Taxis in archaea (original) (raw)

Emerg Top Life Sci. 2018 Dec 12; 2(4): 535–546.

Monitoring Editor: Nicholas P. Robinson

Tessa E.F. Quax

1Molecular Biology of Archaea, Faculty of Biology, University of Freiburg, 79104 Freiburg, Germany

Sonja-Verena Albers

1Molecular Biology of Archaea, Faculty of Biology, University of Freiburg, 79104 Freiburg, Germany

Friedhelm Pfeiffer

2Computational Biology Group, Max-Planck-Institute of Biochemistry, Am Klopferspitz 18, 82152 Martinsried, Germany

1Molecular Biology of Archaea, Faculty of Biology, University of Freiburg, 79104 Freiburg, Germany

2Computational Biology Group, Max-Planck-Institute of Biochemistry, Am Klopferspitz 18, 82152 Martinsried, Germany

Received 2018 Jul 5; Revised 2018 Sep 26; Accepted 2018 Sep 27.

Abstract

Microorganisms can move towards favorable growth conditions as a response to environmental stimuli. This process requires a motility structure and a system to direct the movement. For swimming motility, archaea employ a rotating filament, the archaellum. This archaea-specific structure is functionally equivalent, but structurally different, from the bacterial flagellum. To control the directionality of movement, some archaea make use of the chemotaxis system, which is used for the same purpose by bacteria. Over the past decades, chemotaxis has been studied in detail in several model bacteria. In contrast, archaeal chemotaxis is much less explored and largely restricted to analyses in halophilic archaea. In this review, we summarize the available information on archaeal taxis. We conclude that archaeal chemotaxis proteins function similarly as their bacterial counterparts. However, because the motility structures are fundamentally different, an archaea-specific docking mechanism is required, for which initial experimental data have only recently been obtained.

Keywords: archaeal proteins, archaellum, chemotaxis, flagella, motility

Principles of motility and taxis

Microorganisms respond to changes in the environment in order to optimize growth. One possibility to achieve this is to move towards a location with favorable conditions, a process named taxis. This movement along gradients requires both a sensory system and a motility machinery.

To achieve directed motility in liquid, bacteria and archaea use rotating filamentous motility structures to generate a propulsive force: bacterial flagella and archaella (formerly known as archaeal flagella) (Figure 1) [15]. These may generate different kinds of swimming behavior, such as (i) forward movement powered by a bundle of flagella interspersed with a tumbling non-directed motion (Escherichia coli) [6,7], (ii) polar flagella/archaella that push or pull the cell (Vibrio alginolyticus and Halobacterium salinarum) [810], (iii) unidirectional motor rotation alternating with stops (Rhodobacter sphaeroides) [11], and (iv) wrapping of the flagellum around the cell body resulting in a corkscrew motion of the cell (Burkholderia sp.) [12,13] and other mechanisms.

An external file that holds a picture, illustration, etc. Object name is ETLS-2-535-g0001.jpg

Schematic representation of the chemotaxis system and its signaling cascade to the bacterial and archaeal motility structures.

As examples, a Gram-negative bacterium and a euryarchaeon are shown. Bacteria present the flagellum at their cell surface. New flagellins travel through the hollow interior of the filament to be added at the tip of the growing structure. In archaea, archaellins are N-terminally processed by PibD, which cleaves upstream of the signal peptide H domain. They are then added to the base of the growing structure, in a similar fashion as for type IV pili. A simplified version of the chemotaxis system is depicted, which shows how signals are transferred via the MCPs and CheW, resulting in autophosphorylation of CheA. The phosphate is transferred from CheA to CheY. In bacteria, phosphorylated CheY diffuses to the base of the flagellum where it binds to the switch complex, resulting in a change in the direction of rotation. In archaea, CheY requires the presence of CheF in order to bind to the archaellum. The exact composition and structural organization of the archaeal switch complex is yet not resolved. MCP, methyl-accepting chemotaxis protein; OM, outer membrane; IM, inner membrane; PG, peptidoglycan. Single letters refer to gene names with the prefix che (Che system) or arl (previously fla) (archaellum system).

This swimming behavior depends on three states of the motility structure: forward motion, reverse motion, or stationary (cells are motionless or tumble). The effects of rotation direction on the swimming state vary between species and depend on the positioning and architecture of the motility structures. Motility is not by itself directional. Rather, directionality is generated in the form of a biased random walk [14]. As microorganisms generally are too small to sense spatial gradients, they rely on temporal sensing. In addition, they need some kind of memory in order to allow comparison with the situation a second ago. Bacteria and some archaea utilize the chemotaxis system as a sensory system to achieve tactic movement [1417]. The main effect of the chemotactic system is an influence on the duration between subsequent switches between clockwise and counter-clockwise rotation of the motility structure [1720]. This leads to the bias of the random walk and thus determines the directionality of the movement. Microorganisms may swim towards attractants and swim away from repellents [14]. Commonly, an increase in attractant concentration and a reduction of repellent concentration elicit an equivalent response.

This review has its focus on archaeal motility and chemotaxis. There are similarities and also major differences between this physiological process in bacteria and archaea. As many excellent reviews on taxis in bacteria exist [14,18,21], we cover bacterial systems only to the level necessary to understand the similarities and differences to the archaeal system.

Motility structures of archaea and bacteria

The rotating motility structures of bacteria (flagella) and archaea (archaella) are functionally similar, but have a fundamentally different molecular organization (Figure 1) [2,22]. The archaellum requires only 8–13 proteins, none of which shares homology with the ∼30 proteins constituting the flagellum [2325]. Instead, the assembly mechanism of the archaellum is similar to that of bacterial type IV pili [2,26]. The formation of type IV pili involves N-terminal cleavage of the major filament-forming proteins and their addition to the base of the growing pilus via the type IV pilus assembly system [27,28]. The filament of the flagellum, in contrast, is made up of non-processed proteins, flagellins [29], which are secreted via a type III transport system, located at the base of the flagellum, and travel through the hollow interior of the flagellum to be added at the tip of the growing structure [29]. Rotation of bacterial flagella is driven by a proton motive force (or in some cases, a sodium motive force). The force-generating system is unrelated to the biogenesis mechanism. In contrast, rotation of the archaellum requires ATP hydrolysis [3033]. Also, there is clear evidence that the component responsible for powering archaellar rotation is also involved in archaellar biogenesis [3436]. In short, while bacteria and archaea possess functionally similar motility structures, at a structural level they are fundamentally different.

Diversity of chemotaxis systems

Motility structures are especially useful in concert with a system directing the movement of a cell towards an environment with better conditions. For this purpose, referred to as taxis, bacteria have adopted a two-component sensory system and a protein methylation-dependent system for adaptation. Whereas the motility structures of bacteria and archaea are fundamentally different, the two-component sensory system is conserved between bacteria and some archaea (mainly Euryarchaea, but also some Thaumarchaea) [15,16,37]. The conserved taxis signaling cascade is centered around the histidine kinase CheA and the response regulator CheY [17,18,21,38]. It represents an amplification cascade so that a moderate input signal will result in a strong output signal. Both the bacterial and archaeal chemotaxis system have components for adaptation [15,17,19,21]. However, the details of the adaptation system differ, not only between archaea and bacteria, but also within the bacterial domain [15,18,39]. This is evident when comparing the chemotactic systems of E. coli and Bacillus subtilis, the latter showing more similarities to the archaeal system.

Bacterial chemotaxis: example of E. coli and B. subtilis

Chemotaxis in E. coli has been extensively studied, and because of its simplicity, it serves as an attractive model [19]. E. coli cells are propelled forward by a bundle of about five flagella that rotate in a counter-clockwise direction [6,40]. When the flagella rotate in a clockwise direction, the bundle breaks apart and cells start to tumble, stopping forward movement [7,40]. The tumbling results in reorientation, such that when the flagella start rotating counter-clockwise again, the cell swims off in a different, randomly chosen direction [6,7]. The timing of this switch of rotation of the flagellar bundle relies on the chemotaxis system, in particular the concentration of phosphorylated CheY protein (CheY-P) [41,42]. Signal intensity (e.g. chemical concentration) can be sensed by receptors named methyl-accepting chemotaxis proteins (MCPs), which are organized in trimers of dimers that form large hexagonal arrays of thousands of receptors [21,4345]. Based on the presence or absence of a transmembrane domain in the MCPs, the arrays can be anchored in the membrane or be cytosolic [44,46,47]. Binding of the signaling molecule to MCPs results in a conformational change of the receptors [4850]. The histidine kinase CheA is, via the adaptor protein CheW, coupled to the MCPs and their conformational change can regulate the autophosphorylation activity of CheA. The phosphate on CheA is subsequently transferred to CheY, so that CheA autophosphorylation activity affects the level of phosphorylated response regulator, CheY [42,51]. Therefore, CheA functions as an integrator for diverse stimuli and delivers one unambiguous output. CheY-P diffuses to the flagellar motor where it interacts with proteins in the ‘switch complex’ (FliM, FliN and FliG) resulting in an oppositely rotating flagellum [52,53].

The key reaction for adaptation is a reversible methylation and demethylation of the MCPs, thus regulating their sensitivity. MCPs have a central signaling domain, forming a hairpin turn, which on both sides is embedded in coiled-coil domains consisting of heptad repeats [5456]. Methylation takes place on selected Glu residues within such heptads [5659], some of which are encoded as Gln and are subsequently deamidated to Glu by CheB in E. coli [60]. The CheR protein is responsible for methylation, while CheB is responsible for demethylation [61,62]. Bacterial CheB, which is composed of a catalytic and a CheY-like domain, is activated via phosphorylation by CheA [62]. In B. subtilis, CheD, instead of CheB, catalyzes deamidation of Gln residues to methylatable Glu residues [63]. In labeling studies, demethylation can be studied by analyzing the release of methanol [6466]. In E. coli, a positive stimulus leads to an increase of methanol release and removal of the positive stimulus leads to a decrease [65]. The response of B. subtilis is different as methanol is released upon both, addition as well as removal of a positive stimulus [64]. This may indicate that opposite stimuli lead to adaptation via methylation of alternate sites in the MCPs of B. subtilis [56].

In addition to the aforementioned proteins, in different bacteria, some other chemotaxis proteins play a role in the adaptation of the signal (CheC, CheV) and removal of the phosphate from CheY-P (CheZ, CheX, CheC, FliY) [61,62,6773]. For example, B. subtilis possesses two extra signal adaptation mechanisms in addition to the CheB/CheR methylation system: (i) the CheC/CheD system and (ii) a system involving phosphorylation of CheV, a homolog of CheW, coupled to a response regulator domain [63,74,75]. The CheC/CheD system consists of the CheY-P phosphatase, CheC, and the receptor deamidase, CheD (see above). However, the interaction between these two proteins is more important for adaptation than their enzymatic activities [76]. CheD is thought to activate CheA by binding to the receptors. Increasing levels of CheY-P stimulate CheC–CheD interaction, leading to a reduction of CheA activation by CheD [63,74,75].

About 50% of chemotactic bacterial genomes contain more than one copy of the core chemotaxis genes [77]. Moreover, many organisms contain additional chemotaxis proteins with no homology to those in E. coli [15,37,77]. For example, B. subtilis possesses an extensive chemotaxis system with at least a copy of each characterized chemotaxis protein (except CheZ) [72]. Based on the variation of protein components, chemotaxis systems have been divided into 19 classes [37]. In addition to the limited set of chemotaxis proteins of the E. coli class, chemotaxis systems belonging to other classes contain an extensive number of proteins with additional functions. Therefore, the relative simplicity of the E. coli system is not representative of all bacteria.

Taxis in Halobacterium and other archaea

In contrast with the well-explored bacterial chemotaxis systems, studies on archaeal chemotaxis are limited. It was shown that several euryarchaea display tactic behavior and respond to acetate, some amino acids or light [7880]. However, the majority of these studies seem not yet to have been combined with genetic or biochemical investigations.

Most of our understanding of archaeal taxis has been gleaned from studies of the halophilic model organism, H. salinarum. This microbe can perform aerobic respiration, arginine fermentation and use light as an energy source [81]. The latter requires the activity of light-driven ion pumps bacteriorhodopsin (BR) and halorhodopsin (HR) that pump out protons or import chloride ions, respectively [82]. As a result of these different growth strategies, H. salinarum can sense and respond to different stimuli such as light, oxygen, amino acids, osmolytes, and membrane potential [8393].

H. salinarum contains 18 different MCP homologs named Halobacterial transducer proteins (Htrs), which either possess their own sensing domain or interact with other receptor proteins (Table 1) [56]. Six lack transmembrane helices, while the others contain between two and six transmembrane helices. For eight of these different Htrs, the stimuli have been identified (Table 1). HtrI and HtrII mediate the phototactic response, with light being sensed by their associated retinal proteins SRI and SRII, respectively [86,87,94,95,98]. Htr8 and Htr10 (HemAT) are required for attraction and phobic responses to oxygen, respectively [90,93]. Htr14 (MpcT) can detect changes in membrane potential [91], while the cytosolic Htr11 (Car) mediates taxis to the fermentable amino acid arginine, which is thus measured intracellularly. Htr3 (BasT) is responsible for the detection of branched and sulfur-containing amino acids (Leu, Ile, Val, Met, and Cys) [83,92], and Htr5 (CosT) mediates chemotaxis to compatible osmolytes of the betaine family [84]. Both BasT and CosT require binding proteins in order to transduce environmental stimuli. The binding proteins of BasT and CosT are closely related to periplasmic substrate-binding proteins from ABC transporter systems. They belong to the large set of secreted proteins which carry a lipobox and are retained by a covalently attached lipid anchor [84,99]. Bacterial-binding proteins have a dual function as they initiate solute uptake through substrate binding and interaction with the ABC transport system but also mediate chemotactic responses [100]. However, in H. salinarum, the binding proteins are exclusively involved in detection of substrates [84].

Table 1

Halobacterial transducer proteins: transducer proteins (Htrs, MCPs) from H. salinarum and H. volcanii are listed, together with their protein partners

htr number Gene #TM Code (locus tag) Partner Stimulus Reference
htr1 htrI 2 OE3347F sopI Attractant: orange light; repellent: UV light [87,94,118]
htr2 htrII 2 OE3481R sopII Repellent: blue light; attractant: amino acid (Ser) [89,95,118]
htr3 basT 2 OE3611R; HVO_0554 basB Branched and sulfur-containing amino acids (Leu, Ile, Val, Met, Cys) [83]
htr4 2 OE2189R
htr5 cosT 2 OE3474R cosB Compatible osmolytes of betaine family [84]
htr6 2 OE2168R bdgProt
htr7 3 OE3473F; HVO_1999 3TMprot
htr8 6 OE3167F; HVO_1779 Attractant: oxygen [90]
htr9 0 OE2996R
htr10 hemAT 0 OE3150R; HVO_1484 + HVO_1126 Repellent: oxygen [93]
htr11 car 0 OE5243F Arginine [92]
htr12 0 OE3070R
htr13 0 OE2474R
htr14 mpcT 2 OE1536R; HVO_0420 Membrane potential [91]
htr15 0 OE2392R; HVO_0555 + HVO_3005 arlD (flaD)
htr16 2 OE1929R
htr17 3 OE3436R 3TMprot
htr18 2 OE2195F bdgProt
htr36 2 HVO_2214
htr37 2 HVO_2462 CBSdom
htr38 0 HVO_2220
htr39 1 HVO_0969

Locus tags starting with OE are from H. salinarum strain R1 [96], and those starting with HVO_ are from H. volcanii [97]. Partners are encoded in the same operon or gene cluster. ‘bdgProt’ refers to partners which belong to the ABC-type transport system periplasmic substrate-binding protein superfamily. ‘3TMprot’ refers to an uncharacterized partner having three TM domains, while the MCP also has three TM domains. Htrs with both types of locus tags refer to ortholog sets. Htr numbers htr36–htr39 refer to Htrs from H. volcanii which do not have an ortholog in H. salinarum. Htr numbers between htr18 and htr36 are assigned to other species (e.g. Natronomonas pharaonis).

Phototaxis allows H. salinarum to move in the direction of optimal conditions for the two light-driven ion pumps, BR and HR, and at the same time avoid harmful UV radiation that might cause DNA damage. Phototaxis is mediated by Htr1 and Htr2 that receive signals from the photoreceptors SRI and SRII, respectively [86,87,94,95,98]. SRI is a photochromic receptor, which detects orange (attractant) as well as UV (repellent) light, while SRII detects blue light [101103]. Htr1 and Htr2 are physically connected to the sensory rhodopsin. Besides mediating phototaxis, Htr2 is also involved in sensing of serine (Table 1) [89]. Light activation of SRI and SRII induces the release of membrane-bound fumarate, which was proposed to be an alternative switch factor of archaellum rotation [104,105]. Phototaxis offers boundless experimental possibilities. In contrast with chemical stimuli a light stimulus can be instantly switched on and off, allowing for a dynamic range of the duration of the stimulus as well as the intensity. This well-studied phototactic response is one of the use cases for systems biology applications [20,106,107].

The archaeal signal transduction cascade

Significant studies of the archaeal chemotaxis proteins remain confined to a few archaea: H. salinarum and recently also Haloferax volcanii. Archaea move by alternating forward and reverse swimming motility as facilitated by clockwise (CW) and counter-clockwise (CCW) rotation of the archaellum [8,9,31]. This mode of swimming behavior is different from E. coli, which involves tumbling, and instead appears more similar to the activity displayed by other bacteria such as V. alginolyticus [10]. In the absence of stimuli, H. salinarum and Haloferax volcanii cells were shown to perform a random walk [108,109]. As in bacteria, the concentration of CheY-P determines the switch frequency of the motility structure in archaea. In the absence of CheY, H. salinarum and Haloferax volcanii cells swim preferentially forward [17,108]. This may imply a bias for CW rotation in the absence of switching.

The protein composition of the H. salinarum and Haloferax volcanii chemotaxis systems (containing CheA, Y, W1, W2, R, B, C1, C2, C3, D, F1, and F2, or CheA, Y, W, R, B, C, D, F1, and F2, respectively) are more similar to the extensive set of B. subtilis than to the streamlined version of E. coli [15]. The individual role of some of these chemotaxis components has been genetically analyzed in H. salinarum, and their roles in taxis confirmed [17]. Deletion of cheB abolishes chemotaxis and results in cells with a high frequency of reversals, which is similar to phenotypes observed for an equivalent mutation in bacteria such as E. coli and B. subtilis [17]. While reversals were more frequent, the proportion of time spent in either CW or CCW rotation was still 50 : 50, similar to wild-type H. salinarum. This situation corresponds to that in B. subtilis ΔcheB, but is different from E. coli ΔcheB. A proteomic analysis showed that transducer methylation is absent from a Δ_cheR_ mutant while overmethylation was encountered in a Δ_cheB_ mutation, confirming their homology-assigned functions [56]. H. salinarum CheB is also involved in glutamine deamidation, which is equivalent to the dual activity of the E. coli enzyme, even though H. salinarum encodes a homolog of the B. subtilis glutamine deamidase CheD. Together, the CheB/CheR action is important for signal adaptation [110].

Archaeal chemotaxis operons also encode CheC/CheD homologs, in addition to the CheB/CheR adaptation system, reminiscent of the situation in B. subtilis. Furthermore, CheC and CheD from the euryarchaeon Pyrococcus horikoshii were identified as interaction partners in a large-scale protein interaction study [111]. Deletion of cheC1 (previously cheJ) in H. salinarum led to reduced chemotactic activity, a lower frequency of reversals and the ratio between CW : CCW rotation was perturbed to 88 : 12, which is quite similar to the reported outcome in B. subtilis [17]. Nevertheless, none of the catalytic functions assigned to the bacterial homologs have yet been confirmed in Halobacterium [56].

Deletion of cheY and cheA, encoding the two core proteins of the two-component sensory system, have also been studied in H. salinarum [17,38]. These experiments showed that even though these archaeal chemotaxis proteins are homologous to those of B. subtilis, they lead to different effects on motility. Deletion of cheY and cheA genes in H. salinarum results in a phenotype whereby cells rarely switch swimming direction and progress in a straight trajectory, similar to the behavior seen in the deletion strains of E. coli, but notably different from the situation in B. subtilis [17,38,112,113]. Increasing concentrations of CheY-P lead to CW rotation in E. coli and H. salinarum, but conversely, result in CCW rotation in B. subtilis [15,17,114,115]. It seems that in archaea, like in E. coli, CheA is activated upon repellent binding, while in B. subtilis CheA activation occurs after attractant binding [15,17,114]. It was shown that CheA is responsible for CheY phosphorylation in H. salinarum [38]. CheY phosphorylation is strictly coupled to CheA dephosphorylation and, consequently, the half-life of archaeal CheY-P is probably very short [38]. In addition, the importance of the conserved CheY phosphorylation site was demonstrated by mutating the conserved aspartic acid (D53) of CheY in Haloferax volcanii, which resulted in a complete loss of chemotaxis [108]. Alignment of bacterial and archaeal CheY amino acid sequences indicated a high conservation [108]. Indeed, the recently resolved crystal structure of Methanococcus maripaludis CheY showed that the overall structure of the protein is very similar to that of bacterial CheY [108]. Specifically well conserved are the residues that are involved in the activation of CheY, such as those important for phosphorylation and for Tyr-Thr coupling, which is required for structural rearrangement during activation [108]. Mutation of these residues in Haloferax volcanii indicated that also in archaea, they are important for activation of CheY. In conclusion, archaeal chemotaxis proteins function in a similar fashion as their bacterial homologs, since the phosphorylation-dependent activation mechanism is conserved between them.

Communication between the archaeal motility machinery and the chemotaxis system

Despite the conservation of the bacterial and archaeal chemotaxis system, the dramatic differences in the two disparate motility systems pose the interesting question of how the chemotaxis system interacts with the archaeal motility structure (Figure 1). As components of the bacterial switch complex, such as FliM, are absent from archaea, CheY-P in archaea must have a different binding partner. A protein pull-down approach of different chemotaxis proteins in H. salinarum indicated that CheY can interact with two proteins, named CheF1 and CheF2 [116,117]. These proteins can additionally bind euryarchaeal-specific archaellum proteins, ArlD and ArlCE (previously named FlaD and FlaCE) [116,117]. Thus, the homologous CheF1 and CheF2 proteins were suggested to represent adaptors to connect the archaellum with the motility machinery [116]. Indeed, deletion of CheF1 in H. salinarum and Haloferax volcanii resulted in cells impaired in chemotaxis, with a low frequency of reversals and an almost 100% CW rotation bias [108,116]. CheF1 is conserved in almost all archaea with a chemotaxis system and its gene is usually located within the che operon. A duplication of cheF1 probably gave rise to cheF2, which is only present in a few haloarchaea [116]. Deletion of cheF2 had just a mild effect on chemotaxis [108,116]. In vitro interaction assays between purified CheY and CheF of M. maripaludis showed that phosphorylation of CheY is important for interaction [108]. Moreover, the crystal structure of archaeal CheY showed that, although there is a high conservation of the protein fold, some residues in the archaeal α-4 helix were, in contrast with bacterial CheY, carrying a prominent negative charge. These residues were mutated in Haloferax volcanii and in vivo analysis showed that cells carrying these mutations had impaired chemotaxis and cells displayed an increased frequency of reversals of rotation compared with the wild type [108]. Moreover, in vitro binding studies showed that mutation of these residues diminishes the binding affinity between CheY and CheF. Interestingly, it was shown that CheC2 in H. salinarum interacts with both CheF adaptor proteins, instead of a direct interaction with CheY as is the case in the bacterial system [72,116,117]. This again indicates the important role of CheF in archaeal chemotaxis.

Thus, the structure and activation mechanism of archaeal and bacterial CheY is highly conserved. The only additional requirement for connection of the chemotaxis system to the archaeal motility machinery seems to be a slightly different surface charge of archaeal CheY and the presence of the adaptor protein CheF. The structural organization and composition of the archaeal switch complex has not been elucidated yet.

Future perspectives

Archaea, like bacteria, perform tactic behavior. In the model euryarchaea, H. salinarum and Haloferax volcanii taxis relies on the chemotaxis system, which might have been obtained from bacteria (most likely from Firmicutes or Thermotogae) by horizontal gene transfer [16,37]. The available molecular and genetic analyses of the haloarchaeal chemotaxis system indicate a high similarity with that of bacteria. However, archaeal chemotaxis is far from being completely understood. The structural differences between the archaellum and flagellum necessitate subtle changes in the archaeal response regulator CheY as well as an adaptor protein for connection to the archaeal motility machinery. The structure of CheF and its cellular localization have yet to be elucidated. CheF might be cytosolic or found semi-permanently bound to the archaellum motor. The composition and structural organization of the archaeal switch complex will be an important research topic to understand the functioning of the archaeal chemotaxis system. In recent years, studies on the cellular positioning of bacterial chemosensory arrays and flagella have provided major insights into the regulation of bacterial cell shape and organization. The current development of stable archaea-compatible fluorescent proteins is expected to finally allow the first direct observation of the positioning of the archaeal chemotaxis machinery and to open an exciting new line of research.

A major remaining mystery in this field is the apparent absence of a functional chemotaxis apparatus in several archaeal phyla. Although these are readily identifiable in the euryarchaea and thaumarchaea, these systems have not been found in other phyla, including the crenarchaea. Representative species of almost all archaeal phyla encode the archaellum. As a motility structure is especially useful in combination with a sensory system, there is the exciting possibility that crenarchaea might possess an undiscovered sensory system, developed independently from the classical chemotaxis system.

Summary

Acknowledgements

We thank Mike Dyall-Smith for critical reading of the manuscript.

Abbreviations

BR bacteriorhodopsin
CCW counter-clockwise
CW clockwise
HR halorhodopsin
Htrs Halobacterial transducer proteins
MCPs methyl-accepting chemotaxis proteins

Funding

T.E.F.Q. was supported by a Carl-Zeiss Post-doc Stipendium.

Competing Interests

The Authors declare that there are no competing interests associated with the manuscript.

References

1. Alam M., Claviez M., Oesterhelt D. and Kessel M. (1984) Flagella and motility behaviour of square bacteria. EMBO J. 3, 2899–2903 10.1002/j.1460-2075.1984.tb02229.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

2. Jarrell K.F. and Albers S.-V. (2012) The archaellum: an old motility structure with a new name. Trends Microbiol. 20, 307–312 10.1016/j.tim.2012.04.007 [PubMed] [CrossRef] [Google Scholar]

3. Albers S.-V. and Jarrell K.F. (2018) The Archaellum: an update on the unique archaeal motility structure. Trends Microbiol. 26, 351–362 10.1016/j.tim.2018.01.004 [PubMed] [CrossRef] [Google Scholar]

4. Khan S. and Scholey J.M. (2018) Assembly, functions and evolution of archaella, flagella and cilia. Curr. Biol. 28, R278–R292 10.1016/j.cub.2018.01.085 [PubMed] [CrossRef] [Google Scholar]

5. Silverman M. and Simon M. (1974) Flagellar rotation and the mechanism of bacterial motility. Nature 249, 73–74 10.1038/249073a0 [PubMed] [CrossRef] [Google Scholar]

6. Berg H.C. and Brown D.A. (1972) Chemotaxis in Escherichia coli analysed by three-dimensional tracking. Nature 239, 500–504 10.1038/239500a0 [PubMed] [CrossRef] [Google Scholar]

7. Darnton N.C., Turner L., Rojevsky S. and Berg H.C. (2007) On torque and tumbling in swimming Escherichia coli. J. Bacteriol. 189, 1756–1764 10.1128/JB.01501-06 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

8. Alam M. and Oesterhelt D. (1984) Morphology, function and isolation of halobacterial flagella. J. Mol. Biol. 176, 459–475 10.1016/0022-2836(84)90172-4 [PubMed] [CrossRef] [Google Scholar]

9. Shahapure R., Driessen R.P., Haurat M.F., Albers S.V. and Dame R.T. (2014) The archaellum: a rotating type IV pilus. Mol. Microbiol. 91, 716–723 10.1111/mmi.12486 [PubMed] [CrossRef] [Google Scholar]

10. Xie L., Altindal T., Chattopadhyay S. and Wu X.-L. (2011) From the cover: bacterial flagellum as a propeller and as a rudder for efficient chemotaxis. Proc. Natl Acad. Sci. U.S.A. 108, 2246–2251 10.1073/pnas.1011953108 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

11. Armitage J.P., Pitta T.P., Vigeant M.A., Packer H.L. and Ford R.M. (1999) Transformations in flagellar structure of Rhodobacter sphaeroides and possible relationship to changes in swimming speed. J. Bacteriol. 181, 4825–4833 PMID: [PMC free article] [PubMed] [Google Scholar]

12. Kinosita Y., Kikuchi Y., Mikami N., Nakane D. and Nishizaka T. (2018) Unforeseen swimming and gliding mode of an insect gut symbiont, Burkholderia sp. RPE64, with wrapping of the flagella around its cell body. ISME J. 12, 838–848 10.1038/s41396-017-0010-z [PMC free article] [PubMed] [CrossRef] [Google Scholar]

13. Kühn M.J., Schmidt F.K., Eckhardt B. and Thormann K.M. (2017) Bacteria exploit a polymorphic instability of the flagellar filament to escape from traps. Proc. Natl Acad. Sci. U.S.A. 114, 6340–6345 10.1073/pnas.1701644114 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

14. Bi S. and Sourjik V. (2018) Stimulus sensing and signal processing in bacterial chemotaxis. Curr. Opin. Microbiol. 45, 22–29 10.1016/j.mib.2018.02.002 [PubMed] [CrossRef] [Google Scholar]

15. Szurmant H. and Ordal G.W. (2004) Diversity in chemotaxis mechanisms among the bacteria and archaea. Microbiol. Mol. Biol. Rev. 68, 301–319 10.1128/MMBR.68.2.301-319.2004 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

16. Briegel A., Ortega D.R., Huang A.N., Oikonomou C.M., Gunsalus R.P. and Jensen G.J. (2015) Structural conservation of chemotaxis machinery across Archaea and Bacteria. Environ. Microbiol. Rep. 7, 414–419 10.1111/1758-2229.12265 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

17. Rudolph J. and Oesterhelt D. (1996) Deletion analysis of the che operon in the archaeon Halobacterium salinarium. J. Mol. Biol. 258, 548–554 10.1006/jmbi.1996.0267 [PubMed] [CrossRef] [Google Scholar]

18. Porter S.L., Wadhams G.H. and Armitage J.P. (2011) Signal processing in complex chemotaxis pathways. Nat. Rev. Microbiol. 9, 153–165 10.1038/nrmicro2505 [PubMed] [CrossRef] [Google Scholar]

19. Colin R. and Sourjik V. (2017) Emergent properties of bacterial chemotaxis pathway. Curr. Opin. Microbiol. 39, 24–33 10.1016/j.mib.2017.07.004 [PubMed] [CrossRef] [Google Scholar]

20. Nutsch T., Oesterhelt D., Gilles E.D. and Marwan W. (2005) A quantitative model of the switch cycle of an archaeal flagellar motor and its sensory control. Biophys. J. 89, 2307–2323 10.1529/biophysj.104.057570 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

21. Parkinson J.S., Hazelbauer G.L. and Falke J.J. (2015) Signaling and sensory adaptation in Escherichia coli chemoreceptors: 2015 update. Trends Microbiol. 23, 257–266 10.1016/j.tim.2015.03.003 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

22. Thomas N.A., Bardy S.L. and Jarrell K.F. (2001) The archaeal flagellum: a different kind of prokaryotic motility structure. FEMS Microbiol. Rev. 25, 147–174 10.1111/j.1574-6976.2001.tb00575.x [PubMed] [CrossRef] [Google Scholar]

23. Chaban B., Ng S.Y.M., Kanbe M., Saltzman I., Nimmo G., Aizawa S.-I. et al. (2007) Systematic deletion analyses of the fla genes in the flagella operon identify several genes essential for proper assembly and function of flagella in the archaeon, Methanococcus maripaludis. Mol. Microbiol. 66, 596–609 10.1111/j.1365-2958.2007.05913.x [PubMed] [CrossRef] [Google Scholar]

24. Lassak K., Neiner T., Ghosh A., Klingl A., Wirth R. and Albers S.V. (2012) Molecular analysis of the crenarchaeal flagellum. Mol. Microbiol. 83, 110–124 10.1111/j.1365-2958.2011.07916.x [PubMed] [CrossRef] [Google Scholar]

25. Macnab R.M. (2003) How bacteria assemble flagella. Annu. Rev. Microbiol. 57, 77–100 10.1146/annurev.micro.57.030502.090832 [PubMed] [CrossRef] [Google Scholar]

26. Jarrell K.F., Bayley D.P., Florian V. and Klein A. (1996) Isolation and characterization of insertional mutations in flagellin genes in the archaeon Methanococcus voltae. Mol. Microbiol. 20, 657–666 10.1046/j.1365-2958.1996.5371058.x [PubMed] [CrossRef] [Google Scholar]

27. Berry J.-L. and Pelicic V. (2015) Exceptionally widespread nanomachines composed of type IV pilins: the prokaryotic Swiss army knives. FEMS Microbiol. Rev. 39, 1–21 10.1093/femsre/fuu005 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

28. Craig L., Volkmann N., Arvai A.S., Pique M.E., Yeager M., Egelman E.H. et al. (2006) Type IV pilus structure by cryo-electron microscopy and crystallography: implications for pilus assembly and functions. Mol. Cell 23, 651–662 10.1016/j.molcel.2006.07.004 [PubMed] [CrossRef] [Google Scholar]

29. Chevance F.F.V. and Hughes K.T. (2008) Coordinating assembly of a bacterial macromolecular machine. Nat. Rev. Microbiol. 6, 455–465 10.1038/nrmicro1887 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

30. Streif S., Staudinger W.F., Marwan W. and Oesterhelt D. (2008) Flagellar rotation in the archaeon Halobacterium salinarum depends on ATP. J. Mol. Biol. 384, 1–8 10.1016/j.jmb.2008.08.057 [PubMed] [CrossRef] [Google Scholar]

31. Kinosita Y., Uchida N., Nakane D. and Nishizaka T. (2016) Direct observation of rotation and steps of the archaellum in the swimming halophilic archaeon Halobacterium salinarum. Nat. Microbiol. 1, 16148 10.1038/nmicrobiol.2016.148 [PubMed] [CrossRef] [Google Scholar]

32. Manson M.D., Tedesco P., Berg H.C., Harold F.M. and Van der Drift C. (1977) A protonmotive force drives bacterial flagella. Proc. Natl Acad. Sci. U.S.A. 74, 3060–3064 10.1073/pnas.74.7.3060 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

33. Hirota N. and Imae Y. (1983) Na+-driven flagellar motors of an alkalophilic Bacillus strain YN-1. J. Biol. Chem. 258, 10577–10581 PMID: [PubMed] [Google Scholar]

34. Ghosh A., Hartung S., van der Does C., Tainer J.A. and Albers S.-V. (2011) Archaeal flagellar ATPase motor shows ATP-dependent hexameric assembly and activity stimulation by specific lipid binding. Biochem. J. 437, 43–52 10.1042/BJ20110410 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

35. Reindl S., Ghosh A., Williams G.J., Lassak K., Neiner T., Henche A.-L. et al. (2013) Insights into FlaI functions in archaeal motor assembly and motility from structures, conformations, and genetics. Mol. Cell 49, 1069–1082 10.1016/j.molcel.2013.01.014 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

36. Chaudhury P., Neiner T., D'Imprima E., Banerjee A., Reindl S., Ghosh A. et al. (2016) The nucleotide-dependent interaction of FlaH and FlaI is essential for assembly and function of the archaellum motor. Mol. Microbiol. 99, 674–685 10.1111/mmi.13260 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

37. Wuichet K. and Zhulin I.B. (2010) Origins and diversification of a complex signal transduction system in prokaryotes. Sci. Signal. 3, ra50 10.1126/scisignal.2000724 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

38. Rudolph J. and Oesterhelt D. (1995) Chemotaxis and phototaxis require a CheA histidine kinase in the archaeon Halobacterium salinarium. EMBO J. 14, 667–673 10.1002/j.1460-2075.1995.tb07045.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

39. Bischoff D.S. and Ordal G.W. (1992) Bacillus subtilis chemotaxis: a deviation from the Escherichia coli paradigm. Mol. Microbiol. 6, 23–28 10.1111/j.1365-2958.1992.tb00833.x [PubMed] [CrossRef] [Google Scholar]

40. Macnab R.M. and Ornston M.K. (1977) Normal-to-curly flagellar transitions and their role in bacterial tumbling. Stabilization of an alternative quaternary structure by mechanical force. J. Mol. Biol. 112, 1–30 10.1016/S0022-2836(77)80153-8 [PubMed] [CrossRef] [Google Scholar]

41. Barak R. and Eisenbach M. (1992) Correlation between phosphorylation of the chemotaxis protein CheY and its activity at the flagellar motor. Biochemistry 31, 1821–1826 10.1021/bi00121a034 [PubMed] [CrossRef] [Google Scholar]

42. Hess J.F., Oosawa K., Kaplan N. and Simon M.I. (1988) Phosphorylation of three proteins in the signaling pathway of bacterial chemotaxis. Cell 53, 79–87 10.1016/0092-8674(88)90489-8 [PubMed] [CrossRef] [Google Scholar]

43. Briegel A., Li X., Bilwes A.M., Hughes K.T., Jensen G.J. and Crane B.R. (2012) Bacterial chemoreceptor arrays are hexagonally packed trimers of receptor dimers networked by rings of kinase and coupling proteins. Proc. Natl Acad. Sci. U.S.A. 109, 3766–3771 10.1073/pnas.1115719109 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

44. Briegel A., Ladinsky M.S., Oikonomou C., Jones C.W., Harris M.J., Fowler D.J. et al. (2014) Structure of bacterial cytoplasmic chemoreceptor arrays and implications for chemotactic signaling. eLife 3, e02151 10.7554/eLife.02151 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

45. Salah Ud-Din A.I.M. and Roujeinikova A. (2017) Methyl-accepting chemotaxis proteins: a core sensing element in prokaryotes and archaea. Cell. Mol. Life Sci. 74, 3293–3303 10.1007/s00018-017-2514-0 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

46. Briegel A., Ortega D.R., Tocheva E.I., Wuichet K., Li Z., Chen S. et al. (2009) Universal architecture of bacterial chemoreceptor arrays. Proc. Natl Acad. Sci. U.S.A. 106, 17181–17186 10.1073/pnas.0905181106 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

47. Liu J., Hu B., Morado D.R., Jani S., Manson M.D. and Margolin W. (2012) Molecular architecture of chemoreceptor arrays revealed by cryoelectron tomography of Escherichia coli minicells. Proc. Natl Acad. Sci. U.S.A. 109, E1481–E1488 10.1073/pnas.1200781109 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

48. Lee G.F., Lebert M.R., Lilly A.A. and Hazelbauer G.L. (1995) Transmembrane signaling characterized in bacterial chemoreceptors by using sulfhydryl cross-linking in vivo. Proc. Natl Acad. Sci. U.S.A. 92, 3391–3395 10.1073/pnas.92.8.3391 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

49. Hughson A.G. and Hazelbauer G.L. (1996) Detecting the conformational change of transmembrane signaling in a bacterial chemoreceptor by measuring effects on disulfide cross-linking in vivo. Proc. Natl Acad. Sci. U.S.A. 93, 11546–11551 10.1073/pnas.93.21.11546 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

50. Chervitz S.A. and Falke J.J. (1996) Molecular mechanism of transmembrane signaling by the aspartate receptor: a model. Proc. Natl Acad. Sci. U.S.A. 93, 2545–2550 10.1073/pnas.93.6.2545 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

51. Hess J.F., Bourret R.B. and Simon M.I. (1988) Histidine phosphorylation and phosphoryl group transfer in bacterial chemotaxis. Nature 336, 139–143 10.1038/336139a0 [PubMed] [CrossRef] [Google Scholar]

52. Welch M., Oosawa K., Aizawa S. and Eisenbach M. (1993) Phosphorylation-dependent binding of a signal molecule to the flagellar switch of bacteria. Proc. Natl Acad. Sci. U.S.A. 90, 8787–8791 10.1073/pnas.90.19.8787 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

53. Sarkar M.K., Paul K. and Blair D. (2010) Chemotaxis signaling protein CheY binds to the rotor protein FliN to control the direction of flagellar rotation in Escherichia coli. Proc. Natl Acad. Sci. U.S.A. 107, 9370–9375 10.1073/pnas.1000935107 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

54. Kim K.K., Yokota H. and Kim S.H. (1999) Four-helical-bundle structure of the cytoplasmic domain of a serine chemotaxis receptor. Nature 400, 787–792 10.1038/23512 [PubMed] [CrossRef] [Google Scholar]

55. Park S.-Y., Borbat P.P., Gonzalez-Bonet G., Bhatnagar J., Pollard A.M., Freed J.H. et al. (2006) Reconstruction of the chemotaxis receptor-kinase assembly. Nat. Struct. Mol. Biol. 13, 400–407 10.1038/nsmb1085 [PubMed] [CrossRef] [Google Scholar]

56. Koch M.K., Staudinger W.F., Siedler F. and Oesterhelt D. (2008) Physiological sites of deamidation and methyl esterification in sensory transducers of Halobacterium salinarum. J. Mol. Biol. 380, 285–302 10.1016/j.jmb.2008.04.063 [PubMed] [CrossRef] [Google Scholar]

57. Goy M.F., Springer M.S. and Adler J. (1977) Sensory transduction in Escherichia coli: role of a protein methylation reaction in sensory adaptation. Proc. Natl Acad. Sci. U.S.A. 74, 4964–4968 10.1073/pnas.74.11.4964 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

58. Kehry M.R., Engström P., Dahlquist F.W. and Hazelbauer G.L. (1983) Multiple covalent modifications of Trg, a sensory transducer of Escherichia coli. J. Biol. Chem. 258, 5050–5055 PMID: [PubMed] [Google Scholar]

59. Zimmer M.A., Tiu J., Collins M.A. and Ordal G.W. (2000) Selective methylation changes on the Bacillus subtilis chemotaxis receptor McpB promote adaptation. J. Biol. Chem. 275, 24264–24272 10.1074/jbc.M004001200 [PubMed] [CrossRef] [Google Scholar]

60. Kehry M.R. and Dahlquist F.W. (1982) Adaptation in bacterial chemotaxis: CheB-dependent modification permits additional methylations of sensory transducer proteins. Cell 29, 761–772 10.1016/0092-8674(82)90438-X [PubMed] [CrossRef] [Google Scholar]

61. Springer W.R. and Koshland D.E. (1977) Identification of a protein methyltransferase as the cheR gene product in the bacterial sensing system. Proc. Natl Acad. Sci. U.S.A. 74, 533–537 10.1073/pnas.74.2.533 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

62. Stock J.B. and Koshland D.E. (1978) A protein methylesterase involved in bacterial sensing. Proc. Natl Acad. Sci. U.S.A. 75, 3659–3663 10.1073/pnas.75.8.3659 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

63. Kristich C.J. and Ordal G.W. (2002) Bacillus subtilis CheD is a chemoreceptor modification enzyme required for chemotaxis. J. Biol. Chem. 277, 25356–25362 10.1074/jbc.M201334200 [PubMed] [CrossRef] [Google Scholar]

64. Thoelke M.S., Bedale W.A., Nettleton D.O. and Ordal G.W. (1987) Evidence for an intermediate methyl-acceptor for chemotaxis in Bacillus subtilis. J. Biol. Chem. 262, 2811–2816 PMID: [PubMed] [Google Scholar]

65. Kehry M.R., Doak T.G. and Dahlquist F.W. (1984) Stimulus-induced changes in methylesterase activity during chemotaxis in Escherichia coli. J. Biol. Chem. 259, 11828–11835 PMID: [PubMed] [Google Scholar]

66. Alam M., Lebert M., Oesterhelt D. and Hazelbauer G.L. (1989) Methyl-accepting taxis proteins in Halobacterium halobium. EMBO J. 8, 631–639 10.1002/j.1460-2075.1989.tb03418.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

67. Simms S.A., Stock A.M. and Stock J.B. (1987) Purification and characterization of the _S_-adenosylmethionine:glutamyl methyltransferase that modifies membrane chemoreceptor proteins in bacteria. J. Biol. Chem. 262, 8537–8543 PMID: [PubMed] [Google Scholar]

68. Simms S.A., Keane M.G. and Stock J. (1985) Multiple forms of the CheB methylesterase in bacterial chemosensing. J. Biol. Chem. 260, 10161–10168 PMID: [PubMed] [Google Scholar]

69. Muff T.J. and Ordal G.W. (2007) The CheC phosphatase regulates chemotactic adaptation through CheD. J. Biol. Chem. 282, 34120–34128 10.1074/jbc.M706432200 [PubMed] [CrossRef] [Google Scholar]

70. Karatan E., Saulmon M.M., Bunn M.W. and Ordal G.W. (2001) Phosphorylation of the response regulator CheV is required for adaptation to attractants during Bacillus subtilis chemotaxis. J. Biol. Chem. 276, 43618–43626 10.1074/jbc.M104955200 [PubMed] [CrossRef] [Google Scholar]

71. Park S.-Y., Chao X., Gonzalez-Bonet G., Beel B.D., Bilwes A.M. and Crane B.R. (2004) Structure and function of an unusual family of protein phosphatases: the bacterial chemotaxis proteins CheC and CheX. Mol. Cell 16, 563–574 10.1016/j.molcel.2004.10.018 [PubMed] [CrossRef] [Google Scholar]

72. Szurmant H., Muff T.J. and Ordal G.W. (2004) Bacillus subtilis CheC and FliY are members of a novel class of CheY-P-hydrolyzing proteins in the chemotactic signal transduction cascade. J. Biol. Chem. 279, 21787–21792 10.1074/jbc.M311497200 [PubMed] [CrossRef] [Google Scholar]

73. Hazelbauer G.L., Falke J.J. and Parkinson J.S. (2008) Bacterial chemoreceptors: high-performance signaling in networked arrays. Trends Biochem. Sci. 33, 9–19 10.1016/j.tibs.2007.09.014 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

74. Rao C.V., Glekas G.D. and Ordal G.W. (2008) The three adaptation systems of Bacillus subtilis chemotaxis. Trends Microbiol. 16, 480–487 10.1016/j.tim.2008.07.003 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

75. Kirby J.R., Kristich C.J., Saulmon M.M., Zimmer M.A., Garrity L.F., Zhulin I.B. et al. (2001) Chec is related to the family of flagellar switch proteins and acts independently from CheD to control chemotaxis in Bacillus subtilis. Mol. Microbiol. 42, 573–585 10.1046/j.1365-2958.2001.02581.x [PubMed] [CrossRef] [Google Scholar]

76. Yuan W., Glekas G.D., Allen G.M., Walukiewicz H.E., Rao C.V. and Ordal G.W. (2012) The importance of the interaction of CheD with CheC and the chemoreceptors compared to its enzymatic activity during chemotaxis in Bacillus subtilis. PLoS ONE 7, e50689 10.1371/journal.pone.0050689 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

77. Hamer R., Chen P.-Y., Armitage J.P., Reinert G. and Deane C.M. (2010) Deciphering chemotaxis pathways using cross species comparisons. BMC Syst. Biol. 4, 3 10.1186/1752-0509-4-3 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

78. Migas J., Anderson K.L., Cruden D.L. and Markovetz A.J. (1989) Chemotaxis in Methanospirillum hungatei. Appl. Environ. Microbiol. 55, 264–265 PMID: [PMC free article] [PubMed] [Google Scholar]

79. Sment K.A. and Konisky J. (1989) Chemotaxis in the archaebacterium Methanococcus voltae. J. Bacteriol. 171, 2870–2872 10.1128/jb.171.5.2870-2872.1989 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

80. Scharf B. and Wolff E.K. (1994) Phototactic behaviour of the archaebacterial Natronobacterium pharaonis. FEBS Lett. 340, 114–116 10.1016/0014-5793(94)80183-5 [PubMed] [CrossRef] [Google Scholar]

81. Gonzalez O., Gronau S., Pfeiffer F., Mendoza E., Zimmer R. and Oesterhelt D. (2009) Systems analysis of bioenergetics and growth of the extreme halophile Halobacterium salinarum. PLoS Comput. Biol. 5, e1000332 10.1371/journal.pcbi.1000332 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

82. Oesterhelt D. and Tittor J. (1989) Two pumps, one principle: light-driven ion transport in halobacteria. Trends Biochem. Sci. 14, 57–61 10.1016/0968-0004(89)90044-3 [PubMed] [CrossRef] [Google Scholar]

83. Kokoeva M.V. and Oesterhelt D. (2000) Bast, a membrane-bound transducer protein for amino acid detection in Halobacterium salinarum. Mol. Microbiol. 35, 647–656 10.1046/j.1365-2958.2000.01735.x [PubMed] [CrossRef] [Google Scholar]

84. Kokoeva M.V., Storch K.-F., Klein C. and Oesterhelt D. (2002) A novel mode of sensory transduction in archaea: binding protein-mediated chemotaxis towards osmoprotectants and amino acids. EMBO J. 21, 2312–2322 10.1093/emboj/21.10.2312 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

85. Seidel R., Scharf B., Gautel M., Kleine K., Oesterhelt D. and Engelhard M. (1995) The primary structure of sensory rhodopsin II: a member of an additional retinal protein subgroup is coexpressed with its transducer, the halobacterial transducer of rhodopsin II. Proc. Natl Acad. Sci. U.S.A. 92, 3036–3040 10.1073/pnas.92.7.3036 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

86. Spudich E.N., Hasselbacher C.A. and Spudich J.L. (1988) Methyl-accepting protein associated with bacterial sensory rhodopsin I. J. Bacteriol. 170, 4280–4285 10.1128/jb.170.9.4280-4285.1988 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

87. Yao V.J. and Spudich J.L. (1992) Primary structure of an archaebacterial transducer, a methyl-accepting protein associated with sensory rhodopsin I. Proc. Natl Acad. Sci. U.S.A. 89, 11915–11919 10.1073/pnas.89.24.11915 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

88. Ferrando-May E., Krah M., Marwan W. and Oesterhelt D. (1993) The methyl-accepting transducer protein HtrI is functionally associated with the photoreceptor sensory rhodopsin I in the archaeon Halobacterium salinarium. EMBO J. 12, 2999–3005 10.1002/j.1460-2075.1993.tb05968.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

89. Hou S., Brooun A., Yu H.S., Freitas T. and Alam M. (1998) Sensory rhodopsin II transducer HtrII is also responsible for serine chemotaxis in the archaeon Halobacterium salinarum. J. Bacteriol. 180, 1600–1602 PMID: [PMC free article] [PubMed] [Google Scholar]

90. Brooun A., Bell J., Freitas T., Larsen R.W. and Alam M. (1998) An archaeal aerotaxis transducer combines subunit I core structures of eukaryotic cytochrome c oxidase and eubacterial methyl-accepting chemotaxis proteins. J. Bacteriol. 180, 1642–1646 PMID: [PMC free article] [PubMed] [Google Scholar]

91. Koch M.K. and Oesterhelt D. (2005) Mpct is the transducer for membrane potential changes in Halobacterium salinarum. Mol. Microbiol. 55, 1681–1694 10.1111/j.1365-2958.2005.04516.x [PubMed] [CrossRef] [Google Scholar]

92. Storch K.-F., Rudolph J. and Oesterhelt D. (1999) Car: a cytoplasmic sensor responsible for arginine chemotaxis in the archaeon Halobacterium salinarum. EMBO J. 18, 1146–1158 10.1093/emboj/18.5.1146 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

93. Hou S., Larsen R.W., Boudko D., Riley C.W., Karatan E., Zimmer M. et al. (2000) Myoglobin-like aerotaxis transducers in Archaea and Bacteria. Nature 403, 540–544 10.1038/35000570 [PubMed] [CrossRef] [Google Scholar]

94. Krah M., Marwan W., Verméglio A. and Oesterhelt D. (1994) Phototaxis of Halobacterium salinarium requires a signalling complex of sensory rhodopsin I and its methyl-accepting transducer HtrI. EMBO J. 13, 2150–2155 10.1002/j.1460-2075.1994.tb06491.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

95. Zhang W., Brooun A., Mueller M.M. and Alam M. (1996) The primary structures of the archaeon Halobacterium salinarium blue light receptor sensory rhodopsin II and its transducer, a methyl-accepting protein. Proc. Natl Acad. Sci. U.S.A. 93, 8230–8235 10.1073/pnas.93.16.8230 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

96. Pfeiffer F., Schuster S.C., Broicher A., Falb M., Palm P., Rodewald K. et al. (2008) Evolution in the laboratory: the genome of Halobacterium salinarum strain R1 compared to that of strain NRC-1. Genomics 91, 335–346 10.1016/j.ygeno.2008.01.001 [PubMed] [CrossRef] [Google Scholar]

97. Hartman A.L., Norais C., Badger J.H., Delmas S., Haldenby S., Madupu R. et al. (2010) The complete genome sequence of Haloferax volcanii DS2, a model archaeon. PLoS ONE 5, e9605 10.1371/journal.pone.0009605 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

98. Hoff W.D., Jung K.-H. and Spudich J.L. (1997) Molecular mechanism of photosignaling by archaeal sensory rhodopsins. Annu. Rev. Biophys. Biomol. Struct. 26, 223–258 10.1146/annurev.biophys.26.1.223 [PubMed] [CrossRef] [Google Scholar]

99. Storf S., Pfeiffer F., Dilks K., Chen Z.Q., Imam S. and Pohlschröder M. (2010) Mutational and bioinformatic analysis of haloarchaeal lipobox-containing proteins. Archaea 2010, 1–11 10.1155/2010/410975 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

100. Ehrmann M., Ehrle R., Hofmann E., Boos W. and Schlösser A. (1998) The ABC maltose transporter. Mol. Microbiol. 29, 685–694 10.1046/j.1365-2958.1998.00915.x [PubMed] [CrossRef] [Google Scholar]

101. Marwan W. and Oesterhelt D. (1987) Signal formation in the halobacterial photophobic response mediated by a fourth retinal protein (P480). J. Mol. Biol. 195, 333–342 10.1016/0022-2836(87)90654-1 [PubMed] [CrossRef] [Google Scholar]

102. Wolff E.K., Bogomolni R.A., Scherrer P., Hess B. and Stoeckenius W. (1986) Color discrimination in halobacteria: spectroscopic characterization of a second sensory receptor covering the blue-green region of the spectrum. Proc. Natl Acad. Sci. U.S.A. 83, 7272–7276 10.1073/pnas.83.19.7272 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

103. Spudich J.L. and Bogomolni R.A. (1992) Sensory rhodopsin I: receptor activation and signal relay. J. Bioenerg. Biomembr. 24, 193–200 10.1007/BF00762677 [PubMed] [CrossRef] [Google Scholar]

104. Marwan W., Schäfer W. and Oesterhelt D. (1990) Signal transduction in Halobacterium depends on fumarate. EMBO J. 9, 355–362 10.1002/j.1460-2075.1990.tb08118.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

105. Montrone M., Marwan W., Grünberg H., Musseleck S., Starostzik C. and Oesterhelt D. (1993) Sensory rhodopsin-controlled release of the switch factor fumarate in Halobacterium salinarium. Mol. Microbiol. 10, 1077–1085 10.1111/j.1365-2958.1993.tb00978.x [PubMed] [CrossRef] [Google Scholar]

106. Nutsch T., Marwan W., Oesterhelt D. and Gilles E.D. (2003) Signal processing and flagellar motor switching during phototaxis of Halobacterium salinarum. Genome Res. 13, 2406–2412 10.1101/gr.1241903 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

107. Streif S., Oesterhelt D. and Marwan W. (2010) A predictive computational model of the kinetic mechanism of stimulus-induced transducer methylation and feedback regulation through CheY in archaeal phototaxis and chemotaxis. BMC Syst. Biol. 4, 27 10.1186/1752-0509-4-27 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

108. Quax T.E.F., Altegoer F., Rossi F., Li Z., Rodriguez-Franco M., Kraus F. et al. (2018) Structure and function of the archaeal response regulator CheY. Proc. Natl Acad. Sci. U.S.A. 115, E1259–E1268 10.1073/pnas.1716661115 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

109. Hildebrand E. and Schimz A. (1986) Integration of photosensory signals in Halobacterium halobium. J. Bacteriol. 167, 305–311 10.1128/jb.167.1.305-311.1986 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

110. Lazova M.D., Ahmed T., Bellomo D., Stocker R. and Shimizu T.S. (2011) Response rescaling in bacterial chemotaxis. Proc. Natl Acad. Sci. U.S.A. 108, 13870–13875 10.1073/pnas.1108608108 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

111. Usui K., Katayama S., Kanamori-Katayama M., Ogawa C., Kai C., Okada M. et al. (2005) Protein-protein interactions of the hyperthermophilic archaeon Pyrococcus horikoshii OT3. Genome Biol. 6, R98 10.1186/gb-2005-6-12-r98 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

112. Fuhrer D.K. and Ordal G.W. (1991) Bacillus subtilis CheN, a homolog of CheA, the central regulator of chemotaxis in Escherichia coli. J. Bacteriol. 173, 7443–7448 10.1128/jb.173.23.7443-7448.1991 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

113. Bischoff D.S. and Ordal G.W. (1991) Sequence and characterization of Bacillus subtilis CheB, a homolog of Escherichia coli CheY, and its role in a different mechanism of chemotaxis. J. Biol. Chem. 266, 12301–12305 PMID: [PubMed] [Google Scholar]

114. Garrity L.F. and Ordal G.W. (1997) Activation of the CheA kinase by asparagine in Bacillus subtilis chemotaxis. Microbiology 143, 2945–2951 10.1099/00221287-143-9-2945 [PubMed] [CrossRef] [Google Scholar]

115. Borkovich K.A., Kaplan N., Hess J.F. and Simon M.I. (1989) Transmembrane signal transduction in bacterial chemotaxis involves ligand-dependent activation of phosphate group transfer. Proc. Natl Acad. Sci. U.S.A. 86, 1208–1212 10.1073/pnas.86.4.1208 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

116. Schlesner M., Miller A., Streif S., Staudinger W.F., Muller J., Scheffer B. et al. (2009) Identification of archaea-specific chemotaxis proteins which interact with the flagellar apparatus. BMC Microbiol. 9, 56 10.1186/1471-2180-9-56 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

117. Schlesner M., Miller A., Besir H., Aivaliotis M., Streif J., Scheffer B. et al. (2012) The protein interaction network of a taxis signal transduction system in a halophilic archaeon. BMC Microbiol. 12, 272 10.1186/1471-2180-12-272 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

118. Spudich E.N., Takahashi T. and Spudich J.L. (1989) Seonsory rhodopsins I and II modulate a methylation/demethylation system in Halobacterium halobium phototaxis. Proc. Natl Acad. Sci U.S.A. 86, 7746–7750 10.1073/pnas.86.20.7746 [PMC free article] [PubMed] [CrossRef] [Google Scholar]