首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
ATP-结合盒(ATP-binding cassette,ABC)转运蛋白是目前已知最大、功能最广泛的蛋白质家族。多向耐药性(pleiotropic drug resistance,PDR)蛋白是该家族中仅存于植物和真菌中的一个亚族,结构域与其他亚族相反,即核苷酸结合域(nucleotide-binding domain,NBD)位于跨膜结构域(trans-membrane domain,TMD)的N端。目前已发现PDR型转运蛋白具有转运次生代谢产物和参与胁迫反应等方面的功能。植物PDR基因分为5个亚族:I族基因涉及多种生物和非生物胁迫反应,II ̄V族基因功能研究甚少。植物PDR基因在器官水平、化学及环境因素影响下具有特异性较好的表达谱。本文系统阐述了植物PDR型转运蛋白基因的进化、结构及其功能,为理解植物PDR型转运蛋白在生物分子转运和复杂生理功能方面提供一个基础框架。  相似文献   

2.
ABC transporters involved in the transport of plant secondary metabolites   总被引:10,自引:0,他引:10  
Yazaki K 《FEBS letters》2006,580(4):1183-1191
Plants produce a large number of secondary metabolites, such as alkaloids, terpenoids, polyphenols, quinones and many further compounds having combined structures of those groups. Physiological roles of those metabolites for plants are still under investigation, but they play, at least in part, important functions as protectants for plant bodies against herbivores and pathogens, as well as from physical stresses like ultraviolet light and heat. In order to accomplish these functions, biosyntheses and accumulation of secondary metabolites are highly regulated in a temporal and spatial manner in plant organs, where they can appropriately accumulate. In this mini-review, I introduce the mechanism of accumulation and membrane transport of these metabolites, in particular, focusing on ATP-binding cassette transporters involved.  相似文献   

3.
The ABC transporters (ATP Binding Cassette) compose one of the bigest protein family with the great medical, industrial and economical impact. They are found in all organism from bacteria to man. ABC proteins are responsible for resistance of microorganism to antibiotics and fungicides and multidrug resistance of cancer cells. Mutations in ABC transporters genes cause seriuos deseases like cystic fibrosis, adrenoleucodystrophy or ataxia. Transport catalized by ABC proteins is charged with energy from the ATP hydrolysis. The ABC superfamily contains transporters, canals, receptors. Analysis of the Saccharomyces cerevisiae genome allowed to distinguish 30 potential ABC proteins which are classified into 6 subfamilies. The structural and functional similarity of the yeast and human ABC proteins allowes to use the S. cerevisiae as a model organism for ABC transporters characterisation. In this work the present state of knowleadge on yeast S. cerevisiae ABC proteins was summarised.  相似文献   

4.
Peroxisomes perform a range of different functions, dependent upon organism, tissue type, developmental stage or environmental conditions, many of which are connected with lipid metabolism. This review summarises recent research on ATP binding cassette (ABC) transporters of the peroxisomal membrane (ABC subfamily D) and their roles in plants, fungi and animals. Analysis of mutants has revealed that peroxisomal ABC transporters play key roles in specific metabolic and developmental functions in different organisms. A common function is import of substrates for beta-oxidation but much remains to be determined concerning transport substrates and mechanisms which appear to differ significantly between phyla.  相似文献   

5.
The ATP binding cassette (ABC) superfamily is a large, ubiquitous and diverse group of proteins, most of which mediate transport across biological membranes. ABC transporters have been shown to function not only as ATP-dependent pumps, but also as ion channels and channel regulators. Whilst members of this gene family have been extensively characterised in mammalian and microbial systems, the study of plant ABC transporters is a relatively new field of investigation. Sequences of over 20 plant ABC proteins have been published and include homologues of P-glycoprotein, MRP, PDR5 and organellar transporters. At present, functions have been assigned to a small proportion of these genes and only the MRP subclass has been extensively characterised. This review aims to summarise literature relevant to the study of plant ABC transporters, to review methods of cloning, to discuss the utility of yeast and mammalian systems as models and to speculate on possible roles of uncharacterised ABC transporters in plants.  相似文献   

6.
7.
ABC (ATP-binding cassette) transporters are primary active membrane proteins that translocate solutes (allocrites) across lipid bilayers. The prototypical ABC transporter consists of four domains: two cytoplasmic NBDs (nucleotide-binding domains) and two TMDs (transmembrane domains). The NBDs, whose primary sequence is highly conserved throughout the superfamily, bind and hydrolyse ATP to power the transport cycle. The TMDs, whose primary sequence and protein fold can be quite disparate, form the translocation pathway across the membrane and generally (but not always) determine allocrite specificity. Structure determination of ABC proteins initially took advantage of the relative ease of expression and crystallization of the hydrophilic bacterial NBDs in isolation from the transporter complex, and revealed detailed information on the structural fold of these domains, the amino acids involved in the binding and hydrolysis of nucleotide, and the head-to-tail arrangement of the NBD-NBD dimer interface. More recently, several intact transporters have been crystallized and three types have, so far, been characterized: type I and II ABC importers, and ABC exporters. All three are present in prokaryotes, but only the ABC exporters appear to be present in eukaryotes. Their structural determination has provided insight into the mechanisms of energy and signal transduction between the NBDs and TMDs (i.e. between the ATP- and allocrite-binding sites) and, for some, the nature of the allocrite-binding site(s) within the TMDs. In this chapter, we focus primarily on the ABC exporters and describe the structural, biochemical and biophysical evidence for and against the controversial bellows-like mechanism proposed for allocrite efflux.  相似文献   

8.
ABC transporters form the largest of all transporter families, and their structural study has made tremendous progress over recent years. However, despite such advances, the precise mechanisms that determine the energy-coupling between ATP hydrolysis and the conformational changes following substrate binding remain to be elucidated. Here, we present our thermodynamic analysis for both ABC importers and exporters, and introduce the two new concepts of differential-binding energy and elastic conformational energy into the discussion.We hope that the structural analysis of ABC transporters will henceforth take thermodynamic aspects of transport mechanisms into account as well.  相似文献   

9.
The ATP switch model for ABC transporters   总被引:1,自引:0,他引:1  
ABC transporters mediate active translocation of a diverse range of molecules across all cell membranes. They comprise two nucleotide-binding domains (NBDs) and two transmembrane domains (TMDs). Recent biochemical, structural and genetic studies have led to the ATP-switch model in which ATP binding and ATP hydrolysis, respectively, induce formation and dissociation of an NBD dimer. This provides an exquisitely regulated switch that induces conformational changes in the TMDs to mediate membrane transport.  相似文献   

10.
Candidatus Liberibacter asiaticus (Ca. L. asiaticus) is a Gram‐negative bacterium and the pathogen of Citrus Greening disease (Huanglongbing, HLB). As a parasitic bacterium, Ca. L. asiaticus harbors ABC transporters that play important roles in exchanging chemical compounds between Ca. L. asiaticus and its host. Here, we analyzed all the ABC transporter‐related proteins in Ca. L. asiaticus. We identified 14 ABC transporter systems and predicted their structures and substrate specificities. In‐depth sequence and structure analysis including multiple sequence alignment, phylogenetic tree reconstruction, and structure comparison further support their function predictions. Our study shows that this bacterium could use these ABC transporters to import metabolites (amino acids and phosphates) and enzyme cofactors (choline, thiamine, iron, manganese, and zinc), resist to organic solvent, heavy metal, and lipid‐like drugs, maintain the composition of the outer membrane (OM), and secrete virulence factors. Although the features of most ABC systems could be deduced from the abundant experimental data on their orthologs, we reported several novel observations within ABC system proteins. Moreover, we identified seven nontransport ABC systems that are likely involved in virulence gene expression regulation, transposon excision regulation, and DNA repair. Our analysis reveals several candidates for further studies to understand and control the disease, including the type I virulence factor secretion system and its substrate that are likely related to Ca. L. asiaticus pathogenicity and the ABC transporter systems responsible for bacterial OM biosynthesis that are good drug targets. Proteins 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

11.
Multidrug resistance ABC transporters   总被引:11,自引:0,他引:11  
Chang G 《FEBS letters》2003,555(1):102-105
Clinical multidrug resistance is caused by a group of integral membrane proteins that transport hydrophobic drugs and lipids across the cell membrane. One class of these permeases, known as multidrug resistance ATP binding cassette (ABC) transporters, translocate these molecules by coupling drug/lipid efflux with energy derived from the hydrolysis of ATP. In this review, we examine both the structures and conformational changes of multidrug resistance ABC transporters. Together with the available biochemical and structural evidence, we propose a general mechanism for hydrophobic substrate transport coupled to ATP hydrolysis.  相似文献   

12.
Structural diversity of ABC transporters   总被引:1,自引:0,他引:1  
ATP-binding cassette (ABC) transporters form a large superfamily of ATP-dependent protein complexes that mediate transport of a vast array of substrates across membranes. The 14 currently available structures of ABC transporters have greatly advanced insight into the transport mechanism and revealed a tremendous structural diversity. Whereas the domains that hydrolyze ATP are structurally related in all ABC transporters, the membrane-embedded domains, where the substrates are translocated, adopt four different unrelated folds. Here, we review the structural characteristics of ABC transporters and discuss the implications of this structural diversity for mechanistic diversity.ATP-binding cassette (ABC) transporters are a large group of membrane protein complexes that couple transport of a substrate across the membrane to the hydrolysis of the phosphate bond between the γ- and the β-phosphate of ATP (Ames et al., 1990; Higgins, 1992; Davidson et al., 2008; Rees et al., 2009). The free energy released when ATP is converted into ADP and orthophosphate (Pi; approximately −50 kJ mol−1 in many cells) can be used to accumulate the transported substrates in, or to remove them from, cellular compartments.In prokaryotes, ABC transporters are localized to the plasma membrane, and ATP is hydrolyzed on the cytoplasmic side. In eukaryotes, ABC transporters are also found in organellar membranes. ATP hydrolysis by organellar ABC transporters takes place on the cytosolic side of the membrane, except for transporters from mitochondria and chloroplasts where the ATP-binding domains of the transporters are located on the matrix or stroma side. The side of the membrane where ATP is bound and hydrolyzed is termed the cis-side, and the opposite side is called the trans-side.ABC transporters can be classified as exporters or importers. Exporters move substrates from the cis-side to the trans-side of the membrane, from the hydrophobic core of the lipid bilayer to the trans-side, or transfer substrates between the inner and outer leaflets of the bilayer. In contrast, importers move substrates from the trans-side to the cis-side. There are a few ABC transporters that do not have a bona fide transport function. Notable examples include the CFTR, which is a gated chloride channel (Aleksandrov et al., 2007), and the sulfonylurea receptor SUR1, which is a regulatory complex associated with a potassium channel (Bryan et al., 2007). There has been tremendous interest in ABC transporters not only from a mechanistic point of view but also because malfunctioning of human ABC transporters leads to a plethora of diseases (see for instance, Silverton et al., 2011), and some ABC exporters are involved in the drug resistance of bacteria and cancer cells (Ambudkar et al., 2003; Davidson et al., 2008).Many excellent reviews on ABC transporters have been published over the past few years and cover the history, structure, mechanism, physiology, and pharmacology of these proteins (Davidson et al., 2008; Rees et al., 2009; Parcej and Tampé, 2010; Eitinger et al., 2011; George and Jones, 2012; Lewis et al., 2012). Because a wealth of new crystal structures has been determined lately, here we provide an update on the structural diversity of ABC transporters.

Overview

Similar to other membrane transport proteins, ABC transporters adopt at least two conformations in which the substrate-binding site is accessible from either the cis-side or the trans-side. Alternation between the two conformations allows substrate translocation across the membrane (“alternating access” model; Jardetzky, 1966; Tanford, 1982). The binding of substrate on one side of the membrane and release on the opposite side are coordinated by ATP binding and hydrolysis, and ADP and Pi release. Several ABC transporters have been crystallized in different conformations (see
Name (organism)RemarksSubstrate boundResolution (Å)PDB accession no.Reference
Type I importers
Molybdate transporter ModB2C2-A complex (Archaeoglobus fulgidus)Inward-facing conformation, with SBP boundMg2+, PO43-, WoO42−3.102ONKHollenstein et al., 2007
ModBC (Methanosarcina acetivorans)Inward-facing conformation, without SBP bound, in a trans-inhibited stateMg2+, WoO42−3.003D31Gerber et al., 2008
MalFGK2-MBP maltose uptake transporter (E. coli)Outward-facing conformation stabilized by a mutation in the NBDs (MalK E159Q), with MBPMaltose, ATP2.802R6GOldham et al., 2007
MalFGK2 maltose uptake transporter (E. coli)TM helix 1 deleted, in inward-facing conformation, in resting state, without MBP4.503FH6Khare et al., 2009
MalFGK2-MBP maltose uptake transporter (E. coli)Pre-translocation intermediate state, with mutations in MBP (G69C/S337C) that stabilize the substrate-bound closed conformation by a cross-linkAMP-PNP, Mg2+, maltose2.903PUZOldham and Chen, 2011a
Outward-facing conformation state, with MBPAMP-PNP, Mg2+, maltose3.103PUY
Pre-translocation intermediate state, with mutations in MBP (G69C/S337C) that stabilize the substrate-bound closed conformation by a cross-linkMaltose3.103PV0
MalFGK2-MBP maltose uptake transporter (E. coli)Outward-facing conformation, with MBPAMP-PNP, Mg2+, maltose2.203RLFOldham and Chen, 2011b
Outward-facing conformation, with MBPADP · VO43−, Mg2+, maltose2.403PUV
Outward-facing conformation, with MBPADP · AlF4, Mg2+, maltose2.303PUW
Outward-facing conformation, with MBPADP · BeF3, Mg2+, maltose2.303PUX
MalFGK2-MBP maltose uptake transporter (E. coli)Complex with its regulatory protein EIIAglc, inward-facing conformation3.914JBWChen et al., 2013
MalFGK2-MBP maltose uptake transporter (E. coli)Outward-facing conformation, with MBPMaltoheptaose2.904KHZOldham et al., 2013
Outward-facing conformation, with MBPANP, α-d-glycose2.384KI0
MetNI methionine uptake transporter (E. coli)Inward-facing conformation3.703DHWKadaba et al., 2008
Inward-facing conformation, at higher resolution (detergent Cymal5)ADP2.903TUIJohnson et al., 2012
Inward-facing conformation, C2 domains repositioned (detergent decylmaltoside)4.003TUJ
Type II importers
BtuC2D2 vitamin B12 transporter (E. coli)Outward-facing conformation, no SBP boundV4O124−3.201L7VLocher et al., 2002
BtuC2D2-F complex (E. coli)BtuC in asymmetric conformation. The translocation pore is closed from both sites. BtuF is in an open state.2.602QI9Hvorup et al., 2007
BtuC2D2-F complex (E. coli)Intermediate occluded state, nucleotide boundAMP-PNP3.474FI3Korkhov et al., 2012b
BtuC2D2-F complex (E. coli)E159Q mutation in NBD abolished ATP hydrolysis activity, BtuC in asymmetric conformation. The translocation pore is closed from both sites.3.494DBLKorkhov et al., 2012a
HI1470/1471 putative metal chelate–type ABC transporter (H. influenzae)Inward-facing conformation, without SBP, renamed MolB2C2 (as later was shown to bind WoO42−/MoO42−)2.402NQ2Pinkett et al., 2007
HmuU2V2 heme transporter (Yersinia pestis)Outward-facing conformation3.004G1UWoo et al., 2012
ECF-type importers
RibU S-component for riboflavin (S. aureus)Substrate boundRiboflavin3.603P5NZhang et al., 2010
ThiT S-component for thiamin (L. lactis)Substrate boundThiamin2.003RLBErkens et al., 2011
BioY S-component for biotin (L. lactis)Substrate boundBiotin2.104DVEBerntsson et al., 2012
NikM S-component for Ni2+ (Thermoanaerobacter tengcongensis)Substrate boundNi2+/Co2+1.83–2.5Yu et al., 2013
ECF-FolT transporter (L. brevis)Substrate free, inward-facing conformation3.004HUQXu et al., 2013
ECF-HmpT transporter (L. brevis)Substrate free, inward-facing conformation3.534HZUWang et al., 2013
Exporters
Sav18662 multidrug transporter (S. aureus)Outward-facing conformationADP3.002HYDDawson and Locher, 2006
Outward-facing conformationAMP-PNP3.402ONJDawson and Locher, 2007
Heterodimeric ABC transporter TM287-TM288 (T. maritima)Inward-facing conformationAMP-PNP2.903QF4Hohl et al., 2012
MsbA2 lipid “flippase” (Salmonella typhimurium)Outward-facing conformation, with nucleotide boundANP-PNP4.503B5YWard et al., 2007
Outward-facing conformation, with nucleotide boundANP-PNP3.703B60
Outward-facing conformation, with nucleotide boundADP, VO43−4.203B5Z
MsbA2 lipid “flippase” E. coliInward-facing conformation, open apo structure5.303B5W
MsbA2 lipid “flippase” Vibrio choleraInward-facing conformation, closed apo structure5.503B5X
Multidrug transporter P-glycoprotein (Mus musculus)Inward-facing conformation3.803G5UAller et al., 2009
Inward-facing conformationCyclic-tris-(R)-valineselenazole4.403G60
Inward-facing conformationCyclic-tris-(S)-valineselenazole4.353G61
Re-refined, inward-facing conformation3.804M1MLi et al., 2014
Multidrug transporter P-glycoprotein (Caenorhabditis elegans)Inward-facing conformation3.404F4CJin et al., 2012
ABCB10 mitochondrial ABC transporter (Homo sapiens)Inward-facing conformationAMPPCP2.854AYTShintre et al., 2013
Inward-facing conformation; different crystal form (rod/plate)AMP-PCP/AMP-PNP2.90/3.304AYX/4AYW
Inward-facing conformation2.853ZDQ
Open in a separate windowSBP, substrate-binding protein; MBP, maltose-binding protein; outward and inward facing, the translocation pathway in the TMDs exposed to the trans-side or the cis-side of the membrane, respectively.All ABC transporters have a core with the same modular architecture: two transmembrane (TM) domains (TMDs) or subunits and two nucleotide-binding domains (NBDs) or subunits. The NBDs, which are highly conserved in structure and sequence among all ABC transporters, are the hallmark of the family. NBDs do not always associate with TMDs but may also be involved in various functions that do not occur at the membrane (see for instance, Boël et al., 2014, and references therein). However, the name “ABC transporter” is only used when the NBDs form a complex with TMDs, and NBDs that are not associated with TMDs will not be discussed here.In contrast to the conserved NBDs, several unrelated folds of the TMDs have been found. These different folds, which are defined by the connectivity and three-dimensional arrangement of the secondary structure elements, do not share significant sequence similarity. Because TMDs with the same fold may also lack sequence similarity, structure determination is necessary for fold assignment. So far, four types of ABC transporters have been identified based on the TMDs folds as determined by the crystal structures (Fig. 1).Open in a separate windowFigure 1.Four distinct folds of ABC transporters. All share a similar general architecture: two NBDs (blue and sky blue) are attached to two TMDs (orange and yellow). In some transporters, additional domains are present (green), which often have a regulatory function (C-terminal regulatory domain [CRD]). In Type I and II importers, the transported compounds are delivered to TMDs by SBPs (or SBDs; magenta) located in periplasm (Gram-negative bacteria) or external space (Gram-positive bacteria and Archaea). ECF, energy coupling factor.Three ABC transporter types appear to be associated exclusively with import functions (transport of substrates from the trans-side to the cis-side of the membrane): Type I and Type II importers, and energy coupling factor (ECF) transporters (also named Type III importers). All three types of ABC importers are found only in prokaryotes. The fourth fold is found in all structurally characterized exporters. ABC transporters with the exporter fold are present in both prokaryotes and eukaryotes. ABC transporters with different TMD folds probably also differ in the mechanistic details of transport (see below).Prokaryotic ABC transporters are often assembled from separate protein subunits (two TMDs and two NBDs; Biemans-Oldehinkel et al., 2006). The two NBDs and TMDs can be identical (homodimeric) or different proteins (heterodimeric). In the latter case, the two NBD subunits are invariably structurally similar. The TMDs in single transporters are also usually similar in structure, with the notable exception of the two TMDs in ECF-type importers, which are completely unrelated (Wang et al., 2013; Xu et al., 2013; Slotboom, 2014). Sometimes two, and occasionally three, subunits are fused into a multi-domain protein in the prokaryotic transporters. In particular, the two NBDs are occasionally fused, and in many prokaryotic exporters, the TMDs are fused with the NBDs (for instance, in bacterial exporters MsbA2 [Ward et al., 2007], Sav18662 [Dawson and Locher, 2006], and TM287/TM288 [Hohl et al., 2012], for which crystal structures have been solved). Eukaryotic exporters are generally composed either of one polypeptide chain containing all the domains (e.g., P-glycoprotein) or of a dimer of two polypeptides, each of which contains an NBD and a TMD (as in the bacterial exporters).The Type I and II ABC importers depend on additional soluble substrate-binding domains (SBDs) or substrate-binding proteins (SBPs) (Fig. 1), which capture the transported substrate on the trans-side and deliver it to the TMDs (Quiocho and Ledvina, 1996; Berntsson et al., 2010). In some cases, the SBD is fused with a TMD into a multi-domain subunit (Biemans-Oldehinkel et al., 2006). ECF transporters and exporters do not require SBPs (Rodionov et al., 2009). Many prokaryotic and eukaryotic ABC transporters contain additional domains or subunits, such as regulatory domains (Fig. 1) or extra TMDs of unknown function. These additional domains are very diverse and will not be discussed here (Biemans-Oldehinkel et al., 2006; Parcej and Tampé, 2010).It is unknown whether the different ABC transporter folds have evolved to address specific mechanistic challenges. It is possible that the structural differences between ABC exporters and importers are related to the opposing directions in which the substrate is pumped, which may lead to different mechanistic requirements. Alternatively, the differences between the exporter and the importer folds may be related to the range of transported substrates. ABC exporters are involved in the transport of hydrophobic compounds such as lipids, fatty acids, cholesterol, and drugs, as well as larger molecules such as proteins (toxins, hydrolytic enzymes, S-layer proteins, lantibiotics, bacteriocins, and competence factors). In addition, most drug exporters can transport a large variety of drugs (of different sizes) out of the cell and are therefore called multidrug-resistant transporters. In contrast, importers generally are selective for a single or a few related water-soluble substrates.The three different types of ABC importers have overlapping substrate specificities, and it is therefore not clear why three importer folds have evolved. In general (but not exclusively), the substrates of Type I importers are compounds required in bulk (such as sugars and amino acids), whereas Type II importers and ECF transporters are more often specific for compounds needed in small quantities (metal chelates, vitamins; Davidson et al., 2008; Eitinger et al., 2011). It is possible that Type I ABC importers are more suitable for high capacity, low affinity transport, whereas Type II and ECF importers may better serve high affinity, low capacity transport. However, this distinction is blurred, and it is possible that the variation in kinetics within an importer type may be as large as the variation between the different types.

NBD

All ABC transporters contain two NBDs, also called ATPases or ABCs, which bind and hydrolyze ATP. The NBDs from ABC transporters are a subgroup of the diverse superfamily of P-loop NTPases (Vetter and Wittinghofer, 1999) and depend on magnesium ions for catalysis. Each NBD has a core of ∼200 amino acids and consists of two subdomains: the larger RecA-like domain, which is also found in other P-loop ATPases, and the structurally more diverse α-helical domain, which is unique to ABC transporters (Fig. 2).Open in a separate windowFigure 2.The structure of the NBDs, as exemplified by the MalK dimer of the maltose transporter MalEFGK2 (Protein Data Bank accession no. 3RLF). (A) View along an axis perpendicular to the membrane plane from the trans-side onto the NBDs (The TMDs and SBP have been removed for clarity). Domains and highly conserved sequence motifs are color-coded: green, α-helical domain; light blue, RecA-like domain; faded gray, regulatory C-terminal domain; red, A-loop; magenta, Walker A; orange, Walker B; blue, D-loop; green, H-loop; cyan, ABC motif; yellow, Q-loop. The ATP analogue AMP-PNP is shown in sticks. (B) A closer look onto the nucleotide-binding site. The key amino acids are indicated (see NBD for details).Structures of NBDs in the absence of their corresponding TMDs were determined before the first full ABC transporter structures were solved (Oswald et al., 2006), and have provided crucial insight into the architecture of the catalytic site and mechanism of ATP hydrolysis (Smith et al., 2002; Verdon et al., 2003; Zaitseva et al., 2005). However, to fully understand the catalytic mechanism, high resolution structures of the full complexes are indispensable.NBDs can be identified at the sequence level by a specific set of seven highly conserved motifs (Figs. 2 A and 3 C):Open in a separate windowFigure 3.Schematic representation of NBDs and coupling helices. (A) Side view (from the membrane plane) of an ABC transporter. NBDs (blue and green; colors of the domains are as in Fig. 2) are attached to the TMDs (gray) via so-called coupling helices (red) present in loops of the TMDs. ATP binding and hydrolysis cause rearrangements in the NBDs, which are propagated to the TMDs via the coupling helices. (B) Top view (along an axis perpendicular to the membrane) of the NBDs and the coupling helices from the TMDs. (C) The relative positions of sequence motifs in NBDs (see also Fig. 2).(1) The A-loop contains a conserved aromatic residue (usually a tyrosine) that helps to position the ATP via stacking with the adenine ring (Fig. 2 B).(2) The P-loop or Walker A motif (GXXGXGK(S/T)) is a phosphate-binding loop that contains the highly conserved lysine residue. Backbone amide nitrogens and the ε-amino group of this lysine residue form a network of interactions with β- and γ-phosphate of ATP.(3) The Walker B motif (ϕϕϕϕDE, where ϕ is a hydrophobic amino acid) helps to coordinate the magnesium ion via the conserved aspartate residue. The second acidic residue at the end of the Walker B motif (often a glutamate residue) very likely is the general base that polarizes the attacking water molecule. This role of the glutamate residue has long been under debate, but recent crystal structures of the maltose transporter MalEFGK2 from Escherichia coli strongly favor its function as the general base (Oldham and Chen, 2011b).(4) The D-loop (motif: SALD) directly follows the Walker B motif. The D-loops from the two monomers in the dimeric ensemble run alongside each other. Changes in the conformation of the D-loop affect the geometry of the catalytic site and help to form the ATP hydrolysis site.(5) The H-loop (or switch region) contains a highly conserved histidine residue that forms a hinge between a β strand and an α helix near the C terminus of the NBD. The histidine residue interacts with the conserved aspartate from the D-loop, the proposed general base (glutamate residue of the Walker B motif) and with the γ-phosphate of the ATP. It assists with the positioning of the attacking water, the general base, and the magnesium ion.(6) The Q-loop is a stretch of approximately eight residues with a conserved glutamine residue at its N terminus. It is located at the interface between the RecA-like subdomain and the α-helical subdomain. Conformational changes in the Q-loop allow the conserved glutamine residue to move in and out of the active site during the hydrolysis cycle, forming the active site when Mg-ATP is bound and disrupting it once ATP is hydrolyzed. The Q-loop is also a major site of interaction with the TMDs (see below).(7) The ABC signature motif (or C motif, LSGGQ) is found in the α-helical subdomain and is a characteristic feature of the ABC superfamily, not present in other P-loop NTPases such as the F1-ATPase. This LSGGQ motif is located at the N-terminal end of a long helix that directs the positive charge of the helical dipole toward the γ-phosphate of ATP.The two NBDs in ABC transporters can adopt different orientations relative to each other (Fig. 3): they can tightly pack against each other (closed conformation) or partially dissociate (open conformation). There are two ATP-binding sites at the interface between the two monomers, which are related by twofold (pseudo)symmetry. ATP binding promotes the formation of the closed conformation because each ATP molecule interacts with motifs from both NBDs: the ABC signature motif of one monomer is located close to the Walker A and B motifs and the A, H, and Q loop of the other domain (Fig. 2). Only when the monomers are packed against each other can ATP hydrolysis take place. The release of Pi and ADP after ATP hydrolysis destabilizes the dimer and allows the NBDs to move apart. Moreover, during this catalytic cycle, the RecA-like subdomain and the α-helical subdomain within each NBD rotate toward each other when ATP is bound and away from each other after hydrolysis and ADP and Pi release. In this way, the chemical energy of ATP hydrolysis is transformed into conformational energy that can be transmitted to the TMDs to promote alternating access of the substrate–translocation pathway to the two sides of the membrane.Because NBD dimers have two ATP hydrolysis sites, it is tempting to assume that ABC transporters use the hydrolysis of two ATP molecules for a complete transport cycle. Although a 2:1 stoichiometry (ATP to transported substrate) has indeed been determined experimentally for the glycine-betaine importer OpuA from the bacterium Lactococcus lactis (Patzlaff et al., 2003), it cannot be concluded that this stoichiometry is conserved among all ABC transporters, for two reasons.First, it has been very difficult to accurately measure the stoichiometry of transport, because many purified ABC transporters have basal ATPase activity in the absence of the transported substrate (Lewinson et al., 2010). Although the basal activity may be an artifact of purified ABC transporters, it is also possible that some degree of futile ATP hydrolysis takes place in vivo. The degree of futile cycling may differ for different members of the family.Second, some ABC transporters have heterodimeric NBDs that contain only a single complete ATPase site. The second site is degenerate and cannot hydrolyze ATP because of mutation(s) in the conserved motifs (Procko et al., 2009; Jones and George, 2013). Combinations of canonical and degenerate sites are found frequently in ABC exporters (both prokaryotic and eukaryotic) and possibly indicate that the hydrolysis of a single ATP molecule takes place per transport cycle. Additionally, mutagenesis experiments have shown that a single active ATP hydrolysis site may also be sufficient for transporters with homodimeric NBDs and two canonical ATPase sites (in the histidine transporter, HisP2MQJ from E. coli; Nikaido and Ames, 1999). However, this is not a universal property; for instance, in the maltose transporter MalEFGK2, two functional ATPase sites are required (Davidson and Sharma, 1997). These observations suggest that different transport stoichiometries can be found in ABC transporters, indicating the use of multiple mechanisms of transport, even though all NBDs are structurally related. The functional consequences of differences in stoichiometry between different members of the ABC transport superfamily are not clear. Thermodynamically, the coupled hydrolysis of two ATP molecules rather than one per substrate could lead to greater membrane gradients of the transported substrate.A combination of a consensus and a degenerate ATP-binding site in the NBD dimer introduces obvious asymmetry. But even in the presence of two consensus sites, ATP hydrolysis is not likely to be simultaneous at both sites, and thus asymmetry may be a generic feature of ABC transporters (Mittal et al., 2012; Jones and George, 2013).

TMD

In all four ABC transporter types, the TMDs constitute a translocation pathway, which is alternately accessible from the cis-side and trans-side of the membrane to enable the transport of substrate (Fig. 1).The two TMDs of Type I importers are either identical (homodimers) or structurally similar (e.g., the two TMDs of the maltose transporter MalEFGK2 share only 13% sequence identity but are structurally related), with a core membrane topology of five TM helices per TMD (Fig. 4 A). In many cases, an additional N-terminal helix is present that wraps around the helices of the other TMDs and intertwines the TMDs, making a total of 12 TM helices (Fig. 4 A). However, some TMDs contain up to eight TM helices (Fig. 4 A). The translocation pathway is located at the interface between the two TMDs.Open in a separate windowFigure 4.Arrangement of the membrane helices in ABC transporters. Viewpoints are from the outside (trans-side) along an axis perpendicular to the membrane plane. (A) The MalF and MalG subunits of the maltose transporter MalEFGK2, (B) the BtuC dimer of the vitamin B12 transporter BtuC2D2F, (C) the EcfT subunit (yellow) and the S-component from the ECF-HmpT transporter, and (D) the membrane domains of the TM287 and TM288 subunits of the exporter from T. maritima. TM helices are numbered according to their occurrence in the sequence, with the one located most closely to the N terminus numbered as 1. The two coupling helices in the EcfT subunit are labeled X1 and X2.Type II ABC importers have two identical TMDs, each comprised of 10 TM helices (Fig. 4 B). In the Type II fold, the TMDs are lined up next to each other (Fig. 4 B); they do not have helices that cross over to the other TMD. In each TMD, there is a pseudo twofold symmetry between the segments containing TM helices 2–5 and TM helices 7–10. These two subdomains have a similar helical packing but with opposite orientation with respect to the membrane. The helices of a single TMD are tightly packed together, and the two TMDs line a translocation pore at the interface.In ECF-type ABC importers, the two TMDs are structurally and functionally unrelated. One TMD is termed the EcfT subunit (or T-component). In the available crystal structures of ECF transporters, this subunit has five TM helices (Fig. 4 C). However, in other ECF transporters, EcfT subunits are predicted to have four to eight TM helices (Eitinger et al., 2011). The second TMD is termed the S-component and binds the transported substrate with high affinity (Duurkens et al., 2007; Eudes et al., 2008; Erkens and Slotboom, 2010; Berntsson et al., 2012). In contrast to Type I and II importers, which require water-soluble SBPs for high affinity substrate recognition, ECF transporters only need the hydrophobic integral membrane S-component. S-components have a core of six TM helices, but a few S-components have an additional N-terminal helix (Yu et al., 2013). In ECF transporters, the translocation pathway is probably not located at the interface between the TMDs but confined to the S-component, which uses a unique alternating access mechanism (see below).Crystal structures of ABC exporters show a common structural fold consisting of a core six TM helices per TMD (Fig. 4 D). The two TMDs in the dimer may be identical or structurally similar. The 12 TM helices extend a considerable distance into the cytoplasm, with the NBDs located ∼25 Å away from the membrane surface (Fig. 1). This is very different from the importers, where the NBDs are located very close to the membrane. In exporters, the translocation path most likely is located at the interface of the dimeric assembly. In all crystallized ABC exporters, the NBDs are fused to the TMDs, and two helices of each TMD cross over to the other TMD.

Coupling helix

A crucial mechanistic question is how alternating access in the TMDs is coupled to conformational changes in the NBDs when binding and hydrolysis of ATP and release of Pi and ADP take place. So-called coupling helices have been identified in the TMDs of ABC exporters and Type I and II importers (Dawson et al., 2007). A coupling helix is a short α helix in one of the cytoplasmic protrusions of the TMD that fits into a groove of an NBD monomer. In this way, each NBD is connected to a TMD, and the conformational changes in the NBDs can be transduced to conformational changes in the TMDs, leading to alternating access. Some coupling helices contain a conserved sequence (EAA motif) (Mourez et al., 1997), but in most cases, sequence similarity is lacking in the coupling helices. Coupling helices are found between TM helix 3 and 4 in the core of Type I importers. These helices correspond to helices 4 and 5 in MalG (which has an additional N-terminal helix) and helices 6 and 7 in MalF (three additional helices) in Fig. 4 A. In Type II importers, they are located in TM helices 6 and 7, and in ABC exporters, the coupling helix region is found in the intracellular loop (ICL)2 between TM helices 4 and 5. In exporters where the TMDs are fused to the NBDs, the coupling helix of one TMD interacts with the NBD that is linked to the other subunit. Although the arrangements are different in the different types of transporters, all coupling helices interact in a similar way with the NBDs. The region of the NBDs that interacts with the coupling helix of the TMDs contains the Q-loop. The cleft for the coupling helices in the NBDs is located exactly at the interface between the α-helical subdomain and the RecA-like subdomain, which rotate toward each other in response to ATP binding for ATP hydrolysis.In the Type I maltose importer (MalEFGK2), the coupling helices are not the only site of interaction between the TMDs and the NBDs, because the C-terminal segment of one of the TMDs (MalG) is partially inserted between the two NBDs and seems to further order the Q-loop region. Additional interactions are also seen in Type II importers, where BtuC2D2 has a helical segment next to the coupling helix that also interacts with the NBDs. The crystallized exporters contain the most extensive additional interaction area, with a second cytoplasmic coupling helix. This helix is located between TM helices 2 and 3 (ICL1) and interacts directly with the NBD regions that bind the nucleotide adenine ring. Therefore, they shield the nucleotide and the active site of the NBDs from the bulk solvent in the ATP-bound closed state. The NBDs of exporters contain an additional motif (the X-loop: TEVGERG) that interacts with both coupling helices (ICL1 and ICL2) of the TMDs. This motif is located just before the ABC signature motif. Based on this X-loop motif, it was suggested that the coupling mechanism of ATP hydrolysis and transport would occur through a distinct mechanism in exporters (Dawson and Locher, 2006).In the case of ECF transporters, only one of the two TMDs (the EcfT subunit or T-component) interacts with the NBDs. The EcfT subunit contains two long helices in a single cytoplasmic region that interact with the two NBDs at the similar location as the coupling helices in the Type I and II importers and the exporters. The other TMD (the S-component) interacts extensively with EcfT but barely with the NBDs. The asymmetry in the TMD–NBD interaction in ECF transporters most likely leads to a distinct mechanism of transport.

Substrate binding

Type I and II ABC importers.

The SBD (or SBP) is a soluble constituent of ABC transporters that is located on the trans-side of the membrane. SBPs for different substrates display widely varying binding affinities (Quiocho and Ledvina, 1996; Berntsson et al., 2010). Dissociation constants are often in the range of 0.01 to 1 µM (Davidson et al., 2008) but occasionally are much lower or higher (e.g., OpuBC from Bacillus subtilis has an affinity of 30 µM for choline [Pittelkow et al., 2011], MolA from Haemophilus influenzae has an affinity of ∼100 µM for molybdate and tungstate [Tirado-Lee et al., 2011], and TbpA from E. coli has an affinity of 2.3 nM for thiamin [Soriano et al., 2008]). SBPs are either linked in a single polypeptide with the TMD of the transporter, connected to the membrane via a lipid anchor or separate TM helix, or freely diffusible in the periplasm (the latter only in Gram-negative bacteria; Biemans-Oldehinkel et al., 2006). Even though SBPs can vary considerably in sequence and size, and have very different substrate specificities, they share highly conserved general architecture with two domains or lobes that are connected via a hinge region (Fig. 5). Based on their structure, SBPs can be categorized into six different clusters (Berntsson et al., 2010). It is notable that all Type II importers for which crystal structures have been determined make exclusive use of SBPs from cluster A, whereas all Type I importers use SBPs from group D. It is possible that the use of SBPs from different clusters correlates with the use of different TMD folds, and that more TMD folds remain to be discovered.Open in a separate windowFigure 5.Rearrangements in MBP upon the substrate binding. (A) In the substrate-free form (Protein Data Bank accession no. 1ANF), the cavity between two protein lobes connected by the hinge is accessible. (B) Upon the binding of substrate (dark sticks; Protein Data Bank accession no. 1EZ9), the cavity becomes occluded.

Table 2.

Clusters of soluble SBPs
ClusterSubclusterTypes of ligandsMain feature
AIMetal ionsSingle rigid α helix connects the two domains
IISiderophores
BCarbohydrates, Leu, Ile, Val, autoinducer-2Three interconnecting segments between the two domains
CDi- and oligopeptides, Arg, cellubiose, nickelAn additional domain. These SBPs are significantly larger.
DICarbohydratesTwo relative short hinges between the two domains
IIPutrescine, thiamine
IIITetrahedral oxyanions
ESialic acid, 2-keto acids, ectoine, pyroglutamic acidLarge flexible helix in between the two domains. Only associated with TRAP transporters.
FITrigonal planar anions, unknown ligandsTwo hinges in between the two domains as in cluster D, but these hinges are almost twice as long, giving the SBP more flexibility
IIMethionine
IIICompatible solutes
IVAmino acids
Open in a separate windowData taken from Berntsson et al. (2010).The vast amount of structural information on SBPs has provided profound insight in the mechanism of substrate binding. In the absence of a ligand, the two lobes exist predominantly in an open conformation, but upon substrate binding, they close to trap the ligand (the “Venus fly trap” model [Quiocho and Ledvina, 1996]; Fig. 5) and eventually deliver it to the TMDs. SBPs interact with both TMDs, with each lobe interacting with one of the TMDs.Some of the Type I and II ABC importers can transport more than one substrate, which is possible either because their SBPs can recognize various substrates (e.g., the SBP of the multi-sugar transporter Msm of Streptococcus mutans can recognize melibiose, sucrose, raffinose, isomaltotriose, and isomaltotetraose [Russell et al., 1992]), or because the transporter can interact with various SBPs. Examples of the latter are the His/Lys/Arg transport system in Enterobacteriaceae (Higgins and Ames, 1981), the peptide transporter OppBCDF from Enterococcus faecalis (Leonard et al., 1996), and the oligopeptide/muramyl peptide transport system of E. coli (Park et al., 1998). However, the interaction of different SBPs with the same translocator is relatively rare.In the Type I maltose transporter MalEFGK2, one of the TMDs (MalF) contains a second binding site for maltose (in addition to the binding site in the SBP MalE), located along the translocation path at the center of the bilayer (Oldham et al., 2007). It is likely that substrate moves from the SBP to the central membrane-embedded site during the translocation cycle. To date, this is the only crystal structure where an additional binding site has been identified, but this may be a feature of more Type I importers. In the histidine transporter (HisMQP2-HisJ/LAO), mutations in the NBDs (HisP2) lead to transport of histidine in the absence of the SBP (HisJ/LAO), indicating the presence of a second binding site (Speiser and Ames, 1991). No crystal structure of the His transporter is available, but its TMDs (HisM and HisQ) are predicted to have five TM helices. Therefore it is probably a Type I transporter.No specific binding sites have been found in the translocation pathway between the TMDs of the Type II importers. It has been hypothesized that the translocation pathways of the Type II ABC importers are inert, “Teflon”-like, with little or no affinity for substrate (Korkhov et al., 2012b). This would mean that the specificity of these transporters depends entirely on the SBP.In some Type I transporters, binding sites for the transported substrates are present in additional domains connected to the NBDs. These binding sites are not required for transport but have regulatory functions (Gerber et al., 2008; Kadaba et al., 2008). When cytosolic concentrations of the transported substrate are high, and no further transport is needed, substrate binding to the regulatory sites keeps the NBDs dissociated from each other, in an inhibited state. This type of regulation has been named “trans-inhibition,” a name that may appear confusing because the regulatory site is located on the cis-side of the membrane.

ECF-type ABC importers.

In ECF-type ABC importers, one of the TMDs (the S-component) binds the substrate without the need of an SBP. S-components have very high affinities for their substrates, with dissociation constants in the low or subnanomolar range (Duurkens et al., 2007; Eudes et al., 2008; Erkens and Slotboom, 2010; Berntsson et al., 2012). To date, four crystal structures of individual S-components with bound substrates (RibU, ThiT, BioY, and NikM for uptake of riboflavin, thiamine, biotin, and Ni2+, respectively; Zhang et al., 2010; Erkens et al., 2011; Berntsson et al., 2012; Yu et al., 2013). Access to the binding site from the external solution requires movement of the flexible loop between TM helices 1 and 2, which serves as a lid for the binding site (Fig. 6) (Majsnerowska et al., 2013).Open in a separate windowFigure 6.Structural changes in the S-component of ECF transporters (substrate-binding TM protein of ECF transporters). The loop between TM helices 1 and 2 is highlighted in red. (A) S-component from the ECF–HmpT complex (Protein Data Bank accession no. 4HZU) in the substrate-free form. The shape of the binding pocket is visible, indicated by the arrow. (B) Thiamin-specific ThiT (Protein Data Bank accession no. 3RLB) in the substrate-bound state. The loop between TM helices 1 and 2 closes the binding pocket. The substrate thiamin is shown as sticks.It must be noted though that it is difficult to unambiguously determine the membrane orientation of the S-components in the absence of the NBDs and the EcfT subunit based solely on the crystal structures of these proteins in detergent solution. However, the predicted orientation is supported by molecular dynamics simulations of the solitary S-components, which are consistent with the interpretation that the substrate-binding site is located close to the extracellular side of the membrane (Majsnerowska et al., 2013; Song et al., 2013).No substrates were found bound to the S-components in the structures of the complete (four-subunit) ECF transporters (Wang et al., 2013; Xu et al., 2013). Surprisingly, in the context of the complete complexes, the empty binding sites of the S-components are located close to the cytoplasmic side of the membrane, because the S-components have toppled over compared with the (predicted) membrane orientation of the solitary S-components. The toppling seems to be essential for the alternating access mechanism in ECF transporters.

ABC exporters.

Based on the crystal structures, the substrate-binding sites in ABC exporters are poorly defined compared with those of ABC importers. A possible cause is that some exporters are multidrug transporters, which bind many structurally different substrates, and are unlikely to have a single well-defined binding site (Ambudkar et al., 2003). In addition, the location of the binding site may not be conserved among the different members of the exporter family, because some exporters bind the substrate from the cis-side and others take up the substrate from the lipid bilayer.There are no crystal structures available of exporters in complex with transported substrates, but in P-glycoprotein of mouse, the binding sites for two inhibitors have been located (Aller et al., 2009; Li et al., 2014). These binding sites probably overlap partially with substrate-binding sites. The inhibitor-binding sites are located in the membrane-spanning part of the TMDs and are lined with hydrophobic and aromatic residues (in accordance with the fact that P-glycoprotein substrates are hydrophobic). It must be noted that the original crystal structures of mouse P-glycoprotein contained errors, which had to be corrected (Li et al., 2014).

Mechanistic diversity

The structural diversity in the ABC transporter superfamily suggests differences in transport mechanisms. Below, we will briefly discuss some mechanistic features of a Type I importer (the maltose importer MalEFGK2), a Type II importer (the cobalamin importer BtuC2D2F), ECF transporters (represented by ECF-FolT and ECF-HmpT, and the S-components ThiT, RibU, BioY, and NikM), and exporters (represented by two bacterial proteins, Sav18662 and TM288/287). This selection of ABC transporters is based on the availability of high quality crystal structures and biochemical data, but it is unlikely to cover the entire mechanistic diversity of the ABC transporter superfamily. Notably, it is possible that there are also mechanistic differences between members of the ABC transporter family that share the same fold. Because many separate reviews and overviews have been published over the past few years on the mechanistic details of each of the selected transporters (for instance, Davidson et al., 2008; Rees et al., 2009; Chen, 2013; Zhang, 2013; Slotboom, 2014), we will only discuss the gross differences between the four types.

Type I ABC importers.

The maltose transporter MalEFGK2 from E. coli is one of the best-characterized Type I importers. MalF and MalG are the TMDs, MalK2 is the homodimer of NBDs, and MalE is the periplasmic maltose-binding protein (MBP). The protein has been captured in crystals in several states (Fig. 7), allowing deduction of a tentative mechanism of transport (Chen, 2013).Open in a separate windowFigure 7.The transport mechanism of Type I importers (exemplified by MalEFGK2) based on the available structures (A) and in schematic representation (B). Coloring is as in Fig. 1. Structures have been determined for the inward-facing, pre-translocation, and outward-facing conformations (Protein Data Bank accession nos.: 4JBW, 4KHZ, and 4KI0; see Davidson et al., 1992; Chen et al., 2001; Lu et al., 2005; Grote et al., 2008, 2009; Orelle et al., 2008, 2010; Bordignon et al., 2010; Jacso et al., 2012; Böhm et al., 2013; Chen, 2013), but further refinement of the model may be needed to explain details. It must be noted that Bao and Duong (2013, 2012) has proposed a radically different mechanistic model, which cannot be reconciled with the mechanism described above.

Type II ABC importers.

The vitamin B12 transporter BtuC2D2F from E. coli is the best-characterized Type II importer, with crystal structures available in three states (for a recent overview of the mechanism see Korkhov et al., 2012b). The NBDs (BtuC subunits), TMDs (BtuD subunits), and the periplasmic SBP (BtuF) adopt different conformations depending on whether the transported substrate and nucleotides are present (Fig. 8).Open in a separate windowFigure 8.The transport mechanism of Type II importers (exemplified by the BtuC2D2F transporter) based on the available structures (A) and in schematic representation (B). Coloring is as in Fig. 1. Structures have been determined for an outward-open, occluded nucleotide-bound, and closed ATP-free asymmetric transporter (Protein Data Bank accession nos.: 1L7V, 4DBL, and 2QI9; see Lewinson et al., 2010). It is likely that ATP binding is required to release the BtuF subunit.The tight binding of the SBP is characteristic for Type II transporters. For the maltose (MalFGK2) and other ABC Type I importers, the SBPs have a low affinity for the TMDs of the transporter and only interact with the transporter transiently (Lewinson et al., 2010; Vigonsky et al., 2013).

Type III or ECF-type importers.

These have recently been reviewed in Eitinger et al. (2011), Erkens et al. (2012), Zhang (2013), and Slotboom (2014). Two crystal structures are available for complete ECF transporter complexes: ECF–FolT (specific for folate transport) and ECF–HmpT (predicted to transport pyridoxin) from the bacterium Lactobacillus brevis. One of the TMDs (the EcfT subunit) and the two NBDs (the EcfA and EcfA’ subunits) form a module that is identical in the two complexes and can associate with different S-components (FolT and HmpT, and five more S-components present in the bacterium; Rodionov et al., 2009; Slotboom, 2014). The modular nature (different S-components interacting with the same EcfAA’T complex) is a characteristic feature of many ECF transporters. Both crystallized complexes are in the same conformational state, with neither nucleotides nor transported substrates bound.The orientation of the S-components FolT and HmpT in the complexes is highly unusual. Crystal structures of the S-components in the absence of the EcfAA’T module (Zhang et al., 2010; Erkens et al., 2011; Berntsson et al., 2012; Yu et al., 2013) indicated that the N and C termini of the proteins are exposed to the cytosol, and that the substrate-binding site is located close to the extracellular (or cytoplasmic) space (see above and Fig. 6). Surprisingly, the FolT and HmpT subunits in the complexes have “toppled over” by almost 90°, and helices 1–4 lie parallel to the membrane plane, an unprecedented orientation for membrane proteins. In the crystallized ECF–FolT and ECF–HmpT complexes, the binding sites in the S-components are empty and accessible from the cytoplasm. Thus, it appears that the toppling of the S-components is mechanistically important, as it physically moves the bound substrate from the outside to the cytoplasm (Fig. 9).Open in a separate windowFigure 9.Possible transport mechanism of ECF transporters (Type III importers) based on the available structures (A) and in schematic representation (B). Structures have been determined of the nucleotide- and substrate-free transporter (ECF-HmpT, Protein Data Bank accession no. 4HZU, is shown here; see Henderson et al., 1979; Rodionov et al., 2009; Slotboom, 2014). The solitary S-components may then spontaneously assume an orientation with the substrate-binding site located on the trans-side of the membrane.

ABC exporters

In all crystallized ABC exporters, the NBDs are linked to the TMDs in a single protein. Exporters have been captured in two different states: the outward-facing conformation, represented by Sav18662 from Staphylococcus aureus, which is a homodimer (Dawson and Locher, 2006), and the inward-facing conformation, represented by TM287/288 from Thermotoga maritima, which is a heterodimer (Fig. 10) (Hohl et al., 2012). It must be noted that a few other crystal structure of ABC exporters (MsbA and mouse P-glycoprotein) initially contained large errors and had to be corrected (Ward et al., 2007; Li et al., 2014).Open in a separate windowFigure 10.The transport mechanism of exporters (exemplified by the structures of Sav18662 and TM287/288) based on the available structures (A) and in schematic representation (B). Coloring is as in Fig. 1. Structures of outward-open and nucleotide-bound inward-facing transporter are shown (Protein Data Bank accession nos. 2HYD and 3QF4; see Procko et al., 2009; Hohl et al., 2012). In such exporters, one site is referred to as the consensus site (and contains all the conserved motifs usually found in NBDs; see above), and the second is termed degenerate (because of the deviations from the consensus sequence). The heterodimeric ABC exporter TM287/288 from T. maritima (crystallized in the inward conformation) is one of the transporters with a degenerate and a consensus site. In the reported structure, the two NBDs are not fully disengaged but still interact with each other. A bound nucleotide (AMP-PNP) was found only in the degenerate site, and its presence may prevent further dissociation of the NBDs. Based on analysis of interactions, Hohl et al. (2012) suggest that ATP hydrolysis is blocked at the degenerate site as a result of the increased distance to γ-phosphate caused by replacement of the glutamate from Walker B motif to aspartate. Whether the ABC exporters that have two consensus sites are mechanistically different from the ones with a degenerate site remains to be elucidated.

Concluding remarks

The recent surge in crystal structures of ABC transporters has revealed a remarkable structural diversity and suggests unanticipated mechanistic diversity in the superfamily. Crystal structures alone obviously are not sufficient to elucidate transport mechanisms. They provide snapshots of “states.” The number of states that can be crystallized may not be enough to describe the complete transport cycle. Moreover, what a state represents, and how such a state relates to physiological conditions, is often loosely defined. For example, the name “resting state” has been used to describe the nucleotide-free transporters, but it is unlikely that nucleotide-free conditions are physiologically relevant. In addition, the names “high energy state” and “intermediate state” are sometimes used without solid definition.Because crystal structures are determined in detergent solution, an environment very different than that of the lipid bilayer, it is essential that models based on crystal structures be tested in membrane environments. Spectroscopic techniques such as electron paramagnetic resonance and (single-molecule) FRET are suitable to obtain structural and dynamic information in lipid bilayers (Erkens et al., 2013; Hänelt et al., 2013; Majsnerowska et al., 2013). Such techniques may reveal states that cannot be captured in crystals (Böhm et al., 2013). In addition, classical biochemical experiments in liposomes can be used to test mechanistic models.Finally, a few high profile examples of crystal structures with huge errors have made many non-crystallographers skeptical about the validity and relevance of any crystal structure (Chang et al., 2006). Although such errors have damaged the reputation of crystallography, it is also important to note that other crystallographers have detected these errors, leading to corrections or retractions. Arguably, crystal structures are some of the most scrutinized biochemical data, which—eventually—warrants high standards. The currently available high quality structures of ABC transporters underscore the powerful role of protein crystallography to provide mechanistic insight. Therefore, crystal structures of new states of the four known ABC transporter types, as well as structures of novel ABC transporter folds, are very welcome, and expected to further advance our understanding of this large, important, and fascinating superfamily of membrane transporters.  相似文献   

13.
The ABC transporters and the thickening cholesterol plot     
Getz GS  Reardon CA 《Current opinion in lipidology》2011,22(1):72-73
  相似文献   

14.
The ABC transporters in lipid flux and atherosclerosis     
Voloshyna I  Reiss AB 《Progress in lipid research》2011,50(3):213-224
Atherosclerotic cardiovascular disease is the leading cause of morbidity and mortality in the United States and in many other countries. Dysfunctional lipid homeostasis plays a central role in the initiation and progression of atherosclerotic lesions. The ATP-binding cassette (ABC) transporters are transmembrane proteins that hydrolyze ATP and use the energy to drive the transport of various molecules across cell membranes. Several ABC transporters play a pivotal role in lipid trafficking. They are critically involved in cholesterol and phospholipid efflux and reverse cholesterol transport (RCT), processes that maintain cellular cholesterol homeostasis and protect arteries from atherosclerosis. In this article we provide a review of the current literature on the biogenesis of ABC transporters and highlight their proposed functions in atheroprotection.  相似文献   

15.
The Escherichia coli ABC transporters: an update     
Elie Dassa  Maurice Hofnung  Ian T. Paulsen  & Milton H. Saier Jr 《Molecular microbiology》1999,32(4):887-889
  相似文献   

16.
ABC transporters in lipid transport   总被引:9,自引:0,他引:9  
Borst P  Zelcer N  van Helvoort A 《Biochimica et biophysica acta》2000,1486(1):128-144
Since it was found that the P-glycoproteins encoded by the MDR3 (MDR2) gene in humans and the Mdr2 gene in mice are primarily phosphatidylcholine translocators, there has been increasing interest in the possibility that other ATP binding cassette (ABC) transporters are involved in lipid transport. The evidence reviewed here shows that the MDR1 P-glycoprotein and the multidrug resistance (-associated) transporter 1 (MRP1) are able to transport lipid analogues, but probably not major natural membrane lipids. Both transporters can transport a wide range of hydrophobic drugs and may see lipid analogues as just another drug. The MDR3 gene probably arose in evolution from a drug-transporting P-glycoprotein gene. Recent work has shown that the phosphatidylcholine translocator has retained significant drug transport activity and that this transport is inhibited by inhibitors of drug-transporting P-glycoproteins. Whether the phosphatidylcholine translocator also functions as a transporter of some drugs in vivo remains to be seen. Three other ABC transporters were recently shown to be involved in lipid transport: ABCR, also called Rim protein, was shown to be defective in Stargardt's macular dystrophy; this protein probably transports a complex of retinaldehyde and phosphatidylethanolamine in the retina of the eye. ABC1 was shown to be essential for the exit of cholesterol from cells and is probably a cholesterol transporter. A third example, the ABC transporter involved in the import of long-chain fatty acids into peroxisomes, is discussed in the chapter by Hettema and Tabak in this volume.  相似文献   

17.
The pleitropic drug ABC transporters from Saccharomyces cerevisiae     
Rogers B  Decottignies A  Kolaczkowski M  Carvajal E  Balzi E  Goffeau A 《Journal of molecular microbiology and biotechnology》2001,3(2):207-214
  相似文献   

18.
A fourth gene from the Candida albicans CDR family of ABC transporters     
Renate Franz  Sonja Michel  Joachim Morschh  user 《Gene》1998,220(1-2):91-98
Using primers derived from a region of the Candida albicans CDR1 (Candida drug resistance) gene that is conserved in other ABC (ATP-binding cassette) transporters, a DNA fragment from a previously unknown CDR gene was obtained by polymerase chain reaction (PCR). After screening a C. albicans genomic library with this fragment as a probe, the complete CDR4 gene was isolated and sequenced. CDR4 codes for a putative ABC transporter of 1490 amino acids with a high degree of homology to Cdr1p, Cdr2p and Cdr3p from C. albicans (62, 59 and 57% amino acid sequence identity, respectively). Cdr4p has a predicted structure typical for cluster I.1 of yeast ABC transporters, characterized by two homologous halves, each comprising an N-terminal hydrophilic domain with consensus sequences for ATP binding and a C-terminal hydrophobic domain with six transmembrane helices. In contrast to the CDR1/CDR2 genes, the genetic structure of the CDR4 gene was conserved in 59 C. albicans isolates from six different patients. Northern hybridization analysis showed that the CDR4 gene was expressed in most isolates, but no correlation between CDR4 mRNA levels and the degree of fluconazole resistance of the isolates was found. In addition, a C. albicans mutant in which both copies of the CDR4 gene were disrupted by insertional mutagenesis was not hypersusceptible to fluconazole as compared to the parent strain. Unlike CDR1 and CDR2, CDR4 does not, therefore, seem to be involved in fluconazole resistance in C. albicans.  相似文献   

19.
The controversial role of ABC transporters in clinical oncology     
Tamaki A  Ierano C  Szakacs G  Robey RW  Bates SE 《Essays in biochemistry》2011,50(1):209-232
The phenomenon of multidrug resistance in cancer is often associated with the overexpression of the ABC (ATP-binding cassette) transporters Pgp (P-glycoprotein) (ABCB1), MRP1 (multidrug resistance-associated protein 1) (ABCC1) and ABCG2 [BCRP (breast cancer resistance protein)]. Since the discovery of Pgp over 35 years ago, studies have convincingly linked ABC transporter expression to poor outcome in several cancer types, leading to the development of transporter inhibitors. Three generations of inhibitors later, we are still no closer to validating the 'Pgp hypothesis', the idea that increased chemotherapy efficacy can be achieved by inhibition of transporter-mediated efflux. In this chapter, we highlight the difficulties and past failures encountered in the development of clinical inhibitors of ABC transporters. We discuss the challenges that remain in our effort to exploit decades of work on ABC transporters in oncology. In learning from past mistakes, it is hoped that ABC transporters can be developed as targets for clinical intervention.  相似文献   

20.
The role of ABC transporters in cuticular lipid secretion     
David A. Bird   《Plant science》2008,174(6):563-569
The aerial surfaces of plants are enveloped by a waxy cuticle, which among other functions serves as a barrier to limit non-stomatal water loss and defend against pathogens. The cuticle is a complex three-dimensional structure composed of cutin (a lipid polyester matrix) and waxes (very long chain fatty acid derivatives), which are embedded within and layered on top of the cutin matrix. Biosynthesis of cuticular lipids is believed to take place solely within aerial epidermal cells. Once synthesized, both the waxes and the cutin precursors must leave the cytoplasm, pass through the hydrophilic apoplastic space, and finally assemble to form the cuticle. These processes of secretion and assembly are essentially unknown. Initial steps toward our understanding of these processes were the characterization of CER5/ABCG12/WBC12 and more recently ABCG11/WBC11, a pair of ABC transporters required for cuticular lipid secretion. ABCG12 is involved in wax secretion, as mutations in this gene result in a lower surface-load of wax and a concomitant accumulation of lipidic inclusions within the epidermal cell cytoplasm. Mutations in ABCG11 result in a similar wax phenotype as cer5 and similar cytoplasmic inclusions. In contrast to cer5, however, abcg11 mutants also show significantly reduced cutin, post-genital organ fusions, and reduced growth and fertility. Thus, for the first time, a transporter is implicated in cutin accumulation. This review will discuss the secretion of cuticular lipids, focusing on ABCG12, ABCG11 and the potential involvement of other ABC transporters in the ABCG subfamily.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号