Skip to main content
Human Genomics logoLink to Human Genomics
. 2009 Apr 1;3(3):281–290. doi: 10.1186/1479-7364-3-3-281

Human ATP-binding cassette (ABC) transporter family

Vasilis Vasiliou 1, Konstandinos Vasiliou 1, Daniel W Nebert 2,
PMCID: PMC2752038  NIHMSID: NIHMS134146  PMID: 19403462

Abstract

There exist four fundamentally different classes of membrane-bound transport proteins: ion channels; transporters; aquaporins; and ATP-powered pumps. ATP-binding cassette (ABC) transporters are an example of ATP-dependent pumps. ABC transporters are ubiquitous membrane-bound proteins, present in all prokaryotes, as well as plants, fungi, yeast and animals. These pumps can move substrates in (influx) or out (efflux) of cells. In mammals, ABC transporters are expressed predominantly in the liver, intestine, blood-brain barrier, blood-testis barrier, placenta and kidney. ABC proteins transport a number of endogenous substrates, including inorganic anions, metal ions, peptides, amino acids, sugars and a large number of hydrophobic compounds and metabolites across the plasma membrane, and also across intracellular membranes. The human genome contains 49 ABC genes, arranged in eight subfamilies and named via divergent evolution. That ABC genes are important is underscored by the fact that mutations in at least I I of these genes are already known to cause severe inherited diseases (eg cystic fibrosis and X-linked adrenoleukodystrophy [X-ALD]). ABC transporters also participate in the movement of most drugs and their metabolites across cell surface and cellular organelle membranes; thus, defects in these genes can be important in terms of cancer therapy, pharmacokinetics and innumerable pharmacogenetic disorders.

Keywords: human genome, human ATP-binding cassette (ABC) transporter gene family, genetic polymorphism, evolution, drug transport, cancer chemotherapy

Introduction

Membrane transport proteins can be divided into four types: ion channels; transporters; aquaporins; and ATP-powered pumps http://www.ncbi.nlm.nih.gov/books/bv.fcgi?rid=mcb.figgrp.4031. Genes from all four categories are ancient -- with members present in most, if not all, prokaryotes, as well as in virtually all cell types of all eukaryotes. Transporters in eukaryotic cells move ions, sugars, amino acids and other molecules across all cellular and organelle membranes (cell surface, mitochondrial, endoplasmic reticulum, Golgi apparatus and other vesicles) -- with the possible exception of nuclear membranes (which have pores). The portion of the cell exposed to the lumen is called the apical surface; the rest of the cell (ie sides and base) makes up the basolateral surface. Movement of ions or other molecules into the cell is called influx; movement of ions or other molecules out of the cell is termed efflux.

Membrane transport proteins can be either passive or active. Passive transporters (also called uni-porters or facilitative transporters) transport substrates down a concentration gradient. By contrast, active transporters (or cotransporters) couple the movement of one type of ion or molecule against its concentration gradient, to the movement of another ion or molecule down its concentration gradient. Like ATP pumps, cotransporters mediate coupled reactions in which an energetically unfavourable reaction is coupled to an energetically favourable reaction. When the transported molecule and cotransported ion move in the same direction across a membrane, the transporter is called a symporter; when they move in opposite directions, the transporter is called an antiporter (or exchanger). If the intracellular net charge following transport becomes more negative, the process is termed electronegative; if the intracellular net charge becomes more positive, the process is called electropositive; if the resulting intracellular net charge remains unchanged, the process is termed electroneutral.

Ion channels are pore-forming membrane proteins that help to establish and maintain small-voltage gradients across plasma membrane surfaces of all living cells. As such, they regulate the cell's electric potential by allowing the flow of ions down their electrochemical gradient. Ion channels usually occur in the closed state. Cationic and anionic substrates are transferred down their electrochemical gradients at extremely high efficiencies (as much as 108 sec-1). More than 400 genes are known to encode ion channel subunits [1].

Transporters facilitate the movement of a specific substrate -- either with or against its concentration gradient -- and the conformational change in the transporter protein is important in this transfer process. Transporters move molecules at only about 102-104 sec-1, a rate much slower than that associated with channel proteins. Many of these transporters belong to the solute-carrier (SLC) gene superfamily, and include passive transporters, sym-porters and antiporters, as well as mitochondrial and vesicular transporters. The SLC superfamily comprises 55 gene families, having at least 362 putatively functional protein-coding genes [2].

Aquaporins are a unique class of transporter. These proteins are bi-directional membrane channels which transport water, but they are not ion channels because the H2O is transported as an uncharged molecule and not as an ion. The driving force for aquaporins is the presence of osmotic gradients across membranes [3]. There are 13 putatively functional AQP genes in the human genome http://www.genenames.org/.

ATP-powered pumps include the ATP-binding cassette (ABC) pumps. These pumps use energy released by ATP hydrolysis to move substrates across membranes in or out of cells or into cellular vesicles, against their electrochemical gradient. ABC pumps constitute a large, diverse and ubiquitous superfamily. Most ABC genes encode membrane-bound proteins that participate directly in the transport of a wide range of molecules across membranes [4]. ABC transporters can broadly be categorised as importers or exporters, depending on the direction of transport relative to the cytoplasm [5]. Plants carry a particularly large complement of ABC proteins; these pumps are associated with the need to establish steep concentration gradients of solutes across cellular membranes, as well as metabolic versatility. Plant ABC pumps are not only involved in the transport of hormones, lipids, secondary metabolites, metals and xenobiotics [4], but also contribute to osmolality and ion channel and phytoalexin functions (plant-pathogen interactions). Even the evolution of seed size in the tomato has been associated with an ABC transporter gene [6].

Details of the ABC proteins

By definition, ABC proteins possess an ATP binding cassette, also known as the 'nucleotide-binding domain' (NBD). The NBD contains several highly conserved motifs, including the Walker A and Walker B sequences, the ABC signature motif, the H loop and the Q loop. ABC transporters also contain trans-membrane domains (TMDs), each of which comprises several hydrophobic α-helices. The ABC transporter core unit consists of four domains, two NBDs and two TMDs. The two NBDs together bind and hydrolyse ATP (thereby providing the driving force for transport), while the TMDs participate in substrate recognition and translocation across the lipid membrane [4]. Some ABC genes encode proteins that are 'half-transporters' (meaning that two subunits bind as homodimers or a heterodimer), whereas others are 'full-transporters'.

The human ABC gene family

The human genome carries 49 ABC genes, arranged in seven subfamilies, designated A to G (Figure 1 and Table 1). This diverse transporter family has members that play pivotal roles in many cellular processes. For example, ABC transporters are responsible for the multidrug resistance of cancer cells [7]. ABC proteins also transport a number of substrates, including metal ions, peptides, amino acids, sugars and a large number of hydrophobic compounds and metabolites across the plasma membrane, and also intracellular membranes. The function of each ABC gene product, when known, is listed in the far right column of Table 2. At the present time, 21 ABC pseudogenes have been identified [8] and localised to chromosomal regions (Table 3).

Figure 1.

Figure 1

Clustering dendrogram of the human ATP-binding cassette (ABC) transporter genes. The root 'ABC' is omitted from the figure to simplify it. Thus, the correct gene name for 'Bl' is ABCBI, for 'B4' is ABCB4, and so forth.

Table 1.

Human ABC gene subfamilies

Subfamily name Aliases Number of genes Number of pseudogenes
ABCA ABC1 12 5

ABCB MDR 11 4

ABCC MRP 13 2

ABCD ALD 4 4

ABCE OABP 1 2

ABCF GGN20 3 2

ABCG White 5 2

Total 49 21

Table 2.

Human ABC transporter genes, and their functions, as listed in the HGNC database

Gene Chromosome location Exons AA Accession number Function
ABCA1 9q3l.l 36 2261 NM005502 Cholesterol efflux onto HDL

ABCA2 9q34 27 2436 NM001606 Drug resistance

ABCA3 16pl3.3 26 1704 NM001089 Multidrug resistance

ABCA4 1p22 38 2273 NM000350 N-retinylidene-phosphatidylethanolamine (PE) efflux

ABCA5 17q24.3 31 1642 NM018672 Urinary diagnostic marker for prostatic intraepithelial neoplasia (PIN)

ABCA6 17q24.3 35 1617 NM080284 Multidrug resistance

ABCA7 19p13.3 31 2146 NM019112 Cholesterol efflux

ABCA8 17q24 31 1581 NM007168 Transports certain lipophilic drugs

ABCA9 17q24.2 31 1624 NM080283 Might play a role in monocyte differentiation and macrophage lipid homeostasis


ABCA10 l7q24 27 1543 NM080282 Cholesterol-responsive gene

ABCA12 2q34 37 2595 NM173076 Has implications for prenatal diagnosis

ABCA13 7p12.3 36 5058 NM152701 Inherited disorder affecting the pancreas


ABCB1 7q21.1 20 1280 NM000927 Multidrug resistance

ABCB2 (TAPI) 6p21.3 11 808 NM000593 Peptide transport

ABCB3 (TAP2) 6p21.3 11 703 NM000544 Peptide transport

ABCB4 7q21.1 25 1279 NM000443 Phosphatidylcholine (PC) transport

ABCB5 7p15.3 17 812 NM178559 Melanogenesis

ABCB6 2q36 19 842 NM005689 Iron transport

ABCB7 Xq12-q13 14 753 NM004299 Fe/S cluster transport

ABCB8 7q36 15 718 NM007188 Intracellular peptide trafficking across membranes

ABCB9 12q24 12 766 NM019625 Located in lysosomes

ABCB10 1q42.13 13 738 NM012089 Export of peptides derived from proteolysis of inner-membrane proteins

ABCB11 2q24 26 1321 NM003742 Bile salt transport

ABCC1 16p13.1 31 1531 NM004996 Drug resistance

ABCC2 10q24 26 1545 NM000392 Organic anion efflux

ABCC3 17q22 19 1527 NM003786 Drug resistance

ABCC4 13q32 19 1325 NM005845 Nucleoside transport

ABCC5 3q27 25 1437 NM005688 Nucleoside transport

ABCC6 16p13.1 28 1503 NM001171 Expressed primarily in liver and kidney

ABCC7 (CFTR) 7q31.2 23 1480 NM000492 Chloride ion channel (same as CFTR gene in cystic fibrosis)

ABCC8 11p15.1 30 1581 NM000352 Sulfonylurea receptor

ABCC9 12p12.1 32 1549 NM005691 Encodes the regulatory SUR2A subunit of the cardiac K+(ATP) channel

ABCC10 6p21.1 19 1464 NM033450 Multidrug resistance

ABCC11 16q12.1 25 1382 NM033151 Drug resistance in breast cancer

ABCC12 16q12.1 25 1359 NM033226 Multidrug resistance

ABCC13 21q11.2 6 325 NM00387 Encodes a polypeptide of unknown function


ABCD1 Xq28 9 745 NM000033 Very-long-chain fatty acid (VLCFA) transport

ABCD2 12q11-q12 10 740 NM005164 Major modifier locus for clinical diversity in X-linked ALD (X-ALD)

ABCD3 1p22-p21 16 659 NM002858 Involved in import of fatty acids and/or fatty acyl-coenzyme As into the peroxisome

ABCD4 14q24 19 606 NM005050 May modify the ALD phenotype

ABCE1 4q31 14 599 NM002940 Oligoadenylate-binding protein


ABCF1 6p21.33 19 845 NM001025091 Susceptibility to autoimmune pancreatitis

ABCF2 7q36 14 634 NM005692 Tumour suppression at metastatic sites and in endocrine pathway for breast cancer/drug resistance

ABCF3 3q27.1 21 709 NM018358 Also present in promastigotes (one of five forms in the life cycle of trypanosomes)

ABCG1 21q22.3 13 678 NM004915 Cholesterol transport

ABCG2 4q22 16 655 NM004827 Toxicant efflux, drug resistance

ABCG4 q23.3 15 646 NM022 69 Found in macrophage, eye, brain and spleen

ABCG5 2p2 11 65 NM022436 Sterol transport

ABCG8 2p2 10 673 NM022437 Sterol transport

Abbreviations: HGNC, HUGO Gene Nomenclature Committee; AA, number of amino acids; HDL, high density lipoprotein; CFTR, cysticfibrosis transmembrane conductance regulator gene; ATP adenosine triphosphate; ALD, adrenoleukodystrophy.

Table 3.

Human ABC transporter pseudogenes (adapted and modified from Piehler et al. [8])

Parental gene Pseudogene Chromosomal location Accession number
ABCA3 ABCA17P 16p13.3 DQ266102

ABCA10 ABCA10P 4p16.3 AK024359

Abca14* ABCA14P1 16p12.2

Abca15* ABCA15P1 16p12.2 DR731461
ABCA15P2 16p12.1

ABCB4 ABCB4P 4q32.1

ABCB10 ABCB10P1 15q11.2
ABCB10P2 15q13.1
ABCB10P3 15q13.1

ABCC6 ABCC6P1 16p12.3 DB11925
ABCC6P2 16p13.11

ABCD1 ABCD1P1 2p11.1 AY344117
ABCDIP2 10p11.1
ABCDIP3 16p11.2
ABCDIP4 22q11.1

ABCE1 ABCE1P1 Iq31.2
ABCEIP2 7p15.3

ABCF2 ABCF2P1 3p11.2
ABCF2P2 7q11.2

ABCG2 ABCG2P1 14q24.3
ABCG2P2 15q23

*Abca14, Abca15 and Abca16 are mouse genes, with no human orthologues [9]. The mouse genome contains 52 Abc genes, whereas the human genome carries 49 ABC genes.

Mutations in at least 11 ABC genes to date are clearly associated with inherited diseases:[10] Tangier disease T1 (ABCA1); Stargardt disease, retinitis pigmentosa and age-related macular degeneration (ABCA4); progressive familial intrahepatic cholestasis (ABCB11); Dubin-Johnson syndrome (ABCC2); pseudoxanthoma elasticum (ABCC6); cystic fibrosis (ABCC7); X-linked adrenoleukodystrophy (ALD) (ABCD1 and ABCD2); some forms of Zellweger syndrome (ABCD3 and ABCD2) and sitosterolaemia (a rare lipid metabolic disorder inherited as an auto-somal recessive trait) (ABDG5 and ABCG8). An additional eight ABC genes have been implicated in, or are candidates for, other metabolic inherited diseases http://nutrigene.4t.com:0/humanabc.htm.

Clans: Comments on sequence identity

The ABC family is a member of the P-loop-containing nucleoside-triphosphate hydrolase clan (CL0023). The definition of the clan in the Pfam database is: 'A collection of evolutionarily-related Pfam entries. This relationship may be defined by similarity of protein sequence, tertiary structure or profile, as defined by the Hidden Markov model' [11,12]. At the moment, clan CL0023 contains 55 protein families, including the ABC family.

Subfamily A of the ABC family (ABCA)

Subfamily A contains 12 genes, most of which appear to be involved in lipid trafficking in many diverse organs and cell types. Among the largest of the ABC transporters, some of the ABCA proteins weigh in at more than 2,100 amino acids in length [10]. In fact, the predicted ABCA13 protein contains 5,058 residues, making it the largest ABC protein known. Mutations in specific ABCA genes lead to genetic disorders [13], such as Tangier disease T1, familial high-density lipoprotein (HDL) deficiency, Stargardt disease-1, age-related macular degeneration and retinitis pigmentosa (Table 2).

Subfamily B of the ABC family (ABCB)

This subfamily of 11 genes is unique to mammals; there are four full-transporters and seven half-transporters. Several of the B family members are known to confer multidrug resistance in cancer cells; hence, subfamily B has also been called the 'MDR family of ABC transporters' [10]. Mutations in ABCB genes have been implicated in ankylosing spondylitis, diabetes type 2, coeliac disease, lethal neonatal syndrome, X-linked sideroblastic anaemia with ataxia, and several cholestatic liver diseases of infancy (Table 2). Many of these genotype-phenotype association studies suggesting the involvement of ABCB genes, however, are likely to represent false-positive, inconclusive or under-powered studies, and need further replication in other large cohorts before they can be regarded as 'informative' and conclusive [14].

Recent genome-wide searches for positive selection in the human genome have suggested the possibility that the metabolism and transport of foreign chemicals (foodstuffs, plant metabolites and drugs) might have undergone natural selection [15-17]. In this regard, the 230-kilobase (kb) cluster of four cytochrome P450 3A (CYP3A) functional genes and two pseudogenes, located at chromosome 7, is only 119 kb distant from the ABCB1 gene (the transcript of which spans 210 kb). Moreover, the CYP3A enzymes and ABCB1 transporter appear to have very similar substrate profiles, and both are regulated by the pregnane × receptor (PXR) in the liver, intestine, lung and kidney [18]. In addition, positive selection has recently been noted in the ligand-binding domain of PXR [19], lending further credence to the possible co-evolution of these two nearby loci, in response to dietary pressures. Future investigators should look carefully for similar patterns of possible co-evolution by other drug-metabolising genes and drug-transporter genes.

Subfamily C of the ABC family (ABCC)

Subfamily C includes the cystic fibrosis gene (CFTR, also called ABCC7) plus 12 other genes that encode transporters associated with multidrug resistance. The diverse activities of ABCC transporters include ion-channel and toxin excretion activity and reception on the cell surface; toxin excretion involves various fungal and bacterial toxins [10]. Mutations in one or more of the ABCC genes have been implicated in multidrug resistance, Dubin-Johnson syndrome, congenital bilateral aplasia of the vas deferens, diabetes type 2 and paroxysmal kinesigenic choreoathetosis -- as well as autosomal recessive diseases such as cystic fibrosis, pseudoxanthoma elasticum and hyperinsulinaemic hypoglycaemia of infancy (Table 2).

The CFTR encodes a unique ABC transporter protein. Recent data suggest that CFTR channel activity evolved by converting the conformational changes associated with ATP binding and hydrolysis (found in 'true' ABC pumps) into an open permeation pathway, by means of intra-protein interactions that stabilise the open state [20]. The 'selective pressure' in the environment that might have encouraged this unique CFTR protein to evolve to exhibit this characteristic has been suggested to be severe infection -- that is, cholera epidemics [21].

Subfamily D of the ABC family (ABCD)

Also known as the peroxisomal or ALD transporters, this subfamily contains four genes that encode half-transporters; these subunits form homodimers or heterodimers to make a functional unit. These four human half-transporter genes code for at least 49 distinct proteins by means of 102 alternatively spliced transcripts [10]. Mutations in ADBD genes are known to cause ALD and Zellweger syndrome (Table 2).

Subfamily E of the ABC family (ABCE)

ABCE1 is the single member in this family which is an organic anion-binding protein (its trivial name is OABP), sometimes confused with the 11 SLCO genes that encode solute-carrier organic anion transporters [2]. ABCE1 has an ATP-binding domain but lacks the transmembrane domain, making it unlikely that this protein functions as a transporter. Because of 15 alternatively spliced transcripts, the ABCE1 gene encodes five distinct proteins [10]. ABCE1 has been found to block the activity of ribonuclease L. Activation of ribonuclease L leads to inhibition of protein synthesis in a pivotal pathway involving viral interferon action; hence, ABCE1 functions to promote interferon activity.

Subfamily F of the ABC family (ABCF)

Along with ABCE1, ABCF members also have ATP-binding domains, but no transmembrane domains, making transporter function unlikely. Due to alternatively spliced products, the three ABCF genes encode 26 distinct proteins [10]. Because the ABCF genes appear to be upregulated by tumour necrosis factor-α, it is believed that members of this subfamily might play a role in inflammatory processes. No diseases have been associated, so far, with either the ABCE or ABCF genes (Table 2).

Subfamily G of the ABC family (ABCG)

Subfamily G comprises at least five genes that encode 'reverse half-transporters', meaning that they form the second half of a heterodimer. Mutations in ABCG genes have been implicated in sterol accumulation disorders and atherosclerosis (Table 2). Due to alternative splicing, at least 18 distinct subunit proteins have been identified as products of the five ABCG genes [10]. ABCG1 is involved in cholesterol efflux in macrophages and may regulate cellular lipid homeostasis in other cell types as well. ABCG2 functions in multidrug resistance transport; steroids (cholesterol, oestradiol, progesterone and testosterone) and certain chlorophyll metabolites, as well as organic anions, are transported by ABCG2. ABCG3 is expressed at high levels in the thymus and spleen, suggesting a possible potential role in the transport of specific peptides or hydrophobic compounds from lymphocytes. ABCG5 and ABCG8 both appear to limit intestinal absorption and promote biliary excretion of sterols; their expression is localised primarily in the liver, colon and intestine. Mutations in either of these two genes lead to sitosterolaemia. This is characterised by hyperabsorption plus decreased biliary excretion of dietary sterols, leading to hypercholesterolaemia, tendon and tuberous xanthomata, early-age onset of atherosclerosis, and abnormal blood and liver function test results.

Evolution of the ABC transporter family

Clearly, the ancestral ABC gene first appeared in unicellular organisms and is therefore ancient, being now present in eubacteria, archaebacteria, plants, fungi, yeast and all animals. The nomenclature of the ABC gene family is the same as that developed for more than 150 other gene families and superfamilies in the human and rodent genomes http://www.genenames.org/. All of these nomenclature systems followed the original example described for the CYP genes. The CYP genes were conveniently arranged into families and subfamilies based on percentage amino acid sequence identity [22-27]. Enzymes that share ~40 per cent or higher identity are assigned to a particular family (designated by an Arabic numeral). Protein sequences sharing ~55 per cent or higher identity are grouped into a particular subfamily (designated by a letter). Enzymes that share greater than ~70 per cent amino acid identity are then named as members within that subfamily (and given Arabic numbers, usually in the sequence in which they were discovered).

For example, sterol 27-hydroxylase and 25-hydroxy-vitamin D3 1 α-hydroxylase are both assigned to the CYP27 family because they share >40 per cent sequence identity. Because their protein sequences are < 55 per cent identical, however, sterol 27-hydroxylase is assigned to the CYP27 'A' subfamily and 25-hydroxy-vitamin D3 1 α-hydroxylase to the 'B' subfamily. If another enzyme were to be discovered that shared > 55 per cent identity with sterol 27-hydroxylase, it would be named CYP27A2. Another enzyme was discovered that shared < 55 per cent but > 40 per cent identity with sterol 27-hydroxylase, as well as with the 25-hydroxy-vitamin D3 1 α-hydroxylase, and therefore it was named CYP27C1. The development and application of this pleasingly logical system of nomenclature to the genes of many animals, plants and bacteria http://drnelson.utmem.edu/CytochromeP450.html has eliminated the confusion that often has plagued the naming of gene families and superfamilies. Subsequently, this 'divergent evolution' nomenclature system has been adopted for the ABC gene family, among many others. The human subfamilies and members within each subfamily can be seen (Figure 1) to have diverged over evolutionary time. The nomenclature for both the human ABC gene family and the CYP gene superfamily has been approved by the HUGO Gene Nomenclature Committee (HGNC; http://www.genenames.org).

Conclusions

The family of ABC pumps, along with the super-family of SLC proteins summarised in the last issue of Human Genomics [2], are among the most important large classes of transporters that move inorganic ions, metals, peptides, steroids, nucleosides, sugars and many other small molecules across the cell's surface membrane, as well as organelle membranes within cells. The ABC gene family of transporters comprises 49 and 52 genes (in eight subfamilies) in the human and mouse genomes, respectively; the nomenclature of these genes is based strictly on divergent evolution. The SLC gene superfamily is composed of 362 genes in 55 families in the human genome; the nomenclature of those genes is based largely on biochemical function, rather than divergent evolution from a common ancestor [2].

The ABC transporters are critically important in innumerable physiological functions, underscored by the fact that defects in more than a dozen of these genes have been associated with severe inherited disorders of metabolism -- cystic fibrosis being among the most prominent diseases, caused by mutations in the CFTR gene. The ABC transporters are also of great clinical importance in the transport of cancer chemotherapeutics, numerous other drugs and metabolites, and many chemicals present in foodstuffs. ABC transporters are therefore likely targets for drug therapy. A large number of new studies focused on the ABC transporter genes are anticipated.

Acknowledgements

We thank our colleagues, especially Lei He, for valuable discussions and a careful reading of this manuscript. The writing of this paper was funded, in part, by NIH grants R01 EY11490 (V.V.) and P30 ES06096 (D.W.N.).

References

  1. Camerino DC, Desaphy JF, Tricarico D. et al. Therapeutic approaches to ion channel diseases. Adv Genet. 2008;64:81–145. doi: 10.1016/S0065-2660(08)00804-3. [DOI] [PubMed] [Google Scholar]
  2. He L, Vasiliou K, Nebert DW. Analysis and update of the human solute carrier (SLC) gene superfamily. Hum Genomics. 2009;3:195–206. doi: 10.1186/1479-7364-3-2-195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Benga G. Water channel proteins (later called aquaporins) and relatives: Past, present, and future. IUBMB Life. 2009;61:112–133. doi: 10.1002/iub.156. [DOI] [PubMed] [Google Scholar]
  4. Verrier PJ, Bird D, Burla B. et al. Plant ABC proteins -- A unified nomenclature and updated inventory. Trends Plant Sci. 2008;13:151–159. doi: 10.1016/j.tplants.2008.02.001. [DOI] [PubMed] [Google Scholar]
  5. Dassa E, Bouige P. The ABC of ABCs: A phylogenetic and functional classification of ABC systems in living organisms. Res Microbiol. 2001;152:211–229. doi: 10.1016/S0923-2508(01)01194-9. [DOI] [PubMed] [Google Scholar]
  6. Orsi CH, Tanksley SD. Natural variation in an ABC transporter gene associated with seed size evolution in tomato species. PLoS Genet. 2009;5:e1000347. doi: 10.1371/journal.pgen.1000347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Lage H. An overview of cancer multidrug resistance: A still unsolved problem. Cell Mol Life Sci. 2008;65:3145–3167. doi: 10.1007/s00018-008-8111-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Piehler AP, Hellum M, Wenzel JJ. et al. The human ABC transporter pseudogene family: Evidence for transcription and gene-pseudogene interference. BMC Genomics. 2008;9:165. doi: 10.1186/1471-2164-9-165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Chen ZQ, Annilo T, Shulenin S, Dean M. Three ATP-binding cassette transporter genes, Abca14, Abca15, and Abca16, form a cluster on mouse chromosome 7F3. Mamm Genome. 2004;15:335–343. doi: 10.1007/s00335-004-2281-8. [DOI] [PubMed] [Google Scholar]
  10. Dean M, Rzhetsky A, Allikmets R. The human ATP-binding cassette (ABC) transporter superfamily. Genome Res. 2001;11:1156–1166. doi: 10.1101/gr.GR-1649R. [DOI] [PubMed] [Google Scholar]
  11. Finn RD, Mistry J, Schuster-Bockler B. et al. Pfam: Clans, web tools and services. Nucl Acids Res. 2006;34:D247–D251. doi: 10.1093/nar/gkj149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Finn RD, Tate J, Mistry J. et al. The Pfam protein families database. Nucl Acids Res. 2008;36:D281–D288. doi: 10.1093/nar/gkn226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Albrecht C, Viturro E. The ABCA subfamily --Gene and protein structures, functions and associated hereditary diseases. Pflugers Arch. 2007;453:581–589. doi: 10.1007/s00424-006-0047-8. [DOI] [PubMed] [Google Scholar]
  14. Nebert DW, Zhang G, Vesell ES. From human genetics and genomics to pharmacogenetics and pharmacogenomics: Past lessons, future directions. DrugMetab Rev. 2008;40:187–224. doi: 10.1080/03602530801952864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Voight BF, Kudaravalli S, Wen X, Pritchard JK. A map of recent positive selection in the human genome. PLoS Biol. 2006;4:e72. doi: 10.1371/journal.pbio.0040072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Sabeti PC, Varilly P, Fry B. et al. Genome-wide detection and characterization of positive selection in human populations. Nature. 2007;449:913–918. doi: 10.1038/nature06250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Tang K, Thornton KR, Stoneking M. A new approach for using genome scans to detect recent positive selection in the human genome. PLoS Biol. 2007;5:e171. doi: 10.1371/journal.pbio.0050171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Miki Y, Suzuki T, Tazawa C. et al. Steroid and xenobiotic receptor (SXR), cytochrome P450 3A4, and multidrug resistance gene 1 in human adult and fetal tissues. Mol Cell Endocrinol. 2005;231:75–85. doi: 10.1016/j.mce.2004.12.005. [DOI] [PubMed] [Google Scholar]
  19. Krasowski MD, Yasuda K, Hagey LR, Schuetz EG. Evolution of the pregnane × receptor: Adaptation to cross-species differences in biliary bile salts. Mol Endocrinol. 2005;19:1720–1739. doi: 10.1210/me.2004-0427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Jordan IK, Kota KC, Cui G. et al. Evolutionary and functional divergence between the cystic fibrosis transmembrane conductance regulator and related ATP-binding cassette transporters. Proc Natl Acad Sci USA. 2008;105:18865–18870. doi: 10.1073/pnas.0806306105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Kavic SM, Frehm EJ, Segal AS. Case studies in cholera: Lessons in medical history and science. Yale J Biol Med. 1999;72:393–408. [PMC free article] [PubMed] [Google Scholar]
  22. Nebert DW, Gonzalez FJ. P450 genes: Structure, evolution, and regulation. Annu Rev Biochem. 1987;56:945–993. doi: 10.1146/annurev.bi.56.070187.004501. [DOI] [PubMed] [Google Scholar]
  23. Nebert DW, Adesnik M, Coon MJ. et al. The P450 gene superfamily Recommended nomenclature. DNA. 1987;6:1–11. doi: 10.1089/dna.1987.6.1. [DOI] [PubMed] [Google Scholar]
  24. Nebert DW, Nelson DR, Adesnik M. et al. The P450 superfamily: Updated listing of all genes and recommended nomenclature for the chromosomal loci. DNA. 1989;8:1–13. doi: 10.1089/dna.1.1989.8.1. [DOI] [PubMed] [Google Scholar]
  25. Nebert DW, Nelson DR, Coon MJ. et al. The P450 super-family: Update on new sequences, gene mapping, and recommended nomenclature. DNACell Biol. 1991;10:1–14. doi: 10.1089/dna.1991.10.1. [DOI] [PubMed] [Google Scholar]
  26. Nelson DR, Kamataki T, Waxman DJ. et al. The P450 superfamily: Update on new sequences, gene mapping, accession numbers, early trivial names of enzymes, and nomenclature. DNA Cell Biol. 1993;12:1–51. doi: 10.1089/dna.1993.12.1. [DOI] [PubMed] [Google Scholar]
  27. Nelson DR, Koymans L, Kamataki T. et al. P450 superfamily: Update on new sequences, gene mapping, accession numbers and nomenclature. Pharmacogenetics. 1996;6:1–42. doi: 10.1097/00008571-199602000-00002. [DOI] [PubMed] [Google Scholar]

Articles from Human Genomics are provided here courtesy of BMC

RESOURCES