA by NADPH-dependent acrylyl-CoA reductase (AcuI) belonging to the medium-chain dehydrogenase/reductase (MDR) superfamily [Riveros-Rosas et al., 2003]. Further degradation of propanoyl-CoA can then proceed via the methylmalonyl-CoA pathway to succinyl-CoA and therefrom via the TCA cycle. Acrylate could be channeled into this pathway by a CoA ligase [Schneider et al., 2012]. Evidence for the potential to transform acrylonitrile to acrylate could not be provided from inspecting the genome of A. aromaticum EbN1T.

Fig. 24.

Anaerobic and aerobic degradation of 3-hydroxypropanoate and acrylate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35 and 36, respectively. B Assigned genes. C Protein inventory and proteomic coverage of the pathway.

Fig. 24.

Anaerobic and aerobic degradation of 3-hydroxypropanoate and acrylate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35 and 36, respectively. B Assigned genes. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

No genes for a potential uptake system for 3-hydroxypropanoate and/or acrylate are present in the vicinity of the prpE/acuI genes.

Transcriptional Regulation

No genes coding for a potential sensory/regulatory system, which could control the expression of the prpE/acuI genes, could be detected in their genomic neighborhood.

2-Propanol, 2-Butanol, 2-Propanone (Acetone), and 2-Butanone

Degradation Pathway

The secondary aliphatic alcohols 2-propanol and 2-butanol are metabolized via the same degradation route to acetyl-CoA and propanoyl-CoA, respectively, involving the following three steps (Fig. 25A). (i) A yet to be assigned alcohol dehydrogenase oxidizes the alcohols to the respective ketones, namely 2-propanone (acetone) and 2-butanone. (ii) A subsequent carboxylation at the C1-methyl group yields 2-oxobutanoate and 3-oxopentanoate, respectively. This reaction is catalyzed by an ATP-dependent heterotrimeric (AcxABC) carboxylase that has been intensively studied on the genetic-biochemical [Sluis et al., 1996; Sluis and Ensign, 1997; Sluis et al., 2002; Clark and Ensign, 1999; Boyd et al., 2004; Boyd and Ensign, 2005; Dullius et al., 2011; Schühle and Heider, 2012], proteomic [Wöhlbrand et al., 2007; Oosterkamp et al., 2015] and structural levels [Nocek et al., 2004; Mus et al., 2017] in various bacteria including Alicycliphilus sp., A. aromaticum EbN1T, Xanthobacter autotrophicus Py2, and Rhodobacter capsulatus B10. (iii) The 3-oxoacids formed could then be activated to the respective CoA esters by the predicted acetate-CoA ligase (AcsA) and/or a succinyl-CoA:3-oxoacid CoA transferase (KctAB), which are both encoded by genes proximal to the acxABC genes. Furthermore, substrate-specific formation of KctAB was previously reported for A. aromaticum EbN1T [Wöhlbrand et al., 2007]. This activation would allow subsequent thiolytic cleavage, yielding 2 acetyl-CoA per acetone contrasted by 1 acetyl-CoA and 1 propanoyl-CoA per 2-butanone. While terminal oxidation of acetyl-CoA proceeds via the TCA cycle, the methylmalonyl-CoA pathway is required for processing propanoyl-CoA.

Fig. 25.

Anaerobic and aerobic degradation of 2-propanol and acetone, as well as of 2-butanol and butanone by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35 and 36, respectively. B Assigned genes. The gene coding for the predicted regulator AcxR is marked in black. C Protein inventory and proteomic coverage of the pathway.

Fig. 25.

Anaerobic and aerobic degradation of 2-propanol and acetone, as well as of 2-butanol and butanone by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35 and 36, respectively. B Assigned genes. The gene coding for the predicted regulator AcxR is marked in black. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

No genes for a potential uptake system for the two alcohols and ketones are present in the vicinity of the acx/kct genes. However, it may be assumed that the lipo-philicity of these substrates allows entry into the cell merely by passive diffusion.

Transcriptional Regulation

The zraR gene is located directly upstream of the acx genes and belongs to the atypical two-component ZraP-SR system. Since this has recently been implicated in various functions including “envelope stress response” [Rome et al., 2018; Choudhary et al., 2020], it is unlikely to control expression of the acx/kct genes. However, the σ54-dependent regulator AcxR, which was previously suggested to serve this function [Sluis et al., 2002], is encoded downstream of the kctAB genes.

1-Butanol, 1-Propanol, and Ethanol

Degradation Pathway

The short-chained primary alcohols 1-butanol, 1-propanol, and ethanol are most likely degraded via a shared classical route involving oxidation of the hydroxy group (possibly via the aldehyde intermediate) to the respective carboxylate (Fig. 26A) [Kunau et al., 1995; Houten et al., 2016]. The enzymes involved in these degradative pathway have yet to be identified, but are likely recruited from the large pool of alcohol dehydrogenases encoded in the genome of A. aromaticum EbN1T (Fig. 26B) and belonging to the short- and medium-chain dehydrogenase/reductase (SDR, MDR) superfamilies [Kallberg et al., 2002; Eklund and Ramaswamy, 2008; Jörnvall et al., 2010, 2015]. Further degradation of the carboxylates formed involves activation to the respective CoA esters, followed by conversion via β-oxidation, TCA cycle, and methylmalonyl-CoA pathway (see below).

Fig. 26.

Anaerobic and aerobic degradation of 1-butanol, 1-propanol, and ethanol by A. aromaticumEbN1T. A Proposed reaction sequences; for further degradation see Fig. 35 and 36, respectively. B Inventory of alcohol/aldehyde dehydrogenases for aliphatic compounds.

Fig. 26.

Anaerobic and aerobic degradation of 1-butanol, 1-propanol, and ethanol by A. aromaticumEbN1T. A Proposed reaction sequences; for further degradation see Fig. 35 and 36, respectively. B Inventory of alcohol/aldehyde dehydrogenases for aliphatic compounds.

Close modal

Uptake

No genes for a potential uptake system for the three alcohols are present in the vicinity of the adh genes. However, the lipophilicity of these alcohols should facilitate their uptake by passive diffusion.

Transcriptional Regulation

No genes for a potential sensory/regulatory system are present in the vicinity of the adh genes.

Butanoate, Propanoate, Acetate, and C4-Dicarboxylates

Degradation Pathway

The organic acids butanoate, propanoate, acetate, and C4-dicarboxylates (succinate, fumarate, and malate) are metabolized via conventional routes (Fig. 27). Following activation to the CoA ester, butanoyl-CoA is split via β-oxidation into two acetyl-CoA moieties. Propanoate is activated by CoA and then processed via the methylmalonyl-CoA pathway (Fig. 36). C4-dicarboxylates are directly channeled into the canonical TCA cycle (Fig. 35). Acetate degradation is initiated by acetate-CoA ligase forming acetyl-CoA [Starai and Escalante-Semerena, 2004], which is then channeled into the TCA cycle. Notably, A. aromaticum EbN1T also possesses a phosphotransacetylase and an acetate kinase [Dittrich et al., 2005; Ingram-Smith et al., 2006] that could possibly be used to exploit the energy-rich thioester bond of acetyl-CoA for substrate-level ATP regeneration, yielding acetate as end product.

Fig. 27.

Anaerobic and aerobic degradation of butanoate, propanoate, and acetate by A. aromaticumEbN1T. A Proposed reaction sequences; for further degradation see Fig. 35 and 36, respectively. B Inventory of involved enzymes.

Fig. 27.

Anaerobic and aerobic degradation of butanoate, propanoate, and acetate by A. aromaticumEbN1T. A Proposed reaction sequences; for further degradation see Fig. 35 and 36, respectively. B Inventory of involved enzymes.

Close modal

Uptake

Uptake of n-butanoate, propanoate, and acetate is assumed to be mediated via proton-driven symporters such as ActP1/2, while that of C4-dicarboxylates is achieved by a high-affinity TRAP-transporter (DctPQM) [Rabus et al., 1999; Rosa et al., 2018]. These transporters are encoded in the genome of A. aromaticum EbN1T, albeit not in vicinity of the aforementioned “catabolic” genes.

Transcriptional Regulation

Extracellular occurrence of C4-dicarboxylates is re-cognized by specific sensors in various bacteria [Janausch et al., 2002]. In A. aromaticum EbN1T, the two-component sensory system DctSR is encoded by genes adjacent to the genes encoding the DctPQM uptake systems and known from Rhodobacter spp. to control expression of the dctPQM genes [Hamblin et al., 1993; Sánchez-Ortiz et al., 2021]. Furthermore, benzoate was shown in A. aromaticum EbN1T to negatively affect C4-dicarboxylate utilization, apparently by repressing dctSR expression [Trautwein et al., 2012c].

Lactate and Pyruvate

Degradation Pathway

Lactate and pyruvate are degraded via the classical routes, involving lactate dehydrogenase [Jiang et al., 2014] and pyruvate dehydrogenase [Patel et al., 2014], yielding in conjunction acetyl-CoA and CO2 (Fig. 28). Notably, the genome of A. aromaticum EbN1T also encodes two sets of proteins (LutA1B1C1 and LutA2B2C2) for an alternative pathway of lactate utilization; in each case, the coding genes are organized in an operon-like structure. Such Lut-systems have been demonstrated in B. subtilis to be involved in lactate utilization and biofilm formation [Chai et al., 2009], in Campylobacter jejuni NCTC 11168 to contribute to lactate utilization [Thomas et al., 2011], and to harbor a novel conserved protein domain (LUD) [Hwang et al., 2013]. Notably, proteomic profiling of chemostat studies with C. jejuni NCTC 11168 revealed elevated abundance of the Lut-system with increasing oxygen concentration (aerobiosis) [Guccione et al., 2017], which could be relevant also for facultative anaerobic A. aromaticum EbN1T.

Fig. 28.

Anaerobic and aerobic degradation of lactate, pyruvate, and acetate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. The gene coding for a predicted regulator is marked in black. C Protein inventory and proteomic coverage of the pathway.

Fig. 28.

Anaerobic and aerobic degradation of lactate, pyruvate, and acetate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. The gene coding for a predicted regulator is marked in black. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

No evidence for a known lactate- or pyruvate-specific uptake system could be found during re-annotation of the A. aromaticum EbN1T genome. In E. coli, lactate/pyruvate uptake was reported to be also achievable by peptide transporters [Hwang et al., 2018], several of such systems are encoded in the genome of A. aromaticum EbN1T.

Transcriptional Regulation

The GntR family member LutR is known from B. subtilis to control the expression of the lutABC operon [Chai et al., 2009; Chiu et al., 2014]. While no clear candidate for such a regulator could be predicted from the genome of A. aromaticum EbN1T, an IclR-family member (EbN1_C23190) is encoded directly next to the lutA2B2C2 operon.

Malonate

Degradation Pathway

Malonate degradation has been described meanwhile for a variety of bacterial species, e.g., Rhizobium leguminosarum bv. trifolii [An and Kim, 1998; Kim, 2002], R. capsulatus [Dehning and Schink, 1994], R. palustris [Wang et al., 2020], and other proteobacteria [Suvorova et al., 2012]. The degradation pathway involves a malonyl-CoA synthetase (MatB) and malonyl-CoA decarboxylase (MatA), which convert malonate via its CoA ester to acetyl-CoA (Fig. 29). While the genome of A. aromati­cum EbN1T does not encode a MatB homolog, it harbors a standalone matA gene, which implicates that malonate utilization would require its unspecific activation to the CoA ester by one of the many CoA ligases at the disposal of the strain. At present it is, however, unknown whether A. aromaticum EbN1T can grow with malonate.

Fig. 29.

Anaerobic and aerobic degradation of malonate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned gene. C Protein inventory of the pathway.

Fig. 29.

Anaerobic and aerobic degradation of malonate by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned gene. C Protein inventory of the pathway.

Close modal

Uptake

The mat gene cluster in R. trifolii also encodes a potential carrier (MatC) for the uptake of malonate [Lee et al., 2000] and Malomonas rubra employs the Na+-driven malonate uptake system MadLM [Schaffitzel et al., 1998]. No candidate for such a system could be predicted from the genome of A. aromaticum EbN1T. However, one of the numerous TRAP transporters encoded in the genome of A. aromaticum EbN1T could serve this function, as previously proposed for R. palustris [Wang et al., 2020].

Transcriptional Regulation

In R. leguminosarum bv. trifolii and other bacteria, the GntR family member MatR was previously reported to control expression of the mat genes [Lee et al., 2000; Suvorova et al., 2012]. No MatR homolog could be predicted from the genome of A. aromaticum EbN1T; however, directly next to the matB gene, a PAS domain-containing, predicted diguanylate cyclase/phosphodiesterase is encoded.

Formate and Hydrogen

Reactions

The analysis of the A. aromaticum EbN1T genome revealed the presence of several genes whose products are predicted to be involved in formate and hydrogen metabolism (Fig. 30). First, the fdhA2BC genes encode an ortholog of the periplasm-facing, respiratory formate dehydrogenase-N (Fdh-N) known in E. coli to function together with nitrate reductase [Jormakka et al., 2003]. In proximity to the fdhA2BC genes is the fdhD gene, which codes for a sulfurtransferase assumed to be required for active Fdh-N [Arnoux et al., 2015]. Second, the A. aromaticum EbN1T genome possesses a potential monomeric FdhA1 formate dehydrogenase related to the predicted protein of B. subtilis [Rivolta et al., 1998] and encoded next to the gene for the NADH-binding NuoF component of complex I [Velazquez et al., 2005]. Third, the hyfBCEFGI gene cluster codes for large parts of hydrogen-forming, membrane-localized [NiFe]-hydrogenase HyfA-I of the formate hydrogenlyase (FHL) complex, which is known to disproportionate formate to CO2 and hydrogen (H2) and to be evolutionarily linked to complex I from the respiratory chain [Andrews et al., 1997; McDowall et al., 2014; Pinske and Sawers, 2016; Sargent, 2016]. However, the FdhF component responsible for formate oxidation is apparently missing in A. aromaticum EbN1T. Furthermore, the selABCD genes coding for selenocysteine synthase required for selenocysteine incorporation into Fdh [Böck et al., 1991; Rother et al., 2001; Zorn et al., 2013] are absent in the genome of A. aromaticum EbN1T. Taken together, despite the described genome-based observations, the potential of A. aromaticum EbN1T for formate and hydrogen metabolism remains ambiguous at present and requires future research also to assess loss of function or adoption of another role.

Fig. 30.

Metabolism of formate and hydrogen by A. aromaticumEbN1T. A Proposed reaction sequences. B Assigned genes. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 30.

Metabolism of formate and hydrogen by A. aromaticumEbN1T. A Proposed reaction sequences. B Assigned genes. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

In line with the aforementioned ambiguities, no homolog of the well-known formate transporter FocA [Wang et al., 2009] is predicted from the genome of A. aromaticum EbN1T.

Transcriptional Regulation

Except for a diguanylate cyclase/phosphodiesterase encoded in the neighborhood of the fdh gene cluster, no other genes for potential sensory/regulatory systems are present in the vicinity of either of the two gene clusters.

CoA Ligases and β-Oxidation Potential

Degradation Pathway

Degradation of short-, medium-, and long-chain fatty acids is initiated by CoA ligases followed by the canonical β-oxidation cycle yielding acetyl-CoA moieties (in case of uneven fatty acids also propanoyl-CoA) [Clark and Cronan, 2013] (Fig. 31). The genome of A. aromaticum EbN1T is rich in genes encoding such enzymes. From the 32 predicted CoA ligases, 19 could be assigned to specific substrates including aforementioned aryl and cyclohexane carboxylates. Moreover, 11 enzymes involved in the β-oxidation cycle are encoded, including acyl-CoA dehydrogenases (e.g., FadE) [Campbell and Cronan, 2002] and multifunctional 3-hydroxyacyl-CoA dehydrogenase/enoyl-CoA hydratase (e.g., FadB) [Spratt et al., 1984].

Fig. 31.

Metabolism of fatty acids by A. aromaticumEbN1T. A Canonical β-oxidation cycle. B Inventory of CoA ligases/β-oxidation potential and proteomic coverage.

Fig. 31.

Metabolism of fatty acids by A. aromaticumEbN1T. A Canonical β-oxidation cycle. B Inventory of CoA ligases/β-oxidation potential and proteomic coverage.

Close modal

Uptake

The uptake of the particularly hydrophobic long-chain fatty acids across the cell envelope of Gram-negative bacteria proceeds in a first step via the β-barrel structured FadL, which forms a hydrophobic channel through the outer membrane [van den Berg et al., 2004; Hearn et al., 2009]. Flipping of the fatty acids across the cytoplasmic membrane is then achieved by the membrane localized acyl-CoA ligase [Azizan et al., 1999]. The genome of A. aromaticum EbN1T encodes outer membrane localized FadL and a variety of acyl-CoA ligases.

Transcriptional regulation.

In E. coli, expression of the fad genes is controlled by the transcriptional repressor FadR [Cronan, 2021], which is, however, not encoded in the genome of A. aromaticum EbN1T.

Leucine (and Isovalerate)

Degradation Pathway

Degradation of leucine proceeds via a pathway comprising six steps (Fig. 32A). (i) Leucine is initially deaminated by a branched-chain amino acid aminotransferase (IlvE) [Hutson, 2001; Goto et al., 2003] forming 4-methyl-2-oxo-pentanoate. (ii) The latter is then oxidatively decarboxylated by the branched-chain 2-oxoacid dehydrogenase BdkABC [Sokatch et al., 1981; Sykes et al., 1987; Ævarsson et al., 1999] to isovaleryl-CoA, which is also the entry point for isovalerate upon its activation to the CoA ester. (iii) Isovaleryl-CoA is then oxidized to 3-methylcrotonyl-CoA by the dehydrogenase LiuA [Förster-Fromme and Jendrossek, 2008], followed by (iv) ATP-dependent carboxylation to 3-methylglutaconyl-CoA catalyzed by the methylcrotonyl-CoA carboxylase LiuBD [Höschle et al., 2005]. (v) Addition of water to the double bond of 3-methylglutaconyl-CoA by the hydratase LiuC [Bock et al., 2016] furnishes 3-hydroxy-3-methylglutaryl-CoA, which is (vi) finally split into acetoacetate and acetyl-CoA by the lyase LiuE [Chávez-Avilés et al., 2009]. Notably, most of the liu genes of A. aromaticum EbN1T are organized in an operon-like structure (Fig. 32B, C) as previously described for Pseudomonas aeruginosa PAO1 [Förster-Fromme et al., 2006] and the described pathway is absent in E. coli [Reizer, 2013].

Fig. 32.

Degradation of leucine and isoleucine by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. C Protein inventory and proteomic coverage of the pathway.

Fig. 32.

Degradation of leucine and isoleucine by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Leucine is probably imported by the branched-chain amino acid ABC transporter Bra, which is encoded by the braGFEDC1C2 gene cluster and belongs to the HAAT transporter family [Hoshino and Kose, 1990; Saier, 2000]. Two further similar ABC transporters (EbN1_C15320/40/50/620 and EbN1_C20250-80) are encoded in the genome of A. aromaticum EbN1T.

Transcriptional Regulation

The leucine-responsive regulatory protein Lrp is a well-studied global regulator of metabolism in E. coli and known to control expression of the Liv ABC transporter for branched-chain amino acids including isoleucine [Calvo and Matthews, 1994], which might also be the case in A. aromaticum EbN1T. The transcriptional activator BkdR was shown in Ps. putida to control expression of the bkd operon encoding aforementioned branched-chain 2-oxoacid dehydrogenase [Madhusudhan et al., 1997]. A homolog of BkdR is encoded in the genome of A. aromaticum EbN1T. The LiuR regulator known from Ps. aeruginosa [Kazakov et al., 2009] is apparently not encoded in the genome of A. aromaticum EbN1T.

Histidine

Degradation Pathway

Degradation of histidine in A. aromaticum EbN1T essentially follows the hut (histidine utilization) pathway previously reported for Ps. putida and Ps. aeruginosa, involving a series of four hydrolytic reactions (Fig. 33A). (i) Initially, histidine ammonia lyase (HutH) non-oxidatively eliminates the α-amino group from histidine whereby trans-uroconate is formed [Consevage and Phillips, 1990; Schwede et al., 1999]. (ii) The latter is then converted to 4-imidazolone-5-propanoate via addition of water to the double bond, catalyzed by NAD+-dependent urocanase (HutU) [Klepp et al., 1990; Lenz and Rétey, 1993; Kessler et al., 2004]. (iii) Then, the zinc-dependent imidazolonepropionase (HutI) hydrolyzes via a nucleophilic attack the C−N-bond of the carbonyl carbon of 4-imidazolone-5-propanoate yielding N-formimidoyl-L-glutamate [Yu et al., 2006; Tyagi et al., 2008]. (iv) In the final reaction, the latter is hydrolytically cleaved to L-glutamate and formamide by formiminoglutamate hydrolase (HutG) [Ouzounis and Kyrpides, 1994]. The genes for all aforementioned enzymes are present in the genome of A. aromaticum EbN1T. By contrast, the alternative enzyme N-formimino-L-glutamate iminohydrolase (HutF), which deaminates N-formimidoyl-L-glutamate to N-formyl-L-glutamate and ammonia [Fedorov et al., 2015], is not encoded in the genome of A. aromaticum EbN1T.

Fig. 33.

Degradation of histidine by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. The gene coding for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 33.

Degradation of histidine by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Assigned genes. The gene coding for a predicted regulator is marked in black. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

In Ps. putida, the secondary transporter HutT is responsible for histidine uptake [Wirtz et al., 2021]. Such a system could not be predicted from the genome of A. aromaticum EbN1T, thus, histidine uptake may be performed by one of several available ABC transporters for amino acids.

Transcriptional Regulation

In Pseudomonas spp. transcriptional regulation of the hut operon is controlled by the repressor HutC [Hu et al., 1989; Naren and Zhang, 2020]. No evidence for a hutC homolog could be detected in the genome of A. aromaticum EbN1T.

Proline

Degradation Pathway

Degradation of proline has been elucidated in E. coli and S. typhimurium and shown to involve the peripheral membrane protein proline dehydrogenase PutA performing two consecutive reactions [Surber and Maloy, 1998; Vinod et al., 2002] (Fig. 34A). (i) In the initial reaction, the multifunctional flavoprotein PutA acts as proline dehydrogenase, oxidizing L-proline to (S)-1-pyrroline-5-carboxylate with the released electrons delivered via FAD to the quinone pool. (S)-1-Pyrroline-5-carboxy­late is then assumed to spontaneously hydrolyze to L-glutamate-5-semialdehyde. (ii) The latter is then oxidized to L-glutamate by PutA with NAD+ serving as cofactor. The genome of A. aromaticum EbN1T habors a single putA gene (Fig. 34B, C).

Fig. 34.

Degradation of proline by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Protein inventory and proteomic coverage of the pathway.

Fig. 34.

Degradation of proline by A. aromaticumEbN1T. A Proposed reaction sequence; for further degradation see Fig. 35. B Protein inventory and proteomic coverage of the pathway.

Close modal

Uptake

Uptake of proline is performed by the sodium-driven transporter PutP [Olkhova et al., 2011], which in E. coli and S. typhimurium is encoded directly next to the putA gene. Notably, in A. aromaticum EbN1T, the putA and putP genes are localized at very distant positions in the genome.

Transcriptional Regulation

Remarkably, the PutA protein additionally functions as a redox-dependent transcriptional repressor, controlling in its active form induction of the put operon in the presence of proline in S. typhimurium [Ostrovsky de Spicer and Maloy, 1993; Muro-Pastor and Maloy, 1995].

TCA Cycle

The canonical TCA cycle [Cronan and Laporte, 2013] serves two major functions: First, acetyl-CoA generated from the multitude of aforementioned degradation pathways is completely oxidized to CO2, accompanied by the provision of reduced electron carriers for respiratory energy conservation. Second, precursors for monomer biosynthesis (2-oxoglutarate and oxaloacetate) are generated. The genes encoding all enzymes of the TCA cycle are present in, albeit scattered across, the genome of A. aromaticum EbN1T (Fig. 35). The TCA cycle is constituted by the following eight enzymes: (i) Citrate synthase (GltA) catalyzes the Claisen condensation (C−C-bond formation) between acetyl-CoA and oxaloacetate forming the citryl-CoA intermediate, which is subsequently hydrolyzed to the tricarboxylic acid citrate and coenzyme A [Wiegand and Remington, 1986; Karpusas et al., 1990; Aleksandrov et al., 2014]. (ii) Aconitase (AcnB) is an FeS-enzyme that catalyzes the dehydration of citrate to cis-aconitate followed by the rehydration to isocitrate [Robbins and Stout, 1989; Lloyd et al., 1999]. (iii) Isocitrate dehydrogenase (Icd1) catalyzes, after a random binding of cosubstrates, the oxidation (with NADP+) of isocitrate to the enzyme-bound intermediate oxalosuccinate, which is subsequently decarboxylated to 2-oxoglutarate [Bolduc et al., 1995]. (iv) 2-Oxoglutarate dehydrogenase (SuccAB/Lpd) oxidatively decarboxylates 2-oxoglutarate to succinyl-CoA (NADH-forming). The tripartite enzyme complex is closely related to pyruvate dehydrogenase [de Kok et al., 1998], with E1 (SucA, succinyl-transferring) and E3 (Lpd, dihydrolipoyl dehydrogenase) flexibly tethered to the E2 core (SucB, 2-oxoglutarate dehydrogenase) [Murphy and Jensen, 2005]. (v) Heterodimeric succinyl-CoA ligase (SucCD) converts succinyl-CoA via succinyl-phosphate to succinate, representing the sole substrate-level phosphorylation reaction of the TCA cycle [Wolodko et al., 1994]. (vi) Succinate dehydrogenase (SdhAB2C2D2) oxidizes succinate to fumarate concomitantly feeding the released electrons into the membrane-embedded quinone pool (see section “Respiratory Energy Conservation” below). (vii) Fumarase (Fum) catalyzes the reversible hydration of fumarate via an aci-carboxylate intermediate to S-malate [Weaver et al., 1995; Weaver, 2005]. (viii) Finally, NAD+-dependent malate dehydrogenase (Mdh) oxidizes malate to oxaloacetate, completing the TCA cycle; notably, Mdh has a markedly stricter substrate range than functionally related lactate dehydrogenase [Hall and Banaszak, 1993; Boernke et al., 1995].

Fig. 35.

TCA cycle in A. aromaticumEbN1T. A Reaction sequence. For reasons of simplification, the intermediates cis-aconitate and oxalo-succinate have been omitted. B Assigned genes. The gene coding for a predicted regulator is marked in black. C Protein inventory and proteomic coverage of the cycle.

Fig. 35.

TCA cycle in A. aromaticumEbN1T. A Reaction sequence. For reasons of simplification, the intermediates cis-aconitate and oxalo-succinate have been omitted. B Assigned genes. The gene coding for a predicted regulator is marked in black. C Protein inventory and proteomic coverage of the cycle.

Close modal

Methylmalonyl-CoA Pathway

The methylmalonyl-CoA pathway is required for the further degradation of propanoyl-CoA generated during degradation of uneven fatty acids, 1-propanol, and propanoate [Halarnkar and Blomquist, 1989]. The pathway proceeds via three steps (Fig. 36A): (i) In the initial reaction, propanoyl-CoA is carboxylated to (2S)-methylmalonyl-CoA by the ATP- and biotin-dependent, heteromeric propanoyl-CoA carboxylase (PccAB). The α-subunit of the enzyme harbors the domains for biotin carboxylase and biotin carboxyl carrier protein, while the β-subunit provides the activity for the carboxyl transfer [Huang et al., 2010]. (ii) In the subsequent reaction, (2S)-methylmalonyl-CoA is epimerized by EbN1_C21340 to its (2R)-form. (iii) The latter is finally converted to succinyl-CoA via a C-skeleton rearrangement catalyzed by B12-dependent methylmalonyl-CoA mutase (Sbm) [Ludwig and Matthews, 1997]. The coding genes for all three enzymes of this pathway are present in the genome of A. aromaticum EbN1T (Fig. 36B, C).

Fig. 36.

Methylmalonyl-CoA pathway in A. aromaticumEbN1T. A Reaction sequence. B Assigned genes. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Fig. 36.

Methylmalonyl-CoA pathway in A. aromaticumEbN1T. A Reaction sequence. B Assigned genes. seq %id., amino acid sequence identities (%). C Protein inventory and proteomic coverage of the pathway.

Close modal

As a facultative anaerobic bacterium, A. aromaticum EbN1T can perform respiratory energy conservation either under anoxic conditions via denitrification or under oxic conditions via oxygen (O2) respiration. Electrons are channeled from the NADH pool into the respiratory electron transport chain via complex I, which in concert with complex II reduces the membrane-embedded quinone pool, while complex III subsequently reduces the periplasmic cytochrome c pool. Reduced quinone and cytochrome c pools then deliver electrons for the terminal reduction of NO3 to N2 (via the denitrification apparatus) or of O2 to H2O (via complex IV). In both cases, transmembrane proton motive force (pmf) is generated, which in turn is exploited by the F1F0-type ATP synthase for regeneration of ATP [Price and Driessen, 2010; Unden and Dünnwald, 2013; Vercellino and Sazanov, 2022]. Considering beyond that the presence of ETF:Q OR, Rnf complex, and tetrathionate reductase, the respiratory electron transfer capacities of A. aromaticum EbN1T are even broader.

Complex I, Complex II and ETF:Q OR, and Complex III

NADH: Ubiquinone Oxidoreductase (Complex I)

The respiratory complex I represents the main entry point of respiratory energy conservation. By regenerating NAD+, complex I receives electrons at a low redox potential (E° = −320 mV), which it transports to ubiquinone coupled to energy conservation by means of proton pumping. NADH:ubiquinone oxidoreductase is a large multi-subunit membrane protein complex encoded by the nuoAN genes, organized in a compact cluster in the genome of A. aromaticum EbN1T, as known from other bacteria [Friedrich, 1998]. Complex I represents the largest protein complex of the respiratory chain and harbors an intricate electron transfer path from flavin mononucleotide via seven FeS clusters to the binding site of ubiquinone [Efremov and Sazanov, 2011; Berrisford et al., 2016]. Recent cryo-EM analysis suggests that complex I of E. coli conducts an intriguing coupling between the proton uptake into the Q-cavity (NuoH) and the proton translocations to the periplasm via NuoL, NuoM, and NuoN [Kolata and Efremov, 2021]. In addition to complex I, A. aromaticum EbN1T also possesses a non-proton pumping type II NADH dehydrogenase (Ndh2, encoded by ndh), which is known from E. coli to be localized as homodimer to the cytoplasmic face of the membrane and to directly feed electrons into the ubiquinone pool [Heikal et al., 2014; Salewski et al., 2016]. E. coli shifts between complex I and Ndh2, which differ in efficiency and rate, to optimize growth. This shift is mediated by the global regulator FNR, which represses the expression of the ndh gene in response to anoxic conditions [Green and Guest, 1994; Meng et al., 1997; Unden et al., 2002; Friedrich and Pohl, 2013]. For facultative anaerobic A. aromaticum EbN1T, such a scenario may also be envisioned.

Succinate Dehydrogenase (Complex II) and ETF:Q OR

Succinate dehydrogenase (synonym: succinate: quinone oxidoreductase [SQR]) and electron transfer flavoprotein:quinone oxidoreductase (ETF:Q OR) both feed electrons from catabolic oxidation reactions directly into the transmembrane electron transport chain at the level of ubiquinone. Succinate dehydrogenase is a multisubunit membrane associated enzyme and part of the canonical TCA cycle. It transfers the electrons liberated by oxidizing succinate to fumarate via its FAD and 3 FeS clusters to ubiquinone. Biogenesis of the flavinylated catalytic SdhA subunit requires the chaperone SdhE. The FeS-protein SdhB establishes the interaction between SdhA and the membrane proteins SdhCD that connect to the electron acceptor ubiquinone [Cecchini et al., 2002; Ruprecht et al., 2009; Maher et al., 2018]. In A. aromaticum EbN1T, next to the formed SdhAB2C2D2 encoded in an operon-like structure, also a 2nd gene set (sdhB1C1D1) is present in the genome, albeit not expressed under the hitherto tested conditions.

The electron transfer flavoprotein of ETF:Q OR generally serves as primary electron acceptor for acyl-CoA dehydrogenases (β-oxidation of fatty acids) and catabolism of some amino acids in bacteria and mitochondria [Ghisla and Thorpe, 2004]. ETF has recently also been shown in A. aromaticum EbN1T to receive electrons from (R)-2-benzylsuccinyl-CoA dehydrogenase (BbsG) involved in the anaerobic degradation of toluene (Fig. 3) [Vogt et al., 2019]. The ETF:Q OR from human mitochondria was shown by crystal structural analysis to represent a single integral membrane protein with binding domains for FAD, the [4Fe-4S] cluster, and ubiquinone [Zhang et al., 2006]. Recent comparative genome analysis focusing on the ETF:Q OR family revealed that its members are functionally diverse and widespread, and that electron bifurcating members are characterized by the presence of 2 instead of only a single FAD [Garcia Costas et al., 2017]. The ETF:Q OR encoded in the genome of A. aromaticum EbN1T apparently carries 1 FAD (according to homology and domain structure) and should therefore belong to the non-bifurcating members of this interesting protein family.

Ubiquinol:cytochrome c Oxidoreductase (Complex III)

The membrane integral complex III (PetABC) transfers electrons from the reduced quinone-pool to the periplasmic cytochrome c (Fig. 37A). Concomitantly, protons are pumped via the Q-cycle across the cytoplasmic membrane, thereby contributing to the generation of pmf. Complex III has a dimeric structure, with each monomer consisting of the following three subunits: PetA represents the Rieske FeS protein with a [2Fe-2S] cluster, PetB the cyctochrome b harboring hemes bL and bH, and PetC the cytochrome c with heme c1 as cofactor. Structural studies of the Q-cycle mechanism suggest electron bifurcation from ubiquinol (QH2) into an inner membrane low potential chain via hemes bL and bH back to ubiquinone (Q) and a high potential chain via the [2Fe-2S] cluster and cytochrome c1 to the periplasmic cytochrome c [Hunte et al., 2003; Xia et al., 2013; Esser et al., 2019]. Complex III is suggested to form a supercomplex with complex IV [Steimle et al., 2021]. The petABC genes form an operon-like structure in the genome of A. aromaticum EbN1T and their protein products could be detected (Fig. 37B, C).

Fig. 37.

Respiratory complex I, complex II and ETF:Q OR, and complex III in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Fig. 37.

Respiratory complex I, complex II and ETF:Q OR, and complex III in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Close modal

Rnf Complex

The genome of A. aromaticum EbN1T encodes two variants of an ion-pumping reversible NADH:ferredoxin oxidoreductase termed Rnf complex (for Rhodobacter nitrogen fixation) (Fig. 38). The Rnf complex was first described as energy coupling oxidoreductase in conjunction with N2-fixation in R. capsulatus [Schmehl et al., 1993; Kumagai et al., 1997]. The Rnf complex pumps Na+ and possibly in some cases also H+, conserving the energy of excess reduced ferredoxins produced during anaerobic conversion reactions [Biegel et al., 2011; Buckel and Thauer, 2018]. The topology of the Rnf complex from Vibrio cholerae [Hreha et al., 2015] and Clostridium tetanomorphum [Vitt et al., 2022] revealed the electron transfer path from ferredoxin-binding cytoplasmic RnfB via transmembrane RnfAED and periplasmic RnfG to NAD-binding cytoplasmic RnfC.

Fig. 38.

Rnf complexes in A. aromaticumEbN1T. A Protein complexes together with the electron transfer and ion-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Fig. 38.

Rnf complexes in A. aromaticumEbN1T. A Protein complexes together with the electron transfer and ion-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Close modal

Denitrification

Denitrification viz. the dissimilatory reduction of nitrate (NO3) via nitrite (NO2), nitric oxide (NO), and nitrous oxide (N2O) to molecular nitrogen (N2) is long studied, was comprehensively reviewed by Zumft [1997], and more recently suggested to be performed by an enzyme supercomplex termed the nitrate respirasome [Borrero-de Acuña et al., 2016]. Enzymes and genes associated with denitrification in A. aromaticum EbN1T are depicted in Figure 39.

Fig. 39.

Denitrification apparatus in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. The genes coding for predicted regulators and uptake systems are marked in black and gray, respectively. C Protein inventory and proteomic coverage of the complexes.

Fig. 39.

Denitrification apparatus in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. The genes coding for predicted regulators and uptake systems are marked in black and gray, respectively. C Protein inventory and proteomic coverage of the complexes.

Close modal

Nitrate/Nitrite Transport

Since A. aromaticum EbN1T does not possess genes for periplasmic (napABC) [Stewart et al., 2002], but rather those for a cytoplasm-facing, membrane-localized nitrate reductase (narGHI) reducing NO3 to NO2 [Cole and Richardson, 2013], an efficient uptake system for nitrate is required for functional denitrification. Such systems are encoded in the genome of A. aromaticum EbN1T by the narK1 and narK2 genes, belonging to the MFS superfamily of transporters [Pao et al., 1998]. Previous studies, which were based on the genomic context of their encoding genes, as well as on biochemical and physiological evidence, indicate that these two transporters serve different physiological functions [Wood et al., 2002; Sharma et al., 2006; Goddard et al., 2017]. NarK1 is suggested to represent a NO3/H+-symporter for nitrate uptake in the context of assimilatory reduction of NO3 to NH4+, where all involved proteins are localized in the cytoplasm and the product NH4+ is widely used for anabolic purposes. On the other hand, NarK2 is suggested to function as a NO3/NO2-antiporter associated with dissimilatory nitrate reduction, where NO2 formed by NarGHI in the cytoplasm has to be exported into the periplasm, where it is reduced to N2 by the concerted activity of nitrite, nitric oxide, and nitrous oxide reductases (see paragraphs below). The intriguing mechanism of NO3/NO2-exchange by the NarK2 antiporter was elucidated by structural and mechanistic studies: (i) The five transmembrane α-helices create two domains above and below a rotation axis in the membrane. Conformational change of these domains establishes alternating access (“rocker-switch”) for the two substrates. (ii) The channel for substrate translocation is positively charged while lacking protonable residues. (iii) Differentiation between NO3 and NO2 is achieved by compound-specific hydrogen bond networks [Andrade and Einsle, 2013; Yan et al., 2013; Zheng et al., 2013; Fukuda et al., 2015].

The narK1K2 genes form an operon-like structure together with the narGHJI genes encoding dissimilatory nitrate reductase. In Ps. aeruginosa, transcriptional control of the narK1K2GHJI operon in response to anoxia and presence of nitrate was shown to be excerted at multiple levels by the oxygen regulator Anr, the N-oxide regulator Dnr, and the nitrate-sensing two-component system NarXL [Kolesnikow et al., 1992; Bauer et al., 1999; Härtig et al., 1999; Vollack et al., 1999; Schreiber et al., 2007]. Considering the similar gene organization and availability of regulators, a similar regulatory scenario can be envisioned for A. aromaticum EbN1T.

Nitrate Reductase

The membrane bound dissimilatory nitrate reductase NarGHI of A. aromaticum EbN1T is encoded in a single operon-like structure (together with NarK1K2, see above paragraph) as previously described for other denitrifying bacteria [Blasco et al., 1990; Berks et al., 1995; Philippot et al., 1997]. Nitrate reductase couples the oxidation of the reduced quinone pool (QH2) to the two-electron reduction of NO3 and NO2 with concomitant export of protons whereby it contributes to the generation of pmf. The intricate electron transfer path from ubiquinol to NO3 is assumed to proceed via the following three major steps [Bertero et al., 2003, 2005; Jormakka et al., 2004; Fedor et al., 2014a, b; Coelho and Romão, 2015]: (i) The membrane anchored subunit NarI provides the periplasm oriented Q-site (QD) for binding and stepwise oxidation of ubiquinol. While the two electrons from QH2 oxidation are consecutively accepted by distal heme bD, the two dissociated protons are likewise consecutively released to the periplasm. Electron relay in NarI then proceeds from heme bD to the proximal heme bP and therefrom to the FS4 cluster of NarH. (ii) The electron transfer subunit NarH is sandwiched between NarI and NarG. It harbors one [3Fe-4S] cluster (FS4) and three [4Fe-4S] clusters (FS1−3), which together establish the connecting conduit for transferring the electrons from heme bP of NarI to the final [4Fe-4S] cluster (FS0) of NarG. (iii) The catalytic subunit NarG carries the FS0 cluster that delivers electrons to the Mo-bisMGD, where reduction of NO3 to NO2 finally occurs. The chaperon NarJ is also encoded in the narGHJI gene cluster and is required for the molybdenum cofactor assembly in the catalytic subunit NarG [Blasco et al., 1998]; NarJ is specific for its protein substrate and belongs to the NarJ family of redox enzyme maturation proteins [Chan et al., 2014; Bay et al., 2015]. Transcriptional regulation of the narGHJI genes is assumed to involve the same complex sensory/regulatory circuits as described above for narK1K2.

Nitrite Reductase

Denitrifying bacteria perform the single-electron reduction of NO2 to NO by cytochrome cd1- (NirS) or copper-containing (NirK) dissimilatory nitrite reductases [Jüngst et al., 1991; Ye et al., 1993] of which A. aro-maticum EbN1T possesses only the former (single nirS gene). Also members of the related genus Thauera harbor nirS genes [Song and Ward, 2003]. In accord with its periplasmic location, NirS carries a Sec-leader peptide. Structural and mechanistic analysis of homodimeric cd1 nitrite reductases from Ps. aeruginosa revealed the following electron path from periplasmic reduced cytochrome c to NO2: (i) The heme c carrying domain of NirS serves as electron entry point. (ii) Electron transfer occurs via an intriguing intramolecular path from reduced heme c to oxidized heme d1. (iii) The catalytic, reduced heme d1 has a high affinity for NO2 and its reduction product NO dissociates exceptionally fast [Wherland et al., 2005; Rinaldo et al., 2011; Klünemann and Blankenfeldt, 2020]. In Ps. aeruginosa, expression of the nirS gene was shown to occur only under low oxygen or anoxic conditions, mediated by the NO-sensing, FNR-type regulator DNR (dissimilatory nitrate respiration regulator) [Kuroki et al., 2014].

Nitric Oxide Reductase

The membrane integral, heterodimeric nitric oxide reductase (NorBC) performs the intriguing two-electron reduction of NO via hyponitrite to N2O (2 NO + 2 eON = NO; ON = NO + 2 H+ → N2O + H2O) and is encoded by the colocalizing norBC genes [Zumft et al., 1994], which is also the case for A. aromaticum EbN1T. Depending on the type of primary electron donor used, three types of NOR are distinguished [Hino et al., 2012]: cNOR receives electrons from reduced periplasmic cyto-chrome c, qNOR interacts directly with QH2 for electron transfer, and qCuANOR can be replenished by both electron donors. According to highest sequence similarities with well-studied cNorBCs from Ps. stutzeri [Zumft et al., 1994] and Ps. aeruginosa [Hino et al., 2010], the NorCB of A. aromaticum EbN1T can be classified as a cNOR-type. Based on structural and mechanistic studies, the following electron path is currently assumed for cNOR [Flock et al., 2006; Hino et al., 2010, 2012; Shiro, 2012; Shiro et al., 2012]. The heme c cofactor at the periplasmic face of NorC receives an electron from reduced periplasmic cytochrome c551 and delivers it across the NorCB interface to the heme b cofactor of NorB. From there, the electron is transferred to heme b3, which together with the non-heme FeB constitutes the binuclear center of the catalytic site for NO reduction in the NorB subunit. While cNOR is phylogenetically closely related to oxygen-reducing ccb3 oxidase, it is by contrast non-electrogenic and the protons for conversion of hyponitrite to N2O are obtained from the periplasm. The NorQ and NorD accessory proteins, which are required for the insertion of the non-heme FeB cofactor into the NorB subunit [Kahle et al., 2018], are encoded by genes located in proximity to the norCB genes. The nor operon also contains the dnr gene coding for the NO-responsive Dnr regulator, which controls expression of this operon in Ps. stutzeri [Vollack and Zumft, 2001] and Dinoroseobacter shibae [Ebert et al., 2017]. Such a scenario can also be envisoned for A. aromaticum EbN1T.

Nitrous Oxide Reductase

The final step of denitrification is catalyzed by the soluble periplasmic nitrous oxide reductase (NosZ), which converts N2O in a two-electron reduction to N2 (N2O + 2 H+ + 2 e → N2 + H2O) [Zumft, 1997]. The enzymatic and crystal structural analysis of NosZ from various Pseudomonas and Paracoccus spp. yielded the following current understanding of the catalytic mechanism [Brown et al., 2000; Haltia et al., 2003; Rasmussen et al., 2005; Pomowski et al., 2011; Carreira et al., 2017]. The multicopper enzyme NosZ contains a mixed-valent CuA site (receiving electrons from periplasmic cytochrome c551 or transmembrane NosR) and a tetranuclear CuZ center. The two subunits of homodimeric NosZ are head-to-tail arranged such that the electron storage CuA site from one subunit gets close enough to the CuZ center of the other subunit for electron transfer. After entering the active site through a hydrophobic channel, the N2O substrate is bound in a side-on manner to the [4Cu:2S] (or [4Cu:1S]) cluster of the CuZ (or CuZ*) centre and reduced via two consecutive one-electron steps. While the product N2 is released via the hydrophobic channel, the by-product H2O remains in the hydrophilic part of the active site. Maturation of the NosZ and NosR proteins is a multi-facetted process. Trafficking of copper in the periplasm from the copper chaperone NosL via NosD to NosZ is facilitated by ATP-energized (NosF) conformational changes of the membrane spanning NosDFY maturation machinery [Zhang et al., 2019; Prasser et al., 2021; Müller et al., 2022]. The transmembrane FeS flavoprotein NosR is maturated by the flavinyl transferase ApbE and is assumed to transfer electrons from the reduced quinone pool to NosZ [Wunsch and Zumft, 2005; Zhang et al., 2017, 2019]. In A. aromaticum EbN1T, the nosZRDFYL genes colocalize in an operon-like structure in proximity to the nar gene cluster (Fig. 39B). Expression of the nos gene cluster may be regulated in response to O2 and NO by FNR-type regulators as previously reported for P. denitrificans [Bergaust et al., 2012]. In the case of A. aromaticum EbN1T, semi-inhibitory concentrations of toluene and ethylbenzene, respectively, cause an impairment of denitrification. Concomitantly the flux of acetyl-CoA is re-routed from the TCA cycle toward increasing formation of poly(3-hydroxybutyrate), presumptively as a means to keep alkylbenzene degradation operative despite impaired denitrification [Trautwein et al., 2008; Rabus et al., 2014].

Thiosulfate Dehydrogenase and Tetrathionate Reductase

The genome of A. aromaticum EbN1T contains the tsdAB genes potentially encoding a bifunctional thiosulfate dehydrogenase/tetrathionate reductase belonging to the widespread family of diheme c-type cytochromes (Fig. 40) [Denkmann et al., 2012; Kurth et al., 2016]. The tsdAB genes in A. aromaticum EbN1T are framed by genes encoding conserved proteins of unknown function. The TsdA(B) enzyme is assumed to oxidatively couple 2 molecules of thiosulfate (S2O32−) in the periplasm to tetrathionate (S4O62−) according to the following reaction: 2 S−SO32−O3S−S−S−SO3 + 2 e. The released electrons should then be channeled, e.g., via cytochrome c, into the electron transfer of oxygen respiration, presumably at the level of cytochrome cbb3 (see below) [Kurth et al., 2016]. An intriguing mechanism has been proposed for TsdA from C. jejuni, wherein the dissociation of a heme ligand allows the catalytic Cys-residue to react with the substrate as well as the nearby heme [Jenner et al., 2019, 2022]. The potential of A. aromaticum EbN1T to utilize thiosulfate as electron donor for aerobic or anaerobic growth has not been investigated so far.

Fig. 40.

Thiosulfate dehydrogenase and tetrathionate reductase in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Fig. 40.

Thiosulfate dehydrogenase and tetrathionate reductase in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Close modal

Moreover, the genome of A. aromaticum EbN1T harbors the complete ttrABC gene cluster coding for a tetrathionate reductase, which is neighbored by the ttrRS genes encoding an associated two-component sensory/regulatory system (Fig. 40), as previously reported for S. typhimurium [Hensel et al., 1999; Hinsley and Berks, 2002; James et al., 2013]. The molybdenum-dependent enzyme [Hinojosa-Leon et al., 1986] catalyzes the reduction to tetrathionate to thiosulfate (O3S−S−S−SO3 + 2 e → 2 S−SO32−) in the periplasm with required electrons delivered from the quinol-pool. Export of the TrtAB enzyme via the Tat-pathway [Frain et al., 2019] is supported by the presence of a canonical twin-arginine motive containing signal peptide in TtrA and an atypical one in TtrB (one arginine missing) as well as by the interdependency of TtrAB during export [Hinsley et al., 2001; James et al., 2013]. The functions of the three subunits of tetrathionate reductase are follows: TtrA represents the catalytic subunit of the reductase, carrying a Mo-bisMGD cofactor and a FeS cluster; TtrB employs four FeS clusters for electron transfer; and transmembrane TtrC is assumed to serve as anchor as well as quinol dehydrogenase. Notably, TtrC belongs to the so-called NrfD-like subunits, which were recently also proposed to possess a proton-conducting pathway [Calisto and Pereira, 2021]. Expression of the ttrABC operon is controlled by the TtrRS system in response to the presence of tetrathionate and by the global regulator Fnr under anoxic conditions [Price-Carter et al., 2001].

Oxygen Respiration

In accord with the changing oxygen concentrations A. aromaticum EbN1T is very likely encountering in its natural environment, the genome of this bacterium encodes for a rather large number of alternative terminal oxidases (complex IV, quinol/cytochrome c:O2 oxidoreductase) that should differ by their affinities toward O2 (Cox for high pO2 vs. Cco and Cyd for low pO2) as known from e.g., Ps. aeruginosa [Arai et al., 2014] and that belong to the superfamily of cytochrome c oxidases [Musser and Chan, 1998]. Their common function is to couple the oxi­dation of an electron donor (reduced cytochrome c or ubiquinol) to the four-electron reduction of O2 to water, while concomitantly contributing by proton-pumping to the generation of pmf (Fig. 41). Evidence for the presence of a supercomplex between complexes III and IV, which facilitates electron transfer possibly by improved diffusion of cytochrome c, was initially provided on the basis of copurification [Berry and Trumpower, 1985] and blue-native PAGE [Schägger and Pfeiffer, 2000], and more recently substantiated by means of cryo-EM [Sun et al., 2018; Rathore et al., 2019; Moe et al., 2021].

Fig. 41.

Oxygen respiration in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Fig. 41.

Oxygen respiration in A. aromaticumEbN1T. A Protein complexes together with electron transfer and H+-pumping paths (dashed and solid pink arrows, respectively). B Assigned genes. C Protein inventory and proteomic coverage of the complexes.

Close modal

Cytochrome aa3-Type Cytochrome c Oxidase (Cox)

The genome of A. aromaticum EbN1T encodes in total four cox gene clusters for these heme-copper quinol oxidases, one of which (coxB3A2GC) contains more than the coxAB genes encoding the core enzyme (subunits I and II). The structure and mechanism particularly of the latter have been studied in various organisms [Iwata et al., 1995; Ostermeier et al., 1997; Lyons et al., 2012; Xu et al., 2020], revealing the following insights into the core enzyme. Subunit I (CoxA) carries a low spin heme a and a high-spin binuclear center (heme a3−CuB or heme bo3−CuB); at heme a3 or bo3 oxygen is bound and reduced. In subunit II (CoxB), the binuclear CuA center serves as primary acceptors of electrons delivered by reduced cytochrome c. The enzyme possesses two proton transfer pathways, one for the reduction of O2 to H2O in subunit I and the other for proton pumping through subunit II (up to 4 per O2 reduced). While expression of the cox genes is apparently not controlled by Fnr [van Spanning et al., 1997; Otten et al., 2001], it is positively correlated with starvation conditions [Osamura et al., 2017].

Cytochrome cbb3-Type Cytochrome c Oxidase (Cco)

The genome of A. aromaticum EbN1T harbors a single operon-like structure (ccoN1O1Q1P1G1H1) and a single ccoN gene coding for a cbb3 oxidase. Cytochrome cbb3 oxidase is predominantly found in proteobacteria [Pitcher et al., 2002] and can achieve a proton-pumping efficiency of 1 H+/e [Rauhamäki et al., 2012]. The electron path proceeds from reduced soluble cytochrome c through the multisubunit CcoNOP via three heme c cofactors to heme b and from there to heme b3 and CuB; the latter two provide the electrons for final reduction of O2, which is coupled to proton pumping [Buschmann et al., 2010]. Cytochrome c mediated electron transfer between complex III and cbb3 oxidase (complex IV) was recently shown by cryo-EM to involve super-complex formation between the two complexes [Steimle et al., 2021]. The high O2-affinity of cbb3 oxidase from Pseudomonas stutzeri indicates this oxidase to play a central role under microaerophilic conditions [Pitcher et al., 2002; Xie et al., 2014].

Cytochrome bd-Type Quinol Oxidase (Cyd)

The genome of A. aromaticum EbN1T contains a single set of cydAB genes coding for a cytochrome bd-type quinol oxidase. CydAB has a high affinity toward O2 and its genes are therefore expressed at low pO2, i.e., under O2-limiting conditions [Govantes et al., 2000; Alexeeva et al., 2002]. Structural studies [Borisov et al., 2011; Safarian et al., 2016; Forte et al., 2017; Grund et al., 2021] showed that CydAB harbors three heme cofactors (b558, b595, and d) tightly aligned for direct electron relay from QH2 to O2. While oxygen diffuses freely to heme d, the protons are transferred there from the cytoplasm through a dedicated path in the protein. In contrast to heme-copper oxygen reductases, CydAB is not a proton pump.

ATP Synthase

The F1F0-ATP synthase is a unique macromolecular machine that transforms the electrochemical energy (pmf) via mechanical energy (subunit rotation) into chemical energy (ADP + Pi → ATP) and has been intensively studied over the past decades [Mao and Weber, 2007; Weber, 2006, 2010; Watanabe et al., 2010; Kühlbrandt, 2019; Sobti et al., 2019; Guo and Rubinstein, 2022]. The enzyme of E. coli consists of the membrane embedded F0 subcomplex (ab2c10 composition) and the peripheral F1 subcomplex (α3β3γδε composition). All of these components are encoded in the genome of A. aromaticum EbN1T (Fig. 42). The principle mechanism of a F1F0-ATP synthase is as follows. Pmf energized translocation of protons through a channel between a and c subunits of the F0 subcomplex drive the rotation of the c10 ring. The latter forms the rotor together with the γ and ε subunits of the F1 subcomplex. The γ subunit of the rotor penetrates the F1 cylinder formed by alternating α and β subunits. Rotation of the γ subunit translates into 120° stepwise rotations of this cylinder. Thereby torque is generated, which induces conformational changes in the three catalytic β-subunits yielding the conversion of bound ADP and Pi to ATP. According to the “tri-site” mechanism one catalytic site binds substrate, the second performs the catalysis, and the third releases the product. The peripheral stalk composed of the b2δ-subunits holds together the F1 and F2 subcomplexes during the rotational movements while ensuring elasticity at the same time.

Fig. 42.

ATP synthase in A. aromaticumEbN1T. A Protein complex and H+-translocation path (solid pink arrow). B Assigned genes. C Protein inventory and proteomic coverage of the synthase.

Fig. 42.

ATP synthase in A. aromaticumEbN1T. A Protein complex and H+-translocation path (solid pink arrow). B Assigned genes. C Protein inventory and proteomic coverage of the synthase.

Close modal

We are grateful to many colleagues for their contributions to the field of anaerobic degradation of aromatic compounds.

The authors have no conflicts of interest to declare.

This study was supported by the Deutsche Forschungsgemeinschaft within the framework of the research training group Chemical Bond Activation (GRK 2226).

R.R. conceived the study; P.B., D.W., M.N.-S., D.S., and R.R. conducted reannotation of the genome; L.W. compiled proteomic data; R.R. wrote the manuscript with input from all authors; and P.B. prepared the figures. All authors have agreed to the final version of the manuscript.

1.
Adams
MA
,
Singh
VK
,
Keller
BO
,
Jia
Z
.
Structural and biochemical characterization of gentisate 1,2-dioxygenase from Escherichia coli O157:H7
.
Mol Microbiol
.
2006 Sep
;
61
(
6
):
1469
84
.
2.
Ævarsson A, Seger K, Turley S, Sokatch JR, Hol WGJ. Crystal structure of 2-oxoisovalerate and dehydrogenase and the architecture of 2-oxo acid dehydrogenase multienzyme complexes. Nat Struct Biol. 1999 Aug;6(8):785–92.
3.
Aleksandrov
A
,
Zvereva
E
,
Field
M
.
The mechanism of citryl-coenzyme A formation catalyzed by citrate synthase
.
J Phys Chem B
.
2014 May
;
118
(
17
):
4505
13
.
4.
Alexeeva
S
,
Hellingwerf
KJ
,
Teixeira de Mattos
MJ
.
Quantitative assessment of oxygen availability: perceived aerobiosis and its effect on flux distribution in the respiratory chain of Escherichia coli
.
J Bacteriol
.
2002 Mar
;
184
(
5
):
1402
6
.
5.
An
JH
,
Kim
YS
.
A gene cluster encoding malonyl-CoA decarboxylase (MatA), malonyl-CoA synthetase (MatB) and a putative dicarboxylate carrier protein (MatC) in Rhizobium trifolii. Cloning, sequencing, and expression of the enzymes in Escherichia coli
.
Eur J Biochem
.
1998 Oct
;
257
(
2
):
395
402
.
6.
Anders
H-J
,
Kaetzke
A
,
Kämpfer
P
,
Ludwig
W
,
Fuchs
G
.
Taxonomic position of aromatic-degrading denitrifying pseudomonad strains K 172 and KB 740 and their description as new members of the genera Thauera, as Thauera aromatica sp. nov., and Azoarcus, as Azoarcus evansii sp. nov., respectively, members of the beta subclass of the Proteobacteria
.
Int J Syst Bacteriol
.
1995 Apr
;
45
(
2
):
327
33
.
7.
Anderson
JJ
,
Dagley
S
.
Catabolism of tryptophan, anthranilate, and 2,3-dihydroxybenzoate in Trichosporon cutaneum
.
J Bacteriol
.
1981 Apr
;
146
(
1
):
291
7
.
8.
Andrade
SLA
,
Einsle
O
.
The tricky task of nitrate/nitrite antiport
.
Angew Chem Int Ed
.
2013 Sep
;
52
(
40
):
10422
4
.
9.
Andrews
SC
,
Berks
BC
,
McClay
J
,
Ambler
A
,
Quail
MA
,
Golby
P
, et al
.
A 12-cistron Escherichia coli operon (hyf) encoding a putative proton-translocating formate hydrogenlyase system
.
Microbiology
.
1997 Nov
;
143
(
11
):
3633
47
.
10.
Arai
H
,
Kawakami
T
,
Osamura
T
,
Hirai
T
,
Sakai
Y
,
Ishii
M
.
Enzymatic characterization and in vivo function of five terminal oxidases in Pseudomonas aeruginosa
.
J Bacteriol
.
2014 Dec
;
196
(
24
):
4206
15
.
11.
Arndt
F
,
Schmitt
G
,
Winiarska
A
,
Saft
M
,
Seubert
A
,
Kahnt
J
, et al
.
Characterization of an aldehyde oxidoreductase from the mesophilic bacterium Aromatoleum aromaticum EbN1, a member of a new subfamily of tungsten-containing enzymes
.
Front Microbiol
.
2019 Jan
;
10
:
71
.
12.
Arnoux
P
,
Ruppelt
C
,
Oudouhou
F
,
Lavergne
J
,
Siponen
MI
,
Toci
R
, et al
.
Sulphur shuttling across a chaperone during molybdenum cofactor maturation
.
Nat Commun
.
2015 Feb
;
6
:
6148
.
13.
Asao
M
,
Alber
BE
.
Acrylyl-coenzyme A reductase, an enzyme involved in the assimilation of 3-hydroxypropionate by Rhodobacter sphaeroides
.
J Bacteriol
.
2013 Oct
;
195
(
20
):
4716
25
.
14.
ATSDR
.
Toxicological Profile for Cresols. Agency for Toxic Substances and Disease Registry
.
U.S. Department of Health and Human Services
;
2008
.
15.
Attwood
PV
,
Wallace
JC
.
Chemical and catalytic mechanisms of carboxyl transfer reactions in biotin-dependent enzymes
.
Acc Chem Res
.
2002 Feb
;
35
(
2
):
113
20
.
16.
Azizan
A
,
Sherin
D
,
DiRusso
CC
,
Black
PN
.
Energetics underlying the process of long-chain fatty acid transport
.
Arch Biochem Biophys
.
1999 May
;
365
(
2
):
299
306
.
17.
Bains
J
,
Leon
R
,
Boulanger
MJ
.
Structural and biophysical characterization of BoxC from Burkholderia xenovorans LB400. A novel ring-cleaving enzyme in the crotonase superfamily
.
J Biol Chem
.
2009 Jun
;
284
(
24
):
16377
85
.
18.
Bauer
CE
,
Elsen
S
,
Bird
TH
.
Mechanisms for redox control of gene expression
.
Annu Rev Microbiol
.
1999 Oct
;
53
:
495
523
.
19.
Bay
DC
,
Chan
CS
,
Turner
RJ
.
NarJ subfamily system specific chaperone diversity and evolution is directed by respiratory enzyme associations
.
BMC Evol Biol
.
2015 Jun
;
15
:
110
.
20.
Bayly
RC
,
Chapman
PJ
,
Dagley
S
,
Di Berardino
D
.
Purification and some properties of maleylpyruvate hydrolase and fumarylpyruvate hydrolase from Pseudomonas alcaligenes
.
J Bacteriol
.
1980 Jul
;
143
(
1
):
70
7
.
21.
Becker
P
,
Kirstein
S
,
Wünsch
D
,
Koblitz
J
,
Buschen
R
,
Wöhlbrand
L
, et al
.
Systems biology of aromatic compound catabolism in facultative anaerobic Aromatoleum aromaticum EbN1T
.
mSystems
.
2022a Nov‒Dec
;
7
(
6
):
00685
22
.
22.
Becker
P
,
Döhmann
A
,
Wöhlbrand
L
,
Thies
D
,
Hinrichs
C
,
Buschen
R
, et al
.
Complex and flexible catabolism in Aromatoleum aromaticum pCyN1
.
Environ Microbiol
.
2022b Jul
;
24
(
7
):
3195
211
.
23.
Beltrametti
F
,
Marconi
AM
,
Bestetti
G
,
Colombo
C
,
Galli
E
,
Ruzzi
M
, et al
.
Sequencing and functional analysis of styrene catabolism genes from Pseudomonas fluorescens ST
.
Appl Environ Microbiol
.
1997 Jun
;
63
(
6
):
2232
9
.
24.
Bergaust
L
,
van Spanning
RJM
,
Frostegård
Å
,
Bakken
LR
.
Expression of nitrous oxide reductase in Paracoccus denitrificans is regulated by oxygen and nitric oxide through FnrP and NNR
.
Microbiology
.
2012 Mar
;
158
:
826
34
.
25.
Bergaust
LL
,
Hartsock
A
,
Liu
B
,
Bakken
LR
,
Shapleigh
JP
.
Role of norEF in denitrification, elucidated by physiological experiments with Rhodobacter sphaeroides
.
J Bacteriol
.
2014 Jun
;
196
(
12
):
2190
200
.
26.
Berks
BC
,
Page
MD
,
Richardson
DJ
,
Reilly
A
,
Cavill
A
,
Outen
F
, et al
.
Sequence analysis of subunits of the membrane-bound nitrate reductase from a denitrifying bacterium: the integral membrane subunit provides a prototype for the dihaem electron-carrying arm of a redox loop
.
Mol Microbiol
.
1995 Jan
;
15
(
2
):
319
31
.
27.
Berrisford
JM
,
Baradaran
R
,
Sazanov
LA
.
Structure of bacterial respiratory complex I
.
Biochim Biophys Acta
.
2016 Jul
;
1857
(
7
):
892
901
.
28.
Berry
EA
,
Trumpower
BL
.
Isolation of ubiquinol oxidase from Paracoccus denitrificans and resolution into cytochrome bc1 and cytochrome c-aa3 complexes
.
J Biol Chem
.
1985 Feb
;
260
(
4
):
2458
67
.
29.
Bertero
MG
,
Rothery
RA
,
Palak
M
,
Hou
C
,
Lim
D
,
Blasco
F
, et al
.
Insights into the respiratory electron transfer pathway from the structure of nitrate reductase A
.
Nat Struct Biol
.
2003 Sep
;
10
(
9
):
681
7
.
30.
Bertero
MG
,
Rothery
RA
,
Boroumand
N
,
Palak
M
,
Blasco
F
,
Ginet
N
, et al
.
Structural and biochemical characterization of a quinol binding site of Escherichia coli nitrate reductase A
.
J Biol Chem
.
2005 Apr
;
280
(
15
):
14836
43
.
31.
Biegel
E
,
Schmidt
S
,
González
JM
,
Müller
V
.
Biochemistry, evolution and physiological function of the Rnf complex, a novel ion-motive electron transport complex in prokaryotes
.
Cell Mol Life Sci
.
2011 Feb
;
68
(
4
):
613
34
.
32.
Biegert
T
,
Altenschmidt
U
,
Eckerskorn
C
,
Fuchs
G
.
Enzymes of anaerobic metabolism of phenolic compounds. 4-Hydroxybenzoate-CoA ligase from a denitrifying Pseudomonas species
.
Eur J Biochem
.
1993 Apr
;
213
(
1
):
555
61
.
33.
Biegert
T
,
Altenschmidt
U
,
Eckerskorn
C
,
Fuchs
G
.
Purification and properties of benzyl alcohol dehydrogenase from a denitrifying Thauera sp
.
Arch Microbiol
.
1995 Jun
;
163
(
6
):
418
23
.
34.
Blasco
F
,
Iobbi
C
,
Ratouchniak
J
,
Bonnefoy
V
,
Chippaux
M
.
Nitrate reductase of Escherichia coli: sequence of the second nitrate reductase and comparison with that encoded by the narGHJI operon
.
Mol Gen Genet
.
1990 Jun
;
222
(
1
):
104
11
.
35.
Blasco
F
,
Dos Santos
J-P
,
Magalon
A
,
Frixon
C
,
Guigliarelli
B
,
Santini
C-L
, et al
.
NarJ is a specific chaperone required for molybdenum cofactor assembly in nitrate reductase A of Escherichia coli
.
Mol Microbiol
.
1998 May
;
28
(
3
):
435
47
.
36.
Bock T, Reichelt J, Müller R, Blankenfeldt W. The structure of LiuC, a 3-hydroxy-3-methylglutaconyl CoA dehydratase involved in isovaleryl-CoA biosynthesis in Myxococcus xanthus, reveals insights into specificity and catalysis. ChemBioChem. 2016 Sep;17(17):1658–64.
37.
Böck
A
,
Forchhammer
K
,
Heider
J
,
Leinfelder
W
,
Sawers
G
,
Veprek
B
, et al
.
Selenocysteine: the 21st amino acid
.
Mol Microbiol
.
1991 Mar
;
5
(
3
):
515
20
.
38.
Boernke
WE
,
Sanville Millard
C
,
Wilkins Stevens
P
,
Kakar
SN
,
Stevens
FJ
,
Donnelly
MI
.
Stringency of substrate specificity of Escherichia coli malate dehydrogenase
.
Arch Biochem Biophys
.
1995 Sep
;
322
(
1
):
43
52
.
39.
Bolduc
JM
,
Dyer
DH
,
Scott
WG
,
Singer
P
,
Sweet
RM
,
Koshland
DE
Jr
, et al
.
Mutagenesis and Laue structures of enzyme intermediates: isocitrate dehydrogenase
.
Science
.
1995 Jun
;
268
(
5215
):
1312
8
.
40.
Boll
M
,
Fuchs
G
.
Benzoyl-coenzyme A reductase (dearomatizing), a key enzyme of anaerobic aromatic metabolism. ATP dependence of the reaction, purification and some properties of the enzyme from Thauera aromatica strain K172
.
Eur J Biochem
.
1995 Dec
;
234
(
3
):
921
33
.
41.
Boll
M
,
Albracht
SSP
,
Fuchs
G
.
Benzoyl-CoA reductase (dearomatizing), a key enzyme of anaerobic aromatic metabolism. A study of adenosinetriphosphate activity, ATP stoichiometry of the reaction and EPR properties of the enzyme
.
Eur J Biochem
.
1997 Mar
;
244
(
3
):
840
51
.
42.
Boll
M
,
Laempe
D
,
Eisenreich
W
,
Bacher
A
,
Mittelberger
T
,
Heinze
J
, et al
.
Nonaromatic products from anoxic conversion of benzoyl-CoA with benzoyl-CoA reductase and cyclohexa-1,5-diene-1-carbonyl-CoA hydratase
.
J Biol Chem
.
2000 Jul
;
275
(
29
):
21889
95
.
43.
Boll
M
,
Fuchs
G
,
Meier
C
,
Trautwein
A
,
El Kasmi
A
,
Ragsdale
SW
, et al
.
Redox centers of 4-hydroxybenzoyl-CoA reductase, a member of the xanthine oxidase family of molybdenum-containing enzymes
.
J Biol Chem
.
2001 Dec
;
276
(
51
):
47853
62
.
44.
Boll
M
,
Geiger
R
,
Junghare
M
,
Schink
B
.
Microbial degradation of phthalates: biochemistry and environmental implications
.
Environ Microbiol Rep
.
2020 Feb
;
12
(
1
):
3
15
.
45.
Borisov
VB
,
Gennis
RB
,
Hemp
J
,
Verkhovsky
MI
.
The cytochrome bd respiratory oxygen reductases
.
Biochim Biophys Acta
.
2011 Nov
;
1807
(
11
):
1398
413
.
46.
Borrero-de Acuña
JM
,
Rohde
M
,
Wissing
J
,
Jänsch
L
,
Schobert
M
,
Molinari
G
, et al
.
Protein network of the Pseudomonas aeruginosa denitrification apparatus
.
J Bacteriol
.
2016 May
;
198
(
9
):
1401
13
.
47.
Boyd
JM
,
Ensign
SA
.
ATP-dependent enolization of acetone by acetone carboxylase from Rhodobacter capsulatus
.
Biochemistry
.
2005 Jun
;
44
(
23
):
8543
53
.
48.
Boyd
JM
,
Ellsworth
H
,
Ensign
SA
.
Bacterial acetone carboxylase is a manganese-dependent metalloenzyme
.
J Biol Chem
.
2004 Nov
;
279
(
45
):
46644
51
.
49.
Brackmann
R
,
Fuchs
G
.
Enzymes of anaerobic metabolism of phenolic compounds. 4-Hydroxybenzoyl-CoA reductase (dehydroxylating) from a denitrifying Pseudomonas species
.
Eur J Biochem
.
1993 Apr
;
213
(
1
):
563
71
.
50.
Breese
K
,
Fuchs
G
.
4-Hydroxybenzoyl-CoA reductase (dehydroxylating) from the denitrifying bacterium Thauera aromatica. Prosthetic groups, electron donor, and genes of a member of the molybdenum-flavin-iron-sulfur proteins
.
Eur J Biochem
.
1998 Feb
;
251
(
3
):
916
23
.
51.
Breese
K
,
Boll
M
,
Alt-Mörbe
J
,
Schägger
H
,
Fuchs
G
.
Genes coding for the benzoyl-CoA pathway of anaerobic aromatic metabolism in the bacterium Thauera aromatica
.
Eur J Biochem
.
1998 Aug
;
256
(
1
):
148
54
.
52.
Breinig
S
,
Schiltz
E
,
Fuchs
G
.
Genes involved in anaerobic metabolism of phenol in the bacterium Thauera aromatica
.
J Bacteriol
.
2000 Oct
;
182
(
20
):
5849
63
.
53.
Brown
K
,
Tegoni
M
,
Prudêncio
M
,
Pereira
AS
,
Besson
S
,
Moura
JJ
, et al
.
A novel type of catalytic copper cluster in nitrous oxide reductase
.
Nat Struct Biol
.
2000 Mar
;
7
(
3
):
191
5
.
54.
Buckel
W
,
Thauer
RK
.
Flavin-based electron bifurcation, ferredoxin, flavodoxin, and anaerobic respiration with protons (Ech) and NAD+ (Rnf) as electron acceptors: a historical review
.
Front Microbiol
.
2018 Mar
;
9
:
401
.
55.
Buckel
W
,
Kung
JW
,
Boll
M
.
The benzoyl-coenzyme A reductase and 2-hydroxyacyl-coenzyme A dehydratase radical enzyme family
.
ChemBioChem
.
2014 Oct
;
15
(
15
):
2188
94
.
56.
Buschen R, Lambertus P, Scheve S, Horst S, Song F, Wöhlbrand L, et al. Sensitive and selective phenol sensing in denitrifying Aromatoleum aromaticum EbN1T. Microbiol Spectr. 2023 Dec;11(6):e0210023.
57.
Buschmann
S
,
Warkentin
E
,
Xie
H
,
Langer
JD
,
Ermler
U
,
Michel
H
.
The structure of cbb3 cytochrome oxidase provides insights into proton pumping
.
Science
.
2010 Jul
;
329
(
5989
):
327
30
.
58.
Büsing
I
,
Höffken
HW
,
Breuer
M
,
Wöhlbrand
L
,
Hauer
B
,
Rabus
R
.
Molecular genetic and crystal structural analysis of 1-(4-hydroxyphenyl)-ethanol dehydrogenase from “Aromatoleum aromaticum” EbN1
.
J Mol Microbiol Biotechnol
.
2015a Oct
;
25
(
5
):
327
39
.
59.
Büsing
I
,
Kant
M
,
Dörries
M
,
Wöhlbrand
L
,
Rabus
R
.
The predicted σ54-dependent regulator EtpR is essential for expression of genes for anaerobic p-ethylphenol and p-hydroxyacetophenone degradation in “Aromatoleum aromaticum” EbN1
.
BMC Microbiol
.
2015b Nov
;
15
:
251
.
60.
Cain
RB
.
Anthranilic acid metabolism by microorganisms. Formation of 5-hydroxyanthranilate as an intermediate in anthranilate metabolism by Nocardia opaca
.
Antonie van Leeuwenhoek
.
1968 Dec
;
34
(
1
):
417
32
.
61.
Calisto
F
,
Pereira
MM
.
The ion-translocating NrfD-like subunit of energy-transducing membrane complexes
.
Front Chem
.
2021 Apr
;
9
:
663706
.
62.
Calvo
JM
,
Matthews
RG
.
The leucine-responsive regulatory protein, a global regulator of metabolism in Escherichia coli
.
Microbiol Rev
.
1994 Sep
;
58
(
3
):
466
90
.
63.
Campbell
JW
,
Cronan Jr
JE
.
The enigmatic Escherichia coli fadE gene is yafH
.
J Bacteriol
.
2002 Jul
;
184
(
13
):
3759
64
.
64.
Carreira
C
,
Pauleta
SR
,
Moura
I
.
The catalytic cycle of nitrous oxide reductase – the enzyme that catalyzes the last step of denitrification
.
J Inorg Biochem
.
2017 Dec
;
177
:
423
34
.
65.
Cecchini
G
,
Schröder
I
,
Gunsalus
RP
,
Maklashina
E
.
Succinate dehydrogenase and fumarate reductase from Escherichia coli
.
Biochim Biophys Acta
.
2002 Jan
;
1553
(
1-2
):
140
57
.
66.
Chai
Y
,
Kolter
R
,
Losick
R
.
A widely conserved gene cluster required for lactate utilization in Bacillus subtilis and its involvement in biofilm formation
.
J Bacteriol
.
2009 Apr
;
191
(
8
):
2423
30
.
67.
Champion
KM
,
Zengler
K
,
Rabus
R
.
Anaerobic degradation of ethylbenzene and toluene in denitrifying strain EbN1 proceeds via independent substrate-induced pathways
.
J Mol Microbiol Biotechnol
.
1999 Aug
;
1
(
1
):
157
64
.
68.
Chan
CS
,
Bay
DC
,
Leach
TGH
,
Winstone
TML
,
Kuzniatsova
L
,
Tran
VA
, et al
.
‘Come into the fold’: a comparative analysis of bacterial redox enzyme maturation protein members of the NarJ subfamily
.
Biochim Biophys Acta
.
2014 Dec
;
1838
(
12
):
2971
84
.
69.
Chávez-Avilés
M
,
Díaz-Pérez
AL
,
Reyes-de la Cruz
H
,
Campos-Garcia
J
.
The Pseudomonas aeruginosa liuE gene encodes the 3-hydroxy-3-methylglutaryl coenzyme A lyase, involved in leucine and acyclic terpene catabolism
.
FEMS Microbiol Lett
.
2009 Jul
;
296
(
1
):
117
23
.
70.
Chen
J
,
Li
W
,
Wang
M
,
Zhu
G
,
Liu
D
,
Sun
F
, et al
.
Crystal structure und mutagenic analysis of GDOsp, a gentisate 1,2-dioxygenase from Silicibacter pomeroyi
.
Protein Sci
.
2008 Aug
;
17
(
8
):
1362
73
.
71.
Cheng
M
,
Chen
D
,
Parales
RE
,
Jiang
J
.
Oxygenases as powerful weapons in the microbial degradation of pesticides
.
Annu Rev Microbiol
.
2022 Sep
;
76
:
325
48
.
72.
Chiu
K-C
,
Lin
C-J
,
Shaw
G-C
.
Transcriptional regulation of the L-lactate permease gene lutP by the LutR repressor of Bacillus subtilis RO-NN-1
.
Microbiology
.
2014 Oct
;
160
(
Pt 10
):
2178
89
.
73.
Choudhary
KS
,
Kleinmanns
JA
,
Decker
K
,
Sastry
AV
,
Gao
Y
,
Szubin
R
, et al
.
Elucidation of regulatory modes for five two-component systems in Escherichia coli reveals novel relationships
.
mSystems
.
2020 Nov–Dec
;
5
(
6
):
e00980
20
.
74.
Clark
DP
,
Cronan
JE
.
Two-carbon compounds and fatty acids as carbon sources
.
EcoSalPlus
.
2005
. https://doi.org/10.1128/ecosalplus.3.4.4.
75.
Clark
DD
,
Ensign
SA
.
Evidence for an inducible nucleotide-dependent acetone carboxylase in Rhodococcus rhodochrous B276
.
J Bacteriol
.
1999 May
;
181
(
9
):
2752
8
.
76.
Coelho
C
,
Romão
MJ
.
Structural and mechanistic insights on nitrate reductases
.
Protein Sci
.
2015 Dec
;
24
(
12
):
1901
11
.
77.
Cole
JA
,
Richardson
DJ
.
Respiration of nitrate and nitrite
.
EcoSalPlus
.
2008
. https://doi.org/10.1128/ecosal.3.2.5.
78.
Consevage
MW
,
Phillips
AT
.
Sequence analysis of the hutH gene encoding histidine ammonia-lyase in Pseudomonas putida
.
J Bacteriol
.
1990 May
;
172
(
5
):
2224
9
.
79.
Cronan
JE
.
The Escherichia coli FadR transcription factor: too much of a good thing
.
Mol Microbiol
.
2021 Jun
;
115
(
6
):
1080
5
.
80.
Cronan Jr
JE
,
Laporte
D
.
Tricarboxylic acid cycle and glyoxylate bypass
.
EcoSalPlus
.
2005
. https://doi.org/10.1128/ecosalplus.3.5.2.
81.
Cronin
CN
,
Kim
J
,
Fuller
JH
,
Zhang
X
,
McIntire
WS
.
Organization and sequences of p-hydroxybenzaldehyde dehydrogenase and other plasmid-encoded genes for early enzymes of the p-cresol degradative pathway in Pseudomonas putida NCIMB 9866 and 9869
.
DNA Seq
.
1999 Jul
;
10
(
1
):
7
17
.
82.
Cunane
LM
,
Chen
Z-W
,
Shamala
N
,
Mathews
FS
,
Cronin
CN
,
McIntire
WS
.
Structures of the flavocytochrome p-cresol methylhydroxylase and its enzyme-substrate complex: gated substrate entry and proton relays support the proposed catalytic mechanism
.
J Mol Biol
.
2000 Jan
;
295
(
2
):
357
74
.
83.
Cunane
LM
,
Chen
Z-W
,
McIntire
WS
,
Mathews
FS
.
p-Cresol methylhydroxylase: alteration of the structure of the flavoprotein subunit upon its binding to the cytochrome subunit
.
Biochemistry
.
2005 Mar
;
44
(
8
):
2963
73
.
84.
Cuthbertson
L
,
Nodwell
JR
.
The TetR family of regulators
.
Microbiol Mol Biol Rev
.
2013 Sep
;
77
(
3
):
440
75
.
85.
Datta
S
,
Mori
Y
,
Takagi
K
,
Kawaguchi
K
,
Chen
Z-W
,
Okajima
T
, et al
.
Structure of a quinohemoprotein amine dehydrogenase with an uncommon redox cofactor and highly unusual crosslinking
.
Proc Natl Acad Sci USA
.
2001 Dec
;
98
(
25
):
14268
73
.
86.
de Kok
A
,
Hengeveld
AF
,
Martin
A
,
Westphal
AH
.
The pyruvate dehydrogenase multi-enzyme complex from Gram-negative bacteria
.
Biochim Biophys Acta
.
1998 Jun
;
1385
(
2
):
353
66
.
87.
de Leeuw
JW
,
Versteegh
GJM
,
van Bergen
PF
.
Biomacromolecules of algae and plants and their fossil analogues
.
Plant Ecol
.
2006 Jan
;
182
(
1-2
):
209
33
.
88.
Debnar-Daumler
C
,
Seubert
A
,
Schmitt
G
,
Heider
J
.
Simultaneous involvement of a tungsten-containing aldehyde:ferredoxin oxidoreductase and a phenylacetaldehyde dehydrogenase in anaerobic phenylalanine metabolism
.
J Bacteriol
.
2014 Jan
;
196
(
2
):
483
92
.
89.
Dehning
I
,
Schink
B
.
Anaerobic degradation of malonate via malonyl-CoA by Sporomusa malonica, Klebsiella oxytoca, and Rhodobacter capsulatus
.
Antonie van Leeuwenhoek
.
1994 Dec
;
66
(
4
):
343
50
.
90.
Denef
VJ
,
Patrauchan
MA
,
Florizone
C
,
Park
J
,
Tsoi
TV
,
Verstraete
W
, et al
.
Growth substrate- and phase-specific expression of biphenyl, benzoate, and C1 metabolic pathways in Burkholderia xenovorans LB400
.
J Bacteriol
.
2005 Dec
;
187
(
23
):
7996
8005
.
91.
Denkmann
K
,
Grein
F
,
Zigann
R
,
Siemen
A
,
Bergmann
J
,
van Helmont
S
, et al
.
Thiosulfate dehydrogenase: a widespread unusual acidophilic c-type cytochrome
.
Environ Microbiol
.
2012 Oct
;
14
(
10
):
2673
88
.
92.
Di Gennaro
P
,
Terreni
P
,
Masi
G
,
Botti
S
,
De Ferra
F
,
Bestetti
G
.
Identification and characterization of genes involved in naphthalene degradation in Rhodococcus opacus R7
.
Appl Microbiol Biotechnol
.
2010 Jun
;
87
(
1
):
297
308
.
93.
Ding
B
,
Schmeling
S
,
Fuchs
G
.
Anaerobic metabolism of catechol by the denitrifying bacterium Thauera aromatica – a result of promiscuous enzymes and regulators
.
J Bacteriol
.
2008 Mar
;
190
(
5
):
1620
30
.
94.
Dittrich
CR
,
Bennett
GN
,
San
K-Y
.
Characterization of the acetate-producing pathways in Escherichia coli
.
Biotechnol Prog
.
2005 Jul‒Aug
;
21
(
4
):
1062
7
.
95.
Dörner
E
,
Boll
M
.
Properties of 2-oxoglutarate:ferredoxin oxidoreductase from Thauera aromatica and its role in enzymatic reduction of the aromatic ring
.
J Bacteriol
.
2002 Jul
;
184
(
14
):
3975
83
.
96.
Duboc-Toia
C
,
Hassan
AK
,
Mulliez
E
,
Ollagnier-de Choudens
S
,
Fontecave
M
,
Leutwein
C
, et al
.
Very high-field EPR study of glycyl radical enzymes
.
J Am Chem Soc
.
2003 Jan
;
125
(
1
):
38
9
.
97.
Dudzik
A
,
Snoch
W
,
Borowiecki
P
,
Opalinska-Piskorz
J
,
Witko
M
,
Heider
J
, et al
.
Asymmetric reduction of ketones and β-keto esters by (S)-1-phenylethanol dehydrogenase from denitrifying bacterium Aromatoleum aromaticum
.
Appl Microbiol Biotechnol
.
2015 Jun
;
99
(
12
):
5055
69
.
98.
Dullius
CH
,
Chen
C-Y
,
Schink
B
.
Nitrate-dependent degradation of acetone by Alicycliphilus and Paracoccus strains and comparison of acetone carboxylase enzymes
.
Appl Environ Microbiol
.
2011 Oct
;
77
(
19
):
6821
5
.
99.
Durante-Rodríguez
G
,
Valderrama
JA
,
Mancheňo
JM
,
Rivas
G
,
Alfonso
C
,
Arias-Palomo
E
, et al
.
Biochemical characterization of the transcriptional regulator BzdR from Azoarcus sp. CIB
.
J Biol Chem
.
2010 Nov
;
285
(
46
):
35694
705
.
100.
Durham
DR
,
Perry
JJ
.
Purification and characterization of a heme-containing amine dehydrogenase from Pseudomonas putida
.
J Bacteriol
.
1978 Jun
;
134
(
3
):
837
43
.
101.
Ebenau-Jehle
C
,
Thomas
M
,
Scharf
G
,
Kockelkorn
D
,
Knapp
B
,
Schühle
K
, et al
.
Anaerobic metabolism of indoleacetate
.
J Bacteriol
.
2012 Jun
;
194
(
11
):
2894
903
.
102.
Ebenau-Jehle
C
,
Mergelsberg
M
,
Fischer
S
,
Brüls
T
,
Jehmlich
N
,
von Bergen
M
, et al
.
An unusual strategy for the anoxic biodegradation of phthalate
.
ISME J
.
2017 Jan
;
11
(
1
):
224
36
.
103.
Ebert
M
,
Schweyen
P
,
Bröring
M
,
Laass
S
,
Härtig
E
,
Jahn
D
.
Heme and nitric oxide binding by the transcriptional regulator DnrF from the marine bacterium Dinoroseobacter shibae increases napD promoter affinity
.
J Biol Chem
.
2017 Sep
;
292
(
37
):
15468
80
.
104.
Efremov
RG
,
Sazanov
LA
.
Structure of the membrane domain of respiratory complex I
.
Nature
.
2011 Aug
;
476
(
7361
):
414
20
.
105.
Egland
PG
,
Harwood
CS
.
BadR, a new MarR family member, regulates anaerobic benzoate degradation by Rhodopseudomonas palustris in concert with AdaR, an Fnr family member
.
J Bacteriol
.
1999 Apr
;
181
(
7
):
2102
9
.
106.
Egland
PG
,
Harwood
CS
.
HbaR, a 4-hydroxybenzoate sensor and FNR-CRP superfamily member, regulates anaerobic 4-hydroxybenzoate degradation by Rhodopseudomonas palustris
.
J Bacteriol
.
2000 Jan
;
182
(
1
):
100
6
.
107.
Egland
PG
,
Pelletier
DA
,
Dispensa
M
,
Gibson
J
,
Harwood
CS
.
A cluster of bacterial genes for anaerobic benzene ring biodegradation
.
Proc Natl Acad Sci USA
.
1997 Jun
;
94
(
12
):
6484
9
.
108.
Eklund
H
,
Ramaswamy
S
.
Three-dimensional structures of MDR alcohol dehydrogenases
.
Cell Mol Life Sci
.
2008 Dec
;
65
(
24
):
3907
17
.
109.
Esser
L
,
Zhou
F
,
Yu
C-A
,
Xia
D
.
Crystal structure of bacterial cytochrome bc1 in complex with azoxystrobin reveals a conformational switch of the Rieske iron‒sulfur protein subunit
.
J Biol Chem
.
2019 Aug
;
294
(
32
):
12007
19
.
110.
Fang
T
,
Li
D-F
,
Zhou
N-Y
.
Identification and clarification of the role of key active site residues in bacterial glutathione S-transferase zeta/maleylpyruvate isomerase
.
Biochem Biophys Res Commun
.
2011 Jul
;
410
(
3
):
452
6
.
111.
Fedor
JG
,
Rothery
RA
,
Giraldi
KS
,
Weiner
JH
.
Q-site occupancy defines heme heterogeneity in Escherichia coli nitrate reductase (NarGHI)
.
Biochemistry
.
2014a Mar
;
53
(
11
):
1733
41
.
112.
Fedor
JG
,
Rothery
RA
,
Weiner
JH
.
A new paradigm for electron transfer through Escherichia coli nitrate reductase A
.
Biochemistry
.
2014b Jul
;
53
(
28
):
4549
56
.
113.
Fedorov
AA
,
Martí-Arbona
R
,
Nemmara
VV
,
Hitchcock
D
,
Fedorov
EV
,
Almo
SC
, et al
.
Structure of N-formimino-L-glutamate iminohydrolase from Pseudomonas aeruginosa
.
Biochemistry
.
2015 Jan
;
54
(
3
):
890
7
.
114.
Ferrandez
A
,
Miñambres
B
,
García
B
,
Olivera
ER
,
Luengo
JM
,
García
JL
, et al
.
Catabolism of phenylacetic acid in Escherichia coli. Characterization of a new aerobic hybrid pathway
.
J Biol Chem
.
1998 Oct
;
273
(
40
):
25974
86
.
115.
Ferrández
A
,
García
JL
,
Díaz
E
.
Transcriptional regulation of the divergent paa catabolic operons for phenylacetic acid degradation in Escherichia coli
.
J Biol Chem
.
2000 Apr
;
275
(
16
):
12214
22
.
116.
Fischer-Romero
C
,
Tindall
BJ
,
Jüttner
F
.
Tolumonas auensis gen. nov., sp. nov., a toluene-producing bacterium from anoxic sediments of a freshwater lake
.
Int J Syst Bacteriol
.
1996 Jan
;
46
(
1
):
183
8
.
117.
Flock
U
,
Reimann
J
,
Adelroth
P
.
Proton transfer in bacterial nitric oxide reductase
.
Biochem Soc Trans
.
2006 Feb
;
34
(
1
):
188
90
.
118.
Förster-Fromme
K
,
Jendrossek
D
.
Biochemical characterization of isovaleryl-CoA dehydrogenase (LiuA) of Pseudomonas aeruginosa and the importance of liu genes for a functional catabolic pathway of methyl-branched compounds
.
FEMS Microbiol Lett
.
2008 Sep
;
286
(
1
):
78
84
.
119.
Förster-Fromme
K
,
Höschle
B
,
Mack
C
,
Bott
M
,
Armbruster
W
,
Jendrossek
D
.
Identification of genes and proteins necessary for catabolism of acyclic terpenes and leucine/isovalerate in Pseudomonas aeruginosa
.
Appl Environ Microbiol
.
2006 Jul
;
72
(
7
):
4819
28
.
120.
Forte
E
,
Borisov
VB
,
Vicente
JB
,
Giuffrè
A
.
Cytochrome bd and gaseous ligands in bacterial physiology
.
Adv Microb Physiol
.
2017
;
71
:
171
234
.
121.
Frain
KM
,
van Dijl
JM
,
Robinson
C
.
The twin-arginine pathway for protein secretion
.
EcoSalPlus
.
2019. https://doi.org/10.1128/ecosalplus.ESP-0040-2018.
122.
Frey
PA
.
Radical mechanisms of enzymatic catalysis
.
Annu Rev Biochem
.
2001
;
70
:
121
48
.
123.
Friedrich
T
.
The NADH:ubiquinone oxidoreductase (complex I) from Escherichia coli
.
Biochim Biophys Acta
.
1998 May
;
1364
(
2
):
134
46
.
124.
Friedrich
T
,
Pohl
T
.
NADH as donor
.
EcoSalPlus
.
2007
. https://doi.org/10.1128/ecosalplus.3.2.4.
125.
Fuchs
G
,
Boll
M
,
Heider
J
.
Microbial degradation of aromatic compounds – from one strategy to four
.
Nat Rev Microbiol
.
2011 Nov
;
9
(
11
):
803
16
.
126.
Fujieda
N
,
Mori
M
,
Kano
K
,
Ikeda
T
.
Redox properties of quinohemoprotein amine dehydrogenase from Paracoccus denitrificans
.
Biochim Biophys Acta
.
2003 Apr
;
1647
(
1-2
):
289
96
.
127.
Fukuda
M
,
Takeda
H
,
Kato
HE
,
Doki
S
,
Ito
K
,
Maturana
AD
, et al
.
Structural basis for dynamic mechanism of nitrate/nitrite antiport by NarK
.
Nat Commun
.
2015 May
;
6
:
7097
.
128.
Funk
MA
,
Judd
ET
,
Marsh
ENG
,
Elliott
SJ
,
Drennan
CL
.
Structures of benzylsuccinate synthase elucidate roles of accessory subunits in glycyl radical enzyme activation and activity
.
Proc Natl Acad Sci USA
.
2014 Jul
;
111
(
28
):
10161
6
.
129.
Funk
MA
,
Marsh
ENG
,
Drennan
CL
.
Substrate-bound structures of benzylsuccinate synthase reveal how toluene is activated in anaerobic hydrocarbon degradation
.
J Biol Chem
.
2015 Sep
;
290
(
37
):
22398
408
.
130.
Gao
X
,
Tan
CL
,
Yeo
CC
,
Poh
CL
.
Molecular and biochemical characterization of the xlnD-encoded 3-hydroxybenzoate 6-hydroxylase involved in the degradation of 2,5-xylenol via the gentisate pathway in Pseudomonas alcaligenes NCIMB 9867
.
J Bacteriol
.
2005 Nov
;
187
(
22
):
7696
702
.
131.
Garcia Costas
AM
,
Poudel
S
,
Miller
A-F
,
Schut
GJ
,
Ledbetter
RN
,
Fixen
KR
, et al
.
Defining electron bifurcation in the electron-transferring flavoprotein family
.
J Bacteriol
.
2017 Nov
;
199
(
21
):
e00440
17
.
132.
Gescher
J
,
Zaar
A
,
Mohamed
M
,
Schägger
H
,
Fuchs
G
.
Genes coding for a new pathway of aerobic benzoate metabolism in Azoarcus evansii
.
J Bacteriol
.
2002 Nov
;
184
(
22
):
6301
15
.
133.
Gescher
J
,
Eisenreich
W
,
Wörth
J
,
Bacher
A
,
Fuchs
G
.
Aerobic benzoyl-CoA catabolic pathway in Azoarcus evansii: studies on the non-oxygenolytic ring cleavage enzyme
.
Mol Microbiol
.
2005 Jun
;
56
(
6
):
1586
600
.
134.
Gescher
J
,
Ismail
W
,
Ölgeschläger
E
,
Eisenreich
W
,
Wörth
J
,
Fuchs
G
.
Aerobic benzoyl-coenzyme A (CoA) catabolic pathway in Azoarcus evansii: conversion of ring cleavage product by 3,4-dehydroadipyl-CoA semialdehyde dehydrogenase
.
J Bacteriol
.
2006 Apr
;
188
(
8
):
2919
27
.
135.
Ghisla
S
,
Thorpe
C
.
Acyl-CoA dehydrogenases. A mechanistic overview
.
Eur J Biochem
.
2004 Feb
;
271
(
3
):
494
508
.
136.
Gibson
J
,
Dispensa
M
,
Fogg
GC
,
Evans
DT
,
Harwood
CS
.
4-Hydroxybenzoate-coenzyme A ligase from Rhodopseudomonas palustris: purification, gene sequence, and role in anaerobic degradation
.
J Bacteriol
.
1994 Feb
;
176
(
3
):
634
41
.
137.
Gibson
J
,
Dispensa
M
,
Harwood
CS
.
4-Hydroxybenzoyl coenzyme A reductase (dehydroxylating) is required for anaerobic degradation of 4-hydroxybenzoate by Rhodopseudomonas palustris and shares features with molybdenum-containing hydroxylases
.
J Bacteriol
.
1997 Feb
;
179
(
3
):
634
42
.
138.
Goddard
AD
,
Bali
S
,
Mavridou
DAI
,
Luque-Almagro
VM
,
Gates
AJ
,
Dolores Roldán
M
, et al
.
The Paracoccus denitrificans NarK-like nitrate and nitrite transporters: probing nitrate uptake and nitrate/nitrite exchange mechanisms
.
Mol Microbiol
.
2017 Jan
;
103
(
1
):
117
33
.
139.
Goto
M
,
Miyahara
I
,
Hayashi
H
,
Kagamiyama
H
,
Hirotsu
K
.
Crystal structures of branched-chain amino acid aminotransferase complexed with glutamate and glutarate: true reaction intermediate and double substrate recognition of the enzyme
.
Biochemistry
.
2003 Apr
;
42
(
13
):
3725
33
.
140.
Govantes
F
,
Albrecht
JA
,
Gunsalus
RP
.
Oxygen regulation of the Escherichia coli cytochrome d oxidase (cydAB) operon: roles of multiple promoters and the Fnr-1 and Fnr-2 binding sites
.
Mol Microbiol
.
2000 Sep
;
37
(
6
):
1456
69
.
141.
Green
J
,
Guest
JR
.
Regulation of transcription at the ndh promoter of Escherichia coli by FNR and novel factors
.
Mol Microbiol
.
1994 May
;
12
(
3
):
433
44
.
142.
Grund
TN
,
Radloff
M
,
Wu
D
,
Goojani
HG
,
Witte
LF
,
Jösting
W
, et al
.
Mechanistic and structural diversity between cytochrome bd isoforms of Escherichia coli
.
Proc Natl Acad Sci USA
.
2021 Dec
;
118
(
50
):
e2114013118
.
143.
Guccione
EJ
,
Kendall
JJ
,
Hitchcock
A
,
Garg
N
,
White
MA
,
Mulholland
F
, et al
.
Transcriptome and proteome dynamics in chemostat culture reveal how Campylobacter jejuni modulates metabolism, stress responses and virulence factors upon changes in oxygen availability
.
Environ Microbiol
.
2017 Oct
;
19
(
10
):
4326
48
.
144.
Guo
H
,
Rubinstein
JL
.
Structure of ATP synthase under strain during catalysis
.
Nat Commun
.
2022 Apr
;
13
(
1
):
2232
.
145.
Gupta
A
,
Pande
A
,
Sabrin
A
,
Thapa
SS
,
Gioe
BW
,
Grove
A
.
MarR family transcription factors from Burkholderia species: hidden clues to control of virulence-associated genes
.
Microbiol Mol Biol Rev
.
2019 Mar
;
83
(
1
):
e00039
18
.
146.
Hagel
C
,
Blaum
B
,
Friedrich
T
,
Heider
J
.
Characterisation of the redox centers of ethylbenzene dehydrogenase
.
J Biol Inorg Chem
.
2022 Feb
;
27
(
1
):
143
54
.
147.
Halarnkar
PP
,
Blomquist
GJ
.
Comparative aspects of propionate metabolism
.
Comp Biochem Physiol
.
1989
;
92
(
2
):
227
31
.
148.
Hall
MD
,
Banaszak
LJ
.
Crystal structure of a ternary complex of Escherichia coli malate dehydrogenase citrate and NAD at 1.9 Å resolution
.
J Mol Biol
.
1993 Jul
;
232
(
1
):
213
22
.
149.
Haltia
T
,
Brown
K
,
Tegoni
M
,
Cambillau
C
,
Saraste
M
,
Mattila
K
, et al
.
Crystal structure of nitrous oxide reductase from Paracoccus denitrificans at 1.6 Å resolution
.
Biochem J
.
2003 Jan
;
369
(
1
):
77
88
.
150.
Hamblin
MJ
,
Shaw
JG
,
Kelly
DJ
.
Sequence analysis and interposon mutagenesis of a sensor-kinase (DctS) and response-regulator (DctR) controlling synthesis of the high-affinity C4-dicarboxylate transport system in Rhodobacter capsulatus
.
Mol Gen Genet
.
1993 Feb
;
237
(
1‒2
):
215
24
.
151.
Harpel
MR
,
Lipscomb
JD
.
Gentisate 1,2-dioxygenase from Pseudomonas. Purification, characterization, and comparison of the enzymes from Pseudomonas testosteroni and Pseudomonas acidivorans
.
J Biol Chem
.
1990 Apr
;
265
(
11
):
6301
11
.
152.
Härtig
E
,
Schiek
U
,
Vollack
K-U
,
Zumft
WG
.
Nitrate and nitrite control of respiratory nitrate reduction in denitrifying Pseudomonas stutzeri by a two-component regulatory system homologous to NarXL of Escherichia coli
.
J Bacteriol
.
1999 Jun
;
181
(
12
):
3658
65
.
153.
Harwood
CS
,
Nichols
NN
,
Kim
M-K
,
Ditty
JL
,
Parales
RE
.
Identification of the pcaRKF gene cluster from Pseudomonas putida: involvement in chemotaxis, biodegradation, and transport of 4-hydroxybenzoate
.
J Bacteriol
.
1994 Nov
;
176
(
21
):
6479
88
.
154.
Hearn
EM
,
Patel
DR
,
Lepore
BW
,
Indic
M
,
van den Berg
B
.
Transmembrane passage of hydrophobic compounds through a protein channel wall
.
Nature
.
2009 Mar
;
458
(
7236
):
367
70
.
155.
Heider
J
,
Szaleniec
M
,
Martins
BM
,
Seyhan
D
,
Buckel
W
,
Golding
BT
.
Structure and function of benzylsuccinate synthase and related fumarate-adding glycyl radical enzymes
.
J Mol Microbiol Biotechnol
.
2016a Mar
;
26
(
1‒3
):
29
44
.
156.
Heider
J
,
Szaleniec
M
,
Sünwoldt
K
,
Boll
M
.
Ethylbenzene dehydrogenase and related molybdenum enzymes involved in oxygen-independent alkyl chain hydroxylation
.
J Mol Microbiol Biotechnol
.
2016b Mar
;
26
(
1‒3
):
45
62
.
157.
Heikal
A
,
Nakatani
Y
,
Dunn
E
,
Weimar
MR
,
Day
CL
,
Baker
EN
, et al
.
Structure of the bacterial type II NADH dehydrogenase: a monotopic membrane protein with an essential role in energy conservation
.
Mol Microbiol
.
2014 Mar
;
91
(
5
):
950
64
.
158.
Helmann
JD
,
Chamberlin
MJ
.
Structure and function of bacterial sigma factors
.
Annu Rev Biochem
.
1988 Jul
;
57
:
839
72
.
159.
Hensel
M
,
Hinsley
AP
,
Nikolaus
T
,
Sawers
G
,
Berks
BC
.
The genetic basis of tetrathionate respiration in Salmonella typhimurium
.
Mol Microbiol
.
1999 Apr
;
32
(
2
):
275
87
.
160.
Hermuth
K
,
Leuthner
B
,
Heider
J
.
Operon structure and expression of the genes for benzylsuccinate synthase in Thauera aromatica strain K172
.
Arch Microbiol
.
2002 Feb
;
177
(
2
):
132
8
.
161.
Hillberg
M
,
Pierik
AJ
,
Bill
E
,
Friedrich
T
,
Lippert
M-L
,
Heider
J
.
Identification of FeS clusters in the glycyl-radical enzyme benzylsuccinate synthase via EPR and Mössbauer spectroscopy
.
J Biol Inorg Chem
.
2012 Jan
;
17
(
1
):
49
56
.
162.
Hino
T
,
Matsumoto
Y
,
Nagano
S
,
Sugimoto
H
,
Fukumori
Y
,
Murata
T
, et al
.
Structural basis of biological N2O generation by bacterial nitric oxide reductase
.
Science
.
2010 Dec
;
330
(
6011
):
1666
70
.
163.
Hino
T
,
Nagano
S
,
Sugimoto
H
,
Tosha
T
,
Shiro
Y
.
Molecular structure and function of bacterial nitric oxide reductase
.
Biochim Biophys Acta
.
2012 Apr
;
1817
(
4
):
680
7
.
164.
Hinojosa-Leon
M
,
Dubourdieu
M
,
Sanchez-Crispin
JA
,
Chippaux
M
.
Tetrathionate reductase of Salmonella typhimurium: a molybdenum containing enzyme
.
Biochem Biophys Res Commun
.
1986 Apr
;
136
(
2
):
577
81
.
165.
Hinsley
AP
,
Berks
BC
.
Specificity of respiratory pathways involved in the reduction of sulfur compounds by Salmonella enterica
.
Microbiology
.
2002 Nov
;
148
(
11
):
3631
8
.
166.
Hinsley
AP
,
Stanley
NR
,
Palmer
T
,
Berks
BC
.
A naturally occurring bacterial Tat signal peptide lacking one of the “invariant” arginine residues of the consensus targeting motif
.
FEBS Lett
.
2001 May
;
497
(
1
):
45
9
.
167.
Hirakawa
H
,
Hirakawa
Y
,
Greenberg
EP
,
Harwood
CS
.
BadR and BadM proteins transcriptionally regulate two operons needed for anaerobic benzoate degradation by Rhodopseudomonas palustris
.
Appl Environ Microbiol
.
2015 Jul
;
81
(
13
):
4253
62
.
168.
Hirsch
W
,
Schägger
H
,
Fuchs
G
.
Phenylglyoxylate:NAD+ oxidoreductase (CoA benzoylating), a new enzyme of anaerobic phenylalanine metabolism in the denitrifying bacterium Azoarcus evansii
.
Eur J Biochem
.
1998 Feb
;
251
(
3
):
907
15
.
169.
Höffken
HW
,
Duong
M
,
Friedrich
T
,
Breuer
M
,
Hauer
B
,
Reinhardt
R
, et al
.
Crystal structure and enzyme kinetics of the (S)-specific 1-phenylethanol dehydrogenase of the denitrifying bacterium strain EbN1
.
Biochemistry
.
2006 Jan
;
45
(
1
):
82
93
.
170.
Hong
H
,
Seo
H
,
Kim
K-J
.
Structure and biochemical studies of a pseudomonad maleylpyruvate isomerase from Pseudomonas aeruginosa PAO1
.
Biochem Biophys Res Commun
.
2019 Jun
;
514
(
3
):
991
7
.
171.
Hong
H
,
Seo
H
,
Park
W
,
Kim
K-J
.
Sequence, structure and function-based classification of the broadly conserved FAH superfamily reveals two distinct fumarylpyruvate hydrolase subfamilies
.
Environ Microbiol
.
2020 Jan
;
22
(
1
):
270
85
.
172.
Hopper
DJ
.
The hydroxylation of p-cresol and its conversion to p-hydroxybenzaldehyde in Pseudomonas putida
.
Biochem Biophys Res Commun
.
1976 Mar
;
69
(
2
):
462
8
.
173.
Hopper
DJ
,
Taylor
DG
.
The purification and properties of p-cresol-(acceptor) oxidoreductase (hydroxylating), a flavocytochrome from Pseudomonas putida
.
Biochem J
.
1977 Oct
;
167
(
1
):
155
62
.
174.
Hopper
DJ
,
Jones
MR
,
Causer
MJ
.
Periplasmic location of p-cresol methylhydroxylase in Pseudomonas putida
.
FEBS
.
1985 Mar
;
182
(
2
):
485
8
.
175.
Horswill
AR
,
Escalante-Semerena
JC
.
The prpE gene of Salmonella typhimurium LT2 encodes propionyl-CoA synthetase
.
Microbiology
.
1999 Jun
;
145
(
6
):
1381
8
.
176.
Höschle
B
,
Gnau
V
,
Jendrossek
D
.
Methylcrotonyl-CoA and geranyl-CoA carboxylases are involved in leucine/isovalerate utilization (Liu) and acyclic terpene utilization (Atu), and are encoded by liuB/liuD and atuC/atuF, in Pseudomonas aeruginosa
.
Microbiology
.
2005 Nov
;
151
(
Pt 11
):
3649
56
.
177.
Hoshino
T
,
Kose
K
.
Cloning, nucleotide sequence, and identification of products of the Pseudomonas aeruginosa PAO bra genes, which encode the high-affinity branched-chain amino acid transport system
.
J Bacteriol
.
1990 Oct
;
172
(
10
):
5531
9
.
178.
Houten
SM
,
Violante
S
,
Ventura
FV
,
Wanders
RJA
.
The biochemistry and physiology of mitochondrial fatty acid β-oxidation and its genetic disorders
.
Annu Rev Physiol
.
2016
;
78
:
23
44
.
179.
Hreha
TN
,
Mezic
KG
,
Herce
HD
,
Duffy
EB
,
Bourges
A
,
Pryshchep
S
, et al
.
Complete topology of the RNF complex from Vibrio cholerae
.
Biochemistry
.
2015 Apr
;
54
(
15
):
2443
55
.
180.
Hu
L
,
Allison
SL
,
Phillips
AT
.
Identification of multiple repressor recognition sites in the hut system of Pseudomonas putida
.
J Bacteriol
.
1989 Aug
;
171
(
8
):
4189
95
.
181.
Huang
CS
,
Sadre-Bazzaz
K
,
Shen
Y
,
Deng
B
,
Zhou
ZH
,
Tong
L
.
Crystal structure of the α6β6 holoenzyme of propionyl-coenzyme A carboxylase
.
Nature
.
2010 Aug
;
466
(
7309
):
1001
5
.
182.
Hunte
C
,
Palsdottir
H
,
Trumpower
BL
.
Protonmotive pathways and mechanisms in the cytochrome bc1 complex
.
FEBS Lett
.
2003 Jun
;
545
(
1
):
39
46
.
183.
Hutson
S
.
Structure and function of branched chain aminotransferases
.
Prog Nucleic Acid Res Mol Biol
.
2001
;
70
:
175
206
.
184.
Hwang
WC
,
Bakolitsa
C
,
Punta
M
,
Coggill
PC
,
Bateman
A
,
Axelrod
HL
, et al
.
LUD, a new protein domain associated with lactate utilization
.
BMC Bioinform
.
2013 Nov
;
14
:
341
.
185.
Hwang
S
,
Choe
D
,
Yoo
M
,
Cho
S
,
Kim
SC
,
Cho
S
, et al
.
Peptide transporter CstA imports pyruvate in Escherichia coli K-12
.
J Bacteriol
.
2018 Apr
;
200
(
7
):
e00771
17
.
186.
Ingram-Smith
C
,
Martin
SR
,
Smith
KS
.
Acetate kinase: not just a bacterial enzyme
.
Trends Microbiol
.
2006 Jun
;
14
(
6
):
249
53
.
187.
Ismail
W
,
El-Said Mohamed
M
,
Wanner
BL
,
Datsenko
KA
,
Eisenreich
W
,
Rohdich
F
, et al
.
Functional genomics by NMR spectroscopy. Phenylacetate catabolism in Escherichia coli
.
Eur J Biochem
.
2003 Jul
;
270
(
14
):
3047
54
.
188.
Iwaki
M
,
Yagi
T
,
Horiike
K
,
Saeki
Y
,
Ushijima
T
,
Nozaki
M
.
Crystallization and properties of aromatic amine dehydrogenase from Pseudomonas sp
.
Arch Biochem Biophys
.
1983 Jan
;
220
(
1
):
253
62
.
189.
Iwata
S
,
Ostermeier
C
,
Ludwig
B
,
Michel
H
.
Structure at 2.8 Å resolution of cytochrome c oxidase from Paracoccus denitrificans
.
Nature
.
1995 Aug
;
376
(
6542
):
660
9
.
190.
James
MJ
,
Coulthurst
SJ
,
Palmer
T
,
Sargent
F
.
Signal peptide etiquette during assembly of a complex respiratory enzyme
.
Mol Microbiol
.
2013 Oct
;
90
(
2
):
400
14
.
191.
Janausch
IG
,
Zientz
E
,
Tran
QH
,
Kröger
A
,
Unden
G
.
C4-dicarboxylate carriers and sensors in bacteria
.
Biochim Biophys Acta
.
2002 Jan
;
1553
(
1‒2
):
39
56
.
192.
Jarling
R
,
Sadeghi
M
,
Drozdowska
M
,
Lahme
S
,
Buckel
W
,
Rabus
R
, et al
.
Stereochemical investigations reveal the mechanism of the bacterial activation of n-alkanes without oxygen
.
Angew Chem Int Ed Engl
.
2012 Feb
;
51
(
6
):
1334
8
.
193.
Jenner
LP
,
Kurth
JM
,
van Helmont
S
,
Sokol
KP
,
Reisner
E
,
Dahl
C
, et al
.
Heme ligation and redox chemistry in two bacterial thiosulfate dehydrogenase (TsdA) enzymes
.
J Biol Chem
.
2019 Nov
;
294
(
47
):
18002
14
.
194.
Jenner
LP
,
Crack
JC
,
Kurth
JM
,
Soldánová
Z
,
Brandt
L
,
Sokol
KP
, et al
.
Reaction of thiosulfate dehydrogenase with a substrate mimic induces dissociation of the cysteine heme ligand giving insights into the mechanism of oxidative catalysis
.
J Am Chem Soc
.
2022 Oct
;
144
(
40
):
18296
304
.
195.
Jiang
T
,
Gao
C
,
Ma
C
,
Xu
P
.
Microbial lactate utilization: enzymes, pathogenesis, and regulation
.
Trends Microbiol
.
2014 Oct
;
22
(
10
):
589
99
.
196.
Jobst
B
,
Schühle
K
,
Linne
U
,
Heider
J
.
ATP-dependent carboxylation of acetophenone by a novel type of carboxylase
.
J Bacteriol
.
2010 Mar
;
192
(
5
):
1387
94
.
197.
Jõesaar
M
,
Heinaru
E
,
Viggor
S
,
Vedler
E
,
Heinaru
A
.
Diversity of the transcriptional regulation of the pch gene cluster in two indigenous p-cresol-degradative strains of Pseudomonas fluorescens
.
FEMS Microbiol Ecol
.
2010 Jun
;
72
(
3
):
464
75
.
198.
Johannes
J
,
Unciuleac
M-C
,
Friedrich
T
,
Warkentin
E
,
Ermler
U
,
Boll
M
.
Inhibitors of the molybdenum cofactor containing 4-hydroxybenzoyl-CoA reductase
.
Biochemistry
.
2008 Apr
;
47
(
17
):
4964
72
.
199.
Johnson
HA
,
Pelletier
DA
,
Spormann
AM
.
Isolation and characterization of anaerobic ethylbenzene dehydrogenase, a novel Mo-Fe-S enzyme
.
J Bacteriol
.
2001 Aug
;
183
(
15
):
4536
42
.
200.
Jones
DCN
,
Cooper
RA
.
Catabolism of 3-hydroxybenzoate by the gentisate pathway in Klebsiella pneumoniae M5a1
.
Arch Microbiol
.
1990 Oct
;
154
(
5
):
489
95
.
201.
Jormakka
M
,
Byrne
B
,
Iwata
S
.
Formate dehydrogenase – a versatile enzyme in changing environments
.
Curr Opin Struc Biol
.
2003 Aug
;
13
(
4
):
418
23
.
202.
Jormakka
M
,
Richardson
D
,
Byrne
B
,
Iwata
S
.
Architecture of NarGH reveals a structural classification of Mo-bisMGD enzymes
.
Structure
.
2004 Jan
;
12
(
1
):
95
104
.
203.
Jörnvall
H
,
Hedlund
J
,
Bergman
T
,
Oppermann
U
,
Persson
B
.
Superfamilies SDR and MDR: from early ancestry to present forms. Emergence of three lines, a Zn-metalloenzyme, and distinct variabilities
.
Biochem Biophys Res Commun
.
2010 May
;
396
(
1
):
125
30
.
204.
Jörnvall
H
,
Landreh
M
,
Östberg
LJ
.
Alcohol dehydrogenase, SDR and MDR structural stages, present update and altered era
.
Chem Biol Interact
.
2015 Jun
;
234
:
75
9
.
205.
Jüngst
A
,
Wakabayashi
S
,
Matsubara
H
,
Zumft
WG
.
The nirSTBM region coding for cytochrome cd1-dependent nitrite respiration of Pseudomonas stutzeri consists of a cluster of mono-, di-, and tetraheme proteins
.
FEBS
.
1991 Feb
;
279
(
2
):
205
9
.
206.
Kahle
M
,
ter Beek
J
,
Hosler
JP
,
Ädelroth
P
.
The insertion of the non-heme FeB cofactor into nitric oxide reductase from P. denitrificans depends on NorQ and NorD accessory proteins
.
Biochim Biophys Acta Bioenerg
.
2018 Oct
;
1859
(
10
):
1051
8
.
207.
Kalimuthu
P
,
Heider
J
,
Knack
D
,
Bernhardt
PV
.
Electrocatalytic hydrocarbon hydroxylation by ethylbenzene dehydrogenase from Aromatoleum aromaticum
.
J Phys Chem B
.
2015 Feb
;
119
(
8
):
3456
63
.
208.
Kalimuthu
P
,
Hege
D
,
Winiarska
A
,
Gemmecker
Y
,
Szaleniec
M
,
Heider
J
, et al
.
Electrocatalytic aldehyde oxidation by a tungsten dependent aldehyde oxidoreductase from Aromatoleum aromaticum
.
Chem Eur J
.
2023 Apr
;
29
(
20
):
e202203072
.
209.
Kallberg
Y
,
Oppermann
U
,
Jörnvall
H
,
Persson
B
.
Short-chain dehydrogenase/reductases (SDRs). Coenzyme-based functional assignments in completed genomes
.
Eur J Biochem
.
2002 Sep
;
269
(
18
):
4409
17
.
210.
Karpusas
M
,
Branchaud
B
,
Remington
SJ
.
Proposed mechanism for the condensation reaction of citrate synthase: 1.9-Å structure of the ternary complex with oxaloacetate and carboxymethyl coenzyme A
.
Biochemistry
.
1990 Mar
;
29
(
9
):
2213
9
.
211.
Kaschabek
SR
,
Kuhn
B
,
Müller
D
,
Schmidt
E
,
Reineke
W
.
Degradation of aromatics and chloroaromatics by Pseudomonas sp. strain B13: purification and characterization of 3-oxoadipate:succinyl-Coenzyme A (CoA) transferase and 3-oxoadipyl-CoA thiolase
.
J Bacteriol
.
2002 Jan
;
184
(
1
):
207
15
.
212.
Kazakov
AE
,
Rodionov
DA
,
Alm
E
,
Arkin
AP
,
Dubchak
I
,
Gelfand
MS
.
Comparative genomics of regulation of fatty acid and branched-chain amino acid utilization in Proteobacteria
.
J Bacteriol
.
2009 Jan
;
191
(
1
):
52
64
.
213.
Kessler
D
,
Rétey
J
,
Schulz
GE
.
Structure and action of urocanase
.
J Mol Biol
.
2004 Sep
;
342
(
1
):
183
94
.
214.
Kim
YS
.
Malonate metabolism: biochemistry, molecular biology, physiology, and industrial application
.
J Biochem Mol Biol
.
2002 Sep
;
35
(
5
):
443
51
.
215.
Kim
J
,
Fuller
JH
,
Cecchini
G
,
McIntire
WS
.
Cloning, sequencing, and expression of the structural genes for the cytochrome and flavoprotein subunits of p-cresol methylhydroxylase from two strains of Pseudomonas putida
.
J Bacteriol
.
1994 Oct
;
176
(
20
):
6349
61
.
216.
Klepp
J
,
Fallert-Müller
A
,
Grimm
K
,
Hull
WE
,
Rétey
J
.
Mechanism of action of urocanase. Specific 13C-labelling of the prosthetic NAD+ and revision of the structure of its adduct with imidazolylpropionate
.
Eur J Biochem
.
1990 Sep
;
192
(
3
):
669
76
.
217.
Kloer
DP
,
Hagel
C
,
Heider
J
,
Schulz
GE
.
Crystal structure of ethylbenzene dehydrogenase from Aromatoleum aromaticum
.
Structure
.
2006 Sep
;
14
(
9
):
1377
88
.
218.
Klünemann
T
,
Blankenfeldt
W
.
Structure of heme d1-free cd1 nitrite reductase NirS
.
Acta Crystallogr F Struct Biol Commun
.
2020 Jun
;
76
(
6
):
250
6
.
219.
Knack
D
,
Hagel
C
,
Szaleniec
M
,
Dudzik
A
,
Salwinski
A
,
Heider
J
.
Substrate and inhibitor spectra of ethylbenzene dehydrogenase: perspectives on application potential and catalytic mechanism
.
Appl Environ Microbiol
.
2012 Sep
;
78
(
18
):
6475
82
.
220.
Knack
DH
,
Marshall
JL
,
Harlow
GP
,
Dudzik
A
,
Szaleniec
M
,
Liu
S-Y
, et al
.
BN/CC isosteric compounds as enzyme inhibitors: N- and B-ethyl-1,2-azaborine inhibit ethylbenzene hydroxylation as nonconvertible substrate analogues
.
Angew Chem Int Ed
.
2013 Feb
;
52
(
9
):
2599
601
.
221.
Kniemeyer
O
,
Heider
J
.
Ethylbenzene dehydrogenase, a novel hydrocarbon-oxidizing molybdenum/iron-sulfur/heme enzyme
.
J Biol Chem
.
2001a Jun
;
276
(
24
):
21381
6
.
222.
Kniemeyer
O
,
Heider
J
.
(S)-1-Phenylethanol dehydrogenase of Azoarcus sp. strain EbN1, an enzyme of anaerobic ethylbenzene catabolism
.
Arch Microbiol
.
2001b Jul
;
176
(
1‒2
):
129
35
.
223.
Kolata
P
,
Efremov
RG
.
Structure of Escherichia coli respiratory complex I reconstituted into lipid nanodiscs reveals an uncoupled conformation
.
Elife
.
2021 Jul
;
10
:
e68710
.
224.
Kolesnikow
T
,
Schröder
I
,
Gunsalus
RP
.
Regulation of narK gene expression in Escherichia coli in response to anaerobiosis, nitrate, iron, and molybdenum
.
J Bacteriol
.
1992 Nov
;
174
(
22
):
7104
11
.
225.
Körner
H
,
Sofia
HJ
,
Zumft
WG
.
Phylogeny of the bacterial superfamily of Crp-Fnr transcription regulators: exploiting the metabolic spectrum by controlling alternative gene programs
.
FEMS Microbiol Rev
.
2003 Dec
;
27
(
5
):
559
92
.
226.
Krieger
CJ
,
Roseboom
W
,
Albracht
SPJ
,
Spormann
AM
.
A stable organic free radical in anaerobic benzylsuccinate synthase of Azoarcus sp. strain T
.
J Biol Chem
.
2001 Apr
;
276
(
16
):
12924
7
.
227.
Kube
M
,
Heider
J
,
Amann
J
,
Hufnagel
P
,
Kühner
S
,
Beck
A
, et al
.
Genes involved in the anaerobic degradation of toluene in a denitrifying bacterium, strain EbN1
.
Arch Microbiol
.
2004 Mar
;
181
(
3
):
182
94
.
228.
Kühlbrandt
W
.
Structure and mechanisms of F-type ATP synthases
.
Annu Rev Biochem
.
2019 Mar
;
88
:
515
49
.
229.
Kühner
S
,
Wöhlbrand
L
,
Fritz
I
,
Wruck
W
,
Hultschig
C
,
Hufnagel
P
, et al
.
Substrate-dependent regulation of anaerobic degradation pathways for toluene and ethylbenzene in a denitrifying bacterium, strain EbN1
.
J Bacteriol
.
2005 Feb
;
187
(
4
):
1493
503
.
230.
Kumagai
H
,
Fujiwara
T
,
Matsubara
H
,
Saeki
K
.
Membrane localization, topology, and mutual stabilization of the rnfABC gene products in Rhodobacter capsulatus and implications for a new family of energy-coupling NADH oxidoreductases
.
Biochemistry
.
1997 May
;
36
(
18
):
5509
21
.
231.
Kunau
W-H
,
Dommes
V
,
Schulz
H
.
β-Oxidation of fatty acids in mitochondria, peroxisomes and bacteria: a century of continued progress
.
Prog Lipid Res
.
1995 Dec
;
34
(
4
):
267
342
.
232.
Küppers
J
,
Becker
P
,
Jarling
R
,
Dörries
M
,
Cakić
N
,
Schmidtmann
M
, et al
.
Stereochemical insights into the anaerobic degradation of 4-isopropylbenzoyl-CoA in the denitrifying bacterium strain pCyN1
.
Chem Eur J
.
2019 Mar
;
25
(
18
):
4722
31
.
233.
Kuroki
M
,
Igarashi
Y
,
Ishii
M
,
Arai
H
.
Fine-tuned regulation of the dissimilatory nitrite reductase gene by oxygen and nitric oxide in Pseudomonas aeruginosa
.
Environ Microbiol Rep
.
2014 Dec
;
6
(
6
):
792
801
.
234.
Kurth
JM
,
Brito
JA
,
Reuter
J
,
Flegler
A
,
Koch
T
,
Franke
T
, et al
.
Electron accepting units of the diheme cytochrome c TsdA, a bifunctional thiosulfate dehydrogenase/tetrathionate reductase
.
J Biol Chem
.
2016 Nov
;
291
(
48
):
24804
18
.
235.
Küver
J
,
Xu
Y
,
Gibson
J
.
Metabolism of cyclohexane carboxylic acid by the photosynthetic bacterium Rhodopseudomonas palustris
.
Arch Microbiol
.
1995 Nov
;
164
(
5
):
337
45
.
236.
Lack
A
,
Fuchs
G
.
Carboxylation of phenylphosphate by phenol carboxylase, an enzyme system of anaerobic phenol metabolism
.
J Bacteriol
.
1992 Jun
;
174
(
11
):
3629
36
.
237.
Laempe
D
,
Eisenreich
W
,
Bacher
A
,
Fuchs
G
.
Cyclohexa-1,5-diene-1-carboxyl-CoA hydratase, an enzyme involved in anaerobic metabolism of benzoyl-CoA in the denitrifying bacterium Thauera aromatica
.
Eur J Biochem
.
1998 Aug
;
255
(
3
):
618
27
.
238.
Laempe
D
,
Jahn
M
,
Fuchs
G
.
6-Hydroxycyclohex-1-ene-1-carbonyl-CoA dehydrogenase and 6-oxocyclohex-1-ene-1-carbonyl-CoA hydrolase, enzymes of the benzoyl-CoA pathway of anaerobic aromatic metabolism in the denitrifying bacterium Thauera aromatica
.
Eur J Biochem
.
1999 Jul
;
263
(
2
):
420
9
.
239.
Laempe
D
,
Jahn
M
,
Breese
K
,
Schägger
H
,
Fuchs
G
.
Anaerobic metabolism of 3-hydroxybenzoate by the denitrifying bacterium Thauera aromatica
.
J Bacteriol
.
2001 Feb
;
183
(
3
):
968
79
.
240.
Langkau
B
,
Ghisla
S
,
Buder
R
,
Ziegler
K
,
Fuchs
G
.
2-Aminobenzoyl-CoA monooxygenase/reductase, a novel type of flavoenzyme. Identification of the reaction products
.
Eur J Biochem
.
1990 Jul
;
191
(
2
):
365
71
.
241.
Langkau
B
,
Vock
P
,
Massey
V
,
Fuchs
G
,
Ghisla
S
.
2-Aminobenzoyl-CoA monooxygenase/reductase. Evidence for two distinct loci catalyzing substrate monooxygenation and hydrogenation
.
Eur J Biochem
.
1995 Jun
;
230
(
2
):
676
85
.
242.
Lee
HY
,
An
JH
,
Kim
YS
.
Identification and characterization of a novel transcriptional regulator, MatR, for malonate metabolism in Rhizobium leguminosarum bv. trifolii
.
Eur J Biochem
.
2000 Dec
;
267
(
24
):
7224
9
.
243.
Lenz
M
,
Rétey
J
.
Cloning, expression and mutational analysis of the urocanase gene (hutU) from Pseudomonas putida
.
Eur J Biochem
.
1993 Oct
;
217
(
1
):
429
34
.
244.
Leuthner
B
,
Heider
J
.
A two-component system involved in regulation of anaerobic toluene metabolism in Thauera aromatica
.
FEMS Microbiol Lett
.
1998 Sep
;
166
(
1
):
35
41
.
245.
Leuthner
B
,
Heider
J
.
Anaerobic toluene catabolism of Thauera aromatica: the bbs operon codes for enzymes of β oxidation of the intermediate benzylsuccinate
.
J Bacteriol
.
2000 Jan
;
182
(
2
):
272
7
.
246.
Leuthner
B
,
Leutwein
C
,
Schulz
H
,
Hörth
P
,
Haehnel
W
,
Schiltz
E
, et al
.
Biochemical and genetic characterization of benzylsuccinate synthase from Thauera aromatica: a new glycyl radical enzyme catalyzing the first step in anaerobic toluene metabolism
.
Mol Microbiol
.
1998 May
;
28
(
3
):
615
28
.
247.
Leutwein
C
,
Heider
J
.
Anaerobic toluene-catabolic pathway in denitrifying Thauera aromatica: activation and β-oxidation of the first intermediate, (R)-(+)-benzylsuccinate
.
Microbiology
.
1999 Nov
;
145
(
11
):
3265
71
.
248.
Leutwein
C
,
Heider
J
.
Succinyl-CoA:(R)-benzylsuccinate CoA-transferase: an enzyme of the anaerobic toluene catabolic pathway in denitrifying bacteria
.
J Bacteriol
.
2001 Jul
;
183
(
14
):
4288
95
.
249.
Leutwein
C
,
Heider
J
.
(R)-Benzylsuccinyl-CoA dehydrogenase of Thauera aromatica, an enzyme of the anaerobic toluene catabolic pathway
.
Arch Microbiol
.
2002 Dec
;
178
(
6
):
517
24
.
250.
Li
L
,
Marsh
ENG
.
Mechanism of benzylsuccinate synthase probed by substrate and isotope exchange
.
J Am Chem Soc
.
2006 Dec
;
128
(
50
):
16056
7
.
251.
Li
L
,
Patterson
DP
,
Fox
CC
,
Lin
B
,
Coschigano
PW
,
Marsh
ENG
.
Subunit structure of benzylsuccinate synthase
.
Biochemistry
.
2009 Feb
;
48
(
6
):
1284
92
.
252.
Liu
T-T
,
Xu
Y
,
Liu
H
,
Luo
S
,
Yin
Y-J
,
Liu
S-J
, et al
.
Functional characterization of a gene cluster involved in gentisate catabolism in Rhodococcus sp. strain NCIMB 12038
.
Appl Microbiol Biotechnol
.
2011 Apr
;
90
(
2
):
671
8
.
253.
Lloyd
SJ
,
Lauble
H
,
Prasad
GS
,
Stout
CD
.
The mechanism of aconitase: 1.8 Å resolution crystal structure of the S642A:citrate complex
.
Protein Sci
.
1999 Dec
;
8
(
12
):
2655
62
.
254.
López Barragán
MJ
,
Carmona
M
,
Zamarro
MT
,
Thiele
B
,
Boll
M
,
Fuchs
G
, et al
.
The bzd gene cluster, coding for anaerobic benzoate catabolism, in Azoarcus sp. strain CIB
.
J Bacteriol
.
2004 Sept
;
186
(
17
):
5762
74
.
255.
Ludwig
ML
,
Matthews
RG
.
Structure-based perspectives on B12-dependent enzymes
.
Annu Rev Biochem
.
1997 Jul
;
66
:
269
313
.
256.
Luo
S
,
Adam
D
,
Giaveri
S
,
Barthel
S
,
Cestellos-Blanco
S
,
Hege
D
, et al
.
ATP production from electricity with a new-to-nature electrobiological module
.
Joule
.
2023 Aug
;
7
(
8
):
1745
58
.
257.
Lyons
JA
,
Aragão
D
,
Slattery
O
,
Pisliakov
AV
,
Soulimane
T
,
Caffrey
M
.
Structural insights into electron transfer in caa3-type cytochrome oxidase
.
Nature
.
2012 Jul
;
487
(
7408
):
514
8
.
258.
Maddocks
SE
,
Oyston
PCF
.
Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins
.
Microbiology
.
2008 Dec
;
154
(
12
):
3609
23
.
259.
Madhusudhan
KT
,
Huang
N
,
Braswell
EH
,
Sokatch
JR
.
Binding of L-branched-chain amino acids causes a conformational change in BkdR
.
J Bacteriol
.
1997 Jan
;
179
(
1
):
276
9
.
260.
Maher
MJ
,
Herath
AS
,
Udagedara
SR
,
Dougan
DA
,
Truscott
KN
.
Crystal structure of bacterial succinate:quinone oxidoreductase flavoprotein SdhA in complex with its assembly factor SdhE
.
Proc Natl Acad Sci USA
.
2018 Mar
;
115
(
12
):
2982
7
.
261.
Mao
HZ
,
Weber
J
.
Identification of the βTP site in the x-ray structure of F1-ATPase as the high-affinity catalytic site
.
Proc Natl Acad Sci USA
.
2007 Nov
;
104
(
47
):
18478
83
.
262.
Marsh
M
,
Shoemark
DK
,
Jacob
A
,
Robinson
C
,
Cahill
B
,
Zhou
N-Y
, et al
.
Structure of bacterial glutathione-S-transferase maleylpyruvate isomerase and implications for mechanism of isomerization
.
J Mol Biol
.
2008 Dec
;
384
(
1
):
165
77
.
263.
Marshall
AG
,
Rodgers
RP
.
Petroleomics: chemistry of the underworld
.
Proc Natl Acad Sci USA
.
2008 Nov
;
105
(
47
):
18090
5
.
264.
Mathews
FS
,
Chen
Z-W
,
Bellamy
HD
,
McIntire
WS
.
Three-dimensional structure of p-cresol methylhydroxylase (flavocytochrome c) from Pseudomonas putida at 3.0-Å resolution
.
Biochemistry
.
1991 Jan
;
30
(
1
):
238
47
.
265.
McDowall
JS
,
Murphy
BJ
,
Haumann
M
,
Palmer
T
,
Armstrong
FA
,
Sargent
F
.
Bacterial formate hydrogenlyase complex
.
Proc Natl Acad Sci USA
.
2014 Sep
;
111
(
38
):
E3948
56
.
266.
McIntire
W
,
Hopper
DJ
,
Singer
TP
.
p-Cresol methylhydroxylase. Assay and general properties
.
Biochem J
.
1985 Jun
;
228
(
2
):
325
35
.
267.
McLendon
GL
,
Bagby
S
,
Charman
JA
,
Driscoll
PC
,
McIntire
WS
,
Mathews
FS
, et al
.
Subunit interactions change the heme active-site geometry in p-cresol methylhydroxylase
.
Proc Natl Acad Sci USA
.
1991 Nov
;
88
(
21
):
9463
7
.
268.
Meng
W
,
Green
J
,
Guest
JR
.
FNR-dependent repression of ndh gene expression requires two upstream FNR-binding sites
.
Microbiology
.
1997 May
;
143
(
5
):
1521
32
.
269.
Mergelsberg
M
,
Willistein
M
,
Meyer
H
,
Stärk
H-J
,
Bechtel
DF
,
Pierik
AJ
, et al
.
Phthaloyl-coenzyme A decarboxylase from Thauera chlorobenzoica: the prenylated flavin-K+- and Fe2+-dependent key enzyme of anaerobic phthalate degradation
.
Environ Microbiol
.
2017 Sep
;
19
(
9
):
3734
44
.
270.
Mergelsberg
M
,
Egle
V
,
Boll
M
.
Evolution of a xenobiotic degradation pathway: formation and capture of the labile phthaloyl-CoA intermediate during anaerobic phthalate degradation
.
Mol Microbiol
.
2018 Jun
;
108
(
6
):
614
26
.
271.
Möbitz
H
,
Boll
M
.
A Birch-like mechanism in enzymatic benzoyl-CoA reduction: a kinetic study of substrate analogues combined with an ab initio model
.
Biochemistry
.
2002 Feb
;
41
(
6
):
1752
8
.
272.
Moe
A
,
Di Trani
J
,
Rubinstein
JL
,
Brzezinski
P
.
Cryo-EM structure and kinetics reveal electron transfer by 2D diffusion of cytochrome c in the yeast III-IV respiratory supercomplex
.
Proc Natl Acad Sci USA
.
2021 Mar
;
118
(
11
):
e2021157118
.
273.
Mohamed
ME-S
,
Fuchs
G
.
Purification and characterization of phenylacetyl-coenzyme A ligase from a denitrifying Pseudomonas sp., an enzyme involved in the anaerobic degradation of phenylacetate
.
Arch Microbiol
.
1993 Jun
;
159
(
6
):
554
62
.
274.
Mohamed
ME-S
,
Ismail
W
,
Heider
J
,
Fuchs
G
.
Aerobic metabolism of phenylacetic acid in Azoarcus evansii
.
Arch Microbiol
.
2002Sep
;
178
(
3
):
180
92
.
275.
Muhr
E
,
Schühle
K
,
Clermont
L
,
Sünwoldt
K
,
Kleinsorge
D
,
Seyhan
D
, et al
.
Enzymes of anaerobic ethylbenzene and p-ethylphenol catabolism in “Aromatoleum aromaticum”: differentiation and differential induction
.
Arch Microbiol
.
2015 Nov
;
197
(
9
):
1051
62
.
276.
Müller
C
,
Zhang
L
,
Zipfel
S
,
Topitsch
A
,
Lutz
M
,
Eckert
J
, et al
.
Molecular interplay of an assembly machinery for nitrous oxide reductase
.
Nature
.
2022 Aug
;
608
(
7923
):
626
31
.
277.
Muro-Pastor
AM
,
Maloy
S
.
Proline dehydrogenase activity of the transcriptional repressor PutA is required for induction of the put operon by proline
.
J Biol Chem
.
1995 Apr
;
270
(
17
):
9819
27
.
278.
Murphy
GE
,
Jensen
GJ
.
Electron cryotomography of the E. coli pyruvate and 2-oxoglutarate dehydrogenase complexes
.
Structure
.
2005 Dec
;
13
(
12
):
1765
73
.
279.
Mus
F
,
Eilers
BJ
,
Alleman
AB
,
Kabasakal
BV
,
Wells
JN
,
Murray
JW
, et al
.
Structural basis for the mechanism of ATP-dependent acetone carboxylation
.
Sci Rep
.
2017 Aug
;
7
(
1
):
7234
.
280.
Musser
SM
,
Chan
SI
.
Evolution of the cytochrome c oxidase proton pump
.
J Mol Evol
.
1998 May
;
46
(
5
):
508
20
.
281.
Nakai
T
,
Ono
K
,
Kuroda
S
,
Tanizawa
K
,
Okajima
T
.
An unusual subtilisin-like serine protease is essential for biogenesis of quinohemoprotein amine dehydrogenase
.
J Biol Chem
.
2012 Feb
;
287
(
9
):
6530
8
.
282.
Nakai
T
,
Deguchi
T
,
Frébort
I
,
Tanizawa
K
,
Okajima
T
.
Identification of genes essential for the biogenesis of quinohemoprotein amine dehydrogenase
.
Biochemistry
.
2014 Feb
;
53
(
5
):
895
907
.
283.
Nakai
T
,
Ito
H
,
Kobayashi
K
,
Takahashi
Y
,
Hori
H
,
Tsubaki
M
, et al
.
The radical S-adenosyl-L-methionine enzyme QhpD catalyzes sequential formation of intra-protein sulfur-to-methylene carbon thioether bonds
.
J Biol Chem
.
2015 Apr
;
290
(
17
):
11144
66
.
284.
Naren
N
,
Zhang
X-X
.
Global regulatory roles of the histidine-responsive transcriptional repressor HutC in Pseudomonas fluorescence SBW25
.
J Bacteriol
.
2020 Jul
;
202
(
13
):
e00792
19
.
285.
Narmandakh
A
,
Gad’on
N
,
Drepper
F
,
Knapp
B
,
Haehnel
W
,
Fuchs
G
.
Phosphorylation of phenol by phenylphosphate synthase: role of histidine phosphate in catalysis
.
J Bacteriol
.
2006 Nov
;
188
(
22
):
7815
22
.
286.
Nocek
B
,
Boyd
J
,
Ensign
SA
,
Peters
JW
.
Crystallization and preliminary X-ray analysis of an acetone carboxylase from Xanthobacter autotrophicus Py2
.
Acta Crystallogr B Biol Crystallogr
.
2004 Feb
;
60
(
2
):
385
7
.
287.
Olivera
ER
,
Miñambres
B
,
García
B
,
Muñiz
C
,
Moreno
MA
,
Ferrández
A
, et al
.
Molecular characterization of the phenylacetic acid catabolic pathway in Pseudomonas putida U: the phenylacetyl-CoA catabolon
.
Proc Natl Acad Sci USA
.
1998 May
;
95
(
11
):
6419
24
.
288.
Olkhova
E
,
Raba
M
,
Bracher
S
,
Hilger
D
,
Jung
H
.
Homology model of the Na+/proline transporter PutP of Escherichia coli and its functional implications
.
J Mol Biol
.
2011 Feb
;
406
(
1
):
59
74
.
289.
Oosterkamp
MJ
,
Boeren
S
,
Atashgahi
S
,
Plugge
CM
,
Schaap
PJ
,
Stams
AJM
.
Proteomic analysis of nitrate-dependent acetone degradation by Alicycliphilus denitrificans strain BC
.
FEMS Microbiol Lett
.
2015 Jun
;
362
(
11
):
fnv080
.
290.
Osamura
T
,
Kawakami
T
,
Kido
R
,
Ishii
M
,
Arai
H
.
Specific expression and function of the A-type cytochrome c oxidase under starvation conditions in Pseudomonas aeruginosa
.
PloS One
.
2017 May
;
12
(
5
):
e0177957
.
291.
Ostermeier
C
,
Harrenga
A
,
Ermler
U
,
Michel
H
.
Structure at 2.7 Å resolution of the Paracoccus denitrificans two-subunit cytochrome c oxidase complexed with an antibody FV fragment
.
Proc Natl Acad Sci USA
.
1997 Sep
;
94
(
20
):
10547
53
.
292.
Ostrovsky de Spicer
P
,
Maloy
S
.
PutA protein, a membrane-associated flavin dehydrogenase, acts as a redox-dependent transcriptional regulator
.
Proc Natl Acad Sci USA
.
1993 May
;
90
(
9
):
4295
8
.
293.
Otten
MF
,
Stork
DM
,
Reijnders
WNM
,
Westerhoff
HV
,
van Spanning
RJM
.
Regulation of expression of terminal oxidases in Paracoccus denitrificans
.
Eur J Biochem
.
2001 Apr
;
268
(
8
):
2486
97
.
294.
Ouzounis
CA
,
Kyrpides
NC
.
On the evolution of arginases and related enzymes
.
J Mol Evol
.
1994 Jul
;
39
(
1
):
101
4
.
295.
Pao
SS
,
Paulsen
IT
,
Saier Jr
MH
.
Major facilitator superfamily
.
Microbiol Mol Biol Rev
.
1998 Mar
;
62
(
1
):
1
34
.
296.
Patel
MS
,
Nemeria
NS
,
Furey
W
,
Jordan
F
.
The pyruvate dehydrogenase complex: structure-based function and regulation
.
J Biol Chem
.
2014 Jun
;
289
(
24
):
16615
23
.
297.
Pavel
H
,
Forsman
M
,
Shingler
V
.
An aromatic effector specificity mutant of the transcriptional regulator DmpR overcomes the growth constraints of Pseudomonas sp. strain CF600 on para-substituted methylphenols
.
J Bacteriol
.
1994 Dec
;
176
(
24
):
7550
7
.
298.
Pearce
DA
,
Page
MD
,
Norris
HAC
,
Tomlinson
EJ
,
Ferguson
SJ
.
Identification of the contiguous Paracoccus denitrificans ccmF and ccmH genes: disruption of ccmF, encoding a putative transporter, results in formation of an unstable apocytochrome c and deficiency in siderophore production
.
Microbiology
.
1998 Feb
;
144
(
2
):
467
77
.
299.
Pelletier
DA
,
Harwood
CS
.
2-Hydroxycyclohexane coenzyme A dehydrogenase, an enzyme characteristic of anaerobic benzoate degradation pathway used by Rhodopseudomonas palustris
.
J Bacteriol
.
2000 Mar
;
182
(
10
):
2753
60
.
300.
Pernstich
C
,
Senior
L
,
MacInnes
KA
,
Forsaith
M
,
Curnow
P
.
Expression, purification and reconstitution of 4-hydoxybenzoate transporter PcaK from Acinetobacter sp. ADP1
.
Prot Expr Purif
.
2014 Sep
;
101
:
68‒75
.
301.
Philippot
L
,
Clays-Josserand
A
,
Lensi
R
,
Trinsoutreau
I
,
Normand
P
,
Potier
P
.
Purification of the dissimilative nitrate reductase of Pseudomonas fluorescence and the cloning and sequencing of its corresponding genes
.
Biochim Biophys Acta
.
1997 Feb
;
1350
(
3
):
272
6
.
302.
Pichersky
E
,
Raguso
RA
.
Why do plants produce so many terpenoid compounds
.
New Phytol
.
2018 Nov
;
220
(
3
):
692
702
.
303.
Pinske
C
,
Sawers
RG
.
Anaerobic formate and hydrogen metabolism
.
EcoSal Plus
.
2016
. https://doi.org/10.1128/ecosalplus.ESP-0011-2016.
304.
Pitcher
RS
,
Watmough
NJ
.
The bacterial cytochrome cbb3 oxidases
.
Biochim Biophys Acta
.
2004 Apr
;
1655
(
1‒3
):
388
99
.
305.
Pitcher
RS
,
Brittain
T
,
Watmough
NJ
.
Cytochrome cbb3 oxidase and bacterial microaerobic metabolism
.
Biochem Soc Trans
.
2002 Aug
;
30
(
4
):
653
8
.
306.
Pomowski
A
,
Zumft
WG
,
Kroneck
PMH
,
Einsle
O
.
N2O binding at a [4Cu:2S] copper‐sulphur cluster in nitrous oxide reductase
.
Nature
.
2011 Sep
;
477
(
7363
):
234
7
.
307.
Prasser
B
,
Schöner
L
,
Zhang
L
,
Einsle
O
.
The copper chaperone NosL forms a heterometal site for Cu delivery to nitrous oxide reductase
.
Angew Chem Int Ed
.
2021 Aug
;
60
(
34
):
18810
4
.
308.
Price
CE
,
Driessen
AJM
.
Biogenesis of membrane bound respiratory complexes in Escherichia coli
.
Biochim Biophys Acta
.
2010 Jun
;
1803
(
6
):
748
66
.
309.
Price-Carter
M
,
Tingey
J
,
Bobik
TA
,
Roth
JR
.
The alternative electron acceptor tetrathionate supports B12-dependent anaerobic growth of Salmonella enterica serovar typhimurium on ethanolamine or 1,2-propanediol
.
J Bacteriol
.
2001 Apr
;
183
(
8
):
2463
75
.
310.
Qiao
C
,
Marsh
ENG
.
Mechanisms of benzylsuccinate synthase: stereochemistry of toluene addition to fumarate and maleate
.
J Am Chem Soc
.
2005 Jun
;
127
(
24
):
8608
9
.
311.
Rabus
R
,
Heider
J
.
Initial reactions of anaerobic metabolism of alkylbenzenes in denitrifying and sulfate-reducing bacteria
.
Arch Microbiol
.
1998 Sep
;
170
(
5
):
377
84
.
312.
Rabus
R
,
Widdel
F
.
Anaerobic degradation of ethylbenzene and other aromatic hydrocarbons by new denitrifying bacteria
.
Arch Microbiol
.
1995 Feb
;
163
(
2
):
96
103
.
313.
Rabus
R
,
Jack
DL
,
Kelly
DJ
,
Saier Jr
MH
.
TRAP transporters: an ancient family of extracytoplasmic solute-receptor-dependent secondary active transporters
.
Microbiology
.
1999 Dec
;
145
(
12
):
3431
45
.
314.
Rabus
R
,
Kube
M
,
Beck
A
,
Widdel
F
,
Reinhardt
R
.
Genes involved in the anaerobic degradation of ethylbenzene in a denitrifying bacterium, strain EbN1
.
Arch Microbiol
.
2002 Dec
;
178
(
6
):
506
16
.
315.
Rabus
R
,
Kube
M
,
Heider
J
,
Beck
A
,
Heitmann
K
,
Widdel
F
, et al
.
The genome sequence of an anaerobic aromatic-degrading denitrifying bacterium, strain EbN1
.
Arch Microbiol
.
2005 Jan
;
183
(
1
):
27
36
.
316.
Rabus
R
,
Trautwein
K
,
Wöhlbrand
L
.
Towards habitat-oriented systems biology of “Aromatoleum aromaticum” EbN1. Chemical sensing, catabolic network modulation and growth control in anaerobic aromatic compound degradation
.
Appl Microbiol Biotechnol
.
2014 Apr
;
98
(
8
):
3371
88
.
317.
Rabus
R
,
Boll
M
,
Heider
J
,
Meckenstock
RU
,
Buckel
W
,
Einsle
O
, et al
.
Anaerobic microbial degradation of hydrocarbons: from enzymatic reactions to the environment
.
J Mol Microbiol Biotechnol
.
2016 Mar
;
26
(
1-3
):
5
28
.
318.
Rabus
R
,
Wöhlbrand
L
,
Thies
D
,
Meyer
M
,
Reinhold-Hurek
B
,
Kämpfer
P
.
Aromatoleum gen. nov., a novel genus accommodating the phylogenetic lineage including Azoarcus evansii and related species, and proposal of Aromatoleum aromaticum sp. nov., Aromatoleum petrolei sp. nov., Aromatoleum bremense sp. nov., Aromatoleum toluolicum sp. nov. and Aromatoleum diolicum sp. nov
.
Int J Syst Evol Microbiol
.
2019 Apr
;
69
(
4
):
982
97
.
319.
Rajeev
L
,
Garber
ME
,
Zane
GM
,
Price
MN
,
Dubchak
I
,
Wall
JD
, et al
.
A new family of transcriptional regulators of tungstoenzymes and molybdate/tungstate transport
.
Environ Microbiol
.
2019 Feb
;
21
(
2
):
784
99
.
320.
Rand
JM
,
Pisithkul
T
,
Clark
RL
,
Thiede
JM
,
Mehrer
CR
,
Agnew
DE
, et al
.
A metabolic pathway for catabolizing levulinic acid in bacteria
.
Nat Microbiol
.
2017 Dec
;
2
(
12
):
1624
34
.
321.
Rasmussen
T
,
Brittain
T
,
Berks
BC
,
Watmough
NJ
,
Thomson
AJ
.
Formation of a cytochrome c‒nitrous oxide reductase complex is obligatory for N2O reduction by Paracoccus denitrificans
.
Dalton Trans
.
2005 Nov
;
21
:
3501
6
.
322.
Rather
LJ
,
Knapp
B
,
Haehnel
W
,
Fuchs
G
.
Coenzyme A-dependent aerobic metabolism of benzoate via epoxide formation
.
J Biol Chem
.
2010 Jul
;
285
(
27
):
20615
24
.
323.
Rather
LJ
,
Bill
E
,
Ismail
W
,
Fuchs
G
.
The reducing component BoxA of benzoyl-coenzyme A epoxidase from Azoarcus evansii is a [4Fe−4S] protein
.
Biochim Biophys Acta
.
2011a Dec
;
1814
(
12
):
1609
15
.
324.
Rather
LJ
,
Weinert
T
,
Demmer
U
,
Bill
E
,
Ismail
W
,
Fuchs
G
, et al
.
Structure and mechanism of the diiron benzoyl-coenzyme A epoxidase BoxB
.
J Biol Chem
.
2011b Aug
;
286
(
33
):
29241
8
.
325.
Rathore
S
,
Berndtsson
J
,
Marin-Buera
L
,
Conrad
J
,
Carroni
M
,
Brzezinski
P
, et al
.
Cryo-EM structure of the yeast respiratory supercomplex
.
Nat Struct Mol Biol
.
2019 Jan
;
26
(
1
):
50
7
.
326.
Rauhamäki
V
,
Bloch
DA
,
Wikström
M
.
Mechanistic stoichiometry of proton translocation by cytochrome cbb3
.
Proc Natl Acad Sci USA
.
2012 May
;
109
(
19
):
7286
91
.
327.
Reeve
CD
,
Carver
MA
,
Hopper
DJ
.
The purification and characterization of 4-ethylphenol methylenehydroxylase, a flavocytochrome from Pseudomonas putida JD1
.
Biochem J
.
1989 Oct
;
263
(
2
):
431
7
.
328.
Reizer
L
.
Catabolism of amino acids and related compounds
.
EcoSalPlus
.
2013
. https://doi.org/10.1128/ecosalplus.3.4.7.
329.
Ren
Y
,
Aguirre
J
,
Ntamack
AG
,
Chu
C
,
Schulz
H
.
An alternative pathway of oleate β-oxidation in Escherichia coli involving the hydrolysis of a dead end intermediate by a thioesterase
.
J Biol Chem
.
2004 Mar
;
279
(
12
):
11042
50
.
330.
Rhee
S-K
,
Fuchs
G
.
Phenylacetyl-CoA:acceptor oxidoreductase, a membrane-bound molybdenum-iron-sulfur enzyme involved in anaerobic metabolism of phenylalanine in the denitrifying bacterium Thauera aromatica
.
Eur J Biochem
.
1999 Jun
;
262
(
2
):
507
15
.
331.
Rigali
S
,
Derouaux
A
,
Giannotta
F
,
Dusart
J
.
Subdivision of the helix-turn-helix GntR family of bacterial regulators in the FadR, HutC, MocR, and YtrA subfamilies
.
J Biol Chem
.
2002 Apr
;
277
(
15
):
12507
15
.
332.
Rinaldo
S
,
Giardinà
G
,
Castiglione
N
,
Stelitano
V
,
Cutruzzolá
F
.
The catalytic mechanism of Pseudomonas aeruginosa cd1 nitrite reductase
.
Biochem Soc Trans
.
2011 Jan
;
39
(
1
):
195
200
.
333.
Riveros-Rosas
H
,
Julián-Sánchez
A
,
Villalobos-Molina
R
,
Pardo
JP
,
Piña
E
.
Diversity, taxonomy and evolution of medium-chain dehydrogenase/reductase superfamily
.
Eur J Biochem
.
2003 Aug
;
270
(
16
):
3309
34
.
334.
Rivolta
C
,
Soldo
B
,
Lazarevic
V
,
Joris
B
,
Mauël
C
,
Karamata
D
.
A 35.7 kb DNA fragment from the Bacillus subtilis chromosome containing a putative 12.3 kb operon involved in hexuronate catabolism and a perfectly symmetrical hypothetical catabolite-responsive element
.
Microbiology
.
1998 Apr
;
144
(
4
):
877
84
.
335.
Robbins
AH
,
Stout
CD
.
The structure of aconitase
.
Proteins
.
1989 Jan
;
5
(
4
):
289
312
.
336.
Rome
K
,
Borde
C
,
Taher
R
,
Cayron
J
,
Lesterlin
C
,
Gueguen
E
, et al
.
The two-component system ZraPSR is a novel ESR that contributes to intrinsic antibiotic tolerance in Escherichia coli
.
J Mol Biol
.
2018 Dec
;
430
(
24
):
4971
85
.
337.
Rosa
LT
,
Bianconi
ME
,
Thomas
GH
,
Kelly
DJ
.
Tripartite ATP-independent periplasmic (TRAP) transporters and tripartite tricarboxylate transporters (TTT): from uptake to pathogenicity
.
Front Cell Infect Microbiol
.
2018 Feb
;
8
:
33
.
338.
Rother
M
,
Resch
A
,
Wilting
R
,
Böck
A
.
Selenoprotein synthesis in archaea
.
BioFactors
.
2001 Dec
;
14
(
1‒4
):
75
83
.
339.
Rudolphi
A
,
Tschech
A
,
Fuchs
G
.
Anaerobic degradation of cresols by denitrifying bacteria
.
Arch Microbiol
.
1991 Feb
;
155
(
3
):
238
48
.
340.
Ruprecht
J
,
Yankovskaya
V
,
Maklashina
E
,
Iwata
S
,
Cecchini
G
.
Structure of Escherichia coli succinate:quinone oxidoreductase with an occupied and empty quinone-binding site
.
J Biol Chem
.
2009 Oct
;
284
(
43
):
29836
46
.
341.
Safarian
S
,
Rajendran
C
,
Müller
H
,
Preu
J
,
Langer
JD
,
Ovchinnikov
S
, et al
.
Structure of a bd oxidase indicates similar mechanisms for membrane-integrated oxygen reductases
.
Science
.
2016 Apr
;
352
(
6285
):
583
6
.
342.
Saier Jr
MH
,
.
A functional-phylogenetic classification system for transmembrane solute transporters
.
Microbiol Mol Biol Rev
.
2000 Jun
;
64
(
2
):
354
411
.
343.
Salewski
J
,
Batista
AP
,
Sena
FV
,
Millo
D
,
Zebger
I
,
Pereira
MM
, et al
.
Substrate-protein interactions of type II NADH:quinone oxidoreductase from Escherichia coli
.
Biochemistry
.
2016 May
;
55
(
19
):
2722
34
.
344.
Sánchez-Ortiz
VJ
,
Domenzain
C
,
Poggio
S
,
Dreyfus
G
,
Camarena
L
.
The periplasmic component of the DctPQM TRAP-transporter is part of the DctS/DctR sensory pathway in Rhodobacter sphaeroides
.
Microbiology
.
2021 Mar
;
167
(
3
):
001037
.
345.
Sargent
F
.
The model [NiFe]-hydrogenases of Escherichia coli
.
Adv Microb Physiol
.
2016
;
68
:
433
507
.
346.
Satoh
A
,
Kim
J-K
,
Miyahara
I
,
Devreese
B
,
Vandenberghe
I
,
Hacisalihoglu
A
, et al
.
Crystal structure of quinohemoprotein amine dehydrogenase from Pseudomonas putida. Identification of a novel quinone cofactor encaged by multiple thioether cross-bridges
.
J Biol Chem
.
2002 Jan
;
277
(
4
):
2830
4
.
347.
Schaffitzel
C
,
Berg
M
,
Dimroth
P
,
Pos
KM
.
Identification of an Na+-dependent malonate transporter of Malonomonas rubra and its dependence on two separate genes
.
J Bacteriol
.
1998 May
;
180
(
10
):
2689
93
.
348.
Schägger
H
,
Pfeiffer
K
.
Supercomplexes in the respiratory chains of yeast and mammalian mitochondria
.
EMBO J
.
2000 Apr
;
19
(
8
):
1777
83
.
349.
Schmehl
M
,
Jahn
A
,
Meyer zu Vilsendorf
A
,
Hennecke
S
,
Masepohl
B
,
Schuppler
M
, et al
.
Identification of a new class of nitrogen fixation genes in Rhodobacter capsulatus: a putative membrane complex involved in electron transport to nitrogenase
.
Mol Gen Genet
.
1993 Dec
;
241
(
5‒6
):
602
15
.
350.
Schmeling
S
,
Fuchs
G
.
Anaerobic metabolism of phenol in proteobacteria and further studies of phenylphosphate carboxylase
.
Arch Microbiol
.
2009 Dec
;
191
(
12
):
869
78
.
351.
Schmeling
S
,
Narmandakh
A
,
Schmitt
O
,
Gad’on
N
,
Schühle
K
,
Fuchs
G
.
Phenylphosphate synthase: a new phosphotransferase catalyzing the first step in anaerobic phenol metabolism in Thauera aromatica
.
J Bacteriol
.
2004 Dec
;
186
(
23
):
8044
57
.
352.
Schmitt
G
,
Arndt
F
,
Kahnt
J
,
Heider
J
.
Adaptations to a loss-of-function mutation in the betaproteobacterium Aromatoleum aromaticum: recruitment of alternative enzymes for anaerobic phenylalanine degradation
.
J Bacteriol
.
2017 Oct
;
199
(
20
):
e00383
17
.
353.
Schmitt
G
,
Saft
M
,
Arndt
F
,
Kahnt
J
,
Heider
J
.
Two different quinohemoprotein amine dehydrogenases initiate anaerobic degradation of aromatic amines in Aromatoleum aromaticum EbN1
.
J Bacteriol
.
2019 Aug
;
201
(
16
):
e00281
19
.
354.
Schnaars
V
,
Wöhlbrand
L
,
Scheve
S
,
Hinrichs
C
,
Reinhardt
R
,
Rabus
R
.
Proteogenomic insights into the physiology of marine, sulfate-reducing, filamentous Desulfonema limicola and Desulfonema magnum
.
Microb Physiol
.
2021 Feb
;
31
(
1
):
36
55
.
355.
Schneider
S
,
Fuchs
G
.
Phenylacetyl-CoA:acceptor oxidoreductase, a new α-oxidizing enzyme that produces phenylglyoxylate. Assay, membrane localization, and differential production in Thauera aromatica
.
Arch Microbiol
.
1998 Jun
;
169
(
6
):
509
16
.
356.
Schneider
S
,
Mohamed
ME-S
,
Fuchs
G
.
Anaerobic metabolism of L-phenylalanine via benzoyl-CoA in the denitrifying bacterium Thauera aromatica
.
Arch Microbiol
.
1997 Oct
;
168
(
4
):
310
20
.
357.
Schneider
K
,
Asao
M
,
Carter
MS
,
Alber
BE
.
Rhodobacter sphaeroides uses a reductive route via propionyl coenzyme A to assimilate 3-hydroxypropionate
.
J Bacteriol
.
2012 Jan
;
194
(
2
):
225
32
.
358.
Schreiber
K
,
Krieger
R
,
Benkert
B
,
Eschbach
M
,
Arai
H
,
Schobert
M
, et al
.
The anaerobic regulatory network required for Pseudomonas aeruginosa nitrate respiration
.
J Bacteriol
.
2007 Jun
;
189
(
11
):
4310
4
.
359.
Schühle
K
,
Fuchs
G
.
Phenylphosphate carboxylase: a new C-C lyase involved in anaerobic phenol metabolism in Thauera aromatica
.
J Bacteriol
.
2004 Jul
;
186
(
14
):
4556
67
.
360.
Schühle
K
,
Heider
J
.
Acetone and butanone metabolism of the denitrifying bacterium “Aromatoleum aromaticum” demonstrates novel biochemical properties of an ATP-dependent aliphatic ketone carboxylase
.
J Bacteriol
.
2012 Jan
;
194
(
1
):
131
41
.
361.
Schühle
K
,
Jahn
M
,
Ghisla
S
,
Fuchs
G
.
Two similar gene clusters coding for enzymes of a new type of aerobic 2-aminobenzoate (anthranilate) metabolism in the bacterium Azoarcus evansii
.
J Bacteriol
.
2001 Sep
;
183
(
18
):
5268
78
.
362.
Schühle
K
,
Gescher
J
,
Feil
U
,
Paul
M
,
Jahn
M
,
Schägger
H
, et al
.
Benzoate-coenzyme A ligase from Thauera aromatica: an enzyme acting in anaerobic and aerobic pathways
.
J Bacteriol
.
2003 Aug
;
185
(
16
):
4920
9
.
363.
Schühle
K
,
Nies
J
,
Heider
J
.
An indoleacetate-CoA ligase and a phenylsuccinyl-CoA transferase involved in anaerobic metabolism of auxin
.
Environ Microbiol
.
2016 Sep
;
18
(
9
):
3120
32
.
364.
Schühle
K
,
Saft
M
,
Vögeli
B
,
Erb
TJ
,
Heider
J
.
Benzylmalonyl-CoA dehydrogenase, an enzyme involved in bacterial auxin degradation
.
Arch Microbiol
.
2021 Sep
;
203
(
7
):
4149
59
.
365.
Schwanhold
N
,
Iobbi-Nivol
C
,
Lehmann
A
,
Leimkühler
S
.
Same but different: comparison of two system-specific molecular chaperones for the maturation of formate dehydrogenases
.
PLoS One
.
2018 Nov
;
13
(
11
):
e0201935
.
366.
Schwede
TF
,
Rétey
J
,
Schulz
GE
.
Crystal structure of histidine ammonia-lyase revealing a novel polypeptide modification as the catalytic electrophile
.
Biochemistry
.
1999 Apr
;
38
(
17
):
5355
61
.
367.
Seelmann
CS
,
Willistein
M
,
Heider
J
,
Boll
M
.
Tungstoenzymes: occurrence, catalytic diversity and cofactor synthesis
.
Inorganics
.
2020 Jul
;
8
(
8
):
44
.
368.
Seyhan
D
,
Friedrich
P
,
Szaleniec
M
,
Hilberg
M
,
Buckel
W
,
Golding
BT
, et al
.
Elucidating the stereochemistry of enzymatic benzylsuccinate synthesis with chirally labeled toluene
.
Angew Chem Int Ed
.
2016 Sep
;
55
(
38
):
11664
7
.
369.
Shamala
N
,
Lim
LW
,
Mathews
FS
,
McIntire
W
,
Singer
TP
,
Hopper
DJ
.
Structure of an intermolecular electron-transfer complex: p-cresol methylhydroxylase at 6.0-Å resolution
.
Proc Natl Acad Sci USA
.
1986 Jul
;
83
(
13
):
4626
30
.
370.
Sharma
V
,
Noriega
CE
,
Rowe
JJ
.
Involvement of NarK1 and NarK2 proteins in transport of nitrate and nitrite in the denitrifying bacterium Pseudomonas aeruginosa PAO1
.
Appl Environ Microbiol
.
2006 Jan
;
72
(
1
):
695
701
.
371.
Shaw
JP
,
Harayama
S
.
Purification and characterization of TOL plasmid-encoded benzyl alcohol dehydrogenase and benzaldehyde dehydrogenase of Pseudomonas putida
.
Eur J Biochem
.
1990 Aug
;
191
(
3
):
705
14
.
372.
Shiro
Y
.
Structure and function of bacterial nitric oxide reductases: nitric oxide reductase, anaerobic enzymes
.
Biochim Biophys Acta
.
2012 Oct
;
1817
(
10
):
1907
13
.
373.
Shiro
Y
,
Sugimoto
H
,
Tosha
T
,
Nagano
S
,
Hino
T
.
Structural basis for nitrous oxide generation by bacterial nitric oxide reductases
.
Phil Trans R Soc B
.
2012 May
;
367
(
1593
):
1195
203
.
374.
Sikkema
J
,
DeBont
JAM
,
Poolman
B
.
Mechanisms of membrane toxicity of hydrocarbons
.
Microbiol Rev
.
1995 Jun
;
59
(
2
):
201
22
.
375.
Sluis
MK
,
Ensign
SA
.
Purification and characterization of acetone carboxylase from Xanthobacter strain Py2
.
Proc Natl Acad Sci USA
.
1997 Aug
;
94
(
16
):
8456
61
.
376.
Sluis
MK
,
Small
FJ
,
Allen
JR
,
Ensign
SA
.
Involvement of an ATP-dependent carboxylase in a CO2-dependent pathway of acetone metabolism by Xanthobacter strain Py2
.
J Bacteriol
.
1996 Jul
;
178
(
14
):
4020
6
.
377.
Sluis
MK
,
Larsen
RA
,
Krum
JG
,
Anderson
R
,
Metcalf
WW
,
Ensign
SA
.
Biochemical, molecular, and genetic analyses of the acetone carboxylase from Xanthobacter autotrophicus strain Py2 and Rhodobacter capsulatus strain B10
.
J Bacteriol
.
2002 Jun
;
184
(
11
):
2969
77
.
378.
Sobti
M
,
Ishmukhametov
R
,
Bouwer
JC
,
Ayer
A
,
Suarna
C
,
Smith
NJ
, et al
.
Cryo-EM reveals distinct conformations of E. coli ATP synthase on exposure to ATP
.
Elife
.
2019 Mar
;
8
:
e43864
.
379.
Sokatch
JR
,
McCully
V
,
Roberts
CM
.
Purification of a branched-chain keto acid dehydrogenase from Pseudomonas putida
.
J Bacteriol
.
1981 Nov
;
148
(
2
):
647
52
.
380.
Song
B
,
Ward
BB
.
Nitrite reductase genes in halobenzoate degrading denitrifying bacteria
.
FEMS Microbiol Ecol
.
2003 Apr
;
43
(
3
):
349
57
.
381.
Spratt
SK
,
Black
PN
,
Ragozzino
MM
,
Nunn
WD
.
Cloning, mapping, and expression of genes involved in the fatty acid-degradative multienzyme complex of Escherichia coli
.
J Bacteriol
.
1984 May
;
158
(
2
):
535
42
.
382.
Starai
VJ
,
Escalante-Semerena
JC
.
Acetyl-coenzyme A synthetase (AMP forming)
.
Cell Mol Life Sci
.
2004 Aug
;
61
(
16
):
2020
30
.
383.
Steimle
S
,
van Eeuwen
T
,
Ozturk
Y
,
Kim
HJ
,
Braitbard
M
,
Selamoglu
N
, et al
.
Cryo-EM structures of engineered active bc1-cbb3 type CIII2CIV super complexes and electronic communication between the complexes
.
Nat Commun
.
2021 Feb
;
12
(
1
):
929
.
384.
Stewart
V
,
Lu
Y
,
Darwin
AJ
.
Periplasmic nitrate reductase (NapABC enzyme) supports anaerobic respiration by Escherichia coli K-12
.
J Bacteriol
.
2002 Mar
;
184
(
5
):
1314
23
.
385.
Strijkstra
A
,
Trautwein
K
,
Jarling
R
,
Wöhlbrand
L
,
Dörries
M
,
Reinhardt
R
, et al
.
Anaerobic activation of p-cymene in denitrifying betaproteobacteria: methyl group hydroxylation versus addition to fumarate
.
Appl Environ Microbiol
.
2014 Dec
;
80
(
24
):
7592
603
.
386.
Sun
C
,
Benlekbir
S
,
Venkatakrishnan
P
,
Wang
Y
,
Hong
S
,
Hosler
J
, et al
.
Structure of the alternative complex III in a supercomplex with cytochrome oxidase
.
Nature
.
2018 May
;
557
(
7703
):
123
6
.
387.
Surber
MW
,
Maloy
S
.
The PutA protein of Salmonella typhimurium catalyzes the two steps of proline degradation via a leaky channel
.
Arch Biochem Biophys
.
1998 Jun
;
354
(
2
):
281
7
.
388.
Suvorova
IA
,
Ravcheev
DA
,
Gelfand
MS
.
Regulation and evolution of malonate and propionate catabolism in Proteobacteria
.
J Bacteriol
.
2012 Jun
;
194
(
12
):
3234
40
.
389.
Sykes
PJ
,
Burns
G
,
Menard
J
,
Hatter
K
,
Sokatch
JR
.
Molecular cloning of genes encoding branched-chain keto acid dehydrogenase of Pseudomonas putida
.
J Bacteriol
.
1987 Apr
;
169
(
4
):
1619
25
.
390.
Szaleniec
M
,
Heider
J
.
Modeling of the reaction mechanism of enzymatic radical C−C coupling by benzylsuccinate synthase
.
Int J Mol Sci
.
2016 Apr
;
17
(
4
):
514
.
391.
Szaleniec
M
,
Hagel
C
,
Menke
M
,
Nowak
P
,
Witko
M
,
Heider
J
.
Kinetics and mechanism of oxygen-independent hydrocarbon hydroxylation by ethylbenzene dehydrogenase
.
Biochemistry
.
2007 Jun
;
46
(
25
):
7637
46
.
392.
Szaleniec
M
,
Borowski
T
,
Schühle
K
,
Witko
M
,
Heider
J
.
Ab initio modeling of ethylbenzene dehydrogenase reaction mechanism
.
J Am Chem Soc
.
2010 May
;
132
(
17
):
6014
24
.
393.
Szaleniec
M
,
Dudzik
A
,
Kozik
B
,
Borowski
T
,
Heider
J
,
Witko
M
.
Mechanistic basis for the enantioselectivity of the anaerobic hydroxylation of alkylaromatic compounds by ethylbenzene dehydrogenase
.
J Inorg Biochem
.
2014 Oct
;
139
:
9
20
.
394.
Tataruch
M
,
Heider
J
,
Bryjak
J
,
Nowak
P
,
Knack
D
,
Czerniak
A
, et al
.
Suitability of the hydrocarbon-hydroxylating molybdenum-enzyme ethylbenzene dehydrogenase for industrial chiral alcohol production
.
J Biotechnol
.
2014 Dec
;
192
(
B
):
400
9
.
395.
Tataruch
M
,
Illeová
V
,
Miłaczewska
A
,
Borowski
T
,
Mihal’
M
,
Polakovič
M
.
Inactivation and aggregation of R-specific 1-(4-hydroxyphenyl)ethanol dehydrogenase from Aromatoleum aromaticum
.
Int J Biol Macromol
.
2023 Apr
;
234
:
123772
.
396.
Teufel
R
,
Mascaraque
V
,
Ismail
W
,
Voss
M
,
Perera
J
,
Eisenreich
W
, et al
.
Bacterial phenylalanine and phenylacetate catabolic pathway revealed
.
Proc Natl Acad Sci USA
.
2010 Aug
;
107
(
32
):
14390
5
.
397.
Teufel
R
,
Gantert
C
,
Voss
M
,
Eisenreich
W
,
Haehnel
W
,
Fuchs
G
.
Studies on the mechanism of ring hydrolysis in phenylacetate degradation. A metabolic branching point
.
J Biol Chem
.
2011 Apr
;
286
(
13
):
11021
34
.
398.
Teufel
R
,
Friedrich
T
,
Fuchs
G
.
An oxygenase that forms and deoxygenates toxic epoxide
.
Nature
.
2012 Mar
;
483
(
7389
):
359
62
.
399.
Thiele
B
,
Rieder
O
,
Golding
BT
,
Müller
M
,
Boll
M
.
Mechanism of enzymatic Birch reduction: stereochemical course and exchange reactions of benzoyl-CoA reductase
.
J Am Chem Soc
.
2008 Oct
;
130
(
43
):
14050
1
.
400.
Thomas
MT
,
Shepherd
M
,
Poole
RK
,
van Vliet
AHM
,
Kelly
DJ
,
Pearson
BM
.
Two respiratory enzyme systems in Campylobacter jejuni NCTC 11168 contribute to growth on L-lactate
.
Environ Microbiol
.
2011 Jan
;
13
(
1
):
48
61
.
401.
Trautwein
K
,
Kühner
S
,
Wöhlbrand
L
,
Halder
T
,
Kuchta
K
,
Steinbüchel
A
, et al
.
Solvent stress response of the denitrifying bacterium “Aromatoleum aromaticum” EbN1
.
Appl Environ Microbiol
.
2008 Apr
;
74
(
8
):
2267
74
.
402.
Trautwein
K
,
Lahme
S
,
Wöhlbrand
L
,
Feenders
C
,
Mangelsdorf
K
,
Harder
J
, et al
.
Physiological and proteomic adaptation of “Aromatoleum aromaticum” EbN1 to low growth rates in benzoate-limited, anoxic chemostats
.
J Bacteriol
.
2012a May
;
194
(
9
):
2165
80
.
403.
Trautwein
K
,
Wilkes
H
,
Rabus
R
.
Proteogenomic evidence for β-oxidation of plant-derived 3-phenylpropanoids in “Aromatoleum aromaticum” EbN1
.
Proteomics
.
2012b May
;
12
(
9
):
1402
13
.
404.
Trautwein
K
,
Grundmann
O
,
Wöhlbrand
L
,
Eberlein
C
,
Boll
M
,
Rabus
R
.
Benzoate mediates repression of C4-dicarboxylate utilization in “Aromatoleum aromaticum” EbN1
.
J Bacteriol
.
2012c Jan
;
194
(
2
):
518
28
.
405.
Tropel
D
,
van der Meer
JR
.
Bacterial transcriptional regulators for degradation pathways of aromatic compounds
.
Microbiol Mol Biol Rev
.
2004 Sep
;
68
(
3
):
474
500
.
406.
Tyagi
R
,
Eswaramoorthy
S
,
Burley
SK
,
Raushel
FM
,
Swaminathan
S
.
A common catalytic mechanism for proteins of the HutI family
.
Biochemistry
.
2008 May
;
47
(
20
):
5608
15
.
407.
Unciuleac
M
,
Boll
M
,
Warkentin
E
,
Ermler
U
.
Crystallization of 4-hydroxybenzoyl-CoA reductase and the structure of its electron donor ferredoxin
.
Acta Crystallogr D Biol Crystallogr
.
2004a Feb
;
60
(
2
):
388
91
.
408.
Unciuleac
M
,
Warkentin
E
,
Page
CC
,
Boll
M
,
Ermler
U
.
Structure of a xanthine oxidase-related 4-hydroxybenzoyl-CoA reductase with an additional [4Fe-4S] cluster and an inverted electron flow
.
Structure
.
2004b Dec
;
12
(
12
):
2249
56
.
409.
Unden
G
,
Dünnwald
P
.
The aerobic and anaerobic respiratory chain of Escherichia coli and Salmonella enterica: enzymes and energetics
.
EcoSalPlus
.
2008
. https://doi.org/10.1128/ecosalplus.3.2.2.
410.
Unden
G
,
Achebach
S
,
Holighaus
G
,
Tran
HG
,
Wackwitz
B
,
Zeuner
Y
.
Control of FNR function of Escherichia coli by O2 and reducing conditions
.
J Mol Microbiol Biotechnol
.
2002 May
;
4
(
3
):
263
8
.
411.
Vagts
J
,
Scheve
S
,
Kant
M
,
Wöhlbrand
L
,
Rabus
R
.
Towards the response threshold for p-hydroxyacetophenone in the denitrifying bacterium “Aromatoleum aromaticum” EbN1
.
Appl Environ Microbiol
.
2018 Sep
;
84
(
18
):
e01018
18
.
412.
Vagts
J
,
Weiten
A
,
Scheve
S
,
Kalvelage
K
,
Swirski
S
,
Wöhlbrand
L
, et al
.
Nanomolar responsiveness of an anaerobic degradation specialist to alkylphenol pollutants
.
J Bacteriol
.
2020 Mar
;
202
(
5
):
e00595
19
.
413.
Vagts
J
,
Kalvelage
K
,
Weiten
A
,
Buschen
R
,
Gutsch
J
,
Scheve
S
, et al
.
Responsiveness of Aromatoleum aromaticum EbN1T to lignin-derived phenylpropanoids
.
Appl Environ Microbiol
.
2021 Jun
;
87
(
11
):
e03140
20
.
414.
Valderrama
JA
,
Durante-Rodríguez
G
,
Blázquez
B
,
García
JL
,
Carmona
M
,
Díaz
E
.
Bacterial degradation of benzoate. Cross-regulation between aerobic and anaerobic pathways
.
J Biol Chem
.
2012
Mar;
287
(
13
):
10494
508
.
415.
van den Berg
B
,
Black
PN
,
Clemons Jr
WM
,
Rapoport
TA
.
Crystal structure of the long-chain fatty acid transporter FadL
.
Science
.
2004 Jun
;
304
(
5676
):
1506
9
.
416.
van Spanning
RJM
,
de Boer
APN
,
Reijenders
WNM
,
Westerhoff
HV
,
Stouthamer
AH
,
van der Oost
J
.
FnrP and NNR of Paracoccus denitrificans are both members of the FNR family of transcriptional activators but have distinct roles in respiratory adaptation in response to oxygen limitation
.
Mol Microbiol
.
1997 Mar
;
23
(
5
):
893
907
.
417.
Vandenbroucke
M
,
Largeau
C
.
Kerogen origin, evolution and structure
.
Org Geochem
.
2007 May
;
38
(
5
):
719
833
.
418.
Velazquez
I
,
Nakamaru-Ogiso
E
,
Yano
T
,
Ohnishi
T
,
Yagi
T
.
Amino acid residues associated with cluster N3 in the NuoF subunit of the proton-translocating NADH-quinone oxidoreductase from Escherichia coli
.
FEBS Lett
.
2005 Jun
;
579
(
14
):
3164
8
.
419.
Vercellino
I
,
Sazanov
LA
.
The assembly, regulation and function of the mitochondrial respiratory chain
.
Nat Rev Mol Cell Biol
.
2022 Feb
;
23
(
2
):
141
61
.
420.
Vinod
MP
,
Bellur
P
,
Becker
DF
.
Electrochemical and functional characterization of the proline dehydrogenase domain of the PutA flavoprotein from Escherichia coli
.
Biochemistry
.
2002 May
;
41
(
20
):
6525
32
.
421.
Vitt
S
,
Prinz
S
,
Eisinger
M
,
Ermler
U
,
Buckel
W
.
Purification and structural characterization of the Na+-translocating ferredoxin:NAD+ reductase (Rnf) complex of Clostridium tetanomorphum
.
Nat Commun
.
2022 Oct
;
13
(
1
):
6315
.
422.
Vogt
MS
,
Schühle
K
,
Kölzer
S
,
Peschke
P
,
Chowdhury
NP
,
Kleinsorge
D
, et al
.
Structural and functional characterization of an electron transfer flavoprotein involved in toluene degradation in strictly anaerobic bacteria
.
J Bacteriol
.
2019 Nov
;
201
(
21
):
e00326
19
.
423.
Vollack
K-U
,
Zumft
WG
.
Nitric oxide signaling and transcriptional control of denitrification genes in Pseudomonas stutzeri
.
J Bacteriol
.
2001 Apr
;
183
(
8
):
2516
26
.
424.
Vollack
K-U
,
Härtig
E
,
Körner
H
,
Zumft
WG
.
Multiple transcription factors of the FNR family in denitrifying Pseudomonas stutzeri: characterization of four fnr-like genes, regulatory responses and cognate metabolic processes
.
Mol Microbiol
.
1999 Mar
;
31
(
6
):
1681
94
.
425.
von Horsten
S
,
Lippert
M-L
,
Geisselbrecht
Y
,
Schühle
K
,
Schall
I
,
Essen
L-O
, et al
.
Inactive pseudoenzyme subunits in heterotetrameric BbsCD, a novel short-chain alcohol dehydrogenase involved in anaerobic toluene degradation
.
FEBS J
.
2022 Feb
;
289
(
4
):
1023
42
.
426.
Wang
Y
,
Huang
Y
,
Wang
J
,
Cheng
C
,
Huang
W
,
Lu
P
, et al
.
Structure of the formate transporter FocA reveals a pentameric aquaporin-like channel
.
Nature
.
2009 Nov
;
462
(
7272
):
467
72
.
427.
Wang
Z
,
Wen
Q
,
Harwood
CS
,
Liang
B
,
Yang
J
.
A disjointed pathway for malonate degradation by Rhodopseudomonas palustris
.
Appl Environ Microbiol
.
2020 Jun
;
86
(
11
):
e00631
20
.
428.
Watanabe
R
,
Iino
R
,
Noji
H
.
Phosphate release in F1-ATPase catalytic cycle follows ADP release
.
Nat Chem Biol
.
2010 Nov
;
6
(
11
):
814
20
.
429.
Weaver
T
.
Structure of free fumarase C from Escherichia coli
.
Acta Crystallogr D Biol Crystallogr
.
2005 Oct
;
61
(
10
):
1395
401
.
430.
Weaver
TM
,
Levitt
DG
,
Donnelly
MI
,
Wilkens Stevens
PP
,
Banaszak
LJ
.
The multisubunit active site of fumarase C from Escherichia coli
.
Nat Struct Biol
.
1995 Aug
;
2
(
8
):
654
62
.
431.
Weber
J
.
ATP synthase: subunit–subunit interactions in the stator stalk
.
Biochim Biophys Acta
.
2006 Sep–Oct
;
1757
(
9-10
):
1162
70
.
432.
Weber
J
.
Structural biology: toward the ATP synthase mechanism
.
Nat Chem Biol
.
2010 Nov
;
6
(
11
):
794
5
.
433.
Weidenweber
S
,
Schühle
K
,
Demmer
U
,
Warkentin
E
,
Ermler
U
,
Heider
J
.
Structure of the acetophenone carboxylase core complex: prototype of a new class of ATP-dependent carboxylases/hydrolases
.
Sci Rep
.
2017 Jan
;
7
:
39674
.
434.
Weidenweber
S
,
Schühle
K
,
Lippert
M-L
,
Mock
J
,
Seubert
A
,
Demmer
U
, et al
.
Finis tolueni: a new type of thiolase with an integrated Zn-finger subunit catalyzes the final step of anaerobic toluene metabolism
.
FEBS J
.
2022 Sep
;
289
(
18
):
5599
616
.
435.
Wherland
S
,
Farver
O
,
Pecht
I
.
Intramolecular electron transfer in nitrite reductases
.
ChemPhysChem
.
2005 May
;
6
(
5
):
805
12
.
436.
Wiegand
G
,
Remington
SJ
.
Citrate synthase: structure, control, and mechanism
.
Ann Rev Biophys Biophys Chem
.
1986
;
15
:
97
117
.
437.
Wilkes
H
,
Schwarzbauer
J
.
Hydrocarbons: an introduction to structure, physico-chemical properties and natural occurrence
. In:
Timmes
KN
, editor.
Handbook of hydrocarbon and lipid microbiology
.
Springer-Verlag Berlin Heidelberg
;
2010
. p.
3
48
.
438.
Wilkes
H
,
Buckel
W
,
Golding
BT
,
Rabus
R
.
Metabolism of hydrocarbons in n-alkane-utilizing anaerobic bacteria
.
J Mol Microbiol Biotechnol
.
2016 Mar
;
26
(
1-3
):
138
51
.
439.
Winiarska
A
,
Hege
D
,
Gemmecker
Y
,
Kryściak-Czerwenka
J
,
Seubert
A
,
Heider
J
, et al
.
Tungsten enzyme using hydrogen as an electron donor to reduce carboxylic acids and NAD+
.
ACS Catal
.
2022 Jul
;
12
(
14
):
8707
17
.
440.
Winiarska
A
,
Ramírez-Amador
F
,
Hege
D
,
Gemmecker
Y
,
Prinz
S
,
Hochberg
G
, et al
.
A bacterial tungsten-containing aldehyde oxidoreductase forms an enzymatic decorated protein nanowire
.
Sci Adv
.
2023 Jun
;
9
(
22
):
eadg6689
.
441.
Wirtz
L
,
Eder
M
,
Brand
A-K
,
Jung
H
.
HutT functions as the major L-histidine transporter in Pseudomonas putida KT2440
.
FEBS Lett
.
2021 Aug
;
595
(
16
):
2113
26
.
442.
Wöhlbrand
L
,
Kallerhoff
B
,
Lange
D
,
Hufnagel
P
,
Thiermann
J
,
Reinhardt
R
, et al
.
Functional proteomic view of metabolic regulation in “Aromatoleum aromaticum” strain EbN1
.
Proteomics
.
2007 Jun
;
7
(
13
):
2222
39
.
443.
Wöhlbrand
L
,
Rabus
R
.
Development of a genetic system for the denitrifying bacterium “Aromatoleum aromaticum” strain EbN1
.
J Mol Microbiol Biotechnol
.
2009 May
;
17
(
1
):
41
52
.
444.
Wöhlbrand
L
,
Wilkes
H
,
Halder
T
,
Rabus
R
.
Anaerobic degradation of p-ethylphenol by “Aromatoleum aromaticum” strain EbN1: pathway, regulation, and involved proteins
.
J Bacteriol
.
2008 Aug
;
190
(
16
):
5699
709
.
445.
Wolodko
WT
,
Fraser
ME
,
James
MNG
,
Bridger
WA
.
The crystal structure of succinyl-CoA synthetase from Escherichia coli at 2.5-Å resolution
.
J Biol Chem
.
1994 Apr
;
269
(
14
):
10883
90
.
446.
Wood
NJ
,
Alizadeh
T
,
Richardson
DJ
,
Ferguson
SJ
,
Moir
JWB
.
Two domains of a dual-function NarK protein are required for nitrate uptake, the first step of denitrification in Paracoccus pantotrophus
.
Mol Microbiol
.
2002 Apr
;
44
(
1
):
157
70
.
447.
Wunsch
P
,
Zumft
WG
.
Functional domains of NosR, a novel transmembrane iron-sulfur flavoprotein necessary for nitrous oxide respiration
.
J Bacteriol
.
2005 Mar
;
187
(
6
):
1992
2001
.
448.
Xia
D
,
Esser
L
,
Tang
W-K
,
Zhou
F
,
Zhou
Y
,
Yu
L
, et al
.
Structural analysis of cytochrome bc1 complexes: implications to the mechanism of function
.
Biochim Biophys Acta
.
2013 Nov‒Dec
;
1827
(
11‒12
):
1278
94
.
449.
Xie
H
,
Buschmann
S
,
Langer
JD
,
Ludwig
B
,
Michel
H
.
Biochemical and biophysical characterization of the two isoforms of cbb3-type cytochrome c oxidase from Pseudomonas stutzeri
.
J Bacteriol
.
2014 Jan
;
196
(
2
):
472
82
.
450.
Xu
Y
,
Gao
X
,
Wang
S-H
,
Liu
H
,
Williams
PA
,
Zhou
N-Y
.
MhbT is a specific transporter for 3-hydroxybenzoate uptake by Gram-negative bacteria
.
Appl Environ Microbiol
.
2012 Sep
;
78
(
17
):
6113
20
.
451.
Xu
J
,
Ding
Z
,
Liu
B
,
Yi
SM
,
Li
J
,
Zhang
Z
, et al
.
Structure of the cytochrome aa3-600 heme-copper menaquinol oxidase bound to inhibitor HQNO shows TM0 is part of the quinol binding site
.
Proc Natl Acad Sci USA
.
2020 Jan
;
117
(
2
):
872
6
.
452.
Yan
H
,
Huang
W
,
Yan
C
,
Gong
X
,
Jiang
S
,
Zhao
Y
, et al
.
Structure and mechanism of a nitrate transporter
.
Cell Rep
.
2013 Mar
;
3
(
3
):
716
23
.
453.
Yang
S-Y
,
He
X-Y
,
Schulz
H
.
Glutamate 139 of the large α-subunit is the catalytic base in the dehydration of both D- and L-3-hydroxyacyl-coenzyme A but not in the isomerization of Δ32-enoyl-coenzyme A catalyzed by the multienzyme complex of fatty acid oxidation in Escherichia coli
.
Biochemistry
.
1995 May
;
34
(
19
):
6441
7
.
454.
Ye
RW
,
Fries
MR
,
Bezborodnikov
SG
,
Averill
BA
,
Tiedje
JM
.
Characterization of the structural gene encoding a copper-containing nitrite reductase and homology of this gene to DNA of other denitrifiers
.
Appl Environ Microbiol
.
1993 Jan
;
59
(
1
):
250
4
.
455.
Yu
Y
,
Liang
Y-H
,
Brostromer
E
,
Quan
J-M
,
Panjikar
S
,
Dong
Y-H
, et al
.
A catalytic mechanism revealed by the crystal structures of the imidazolonepropionase from Bacillus subtilis
.
J Biol Chem
.
2006 Dec
;
281
(
48
):
36929
36
.
456.
Zaar
A
,
Eisenreich
W
,
Bacher
A
,
Fuchs
G
.
A novel pathway of aerobic benzoate catabolism in the bacteria Azoarcus evansii and Bacillus stearothermophilus
.
J Biol Chem
.
2001 Jul
;
276
(
27
):
24997
5004
.
457.
Zaar
A
,
Gescher
J
,
Eisenreich
W
,
Bacher
A
,
Fuchs
G
.
New enzymes involved in aerobic benzoate metabolism in Azoarcus evansii
.
Mol Microbiol
.
2004 Oct
;
54
(
1
):
223
38
.
458.
Zhang
J
,
Frerman
FE
,
Kim
J-JP
.
Structure of electron transfer flavoprotein-ubiquinone oxidoreductase and electron transfer to the mitochondrial ubiquinone pool
.
Proc Natl Acad Sci USA
.
2006 Oct
;
103
(
44
):
16212
7
.
459.
Zhang
L
,
Trncik
C
,
Andrade
SLA
,
Einsle
O
.
The flavinyl transferase ApbE of Pseudomonas stutzeri matures the NosR protein required for nitrous oxide reduction
.
Biochim Biophys Acta Bioenerg
.
2017 Feb
;
1858
(
2
):
95
102
.
460.
Zhang
L
,
Wüst
A
,
Prasser
B
,
Müller
C
,
Einsle
O
.
Functional assembly of nitrous oxide reductase provides insights into copper site maturation
.
Proc Natl Acad Sci USA
.
2019 Jun
;
116
(
26
):
12822
7
.
461.
Zheng
H
,
Wisedchaisri
G
,
Gonen
T
.
Crystal structure of a nitrate/nitrite exchanger
.
Nature
.
2013 May
;
497
(
7451
):
647
51
.
462.
Zhou
N-Y
,
Fuenmayor
SL
,
Williams
PA
.
nag Genes of Ralstonia (formerly Pseudomonas) sp. strain U2 encoding enzymes for gentisate catabolism
.
J Bacteriol
.
2001 Jan
;
183
(
2
):
700
8
.
463.
Zorn
M
,
Ihling
CH
,
Golbik
R
,
Sawers
RG
,
Sinz
A
.
Selective selC-independent selenocysteine incorporation into formate dehydrogenase
.
PLoS One
.
2013 Apr
;
8
(
4
):
e61913
.
464.
Zumft
WG
.
Cell biology and molecular basis of denitrification
.
Microbiol Mol Biol Rev
.
1997 Dec
;
61
(
4
):
533
616
.
465.
Zumft
WG
,
Braun
C
,
Cuypers
H
.
Nitric oxide reductase from Pseudomonas stutzeri. Primary structure and gene organization of a novel bacterial cytochrome bc complex
.
Eur J Biochem
.
1994 Jan
;
219
(
1‒2
):
481
90
.