
Chemistry of the Retinoid (Visual) Cycle
PhilipD Kiser
Marcin Golczak
Krzysztof Palczewski
K.P.: phone, 216-368-4631; fax, 216-368-1300; e-mail,kxp65@case.edu. M.G.:phone, 216-368-3063; fax, 216-368-1300;e-mail,mxg149@case.edu. P.D.K.: phone, 216-368-8794; fax, 216-368-1300; e-mail,pdk7@case.edu.
Received 2013 Feb 15; Issue date 2014 Jan 8.
1. Introduction
Assuccinctly summarizedby Wolf,1 lackof vitamin A (all-trans-retinol) was recognized byancient Egyptians as causing a visual deficiency involving the retinaand cornea that could be cured by eating liver. One of the symptomsof vitamin A deficiency is night blindness or nyctalopia (from Greekνύκτ-, nykt – night; and αλαός,alaos – blindness), recognized by ancient Greeks, includingHippocrates, as affecting the retina.2 In1913 McCollum showed that “fat-soluble factor A” wasessential for growth of a rat colony (reviewed in ref (3)). The treatment of factorA-deficiency included liver or liver extracts, but later in 1930 Moorefound that yellow pigment (carotene) was a good substitute for thistherapy.4 A major breakthrough occurredin 1931 when the chemical structures for β,β-caroteneand retinol (itsall-trans isomer now known as vitaminA) were determined by Karrer and colleagues.5 But, it was Wald who discovered that retinol derivatives constitutethe chemical basis of our vision,6 a contributionsubsequently recognized by a Nobel Prize award in 1967. In 1950–1960,a variety of vitamin A metabolic transformations, including oxidation/reductionand esterification, were elucidated by Olson,7−20 Goodman,21−32 Chytil and Ong,33−41 and Norum and Blomhoff.42−57 The discovery that one set of these metabolites, namely retinoicacids, plays a key role in the nuclear regulation of a large numberof genes added a notable dimension to our knowledge of gene expression.This mechanism is also a critical player in the successful healingof corneal wounds,58 a second manifestationof vitamin A-deficiency recognized earlier.
Further progressin understanding the multiple physiological rolesof retinoids has been made in recent years, due mainly to the successfulapplication of modern scientific technology. Examples include enzymologycombined with structural biology,in vivo imagingbased on retinoid fluorescence, improvements in analytical methods,generation and testing of animal models of human diseases with specificpathogenic genetics, genetic analysis of human conditions relatedto changes in vitamin A metabolism, and pharmacological approachesto combat these diseases.
In this review we focus on the involvementof retinoids in supportingvisionvia light-sensitive rod and cone photoreceptorcells in the retina. We begin with a brief description of isopentenyldiphosphate (IPP) biosynthesis, which is essential for carotenoid(C40 isoprenoid) production. Certain of these colored compounds, suchas lutein, are deposited in our retina’s macula, appearingas a “yellow” spot. Other carotenoids containing atleast one unmodified β-ionone ring (represented by β,β-caroteneand cryptoxanthin) serve as precursors ofall-trans-retinal. Many different compounds can be generated from this monocyclicditerpenoid, which contains a β-ionone ring and polyene chainwith a C15 aldehyde group. Among the numerous enzymatic activitiesthat contribute to retinoid metabolism, polyenetrans/cis isomerization is a particularly fascinatingreaction that occurs in specialized structures of the retina basedon a two cell system comprised of retinal photoreceptor cells andthe retinal pigment epithelium (RPE). A specific enzyme system, calledthe retinoid (visual) cycle, has evolved to accomplish retinoid isomerizationthat is required for visual function in vertebrates. Individual enzymesof this pathway harbor secrets about the molecular mechanisms of thischemical transformation. Malfunctions of these processes or otherpathological reactions often precipitate severe retinal pathologies.This review attempts to balance contributions that have been publishedover the past decades and does not intend to replace the views ofinvestigators with different perspectives of retinoid chemistry inthe eye.59−85
2. Chemistry of Isoprenoids
In animals, carotenoidsand retinoids must be acquired throughthe diet, as they cannot be synthesizedde novo.These compounds are involved in critical functions of many organsin addition to their vital involvement in vision.86
2.1. Mevalonate and Nonmevalonate Pathways
Naturally occurring carotenoids are all synthesized from two basic5-carbon precursors: isopentenyl diphosphate (IPP) and its isomerdimethylallyl diphosphate (DMAPP).87 Amultitude of additional compounds, including steroids, ubiquinones,and chlorophylls, are also synthesized from these isoprene precursorsand are thus collectively referred to as isoprenoids or terpenoids.88,89 Until the early 1990s it was believed that isoprenoid precursorswere synthesized exclusively through the mevalonate pathway, a seriesof enzymatically catalyzed reactions in which three molecules of acetate,in the form of acetyl-coenzyme A (acetyl-CoA), are condensed and modifiedby reduction, phosphorylation, and decarboxylation to generate IPP89 (Figure1). Researchin eubacteria and plants then revealed a second metabolic route forIPP synthesis referred to as the methylerythritol 4-phosphate (MEP)or nonmevalonate pathway, which utilizes the triose derivative,d-glyceraldehyde 3-phosphate along with pyruvate as startingmaterials90−94 (Figure2). Following condensation of thesetwo molecules, a methyl isomerization reaction occurs that convertsthe initially linear carbon chain into an isopentyl linkage.95,96 Subsequent reduction, CDP transfer, phosphorylation, and reductionserve to eliminate hydroxyl functional groups and introduce a diphosphategroup to generate IPP as well as DMAPP in about a five to one ratio.97 In many cases, a particular organism encodesthe enzymes for only one of these two metabolic pathways in its genome.90 Metazoans, fungi, and archaea rely exclusivelyon the mevalonate pathway for isoprenoid biosynthesis, whereas cyanobacteriaand algae make sole use of the MEP pathway.98 Depending on the species, eubacteria and protozoa can employ eitherof the two pathways in a mutually exclusive fashion.99 However, most eubacteria utilize the MEP pathway.99 In higher plants, most cytosolic IPP and DMAPPare derived from the mevalonate pathwayvia enzymesencoded by genomic DNA. By contrast, these isoprene compounds aresynthesized through the MEP pathway in plastids of plants as a consequenceof the evolutionary relationship of plastids to cyanobacteria.100 However, mixing of cytosolic and plastid isoprenoidprecursors has been reported.101,102 Most carotenoids consumedby humans are synthesized from MEP-derived IPP and DMAPP.
Figure 1.
Mevalonatepathway for the synthesis of IPP. Two molecules of acetyl-CoAare joined together to form acetoacetyl-CoA (i) in areaction catalyzed by thiolase with the release of free CoA (CoASH).A third molecule of acetyl-CoA is added by 3-hydroxy-3-methylglutaryl(HMG)-CoA synthase to form HMG-CoA (ii). Compoundii is reduced by HMG-CoA reductase in an NADPH-dependent mannerto form (R)-mevalonate (iii), whichis the rate-limiting step of the pathway. Compoundiii is sequentially phosphorylated by mevalonate kinase and phosphomevalonatekinase to form (R)-mevalonate-5-diphosphate (iv). Finally,iv is decarboxylated by mevalonate-5-diphosphatedecarboxylase in an ATP-dependent manner to form IPP (v). Coloring of oxygen atoms is intended to assist in tracking ofthe chemical origin of the carbon skeleton. P, phosphoryl group; OPi, inorganic phosphate.
Figure 2.
MEP (nonmevalonate) pathway for the synthesis of IPP and DMAPP.First,d-glyceraldehyde-3-phosphate (i) is condensedwith pyruvate to form 1-deoxy-d-xylulose-5-phosphate (DOXP,ii) catalyzed by 1-deoxy-d-xylulose-5-phosphate synthase(DXS) using a thiamine diphosphate cofactor with the loss of CO2.ii is isomerized and reduced by DOXP-isomeroreductase(IspC) in an NADPH-dependent manner to form 2C-methyl-d-erythritol-4-phosphate, which is then conjugated with CTPto form 4-diphosphocytidyl-2C-methyld-erythritol(iv) in a reaction catalyzed by 2C-methyl-d-erythritol cytidylyltransferase (IspD).iv isthen phosphorylated to form 4-diphosphocytidyl-2C-methyl-d-erythritol-2-phosphate (v) by 4-diphosphocytidyl-2C-methyl-d-erythritol kinase (IspE).v is cyclized with the loss of CMP to form 2C-methyl-d-erythritol-2,4-cyclodiphosphate (vi) by 2C-methyl-d-erythritol-2,4-cyclodiphosphate synthase(IspF).vi is then reduced with cleavage of the cyclodiphosphatemoiety to form 1-hydroxy-2-methyl-2-(E)-butenyl diphosphate(vii) by the iron-sulfur enzyme IspG using ferrodoxinas a cofactor. Finally,vii is further reduced by a secondiron–sulfur protein IspH, giving a mixture of IPP (viii) and dimethylallyl diphosphate (DMAPP,ix). The redcolored oxygen atom is intended to assist in tracking of the chemicalorigin of the carbon skeleton. P, phosphoryl group; PPi, inorganic pyrophosphate.
IPP and DMAPP differ only in the position of an alkene doublebondand are thus classified as geometric isomers (Figure3). Enzymes known as IPP isomerases (IDIs) catalyze the reversibleisomerization of IPP into DMAPP. For organisms that solely use themevalonate pathway for isoprene biosynthesis, IDIs are essential forproduction of DMAPP because only IPP is generated through this pathway.Although both IPP and DMAPP are producedvia theMEP pathway, IDIs are still required to generate the proper IPP/DMAPPratio for isoprenoid biosynthesis.103,104 Two structurallyunrelated enzymes, called IDI type 1 and IDI type 2, can each carryout the reaction.105 Type 1 IDIs were thefirst to be characterized and rely on an active site Cys residue aswell as divalent cations for their catalytic function.106 Biochemical and crystallographic analyses oftype I IDI indicate that an active site, metal-binding Glu residuetransfers a proton to the alkene π bond of IPP, generating atertiary carbocation or carbocation-like intermediate.107,108 Abstraction of a proton from C2 of the isoprene skeleton by an activesite Cys residue regenerates the alkene forming DMAPP.109,110 Isotope labeling studies demonstrated that the proton transpositionoccurs in an antarafacial manner, a mechanism consistent with theplacement of catalytic groups in the active site of type I IDIs.111 Type II IDIs are a more recently discoveredclass of enzymes that catalyze the same reaction as type I IDIs butare structurally and evolutionarily unrelated.105,106 In contrast to type I IDIs, type II enzymes rely on flavin mononucleotide(FMNred) and NADPH as well as divalent cations as cofactorsto perform catalysis.105 The dependenceon these nucleotide cofactors was somewhat surprising given that IPPto DMAPP isomerization does not involve net redox changes. Free radicalmechanisms involving transient abstraction of a hydrogen atom112 as well as protonation–deprotonation(carbocation intermediate) mechanisms113,114 have beenproposed as a means to effect the double bond shift. The involvementof carbocation intermediates turns out to be a major theme in theenzymatic isomerization of isoprenoids.
Figure 3.
Reversible isomerizationof IPP (i) into DMAPP (ii). The reaction,catalyzed by IPP/DMAPP isomerase, is thoughtto proceedvia a protonated carbocationic intermediateshown in brackets.
2.2. CarotenoidBiosynthesis
The biosynthesisof carotenoids from IPP and DMAPP begins with the condensation ofone molecule of DMAPP with three molecules of IPP catalyzed by theenzyme geranylgeranyl diphosphate synthase to form geranylgeranyldiphosphate (GGPP) (Figure4).87 In some plant species, geranyl diphosphate is synthesizedfirst by the enzyme geranyl diphosphate synthase and then elongatedby the addition of two IPP molecules in a GGPP synthase-catalyzedreaction. These enzymes belong to the prenyl transferase family, whichuse divalent cations, such as Mg2+ or Mn2+,to carry out the condensation of isoprenoid precursors.115 Mechanistically, the divalent cation polarizesthe diphosphate moiety of DMAPP to facilitate its dissociation withconsequent formation of an electrophilic carbocation intermediate.116,117 Proper positioning of IPP in the active site facilitates a nucleophilicattack of the alkene π bond electrons on C1 of the isoprenecarbocation, resulting in formation of a new carbon–carbonsingle bond. Deprotonation of the diphosphate-containing unit resultsin C1–C2 π bond formation, allowing chain elongationto continue in the presence of appropriate enzymes. Two moleculesof GGPP are then joined in a head to head fashion to form phytoenein a reaction catalyzed by the enzyme phytoene synthase, which isthe first committed step in the synthesis of carotenoids.100 This reaction again features a carbocationor carbocation-like intermediate that reacts with a second GGPP tofrom a cyclopropylcarbonyl diphosphate compound (prephytoene diphosphate).118−120 This compound breaks down with loss of pyrophosphate and a protonto produce 15-cis-phytoene.121 The colorless, C40 tetraterpenoid product is then subjectedto a series of desaturation and isomerization reactions that culminatein the production ofall-trans-lycopene, the immediateprecursor of β,β-carotene. In bacteria, a single enzymecalled carotene desaturase (CrtI) is responsible for conversion ofphytoene into lycopene. In plants, two desaturase enzymes called phytoenedesaturase (PDS) and ξ-carotene desaturase (ZDS)122 and two isomerases called ξ-caroteneisomerase123 and carotenoid isomerase (CrtIso)124 together convert phytoene into lycopene.87 Interestingly, CrtIso shares significant sequencehomology with the carotene desaturase (CrtI) as well as a mammalianenzyme known as retinol desaturase (RetSat).63,125 Lycopene, a compound with a red hue conferred by a set of 11 conjugateddouble bonds, is transformed into β,β-carotene by an enzymeknown as lycopene-β-cyclase.126,127 Althoughnot a reaction that involves net redox changes, the cyclization performedby lycopene-β-cyclase is dependent on an NAPDH cofactor.128 In keeping with the general theme of terpenoidisomerization, lycopene cyclization reactions also occur through carbocationintermediates.129
Figure 4.
The carotenoid branchof isoprenoid biosynthesis. Synthesis ofβ,β-carotene begins with the sequential condensation ofa single DMAPP molecule (i) with three IPP moleculesto form C20 geranylgeranyl diphosphate (ii, GGPP) catalyzed by GGPP synthase (CrtE). Next, two GGPP moleculesare combined in a head-to-head fashion to form C40 15-cis-phytoene (iii, shown in theall-trans configuration for ease of presentation) in a reaction catalyzedby phytoene synthase (CrtB), which is the first committed step incarotenoid biosynthesis. In bacteria, phytoene is converted toall-trans-lycopene (iv) by a series of desaturationand isomerization steps catalyzed by CrtI. In plants this conversionis catalyzed by phytoene desaturase and ζ-carotene desaturasetogether with the isomerases ζ-carotene isomerase and CrtIso.Finally, lycopene is converted to β,β-carotene (v) in two steps by lycopene β-cyclase. Lycopene is alsoa substrate for lycopene ε-cyclase, which catalyzes formationof δ-carotene (not shown).
2.3. Retinoid Metabolism in Vertebrates
Mammals efficiently utilize both preformed vitamin A in the formofall-trans-retinyl esters and pro-vitamin A carotenoids(mainly β,β-carotene) to sustain the body pool of vitaminA.130 Unlike retinyl esters, which arehydrolyzed to retinol in the small intestine followed by rapid absorptionby enterocytes, uptake of carotenoids is mediated and regulated byscavenger receptor class-B type-I (SR-BI).131,132 Upon absorption by enterocytes, β,β-carotene undergoesoxidative cleavage catalyzed by β,β-carotene 15,15′-monooxygenase(BCMO1).133,134 This symmetric split of thecarotenoid produces two molecules ofall-trans-retinalthat subsequently enter vitamin A metabolic pathways (Figure5).
Figure 5.
Retinoid metabolism in vertebrates. Dietaryall-trans-β,β-carotene (i), obtained primarily fromplants, is oxidatively cleaved in a symmetric manner by β-carotenemonooxygenase I (BCMO I), yielding two molecules ofall-trans-retinal (ii). Retinal can reversibly combine with anamino group to form a retinyl imine (Schiff base) (iv). Retinal is also subject to oxidation and reduction to form retinoicacid (iii) and retinol (vitamin A)v, respectively,the latter in a physiologically reversible manner. Retinoic acid canbe converted into several conjugated and/or oxidized derivatives,some of which exert biological effects. Retinol also can be convertedinto several derivatives including retro-retinoids, saturated retinols,and phosphate conjugates. Retinol is also reversibly esterified toproduce retinyl esters (vi), the main storage form ofvitamin A in the body.
The diverse functions of retinoids are carried out by a fewphysiologicallyactive metabolites ofall-trans-retinol that areproduced by enzymatic modification of the functional groups of thisvitamin and geometric isomerization of its polyene chain. Oxidationofall-trans-retinal catalyzed by retinaldehyde dehydrogenases(RALDHs) leads to formation ofall-trans-retinoicacid, a ligand for nuclear retinoic acid receptors (RARs) that, coupledwith retinoid X receptors (RXRs), bind to retinoic acid response elementson the promoter region of target genes and regulate their transcription.135,136 Becauseall-trans-retinoic acid formation is irreversible,an excess of this active retinoid can be cleared only by its furtherconversion to more polar metabolites through oxidation by cytochromeP450 (CYP26) and/or glucuronidation by UDP-glucuronosyl transferases(UGTs).137,138 An alternative metabolic pathway ofall-trans-retinal leads to formation of 11-cis-retinal, a visual chromophore that couples to rod and cone opsinsto form photosensitive pigments.139−141 The thermodynamicallyunfavorable isomerization of the 11–12 double bond does notoccur at the aldehyde levelin vivo.142 Instead,all-trans-retinalis first reduced toall-trans-retinol by short-chaindehydrogenase/reductase (SDR) or alcohol dehydrogenase enzymes.136,143,144 Subsequent esterification ofall-trans-retinol, mainly by lecithin/retinol acyltransferase(LRAT), provides both a major storage form of retinoids in the bodyand a direct substrate for RPE65-dependent enzymatic isomerizationof the retinoid polyene chain.145−148 Multiple additional endogenous retinoidmetabolites are derived fromall-trans-retinol. Theseincludeall-trans-13,14-dihydroretinol produced bysaturation of the double bond by RetSat. Other examples are retro-retinoids,such as anhydroretinol and 14-hydroxy-4,14-retro-retinol. Though themolecular identities of enzymes involved in the production of retro-retinoidsin vertebrates are currently unknown, these metabolites have beenshown to regulate lymphocyte proliferation.149−151 Transfer of a phospho group ontoall-trans-retinolresults in formation of retinyl monophosphate that has been detectedin the liver of rodents.152,153 Glycophospholipidsconsisting of a retinol moiety linked by a phosphodiester bond tomannose or galactose were postulated to function in the transfer ofsugar units onto proteins to form some glycoproteins.154
3. Isomerization of Retinoids
Retinoids are reactive compounds that readily isomerize. Here wetake a close look at their important chemical transformations.
3.1. Geometric Isomers of Retinoids
Progressin understanding vitamin A metabolism would not be possible withoutdevelopment of adequate analytical methods that allow separation,detection, and quantification of retinoids. Such methods have graduallyadvanced since the discovery of vitamin A about 100 years ago. A complicationis that retinoids exist in several geometrical configurations withdifferently modified functional groups (Figure6A). Lipophilic compounds soluble in organic solvents, including retinoland its esters, were initially separated by thin-layer chromatographyon alumina- and silica-based stationary phases.155,156 However, introduction of modern high performance liquid chromatography(HPLC) techniques in the early 1970s together with standardized commerciallyavailable compact stationary phase columns enabled precise, reproducibleanalyses of vitamin A metabolites and, most importantly, determinationof their isomeric composition.157−160 Today retinoids can be separated under numerouschromatographic conditions (summarized in refs (160 and161)) optimized fornormal and reverse-phase columns. Selection of the most appropriatemethodology depends on the chemical properties of the particular retinoidas well as the source of the biological sample. The eye exhibits anespecially complex retinoid composition. The highest resolution methodfor separating retinyl esters, retinal, and retinol as well as theirisomers present in the eye is normal phase HPLC (Figure6B).161 The strength of an analyte’sinteraction with the silica stationary phase depends not only on itsfunctional groups but also on steric factors and the structure ofthe molecule. This feature is especially important for separationof molecules that are chemically similar but physically different,e.g. geometric isomers. Moreover, this methodology provides uniqueflexibility in tuning chromatographic conditions by adjusting thepolarity of the mobile phase, routinely composed of hexane and ethylacetate. Highly hydrophobic hexane simplifies the tissue homogenateextraction procedure and greatly reduces sample complexity withoutsacrificing overall analytical performance.162
Figure 6.
HPLC-basedseparation and detection of retinoids. (A) The mainclasses of retinoid isomers commonly found in experimental samplesthat can be distinguished by analytical methods. (B) Elution profilesof retinol, retinal, and retinal oxime isomers from a normal phaseHPLC column with an isocratic flow of 10% ethyl acetate/hexane. Primaryanalytical methods of retinoid identification and quantification includeUV/vis spectroscopy and mass spectrometry. (C) UV/vis absorbance spectraof selected retinoids reveal characteristic differences in absorbancemaxima and overall shape of the spectra that are used to classifythe chemical and geometric form of retinoids. (D) Electrospray ionizationof retinol and retinyl esters triggers water or carboxylate dissociation,resulting in the predominant parent ion ofm/z = 269 [M – 17]+, whereas retinal exhibitsthe expected molecular ion ofm/z = 285 [M + H]+. Characteristic MS/MS fragmentation patternsof the parent ions are shown in the bottom panels. The typicalm/z = 161 fragment of retinal in MS/MSspectra is indicative of ionone ring loss from the parent ion.
Retinoid detection, identification,and quantification are simplifiedby their characteristic spectral properties (Figure6C and D). The conjugated polyene chain contributes to relativelystrong absorption at ultraviolet (UV) and visible (Vis) wavelengths.Thus, absorbance maxima as well as the overall shape of the spectraprovide valuable information about the number of conjugated doublebonds and allow identification of the isomeric states of the compound.UV/Vis detection offers a limit of retinoid quantification at a lowpicomolar range through most photodiode array detectors and providesexcellent linearity over a wide range of concentrations (2–1500pmol) (Figure6C).163 A complementary analytical method allowing precise molecular identificationand quantification of retinoids is mass spectrometry (MS). The greatestadvantage of MS coupled to HPLC is sensitivity in the low femtomolarrange. Moreover, modern tandem mass spectrometry offers definitivemolecular identification based on the induced fragmentation patternof the precursor ion (Figure6D).
3.2. Retro- and anhydro-retinols
In the1920s two groups164,165 made the discovery that vitaminA-containing solutions, in the presence of Brønsted and Lewisacids, undergo a color change from pale yellow (λmax ∼ 325 nm) to brilliant blue (λmax ∼600 nm). The blue color is semistable with an effective lifetime ofup to ∼3 min when chloroform is used as a solvent.166 This color change, when initiated using antimonytrichloride as the Lewis acid, is known as the Carr–Price reactionand was a primary means of retinol detection and quantification priorto the advent of chromatographic methods.167 Investigation into the mechanism of the reaction underlying thecolor change revealed that the hydroxyl moiety of retinol readilycombines with acids (e.g., protons or antimony), converting it intoa good leaving group.166 Hydroxyl dissociationgenerates a short-lived retinylic cation that can undergo proton eliminationto form anhydroretinol (Figure7). Anhydroretinolcan then undergo further chemical reactions to generate relativelylong-lived, blue-color species.168 In additionto anhydroretinol, direct protonation of the 13,14 retinol doublebond followed by proton elimination at C4 leads to formation retro-retinol(Figure7). Generation of the retinylic cationcan also be accomplished by flash photolysis of retinyl acetatevia heterolytic carbon–oxygen bond dissociation169 or by reactions of protons, released by radiolyticpulses, with retinol or retinyl acetate.170 The special location of the retinol hydroxyl group with respectto the conjugated polyene chain makes it a much better leaving groupthan is typical for regular aliphatic alcohols. Upon dissociation,the resulting retinylic cation is stabilized by extensive delocalizationalong the polyene chain with a half-life on the nanosecond time scale.169,171 Hydroxyl dissociation can also be facilitated by its conjugationwith an electron-withdrawing group, for example a sulfo moiety, asoccurs catalyticallyvia the enzyme retinol dehydratase.This member of the sulfotransferase family is responsible for thecatalytic conversion of retinol into anhydroretinol.172 In this reaction, a sulfo moiety is first transferred from3′-phosphoadenosyl-5′-phosphosulfate (PAPS) onto theretinol hydroxyl group to produce retinyl sulfate. Sulfate then readilydissociates, generating a resonance-stabilized carbocation that isquenched by proton loss from carbon 4 of the ionone ring to yieldanhydroretinol172 (Figure8).
Figure 7.
Conversion ofall-trans-retinol (i) into anhydroretinol (ii) and retro-retinol (iii) in the presence of acidvia carbocationicintermediates.
Figure 8.
Production of anhydroretinolfromall-trans-retinolcatalyzed by retinol dehydratase. First, a sulfo group is transferredtoall-trans-retinol (i) to form retinylsulfate (ii). Loss of sulfate then generates a carbocationicintermediate (iii) that, in the confines of the enzymeactive site, preferentially rearranges with loss of a proton to formanhydroretinol (iv).
3.3. Chemical/photochemical isomerization of retinoids
Spectroscopic properties of retinoids can be directly related totheir conjugated double bond system. Mobile π electrons of polyenesare delocalized over the entire molecule, resulting in resonance stabilizationof the compound.173 Thus, polyene singlebonds display some double-bond characteristics contributing to theirpreferred planar conformation. The most stable conformation for retinoidsisall-trans.174,175 Alternative geometricalisomers introduce some steric hindrance that increases the conformationalenergy compared to theall-trans isomer.176 Although some retinal isomers introduce onlymild steric clashes, e.g. between 12H and 15H in 13-cis or between 8H and 11H in the 9-cis conformation,other isomers such as 11-cis and 7-cis constitute examples of severely hindered isoprenoids in which theplanar conformation of the polyene chain cannot be sustained.174 To eliminate the strain, a twist is introducedthat further contributes to the increased conformational energy ofthe retinoid.177
Cis/trans isomerization of retinals does not occur efficiently in the darkat room temperature. However, an equilibrium between retinoid isomerscan be introduced chemically or photochemically (Figure9). The radical-mediated, iodine-catalyzed isomerization ofpolyenes is a three-step process that involves binding of an iodineatom to a carbon, an internal rotation, and then detachment of theiodine.178,179 Because iodine atoms are easily formed thermallyfrom I2, retinoid isomerization can be studied independentlyof light. Starting fromall-trans-retinal, the compositionof retinal isomers at equilibrium was consistent among multiple reportsthat includedall-trans, 13-cis,9-cis, 11-cis, and 9,13-di-cis isomers in ratios of about 0.62, 0.23, 0.11, 0.001,and 0.04, respectively.177,180,181 The same results were obtained whenall-trans-retinolor its palmitoyl ester was used as the starting material.181 Interestingly, kinetic studies revealed thatthe 9,13-di-cis isomer was a favored intermediatein the isomerization of 9-cis-retinal, whereas 13-cis-retinal was directly converted into theall-trans conformation without the contribution of any other isomeric intermediates.181 Consequently, different isomers lead to a diverseretinoid composition in the equilibrium state. Nevertheless, 11-cis-retinal was never preferably formed upon I2 or acid-catalyzed isomerization. In contrast, early work by Waldindicated that photolysis of dilute ethanolic solutions ofall-trans-retinal resulted in the formation of about 50%cis-retinals, with an exceptionally high contribution of11-cis-retinal representing 25% of the mixture.182−184 Since then, quantum yields fortrans/cis andcis/trans photoisomerization upon light excitation of retinalhave been studied under a variety of conditions demonstrating bothisomer and solvent dependence. The quantum yields forall-trans-retinal were considerably lower in polar (0.1–0.2) than innonpolar (0.4–0.7) solvents.185−189
Figure 9.
Generation of retinoid isomers fromall-trans-retinal.Theall-trans form of retinoids is the lowest infree energy and thus predominates at equilibrium. Formation of retinoidisomers can be facilitated chemically by treatment with I2, sulfhydryls, and trifluoroacetic acid, or by exposure to light.The composition of retinal isomers found at equilibrium is reportedafter Rando et al.181 and Deval et al.192
The influence of solvents on the efficiency and compositionofisomers at photoequilibrium has been attributed to solvent-dependentshifting of the nπ* and ππ* excited states.190−192 For example, in hexane light illumination led to almost exclusiveformation of the 13-cis isomer whereas increasedsolvent polarity contributed to efficient production of 11-cis, 9-cis, and 7-cis-retinalsin 0.19, 0.06, and 0.005 ratios, respectively.192,193 Many studies were dedicated to elucidate the mechanism(s) of photoisomerization,particularly the potential role of singlet and triplet excitationstates of retinals that could provide a mechanistic explanation forearly stages of rhodopsin-mediated visual excitation.186,190,194−196 Although the directly formed singlet excited state had a major rolein the photoisomerization of eitherall-trans or13-cis-retinal, in the case of 11-cis-retinal and 7-cis-retinal, isomerization occurredalso after an intersystem crossing to an excited triplet state, sothe triplet state was responsible for up to 50% of the observed isomerizationwith the other half of isomerized retinoid arising from the singletstate.188,193,197,198 Because the triplet state of chromophore bound torhodopsin has never been spectroscopically observed, it is not clearwhether these observations can be translated into a biologically relevantmodel. Nevertheless, these early studies on retinoid isomerizationclearly showed that the mixture of isomers obtained by any of theavailable methods did not recapitulate the equilibrium observed inliving tissue. Thus, they indirectly facilitated a shift in researchfocus toward the potential role of enzymes and specific binding proteinsin maintaining the composition of retinoid isomersin vivo.
3.4. Dihydroretinol
As mentioned above,retinoid isomers can be directed toward equilibrium by free radicalsvia addition–elimination processes which implicatetransient disintegration of the double bond prior to an internal rotation.178,181 By analogy, reduction of a selected double bond that allows freerotation followed by its desaturation could represent an alternativeroute to polyene isomerization (Figure10).Although the conjugated double bond system of retinoids is resistantto chemical reduction, this scenario can be effectively completedin vivo.199 Cyanobacterial andplant CrtIso catalyze the isomerization of (7Z,9Z,9′Z,7′Z)-tetra-cis-lycopene (prolycopene) toall-trans-lycopene.200−203 This enzyme-mediated isomerization employing redox cofactors occursthrough a reversible saturation–desaturation reaction of thecis-double bonds.204 Animal homologuesof the plant enzyme are inactive toward prolycopene.125,205 Instead, the mouse CrtIso-related enzyme acceptsall-trans-retinol as a substrate in carrying out the saturation of the retinoid13–14 double bond.125,206 However, in contrastto plant CrtIso, this reaction is not accompanied by isomerizationof the retinoid, yielding (R)-all-trans-13,14-dihydroretinolas a final product (Figure11).206 Consequently, this animal enzyme has been namedretinol saturase or RetSat.125 The enzymaticactivity of RetSat does not contribute to visual chromophore regenerationin the eye but rather influences processes involvingperoxisomeproliferator-activated receptor γ (PPARγ) activity,which regulates lipid accumulation in mice.125,207,208
Figure 10.
Retinoid isomerizationby sequential saturation–desaturationchemistry. This strategy is used, for example, by plant lycopene isomerase(CrtIso).
Figure 11.
Conversion ofall-trans-retinol (i) intoall-trans-(R)-13,14-dihydroretinol(ii) by RetSat, an enzyme evolutionarily related to CrtIso.
4. PhotoreceptorCells and Pigmented Helper Cells
To place retinoid chemicaltransformations in response to lightin a proper context, we must review the evolution of the visual system.Ancient photoreceptors composed of light-sensing cells (proto-eyes)mediated phototaxis with the assistance of pigmented cells, precursorsof the retinal pigment epithelium (RPE) in the retina of the eye.209,210 The proto-eye could likely detect directionality of light and couldalso be considered a precursor of a primitive circadian clock. Asliving organisms evolved over time, the cell numbers, complexity,and functional capability of the light-sensing organ increased. Forexample, whereasDrosophila melanogaster eye has10 different cells, that number has grown to 400 in humans, with aparallel development of retinal circuitry.211−215 Darwin wrote “Reason tells me, that if numerous gradationsfrom a simple and imperfect eye to one complex and perfect can beshown to exist, each grade being useful to its possessor, as is certainlythe case; if further, the eye ever varies and the variations be inherited,as is likewise certainly the case and if such variations should beuseful to any animal under changing conditions of life, then the difficultyof believing that a perfect and complex eye could be formed by naturalselection, though insuperable by our imagination, should not be consideredas subversive of the theory”.216
A large number of light sensitive molecules are needed toincreasethe probability of photon capture. These light-sensitive moleculesmust also couple to a receptor protein capable of initiating the requiredbiochemical chain of events. Such 30–60 kDa proteins were giventhe name “opsins”, and in bacteria they function ascation and anion pumps,217−223 whereas in higher organisms they are coupled to G proteins and thereforeare called G protein-coupled receptors (GPCRs). In every case, thechromophores are retinals bound to opsinvia a Schiffbase (retinylidene) group with a transmembrane domain (TMD) Lys residue.60,61,79,84,224−228 Because opsins are membrane proteins, theyare densely packed into specific regions of the cell to prevent theirrandom diffusion.229 For example, in bacteriathey form 2-dimensional crystalline membrane patches, whereas invertebrateopsins occupy cellular membrane protrusions called microvilli of rhabdomericphotoreceptor cells (Figure12). A differentsolution is employed by vertebrate photoreceptors which are modifiedciliary elongated protrusions called outer segments (OS).230−232 Rhabdomeric and ciliary photoreceptors also employ fundamentallydifferent mechanisms for regeneration of their visual pigments followinglight stimulation.233 Light-activated rhabdomericopsins can be restored to their ground state by absorption of a second,lower energy photon in a process known as photoreversal. Light activatedciliary opsins, on the other hand, do not undergo photoreversal butinstead rely on biochemical regeneration of the visual chromophorevia the retinoid (visual) cycle. Notably, there are exceptionsto these typical routes of visual chromophore regeneration, as hasbeen described, for example, inDrosophila.234 Rods and cones are the two morphologicallydistinct types of ciliated photoreceptor cells found in human retina(Figure12). Cones are further divided basedon the opsin they express, such as short-wavelength responsive cells(S cones that express blue opsin with an absorption peak at 420–440nm), medium-wavelength responsive cells (M cones that express so-called“green” opsin with an absorption peak at 530–540nm), and long-wavelength responsive cells (L cones that express “red”opsin with a maximum absorbance at 560–580 nm).66,84,235−241
Figure 12.
Comparison of photoreceptor structure between invertebrates representedbyDrosophila and vertebrates represented by man.Invertebrates utilize a rhabdomeric photoreceptor cell whereas vertebratephotoreceptors are modified ciliary cells. Notably, a few invertebrates,such as Amphioxus, employ ciliary photoreceptors.
4.1. Structure of the Mammalian Retina
The vertebrateeye evolved to utilize key optical principles alongwith the physicochemical properties of retinoids. First, light isfocused by passing through two proteinaceous biological lenses atthe front of the eye, namely the cornea and lens.242 Then the focused wave of photons is absorbed by photoreceptorsof the retina (Figure13), a highly layeredtissue, where rod and cone photoreceptors constitute more than 75%of the cells. Detailed phylogenetic analyses across species demonstratedthat the last common ancestor of all jawed vertebrates evolved ∼400million years ago with a key role for transcriptional master regulatorpaired box gene 6 (PAX6) in controlling development of eyes and othersensory organs.243−245 Vision is initiated by absorption of lightby photoreceptors, with rod cells acting as photon counters becauseunder dim illumination they can respond to a single photon.246,247 The extreme photosensitivity of the 11-cis-retinylidenechromophore is modulated by the chemical environment of the opsinchromophore binding pocket, yielding different visual pigments thatrespond maximally to light at different wavelengths, causing the so-called“opsin shift” that allows us to discern differencesin color.235,248 Forced twisting of the polyenebackbone away from its normally planar configuration induced by thebinding pocket as well as the electrostatic environment surroundingthe chromophore are major factors that influence its absorbance maximum.249 The key role of binding pocket electrostaticsin spectral tuning of the retinylidene chromophore has been vividlydemonstrated in a rationally engineered cellular retinol-binding protein(CRBP).250
Figure 13.
Structure of the mammalianretina. The retina consists of severallayers of neuronal cells. The innermost photoreceptor layer is embeddedin an epithelial monolayer, known as the RPE. Electron micrographsshow stacks of membranous disks within ROS (left) and COS (right),which contain visual pigments and the associated phototransductionmachinery. The leftmost and rightmost electron micrographs displaythe bacillary structure of the outer segments and their interactionwith the RPE, respectively. A portion of this figure is reproducedwith permission from ref (235). Copyright 2009 Elsevier.
There are similarities and differences between rods and cones.Rod cells are more light-sensitive than cones but saturate at relativelylow levels of light. At the light level at which rods reach saturation,cones generate measurable responses but have an extremely high photonsaturation threshold.242 The number andtypes of cones that collect photons in ambient light differ dramatically,from less than 1% in rats to more than 95% in the ground squirrel.251,252 Both rod and cone cells feature the same structural design. Themost distal portions, called rod/cone outer segments (ROS/COS), arein close contact with the RPE cell layer. These ciliary structuresconnect with the somavia an inner segment endowedwith a high density of mitochondria that supply energy to the highlymetabolically active photoreceptor. The soma in turn connects withsynaptic terminivia inner fibers (Figure13). The specialized photoreceptor cilia containvisual pigments responsible for absorption of light and house allphototransduction proteins needed for amplification and quenchingof the light signal. Here there is yet another difference betweenrods and cones. ROS are made up of a stack of individualized diskssurrounded by the ROS plasma membrane, whereas COS differ by havinga series of invaginations continuously connected with the COS plasmamembrane.
The structure of ROS is better known than that ofCOS because moreadvanced methods were developed for its isolation,253,254 including mouse ROS, which can be altered genetically.255−258 ROS structure has recently been reviewed.66 In short, a mouse ROS with a length of ∼24 μm and adiameter of ∼1.2 μm259 containsabout 600–800 membranous stacked disks which increase the densityof rhodopsin available for photon absorption. ROS extend up to theapical part of the RPE and are tightly enveloped by microvilli ofRPE cells that increase the contact area of these two cell types tofacilitate transfer of substances, including retinoids.
4.2. Structure of the RPE
Without RPEcells, our vision would not be sustainable. The RPE is a monolayerof highly polarized, quasi-hexagonal, epithelial cells. The apicalmembrane of RPE cells lies adjacent to the photoreceptor OS, whereasthe RPE cell basal surface faces Bruch’s membrane.260−262 Contact with ROS is maintained by a highly elaborate network ofmicrovilli visible under the electron microscope (Figure14) which in some species changes shape and elongateswhen the eye is exposed to light. About 20–40 photoreceptorcells project toward a single RPE cell. In humans, this ratio dependson the location of these cells in the eye because the gradient ofrod/cone cells and also slight differences in the dimension of photoreceptorsdictates their packing density. In the periphery, the rod to RPE cellratio is 29 whereas, in the fovea, the cones to RPE cells ratio is22.263 Functions of the RPE cell layerare diverse,59,264 but from the perspective ofthis review, two roles are the most critical.
Figure 14.
Structure of the RPE.Electron micrographs show the apical processesthat extend out from the cell body and interdigitate with photoreceptorouter segments (A, lower resolution; B, higher resolution). In partC, a cross section though an RPE cell shows its cuboidal morphologyand numerous melanin granules. Panel D depicts interactions betweenan RPE cell and photoreceptor outer segments at high resolution. Transmissionelectron micrographs from a C57BL/6J mouse retina were taken at postnataldays 60–66. The RPE cell intimately interacts with the photoreceptorouter segmentvia apical microvilli, thereby supportingphotoreceptor cell function.
The first function of the RPE is to facilitate photoreceptorcellrenewal. To provide optimal signal amplification, membranes of ROSand COS contain high levels of unsaturated lipids.265 These lipids are prone to oxidation in the presence oflight, (photo)reactive retinal, and high oxygen tension,266,267 which are physiological conditions encountered in the retina. Photochemicalreactions induced by light in transparent retinal tissue require adelicate balance between protein, lipid, and metabolite renewal anddamaged component disposal, which when disturbed can lead to rapidand massive retinal degeneration. Impressively, postmitotic rod andcone photoreceptor cells undergo a daily regeneration process wherein∼10% of their OS volume is shed, subsequently phagocytosedby adjacent RPE cells268 and replaced withnewly formed outer segments. Thus, RPE cells dispose of but also accumulatean immense amount of oxidized cellular debris. Indeed it was estimatedthat each RPE cell phagocytoses hundreds of thousands of OS disksover a human lifetime.269 Several potentiallytoxic byproducts are condensation compounds derived fromall-trans-retinal.69,270 Dysfunction of such processesas phagocytosis, lysosomal degradation, and removal of waste productsby the RPE can lead to severe retinopathies, including age-relatedmacular degeneration (AMD).270−274
As a second major function, the RPE expresses key metabolicenzymesrequired for production of the visual chromophore, 11-cis-retinal, and thus comprises an integral part of the retinoid cycle.59,63−65,68,69,71−78,80,81,83,85,261,275−281 In the first step of this metabolic pathway, LRAT, through its abilityto catalyze the formation of retinyl esters that are readily sequesteredby aggregation into lipid droplets called retinosomes, changes themass action ratio to favor retinoid uptake from both photoreceptorsand the choroidal circulation.278,282,283 In addition to LRAT, another critical enzyme called RPE65 (retinoidisomerase) catalyzes the conversion ofall-trans-retinylesters into 11-cis-retinol. Thiscis-retinol is oxidized and sent back to ROS/COS to re-form photoactivevisual pigments. Diffusion likely suffices for this process becausethe chromophore forms, especially in rods, a highly stable covalentcomplex with opsin, thereby driving the transfer of retinoids to photoreceptors.
5. Transformation of Retinoids within the Eye
The photosensitive active retinoid, 11-cis-retinal,is produced in the RPE and delivered to the photoreceptors.59,64,68,69,71,83,278,279,284−287 It has been postulated that additional but less understood transportprocesses take place between Müller cells and cone photoreceptors.85,285,288 Because retinols and retinylesters (but not retinals) are intrinsically fluorescent, their transformationcan be followed by fluorescence induced by two-photon excitation.278,282,283,289−293 Further development of this method guarantees improved understandingof retinoid flow at the subcellular level.
5.1. CanonicalRetinoid (Visual) Cycle
The initial discovery of a light-sensitivepigment in the retinais generally attributed to a German physiologist by the name of Böll,who, in ca. 1876, observed the “purple red color of the bacillary(i.e. rod) layer of the retina” while dissectingthe retina of a frog kept in the dark just prior to the procedure.1,294 The red color of frog rod cells had actually been noted some 25years earlier by Müller, who incidentally first described retinalMüller cells,295 but the color wasattributed to hemoglobin and not further explored.1 Böll made the key observation that the red hue ofthe retina was fleeting, gradually transforming to a yellow colorand then fading over the course of several seconds, leaving the tissuecolorless. Böll insightfully inferred that this light-inducedbleaching of the retina must be reversed when the animal was maintainedin the dark.
A second German physiologist by the name of Kühnegreatly expanded on Böll’s research by showing thata photobleached retina regained its red color when placed in contactwith the RPE and stored in the dark.296,297 This criticalexperiment demonstrated that at least two different tissues were requiredfor the bleach–regeneration cycle proposed by Böll.Kühne used bile salts, which he had on hand from his experimentson digestion, to solubilize the red pigment and then used the preparationto demonstrate a correspondence between its absorption spectrum andthe spectral sensitivity of the retina to light, firmly establishingthe red substance as the visual pigment of rod cells. Kühnenamed the pigment “sehpurpur” or “visual purple”and later referred to it as “rhodopsin”.297
Despite the rapid progress made by Bölland Kühneon the mechanism of light perception, the field was essentially dormantfor the next 50 years. In the 1930s the task of identifying the molecularcomponents of the visual system was taken up by Wald and his colleagues.Among his many achievements, Wald discovered that the chromophoreimparting a red color to rhodopsin was a vitamin A-derived compoundcalled 11-cis-retinal. Moreover, he found photonsstriking this chromophore could induce a change in its configurationfrom an 11-cis, through a series of photointermediates,to anall-trans state, and that this photochemicalreaction, which initiates the series of retinal color changes observedby Böll and Kühne, represents the first step in thesensing of light by the eye.6 He also showedthat, following photoactivation, rhodopsin decomposes into its proteinand retinal components. Wald was the first to lay out a general schemeof chemical reactions, termed the visual or retinoid cycle, whichunderlies visual perception and regeneration.6 Dowling, a student of Wald, through detailed measurements of retinoidflow between photoreceptors and RPE during light exposure and darkadaption, established that the retinal liberated from photooactivatedrhodopsin is rapidly taken up by the RPE and esterified. During darkadaption, the flow of retinoids proceeds in a reverse manner fromthe RPE to photoreceptors for rhodopsin regeneration to occur. Thesefindings firmly established the role of the RPE in visual chromophoreregeneration.298 This visual cycle schemewas refined over the years, and the enzymes responsible for catalyzingthe individual reactions were identified. Our current understandingof the retinoid (visual) cycle is summarized in Figure15.
Figure 15.
Retinoid (visual) cycle. Enzymes (red) and binding proteins (blue)involved in 11-cis-retinal regeneration are foundin both photoreceptor and RPE cells. Metabolic transformations occurringin the RPE take place in the smooth ER, where key enzymes of the visualcycle are located. PC, phosphotidylcholine.
5.2. Cone Visual Cycle
In addition tothe canonical visual cycle, several lines of evidence indicate thatcone photoreceptors might have access to a special source of visualchromophore not available to rods85 (Figure16). For example, after equal levels of bleaching,cones dark-adapt faster than rods by a factor of ∼10.299 In contrast to rods, cones operate in brightlight without saturating, implying that the rate of visual chromophoredelivery to these two cell types must be substantially different.300,301 Additionally, studies have found that the canonical visual cycleis too slow to provide enough visual chromophore to maintain conelight responsiveness under bright light conditions.302 The distribution of retinoids in cone-dominant retinasis substantially different from that of rod-dominant species, withan abundance of 11-cis-retinyl esters found in theretina as opposed to the stores ofall-trans-retinylesters in the RPE.303,304 Three enzymatic activities,namely isomerase, 11-cis- andall-trans-retinyl ester synthase, and retinol dehydrogenase/reductase, areassociated with membrane fractions from chicken neural retina.302 Cultured primary Müller cells from chickenswere found to convertall-trans-retinol added tothe media into 11-cis-retinol and 11-cis-retinyl esters, suggesting that this cell type could be a site ofvisual chromophore productionin vivo.305 Moreover, the 11-cis-retinol-bindingprotein, cellular retinaldehyde-binding protein (CRALBP), is knownto be expressed in Müller cells, providing additional supportfor the above hypothesis.306,307 Although significantprogress has been made in elucidating this alternative pathway, mostof the responsible enzymes have not yet been molecularly characterized.A candidate protein responsible for the alternative retinoid isomerization(all-trans-retinol to 11-cis-retinol)activity found in chicken retinas was identified as dihydroceramidedesaturase-1 (DES-1), a member of the integral membrane hydroxylase/desaturaseenzyme family that contains an eight-His-coordinated di-iron activesite.308,309 This enzyme is a sphingolipid Δ4-desaturasethat converts dihydroceramide into ceramide, with the latter beinga more active signaling molecule.310,311 Interestingly,DES-1 produces mainly 9-cis- and 13-cis-retinoids rather than 11-cis-retinoids fromall-trans-retinol, which is unexpected given the abundanceof 11-cis-retinoids in the chicken retina.308 Further research is required to determine thephysiological relevance of DES-1 to the synthesis of a visual chromophore.
Figure 16.
Putativecone-specific retinoid (visual) cycle. This metabolicpathway is postulated to involve enzymes located in cone photoreceptorand Müller glial cells. The proposed direct isomerization ofall-trans-retinol into 11-cis-retinol isa key difference between this pathway and the canonical retinoid cycle.
5.3. Retinoid-ContainingStructures: Retinosomesand Retinal Condensation Products Found in Healthy and Diseased EyesImaged by Two-Photon Microscopy (TPM)
Once released fromROS,all-trans-retinal is reduced toall-trans-retinol which then diffuses into the RPE. There, this alcohol isesterified by fatty acid in a reaction catalyzed by LRAT. As highlyhydrophobic substances, in part because of the physicochemical propertiesof their long fatty acid esters, these retinoid ester products thencoalesce into lipid droplets termed retinosomes, which are found invertebrate RPE.291 These esters can beminor or major components of lipid droplets that along with phospholipidstotal about 160 molecular species including other fatty acid esters,triglycerides, cholesterol, cholesterol esters, and various proteins.312 Retinosomes are reminiscent of lipid dropletsin other tissues that store other hydrophobic substances. The highlyefficient enzymatic activity of LRAT traps retinol delivered fromthe photoreceptors or from the circulation, whereas retinosomes areabsent in the eyes ofLrat–/– mice deficient in retinyl ester synthesis. The lipid corecontent of retinosomes appears to be homogeneous.312 Retinosomes recruit proteins such as caveolin-1, perilipinsPLIN1-3, sterol carrier protein-2, structural proteins, chaperoneproteins, and redox enzymes.290,312 We postulated thatretinosomes in RPE cells perform functions similar to those of lipiddroplets in other types of cells, suggesting that they are ratherdynamic tissue-specific organelles that change their composition inresponse to fatty acid, cholesterol, andall-trans-retinol availability. Because retinoids are naturally fluorescent,the study of retinosomes could provide important insights into theformation and metabolism of lipid droplets in general, especiallybecause these structures are accessible for real time TPM imagingin vivo.282,289,291,313
Capitalizing on the intrinsicfluorescence ofall-trans-retinyl esters, noninvasiveTPM revealed that retinosomes are elongated structures approximately8 μm long and 1 μm wide with their long axis orientedperpendicularly to the RPE basal surface (Figure17).291 Retinyl esters in retinosomesaccumulate inRpe65–/– mice lacking retinoid enzymatic isomerization. Retinosomesare located close to the RPE plasma membrane and are essential componentsfor 11-cis-retinal production.282,312,313
Figure 17.
Retinoid-containingstructures found in healthy and diseased retinasimaged by two-photon microscopy. The top panel shows a schematic ofthe photoreceptor outer segment–RPE interaction with retinoid-containingretinosomes (red ovals) and retinoid conjugate-containing particles(orange circles) shown with their approximate dimensions (bottom left).In healthy eyes (WT), numerous peripherally located retinosomes (punctuategreen spots) can be visualized. The number and size of these vesiclesare elevated inRpe65–/– mice, owing to excessive accumulation of retinyl esters (bottomcenter). InAbca4–/–Rdh8–/– mice with delayedall-trans-retinal clearance,retinoids are diffusely present throughout the cell, presumably inthe form ofall-trans-retinal-conjugates (bottom right). Scale bars represent 20 μm.
The RPE also accumulates other retinoids, includingretinal condensationproducts.314,315 These metabolites also can bedetected by TPM, and since they are characterized by unique spectralproperties, the intracellular accumulation and distribution of thesecompounds can be monitored independently from other retinoids (Figure17).313 Resulting frominadequate reduction/clearance ofall-trans-retinal,they are often detected as small 1 μm deposits scattered throughoutRPE cells. These condensation products accumulate most prominentlyin mice lacking the ABCA4 transporter and retinol dehydrogenase 8(RDH8) (Figure17).313 They also increase with age because they are delivered from photoreceptors(at least the first steps of condensation reaction occur in ROS) bydaily phagocytosis.270,314 Some investigators considerthese retinoid condensation products as major contributors to retinaldysfunction in Stargardt and age-related macular degeneration (AMD)diseases.270,316,317 Others propose that these biomarkers are merely indicative of inadequateclearance ofall-trans-retinal, which at elevatedlevels may contribute to photoreceptor/RPE degeneration.318−320
6. Proteins and Enzymes of the Retinoid Cycle
Here we describe the molecular properties of proteins involvedin retinoid transformation in the retina including the light receptorrhodopsin.
6.1. Rhodopsin
At the center stage ofphototransduction is rhodopsin, the most extensively studied GPCR.Rhodopsin is the main component of disk and plasma membranes of ROS,accounting for 90 and 75% of their protein content. In disk membranes,the density of rhodopsin translates into about 50% of the entire volumeor surface area.61,66,79,225,226,235,321−324 Thus, it is not surprising that the expression level of rhodopsindictates the size of the ROS.259,325−327 Although rhodopsin is not uniformly distributed throughout disks,84,328 its local high density within each disk allows efficient absorptionof light. Corresponding visual pigments in cone cells of frog retinaare so dense that they form crystalline structures.329 Physiologically, the minimal building block of rhodopsinis a dimer in which only one monomer is activated under normal lightingconditions330−335 (reviewed in refs (336−338)).
The fundamental photochemical reaction of our visual systemis isomerization of 11-cis-retinylidene toall-trans-retinylidene (Figure18). The Schiff base between 11-cis-retinal and Lys296of opsin is protonated to allow a spectral shift to longer wavelengths(at least in rhodopsin, green and blue pigments). It is remarkablethat a ligand, retinal, which is just slightly larger than tryptophan,when photoisomerized causes a reliable change in the conformationof rhodopsin to its activated form that couples with G protein. Thisactivation is accomplished through geometriccis/trans isomerization of the chromophore,182 deprotonationof the Schiff linkage,339 and reorganizationof water molecules within the TMD of this receptor.333,340−342
Figure 18.
Structure and photoactivation of rhodopsin.(a) Crystal structureof ground-state bovine rhodopsin. The Schiff base-linked 11-cis-retinal chromophore is shown in stick representation(red). (b) Photoactivation and regeneration of rhodopsin. (c) Primaryconformational changes observed between ground-state (red, PDB accessioncode 1U19) andactivated, meta II-like (yellow, PDB accession code 3PXO) rhodopsin.
Photoactivation of rhodopsin causesconformational changes thatprovide a binding site for the rod G protein called transducin.343,344 No high resolution structure of the complex is yet available, butlower resolution methods have been informative. On the basis of structuralmass spectrometry techniques, we found that the transition of groundstate rhodopsin to its photoactivated state causes a structural relaxationthat then tightens upon transducin binding.333 Using affinity chromatography, we trapped and purified the photoactivatedrhodopsin–transducin complex. Scanning transmission electronmicroscopy demonstrated about a 221 kDa molecular weight for thiscomplex. A 22 Å structure was calculated from projections ofnegatively stained photoactivated rhodopsin–transducin complexes.The determined molecular envelope accommodated two rhodopsin moleculestogether with one transducin heterotrimer, indicating a heteropentamericstructure for the photoactivated rhodopsin–transducin complex.334 The dimeric structure of rhodopsin in the complexwas confirmed using succinylated concanavalin A as a labeling probe.345 Recently, we used the retinoid chromophores,11-cis-retinal, 9-cis-retinal, andall-trans-retinal to monitor each dimeric rhodopsin monomerwithin a stable complex with transducin. We found that each of thedimeric rhodopsin monomers contributed differently to the pentamericcomplex, indicating a functional distinction between rhodopsin monomersin their oligomeric form.332 For a moredetailed description of the activation events, recent reviews areavailable (refs (66,79, and346−349)).
Two other importantquestions need to be answered about the rhodopsincycle. First, how is the chromophore released and how do opsins recombinewith 11-cis-retinal to regenerate rhodopsin and conevisual pigments? Mechanistically, more information is available forrhodopsin than for cone pigments. Key residues in rhodopsin’sactive site are Lys296, Glu113, and Glu181 (Figure19).60,139,226,227,240,350−355 The 11-cis-retinylidene bond is protonated, andGlu113 is the counterion of this linkage. A counterion is essential,as positively charged groups are extremely rare in the TMD of membraneproteins. Upon illumination of rhodopsin (λmax =500 nm), the chromophore undergoes geometrical isomerization.356,357 Next, rapidly formed and decaying intermediates have been detectedbefore Meta I is observed (λmax = 478 nm). Thoughthe difference in absorption results from relaxation of the chromophore,only small changes in the protein moiety take place at this stage358−360 and it is believed that a switch in the counterion occurs from Glu113to Glu181.324,361−364 Further relaxation combined with Schiff base deprotonation resultsin a signaling form called Meta II (λmax = 382 nm)343 which then interacts with G protein-coupledreceptor kinase 1 and arrestin.60,62,239,365,366 Especially in cold blooded vertebrates, a prominent fraction ofMeta I relaxes into Meta III (λmax = 465 nm), thedifference in the latter’s light absorption derived from ananti tosyn thermal isomerization of theSchiff base double bond.367 From both MetaII and Meta III, a protonated carbinol ammonium ion is formed beforeall-trans-retinal is released from opsin. When regenerationwith 11-cis-retinal subsequently occurs, formationof the Schiff base requires polarization of the carbonyl group of11-cis-retinal and deprotonation of the Lys296 sidechain amine as well as exclusion of water from the active site (Figure19). Regeneration and recombination of 11-cis-retinal with opsin restores the dark state conditionneeded for subsequent photon absorption. Although there is convincingevidence for the roles proposed for Glu113 and Glu181, independentverification of this switch and the whole mechanism of chromophoreregeneration is still lacking.
Figure 19.
Chemical changes in the rhodopsin chromophoreduring photoactivation.The pathway is initiated when 11-cis-retinylidene(i) absorbs a photon, leading tocis/trans isomerization. Then the Glu113 counterion of the protonated Schiffbase becomes protonated, leading to the formation of Meta I rhodopsin(ii). Meta I, in turn, can convert to Meta II rhodopsin(iii), the active signaling form of rhodopsin, or, rarely,to Meta III rhodopsin (iv), a non-signaling form of rhodopsin.Both forms decay through a carbinol ammonium intermediate (v) to form a non-covalent opsin–all-trans-retinalcomplex (vi), which then dissociates to yield freeall-trans-retinal and opsin (vii).
How the chromophore migrates into and out of opsinremains an openquestion. Taking into account the hydrophobic nature of retinal, doesit dissociate into the lipid bilayer or the cytoplasm for the reductionreaction, or could it enter the intradiscal space where condensationproducts of retinal start forming? Similar to other GPCRs, which inaddition to orthosteric-binding sites contain other well-defined allostericligand-binding sites, rhodopsin also has two other retinoid-bindingsites within opsin368,369 in addition to the retinylidenepocket (site 1). Site II is called an entrance site, and the exitsite (site III) is occupied by retinal after its release from siteI. The crystal structure of opsin,370 anopsin structure with another retinoid bound,371 and mutagenesis studies372 all suggestan escape and entrance route for retinal.
6.2. RetinolDehydrogenases (RDH’s)
An integral part of the retinoidcycle is the interconversion ofretinals and retinols.64,287 Enzymes that catalyze this processcan be classified into three major protein families: cytosolic alcoholdehydrogenases (ADH’s) that belong to a medium-chain dehydrogenase/reductasefamily and selected members of the aldo–keto reductase family(AKR’s)373 and microsomal RDH’sthat represent the short-chain dehydrogenase/reductase (SDR) group.374 However, only RDH’s contribute to vision-relatedmetabolism of retinoids.287 The universalredox carriers for these reactions are the dinucleotide cofactorsNAD(H) and NADP(H). In ADH’s and RDH’s they are boundby a Rossmann fold, a classic structural element composed of 6 to7 parallel β-strands flanked by 3–4 α-helices presentin these enzymes (Figure20).375 This structural motif contains a Gly-rich sequence (TGXXXGXG)responsible for its structural integrity and binding of the diphosphateportion of the nucleotide cofactors. An acidic residue binding tothe 2′ and 3′ hydroxyls of the adenine ribose and locateddownstream of the Gly-rich motif confers NAD(H) specificity, whereasNADP(H) binding is dictated by the presence of a basic residue withinthe Gly-rich segment.375 By contrast, AKR’sdo not contain the canonical Rossmann fold. Preferable NADPH bindingoccurs within the characteristic (β/α)8 motifof this protein family.376 Despite theirconserved mode of cofactor binding, ADH and RDH families of oxidoreductasesreveal diverse protein domain architectures and mechanisms of catalysis.ADH’s depend on a catalytic Zn atom bound in the active site,which electrostatically stabilizes the substrate’s oxygen andthus increases the acidity of the alcohol proton.377 In contrast, SDR’s show a Tyr-based catalytic centerwith adjacent Ser and Lys residues.378−381 Here, the deprotonated phenolicgroup of Tyr initially forms a hydrogen bond with the alcohol hydroxylgroup and the deprotonated Tyr residue acts as a catalytic base toextract a proton from the substrate’s hydroxyl group. The hydrideion, extracted from the substrate, can be directly transferred toposition 4 of the nicotinamide ring. In addition to the interactionwith the nicotinamide ring of the cofactor, the reaction intermediateis stabilized by the hydroxyl group of an adjacent serine residue(Figure20).382 Becausethe phenolic group of the Tyr side chain has a pKa value around 10, the ε-amino group of Lys is neededto convert tyrosine to tyrosinate (pKa 7.6), to facilitate catalysis at neutral pH. Additionally, a Lysresidue forms hydrogen bonds to both the 2′- and the 3′-hydroxylgroups of the cofactor’s ribose moiety and thus enforces aproper orientation of the nicotinamide ring to allow a pro-S hydride transfer only. Recently, an Asn residue was shownto stabilize the position of this catalytic Lysvia a conserved water molecule. Thus, the final sequence of the catalytictetrad of SDR’s is composed of Asn, Ser, Tyr, and Lys residues(Figure21).383 Thecatalytic mechanism of AKR’s, in principle, is similar to thatfound in SDR’s with an active site Tyr residue and an assistingLys residue facilitating the deprotonation of the Tyr hydroxyl group.384
Figure 20.
Structure and catalytic mechanism of RDHs.(a) Cartoon representationof a representative SDR family member (type 1 17-β-hydroxysteroiddehydrogenase, PDB accession code 1A27). (b) Hypothetical structure of an RDHwith bound nucleotide (NAD(H) or NADP(H)) and retinoid (all-trans-retinol orall-trans-retinal) substrate. The structuresin panels a and b are depicted in the same orientation. (c) Reversibletransfer of hydride from theS4-face of the nucleotidetoall-trans-retinal to produce pro-R-all-trans-retinol.
Figure 21.
Sequence alignment of known vertebrate RDH’s of the SDRfamily. Glycine residues of the conserved TGXXXGXG, nucleotide-bindingmotif are highlighted in blue, whereas residues comprising the catalytictetrad are highlighted in orange.
The RDH’s represent a microsomal SDR group with anoverallsequence similarity of at least 30% (Figure21) although the catalytic core of these enzymes reveals a much higherhomology with nearly identical folding. Despite these similarities,the mode of membrane binding and membrane topology of specific RDH’sis a matter of controversy. Based on biochemical studies, RDH1 isanchored in ER membranes with the catalytic domain facing the cytoplasm.385 The N-terminal residues are essential for themembrane localization and topology of this enzyme, whereas the C-terminuswas postulated to be involved in stabilization of the protein’smembrane orientation.385 In contrast tothis model, RDH12 was reported to be a glycoprotein carrying an endoglycosidaseH-sensitive sugar modification, which suggests a luminal orientationof this enzyme.386 Yet other models indicatethat the hydrophobic stretch of the catalytic domain in RDH1 and RDH4can contribute to membrane binding.387,388 An interestingcase is RDH8, which localizes to ROS/COS. Efficient transport to theROS/COS is mediated by a signaling sequence at the C-terminus of thisenzyme whereas membrane anchoring is achieved by fatty acylation ofconserved Cys residues.389
Althoughall RDH’s identified in vertebrates can utilizeretinol or retinal, some have wide substrate specificities, and retinoidsare not their preferred physiological substrates.374,390 For example, RDH1, RDH3, RDH4, RDH6, and RDH7 reveal 25–60times higher affinities for androgens thanall-trans-retinol.391−393 In fact only RDH5, RDH8, RDH10, RDH11, RDH12,RDH13, RDH14, and retSDR1 were proven to be expressed in the retinaor RPE, and their roles have been studied in the context of the visual(retinoid) cycle.68,287 Based on their preferred substrategeometry and role in the retinoid cycle, this group of enzymes canbe classified intoall-trans and 11-cis-retinol dehydrogenases. RDH8 and RDH12 belong to the first class,whereas RDH5, RDH11, and RDH10 act on 11-cis-retinoids.Functions of the two remaining enzymes, RDH13 and RDH14, have yetto be adequately assigned.394 Reactionscatalyzed by RDH’s are fully reversible. In a test tube, thenet direction of retinoid interconversion depends on the oxidationstate of the provided cofactor and the ratio between the concentrationsof substrate and product. However, in more complexin vivo systems, the direction of the enzymatic reaction is determined byenzyme specificity for binding either NAD(H) or NADP(H). Under physiologicalconditions the ratio between NAD/NADH is close to 1000;287 thus, RDH’s that bind this cofactorcan contribute significantly only to retinol oxidation. In contrast,the ratio of NADP to its reduced form is about 0.005,395 such that enzymes utilizing this dinucleotidereduce retinal to retinol.
The physiological role and significanceof particular RDH’sin vitamin A homeostasis, retinoic acid signaling, and visual chromophoreregeneration has recently been extensively reviewed68,287,396 and thus will not be describedhere. However, it is worth noting that, in addition to genetic andbiochemical identification of RDH’s, many of these enzymeswere characterizedin vivo in the past few years.397−400 These studies not only provided detailed information about the physiologicalfunction of many RDH’s but also led to the development of animalmodels to investigate their roles in human pathological conditions,including retinal degenerative diseases.401
6.3. Lecithin/Retinol Acyl Transferase (LRAT)
LRAT is the main enzyme that catalyzes retinyl ester formationin most tissues,147,402,403 with the exception being adipocytes that instead exhibit acyl-CoA-dependenttransferase activity of a protein (acyl-CoA:retinol acyltransferaseor ARAT) yet to be identified.404 Consequently,LRAT is critical for uptake and storage of retinoids in peripheraltissues, including RPE cells where it plays a pivotal role in providingsubstrate for visual chromophore regenerationvia the enzymatic activity of RPE65.147,148,405
Localized in the endoplasmic reticulum (ER),LRAT is a 25 kDa bitopic integral membrane protein with a single membrane-spanninghelix localized at the C-terminus406 anda potential membrane-interacting N-terminal domain.407 The C-terminal domain is critical for post-translationaltargeting of the enzyme to the endoplasmic reticulum (ER) in a cytosolicTMD recognition complex-dependent manner.406,408 On the basis of its amino acid sequence and predicted tertiary structure,LRAT is classified as a member of the NlpC/P60 thiol peptidase proteinsuperfamily (Figure22).409 Besides LRAT, there are seven genes in the human genomethat encode proteins belonging to the NlpC/P60 family: two neurologicalsensory proteins (NSE1-2) and five H-ras-like tumor suppressors (HRASLS1-5).409,410 The common feature of LRAT and HRASLS proteins is a 6 amino acidsequence that contains a conserved catalytic Cys residue (NCEHFV).411 Although the structure of LRAThas yet to be determined, recently solved structures of two LRAT-likeproteins, human HRASLS2 and HRASLS3, provide important insights intothe molecular organization of this enzyme.411,412 By analogy to HRASLS proteins, LRAT’s basic structural motifis largely reminiscent of papain-like proteases and consists of afour-strand antiparallel β-sheet, three α-helices, andconserved catalytic residues Cys161, His60, and His72 that definethe active site located at the N-terminus of helix α3 and β-sheetsβ2 and β3, respectively (Figure22).411 Although the overall folding ofLRAT is similar to that of other NlpC/P60 peptidases, there are significanttopological differences derived from a circular permutation withinthe catalytic domain of classical NlpC/P60 proteins.409,413 Consequently, alternatives to peptidase activity evolved in LRAT-likeproteins. The best characterized example is the acyltransferase activityof LRAT, which catalyzes the formation of retinyl esters by transferringan acyl group directly from the sn-1 position of phosphatidylcholine(PC) ontoall-trans-retinol.145,146,414
Figure 22.
Sequence alignment andstructure of LRAT-like acyltransferase enzymes.A protein sequence alignment of all LRAT-like proteins encoded inthe human genome is displayed on the left, showing conserved His andCys residues (orange) that constitute the catalytic triad of thisenzyme family. Hydrophobic C-terminal membrane-anchoring sequencesare colored green. The crystallographic structure of human HRASLS3is shown on the right with the carbon atoms of residues comprisingthe catalytic triad colored orange.
As a consequence of its structural relationship to thiolpeptidases,LRAT adopts an analogous catalytic strategy whereby the deprotonatedCys161 serves as a nucleophile attacking the carbonyl carbon of anester bond in the lipid substrate, both forcing the carbonyl oxygento accept a pair of electrons and transforming the sp2-hybridized carbon into an sp3-hybridized tetrahedral intermediate (Figure23).411,415,416 Collapseof this intermediate results in transient acylation of the proteinby formation of a thioester bond at the Cys161 sulfhydryl group. Concomitantly,1-hydroxy-2-acyl-sn-glycero-3-phosphocholine (Lyso-PC)is liberated as a first product of the reaction. Deprotonation ofthe hydroxyl group of retinol then permits decomposition of the thioesterintermediate and transfer of the acyl group from the enzyme onto retinolto form retinyl esters. Several lines of evidence support this LRATmodel. Initially, a role for Cys161 and His60 in catalysis was derivedfrom site-directed mutagenesis studies where replacement of eitherof these two amino acids abolished retinyl ester formation.417,418 Recently, the proposed mechanism was proved by trapping the catalyticintermediate in the absence of an acyl acceptor and directly detectingthe covalent thioester protein modification by mass spectrometry.415 An identical enzymatic mechanism holds forother LRAT-like proteins.411 Importantly,despite their highly conserved catalytic domains, HRASLS proteinscannot employall-trans-retinol as an acyl acceptorand they lack specificity toward the sn-1 ester cleavage site as well.411 Instead, they catalyze both lipid hydrolysisand acyl transfer onto a variety of enzyme-specific substrates, suchas lyso-phospholipids or phosphatidylethanolamine (PE).411,419−422 These fundamental differences in enzymatic activity do not arisefrom changes in the catalytic mechanism but rather are determinedby subtle modifications in the primary sequence and structure of theseproteins. Despite recent advances in understanding the principlesof catalysis, several critical questions remain, including the evolutionaryand structural basis for adaptation ofall-trans-retinolprocessing by LRAT.
Figure 23.
Catalytic mechanism of LRAT. The enzyme utilizes a pingpong bibi catalytic mechanism.416 In this reaction,the active site Cys nucleophile, which was crystallographically observedto exist in two conformations (i), attacks the sn-1 estergroup to form a tetrahedral intermediate (ii) that collapsesinto a stable acyl-enzyme intermediate with liberation of Lyso-PC(iii).411 The negatively chargedoxygen is stabilized by an oxyanion hole (dotted curve inii). Next,all-trans-retinol binds to the active siteand is activated to produce a nucleophilic attack on the acyl-enzymethioester bond (iv), resulting in formation of a tetrahedralintermediate again stabilized by an oxyanion hole (v)that collapses to release theall-trans-retinyl esterand regenerate the nucleophilic Cys residue. Catalytic His residueslikely promote catalysis by increasing the nucleophilicity of theactive site Cys and by serving as general proton donors/acceptors.
6.4. Acyl-CoA/RetinolAcyltransferase
Early studies concerning the esterificationofall-trans-retinol revealed two independent enzymaticactivities that led toformation of retinyl esters. In addition to lecithin-dependent acyltransfer facilitated by LRAT described above, profound acyl-CoA-dependentactivity has been reported to exist in a variety of tissues, includingsmall intestine, liver, adipocytes, skin, testis, and retina.423−429 In contrast to LRAT, which efficiently utilizes phospholipids asacyl donors, acyl-CoA/retinol acyltransferase (ARAT) requires a preactivatedacyl moiety coupled to coenzyme A (Figure24).
Figure 24.
Formation of retinyl esters catalyzed by acyl-CoA/retinol acyltransferase.
ARAT has never been purified orcloned. However, studies ofLRAT-deficient mice indicatethat the intestinal absorptionof vitamin A decreased to 50% that of wild type animals challengedwith a physiologic dose of retinol.402 Onthe basis of this evidence, ARAT may contribute to retinyl ester formationwithin the intestine and thus facilitate retinoid uptake and packaginginto chylomicrons. A distinct role of acyl-CoA-dependent retinoidesterification has been proposed for the retina. On the basis of enzymaticstudies, retinas isolated from cone-dominant species such as ground-squirreland chicken revealed retinoid isomerase activity that, in contrastto RPE65, convertsall-trans-retinol directly into11-cis-retinol.302,430 This enzymaticreaction is thermodynamically driven by secondary esterification ofnewly produced 11-cis-retinol in a palmitoyl-CoA-dependentmanner. Studies involving primary cultures of chicken Müllercells indicate that inner retinal ARAT enzymatic activity is associatedwith this cell type.85,302,431 Interestingly, the retinal ARAT preferably synthesized 11-cis-retinyl esters428 distinguishingit from intestinalall-trans-ARAT and suggestingthe existence of two or more separate enzymes responsible for acyl-CoA-dependentall-trans-retinol esterification. Intestinal acyl-CoA-diacylglycerolacyltransferase 1 (DGAT1) has been shown to catalyze formation ofretinyl esters in an acyl-CoA-dependent mannerin vitro.402432,433 However,later detailedin vivo studies indicated that DGAT1deficiency did not cause a decline in retinol esterification but rathermarkedly reduced postprandial plasma trigliceride and retinyl esterexcursions by inhibiting chylomicron secretion.434
6.5. Retinoid Isomerase (RPE65)
Our understandingof RPE65 biochemistry has changed dramatically since the lastChemical Reviews article covering the visual cycle.435 Once thought to function solely as a retinoid-bindingprotein, RPE65 has now been conclusively identified as the retinoidisomerase of the canonical retinoid cycle.
6.5.1. Historyand Functional Characterization
RPE65 was identified in theearly 1990s as a conserved, developmentallyregulated, RPE-specific, microsomal membrane protein with an apparentmolecular mass of 65 kDa.436−438Rpe65–/– mice, generated in the Redmond laboratory,exhibited early onset blindness and ocular retinoid abnormalitiesconsisting of a lack of 11-cis-retinoids and overaccumulationofall-trans-retinyl esters, indicating an importantphysiological role for RPE65 in the visual cycle.439 Furthermore, humanRPE65 mutations wereshown to cause Leber Congenital Amaurosis (LCA), a recessive, severechildhood blinding disease.440,441 These two importantfindings established a key role for RPE65 in retinal physiology.
At about the same time, in the seemingly unrelated field of plantbiology, a carotenoid cleavage enzyme from maize called viviparous14 (VP14), involved in the production of abscisic acid from 9-cis-violaxanthin,442 was identifiedand shown to possess significant sequence homology to RPE65.443 Despite the clear relationship of RPE65 toan established enzyme family, efforts to demonstrate either carotenoidoxygenase133 or retinoid isomerase activity444,445 for purified RPE65 were initially unsuccessful, leading to the conclusionthat RPE65 was not an enzyme but instead a retinoid-binding protein.Through visual cycle complementation experiments using unbiased RPEcDNA libraries405 or candidate gene approaches,446,447 RPE65 was finally identified as the visual cycle retinoid isomerasein 2005, about 12 years after the gene was first cloned. This delayin identifying the catalytic function of RPE65 was primarily due toits extreme sensitivity to detergents that are required for its solubilizationand purification.
6.5.2. Evolution
RPE65is evolutionarilyrelated to carotenoid cleavage oxygenases (CCOs), a group of enzymesthat catalyze the oxidative cleavage of various carotenoids and apocarotenoidsas well as certain other olefin-containing compounds, such as lignostilbenes.443,447 However, RPE65 is the black sheep of this family, as it is not knownto possess such oxygenase activity.133 Sequenceidentity between RPE65 and other CCO members varies from ∼20to 40%, but all of these enzymes contain an absolutely conserved setof four His and three Glu residues that are involved in binding ofa required iron cofactor448,449 (Figure25). RPE65 is found only in vertebrates and can usuallybe discriminated from other CCOs on the basis of its characteristicchain length of 533 (±1–2) residues and high (≥70%)sequence conservation. Additionally, sequence alignments have revealedseveral positions where residue type can be predictive of whethera protein likely has RPE65 activity.410,450 The stepsin RPE65 evolution from the true CCOs remain unclear, but this processwas undoubtedly critical for the establishment of the vertebrate visualcycle.451
Figure 25.
Structural alignment of CCO family members.(A) Iron-binding Hisresidues are highlighted in orange, and second sphere Glu residuesare highlighted in blue. (B) Structural superposition of CCO membersof known structure (RPE65, orange; ACO, blue; VP14, pink). These enzymesadopt a 7-bladed β-propeller fold (blades labeled with Romannumerals) with a helical cap on the top face of the propeller thathouses the active site and membrane-binding domain (curved, dashedline), which surrounds the active site entrance indicated by a yellow-greenarrow. The iron cofactor located at the center of the propeller iscoordinated by four conserved His residues (green). The two viewsin panel B differ by a 90° horizontal rotation.
6.5.3. Structure
Thecrystal structureof RPE65, determined by Kiser in 2009 in the Palczewski laboratory,has provided a basis for understanding the substantial biochemicaldata obtained for this enzyme as well as structural features thatdistinguish it from true CCOs452 (Figure26). The basic fold is a seven bladed β-propellercapped on one face by a group of α-helices. This same fold wasalso found for a cyanobacterial apocarotenoid oxygenase (ACO) enzyme,thus confirming that the entire CCO family shares a similar three-dimensionalstructure453 (Figure25). The propeller is sealed within blade VII in a “Velcro”mannervia interactions between the first and laststrands of the core propeller fold. An iron cofactor is located onthe propeller axis directly coordinated by the four conserved Hisresidues mentioned above. Retinyl esters gain access to the iron centerthrough a tunnel that runs along the interface between the helicalcap and the top face of the β-propeller (Figure27). The tunnel terminates within the protein interior, suggestingthat the substrate uptake and product egress pathways are the same.By contrast, the structures of twobona fide CCOs,ACO, and VP14 show two pathways to the iron center, possibly reflectingthe need for the two products of the cleavage reaction to be releasedinto different cellular compartments (i.e., membrane vs cytosol).453,454 In all crystal forms reported to date, RPE65 forms a dimeric assemblyprimarily mediated by an extension to the β-propeller structure455 (Figure26). The otherstructurally characterized CCOs lack this extension and consequentiallyare not dimeric (Figure25).
Figure 26.
Crystal structure ofRPE65 obtained in the presence of native microsomalmembranes. Conserved His residues (green sticks) are shown coordinatingthe catalytic iron (orange spheres). The dimeric structure of RPE65has been observed in multiple crystal forms. This results in a parallelorientation of the membrane-binding surfaces (brown sticks), whichlikely promotes membrane attachment. The membrane-binding surfacesurrounds the entrance to the active site cavity outlined in magentamesh.
Figure 27.
RPE65 active site cavity. The cavity(gray mesh) is predominantlylined by hydrophobic residues that facilitate retinyl ester uptakefrom the membrane. The cavity passes by the catalytic iron and terminatesdeep inside the enzyme core. Residues colored green have been shownthrough mutagenesis studies to be important in maintaining the 11-cis specificity of RPE65 isomerase activity.
6.5.4. Expression and MembraneBinding
RPE65 is expressed almost exclusively in the RPE.436,437 It localizes to the abundant smooth ER of the RPE where other retinoidprocessing enzymes such as LRAT and RDH5 are found.436,456−458 Independent studies have shown that RPE65can associate with RDH5, possibly forming a functional complex.459,460 Our understanding of the interaction of RPE65 with the ER membranehas evolved considerably.284 RPE65 wasonce thought to be peripherally anchored to the ER by reversible post-translationalpalmitoylation.461 Although palmitoylationof a particular Cys residue (position 112 in the bovine sequence)might promote membrane attachment,452,462 more recentbiochemical studies indicate that RPE65 has substantial integral membraneprotein character even though it has no transmembrane spanning segments.284,459 Inspection of the RPE65 crystal structure reveals a surface enrichedin residues capable of interacting with the lipophilic core as wellas headgroups of a phospholipid membrane indicating a monotopic modeof membrane attachment452,455,463 (Figure26). This surface surrounds and helpsform the entrance to the active site tunnel, allowing the enzyme toextract hydrophobic retinoids from the membrane. The structure ofRPE65 determined in the presence of native microsomal phospholipids455 revealed that these membrane-binding residuesundergo major conformational changes upon phospholipid removal bydetergents, consistent with prior biochemical data showing that RPE65adopts structurally and functionally different conformations in itsmembrane-bound versus detergent-solubilized states.464 These findings provide a structural explanation for thewell-known inhibitory effects of detergents on RPE65 activity.465 The membrane-interacting regions are arrangedin parallel (i.e., on the same side) in the RPE65 dimer, allowingthem to reinforce each other in anchoring the protein to the membrane466 (Figure26). This hydrophobicpatch is conserved in all CCOs reported to date, indicating that itis a general structural feature of these enzymes, which allows theextraction of hydrophobic substrates from lipid membranes (Figure25).
6.5.5. Active Site and IronCofactor Binding
The active site cavity of RPE65 is linedprimarily by nonpolarside chains, including several aromatics (Tyr, Phe, and Trp residues),consistent with the hydrophobic nature of the retinyl ester substrate(Figure27). The relatively narrow width ofthe tunnel indicates that a retinyl ester would have to enter theactive site in an extended conformation. A complex structure betweenRPE65 and an intact retinyl ester has not yet been experimentallydetermined, so the orientation of retinyl ester entry remains uncertain.However, modeling studies467,468 as well as the locationof a putative fatty acid molecule observed in the active site of severalRPE65 crystal structures455 (vidainfra) suggest that the retinoid part of the retinyl esterenters first. Like other CCOs, RPE65 contains a ferrous iron cofactorwithin its active site directly coordinated in a distorted octahedralor trigonal bipyramidal fashion by His residues 180, 241, 313, and527 (bovine and human sequence numbering) with Fe–Nε bond lengths between 2.1 and 2.2 Å. A second sphere of highlyconserved and functionally important Glu residues 148, 417, and 469hydrogen bond with the Nδ atoms of His residues 241, 313, and527, respectively. The open iron coordination sites (i.e., sites notoccupied by protein ligands) contain electron density that has beenattributed to a bound fatty acid molecule coordinating the iron throughits free carboxylate moiety based on crystallographic and extendedX-ray absorption spectroscopy (EXAFS) data.455 Notably, iron–carboxylate interactions have been observedin a number of proteins containing the 4-His iron-binding motif, includingphotosystem II, 15-lipoxygenase, and the photosynthetic reaction center.469−471 The functional importance of several RPE65 active site residueshas been probed either directly through site-directed mutagenesisor indirectly by identifying RPE65 mutations that give rise to typeII (RPE65-associated) LCA. Mutations in the iron-coordinating Hisresidues,447 the second sphere Glu residues,472−474 and even a third sphere of residues that form hydrogen bonding interactionswith the second-sphere Glu residues468,475 either abolishor greatly reduce RPE65 activity. In general, any change in activesite amino acid composition is detrimental for activity. Interestingly,mutations of a few residues in close proximity to but not directlybinding the iron cofactor, namely Phe103, Thr147, Tyr338, and Phe526,can alter RPE65 product specificity by changing the ratio of 11-cis to 13-cis-retinol isomers produced(Figure27) (vide infra).467,468 Moreover, even wild-type RPE65 in native membranes produces some13-cis-retinol.171,476
7. Mechanisms of Retinoid Isomerization
The heart of theretinoid cycle is the isomerization reaction.The reaction catalyzed by RPE65 consists of two steps: atrans/cis alkene bond isomerization and an ester bond cleavage. Because ofthis dual activity, RPE65 is frequently referred to as an isomerohydrolase,although it is not hydrolytic water that mediates this process.
7.1. Acyl versusO-Alkyl Cleavagein the Hydrolysis of Esters
The ester cleavage reaction catalyzedby RPE65 is not an ordinary ester bond hydrolysis whereby attack ofwater on the acyl carbon generates a tetrahedral intermediate thatcollapses to form a carboxylic acid and an alcohol (Figure28). Instead, the ester dissociates by cleavage oftheO-alkyl bond. Owing to this distinction, we preferto refer to the process as ester “cleavage” rather than“hydrolysis.” This unusual reaction was identified byusing various isotopically labeled retinols as substrates for theisomerization reaction. It was found that a stereochemical inversionof carbon 15 occurs477,478 and that the 15-hydroxy oxygenis lost171 and replaced by bulk water-derivedoxygen during the isomerization.452 Because11-cis-retinol is thermodynamically less stable thanall-trans-retinol by ∼4 kcal/mol,181 it has been argued that ester cleavage, which typicallyreleases ∼5 kcal/mol of free energy, could be used to drivethe isomerization reaction.479 The requirementof CRALBP and other binding proteins for robust 11-cis-retinol production by RPE65 indicates that product release couldbe rate limiting.476 The unusual estercleavage reaction instead seems to be more important in overcomingthe ∼36 kcal/mol activation energy barrier of the double bondisomerization.171 The retinoid polyenechain is relatively rigid, owing to favorable continuous π-orbitaloverlap. Thus, in order for isomerization of these double bonds tooccur at physiological temperatures, the carbon–carbon bondorder must be temporarily lowered.
Figure 28.
Acyl versusO-alkylester cleavage.
7.2. BinuclearNucleophilic Substitution Mechanismfor Retinoid Hydrolysis/Isomerization
Taking into accountthese biochemical data, three different mechanisms of RPE65-dependentretinoid isomerization have been proposed. The first, proposed byRando and colleagues, involves the attack of an enzyme-associatednucleophile on C11 of the retinyl ester with simultaneous ester dissociationand shuffling of the double bonds (Figure29). The enzyme-linked retinoid intermediate could undergo low-energyrotation around the 11–12 single bond to acis-like conformation. Attack of water or hydroxide on C15 would leadto a reshuffling of the double bonds and dissociation of the retinoid–enzymecovalent intermediate locking the retinoid in an 11-cis configuration. This mechanism, which can be classified as a dualSN2′ nucleophilic substitution reaction, predictsa high degree of 11-cis-retinol product specificityfor RPE65. However, RPE65 does not exhibit such specificity468 and can readily produce 13-cis orall-trans-retinol, depending on the reactionconditions and retinoid-binding proteins used for the assay.476 Structurally, RPE65 also does not possess asuitable nucleophile such as a Cys residue in its active site.452
Figure 29.
Binuclear nucleophilic substitution mechanismof retinoid hydrolysis/isomerization.The key feature is formation of a covalent enzyme–retinoidintermediate that allows rotation around the 11–12 bond.
7.3. UnimolecularNucleophilic Substitution Mechanismfor Retinoid Isomerization
A second mechanism, postulatedby Palczewski and colleagues, proposes a key role for a carbocationintermediate in the isomerization reaction (Figure30). In this reaction involving an SN1′ nucleophilicsubstitution mechanism, ester dissociationvia O-alkylcleavage removes electrons from the polyene, thereby generating aretinylic cation with reduced carbon–carbon bond order allowingbond rotation to take place. Quantum mechanical calculations indicatethat the activation energy oftrans/cis isomerization at the 11–12 bond is reduced to about 18 kcal/mol,consistent with the experimentally determined energy of activation.171 The aromatic residues lining the active sitecavity of RPE65 are ideal for stabilizing carbocation intermediates.480 Following rotation of the 11–12 bond,likely under steric influence by the enzyme active site, water orhydroxide attacks C15, quenching the carbocation with the 11–12double bond in acis configuration. The interiorcavity of RPE65 indeed has a curved shape that can accommodatecis retinoid isomers (Figure27).Importantly, this mechanism allows for the generation of alternativeisomers such as the experimentally observed 13-cis-retinol.171,468 As discussed earlier, retinyliccation formation from retinol or retinyl esters is facilitated byBrønsted or Lewis acids. The iron-coordinated fatty acid observedvia crystallographic and XAS data suggests that the RPE65iron cofactor could promote catalysis by coordinating the ester moietyand facilitating its dissociation.455 Thismechanism thus bears substantial similarity to the well-characterizedCarr–Price type reactions described earlier, with the exceptionthat 11-cis-retinol rather than anhydroretinol isgenerated in the RPE65 active site.
Figure 30.
Unimolecular nucleophilic substitutionmechanism of retinoid isomerization.The key feature is the generation of a carbocation (retinylic cation)intermediate with lowered bond order that allows rotation around the11–12 bond to occur. Dissociation of the ester moiety can befacilitated by a Brønsted or Lewis acid catalyst (X).
7.4. Radical Cation Mechanismfor Retinoid Isomerization
A third mechanism, recently proposedby Redmond and colleagues,posits a role for a retinyl radical cation intermediate in the isomerizationreaction (Figure31). This proposal is basedon the finding that certain radical spin-trap compounds, for exampleN-tert-butyl-α-phenylnitrone (PBN)can inhibit RPE65-mediated isomerization in an uncompetitive manner.481,482 In this proposal, removal of a single π electron from thepolyene by an unknown electron acceptor generates a retinyl radicalcation intermediate that can undergo double bond rotation to formcis isomers followed by quenching of the radical cationto form an 11-cis-retinyl ester. Ester cleavage thenoccurs to produce 11-cis-retinol. Reversal of theisomerization and ester cleavage steps in relation to the carbocationmechanism was proposed based on the finding thatall-trans-retinyl ester could not easily be accommodated in the RPE65 activesite whereas 11-cis- and 13-cis-retinylesters could.467 However, more researchis required to develop the mechanistic details of this scheme.
Figure 31.
Radical cationmechanism of retinoid isomerization.
8. Retinoid-Binding Proteins Relevant to RetinoidTransport to the Eye and the Visual Cycle
Because retinoidsquickly equilibrate between membranes by passagethrough the aqueous phase,483,484 part of their transportlikely takes place by passive diffusion. But this process could befacilitated by ATP-dependent transporters such as ABCA476,485−487 or the channel-like property of STRA6.488−492 Finally retinoids are readily oxidized and isomerized; thus, a setof binding proteins evolved to protect them from these undesirablereactions as well as defend surrounding molecules from retinoid (photo)reactivity.59,73,171,279,285 STRA6, a key transporter involvedin retinol transfer between the RPE and choroidal circulation. andits homologue RBPR2 involved in retinol uptake by the liver493 will be reviewed in detail in another articlein this partial thematic issue ofChemical Reviews. Thus, they will not be discussed here.
8.1. Retinoid-BindingProteins Involved in theRetinoid Cycle
The hydrophobic nature of retinoids limitstheir ability to diffuse in aqueous environments. This property presentsa barrier to retinoid transport from storage sites to sites of utilizationas well as between cellular membranes. Retinoid-binding proteins overcomethis barrier by reversibly binding and sequestering retinoids awayfrom water in internal cavities, greatly facilitating their diffusion.A number of retinoid-binding proteins have been characterized.494 Here, the discussion focuses on those bindingproteins involved in transport of retinols and retinals.
Whenrequired by the body, vitamin A liberated from its hepatic storagesites in Stellate cells by retinyl ester hydrolysis is complexed witha 21 kDa retinoid-binding protein belonging to the lipocalin familycalled retinol-binding protein 4 (RBP4) to form holo-RBP4.24 The main structural feature of holo-RBP4, determinedcrystallographically in the early 1980s, is an eight-stranded β-barrelcore containing a single retinol-binding site (Figure32).495all-trans-Retinol is oriented in the binding pocket with its hydroxyl groupfacing the solvent and the retinyl moiety snuggly bound by a numberof nonpolar residues in a highly complementary fashion, explainingthe high binding specificity toward theall-trans isomer. Holo-RBP4 circulates in the plasma in complex with a secondprotein called transthyretin, which by increasing the molecular weightof the overall complex prevents holo-RBP4 from being excreted in theurine.24,496
Figure 32.
Structures of retinoid-binding proteins involvedin the retinoidcycle. (A) RBP4, (B) Module 2 ofX. laevis IRBP,(C) CRBP I, and (D) CRALBP. Bound retinoids are depicted as stickswith carbon atoms colored orange.
Inside cells, retinol is transported between membranes incomplexwith a second binding protein called cellular retinol-binding protein(CRBP). A member of the intracellular lipid-binding protein family,this compact 15 kDa protein is evolutionarily unrelated to RBP4.497 Two isoforms of the protein, CRBP I and CRBPII with ∼50% sequence identity depending on the species, exhibitdifferential tissue expression.494 CRBPI is expressed in many tissues, including the eye, whereas CRBP IIexpression is restricted to the intestine. The structure of CRBP,similar to that of RBP4, features a β-barrel fold that housesthe retinoid-binding site.495 However,the orientation of retinol in the binding pocket is reversed in relationto that in holo-RBP4 with the β-ionone nearest the mouth ofthe pocket and the hydroxyl group hydrogen bonding with a Glu residueat the base of the cavity. An important role for CRBP inall-trans-retinol trafficking in the retina has been demonstrated inCRBP I–/– mice, which show defectivetransport ofall-trans-retinol from photoreceptorsto ER membranes of the RPE. This is evidenced by accumulation ofall-trans-retinol in the retina and reduced formation ofretinyl esters in the RPE.498
Tworetinoid-binding proteins are specifically expressed in theeye.494 The first of these is an intracellular-bindingprotein CRALBP that displays a binding preference for 11-cis-retinoids.499 CRALBP is expressed predominantlyin two retinal locations, the RPE and Müller glial cells, whereit plays a key role in visual chromophore regeneration by binding11-cis-retinoids, protecting them from esterificationor from photo- or thermal isomerization and facilitating their intracellulartransport.476,500,501 11-cis-Retinal and 11-cis-retinolare the main retinoids associated with CRALBP isolated from bovineretina.499,502 It can also effectively bind 9-cis-retinoids as well but not 13-cis-retinoids.502 The gene encoding CRALBP,RLBP1, was cloned in 1988503 and was localizedto human chromosome 15.504 CRALBP is oneof the founding members of the CRAL-TRIO protein family, members ofwhich share a common three-dimensional architecture.505−507 The crystallographic structure of the CRALBP–11-cis-retinal complex has been determined, revealing a curved bindingpocket with high shape complementarity to 9-cis and11-cis retinoid isomers, consistent with its ligand-bindingpreferences.507 The binding pocket completelyshields the retinoid from solvent. Bound 11-cis-retinalis found in a 6-s-trans, 11-cis,twisted 12-s-cis configuration. The near perfectcis configuration of the 11–12 double bond imposedby the retinoid-binding pocket is likely important to prevent unwantedphotoisomerization and thermal isomerization. A patch of basic aminoacid residues on the protein’s surface probably mediates theinteraction of the protein with acidic phospholipids, which inducedissociation of the bound retinoid.508 Mutationsin the CRALBP gene are associated with several retinal diseases, includingretinitis pigmentosa, Newfoundland rod-cone dystrophy, fundus albipunctatus,and Bothnia retinal dystrophy.509
The second eye-specific retinoid-binding protein is a soluble lipoglycoproteincalled interphotoreceptor-binding protein (IRBP). This large, 136kDa protein is produced by photoreceptors and secreted into the interphotoreceptormatrix, where it is the most abundant extracellular protein. Unlikeother binding proteins that contain a single retinoid-binding site,IRBP has at least three high affinity sites. The protein can alsobind several isomeric forms of retinol and retinal but has a preferenceforall-trans and 11-cis-retinoids.510,511 IRBP can also bind a number of nonretinoid, hydrophobic ligands,but the physiological significance of this capability is not currentlyunderstood.512 The retinoid-binding preferencesand localization of IRBP indicate that it participates in the retinoidcycle by transporting retinoids between photoreceptors, RPE, and Müllercells. Interestingly, IRBP knockout mice do not exhibit acute problemswith vision or dark adaption, and no human disease has been attributedto mutations in the gene encoding IRBP. The protein consists of fourhomologous “modules” generated by gene duplication.512 Full-length IRBP, as observed by negative stainelectron microscopy, adopts a rod-shaped structure that is flexibleand undergoes major conformational changes in the presence of retinoids.513 The crystal structure of an isolated modulefromX. laevis IRBP has been determined, revealinga two-domain architecture.514 Domain Aadopts a ββα-spiral fold whereas domain B formsan αβα sandwich. Although a retinoid-bound structurehas not yet been obtained, two candidate-binding sites have been identifiedthrough molecular modeling: one in the hinge region connecting thetwo domains and a second in domain B. Notably, domain B of the IRBPmodule shares structural similarity with the ligand-binding domainof CRALBP.83
8.2. Structureand Function of ATP-Binding CassetteTransporter Member 4 (ABCA4)
The metabolism of vitamin Ain the eye involves a complex interplay between over twenty differentproteins. The flow of retinoid substrates and products between componentsof the cycle depends on specific binding proteins and transporters.The retinal-specific ATP-binding cassette transporter, ABCA4, playsa special role in this process. ABCA4, a member of the ABCA transportersubfamily, is predominantly expressed in the outer segments of photoreceptorcells where it is located in the rims of rod disk membranes and coneincisures.515−517 As indicated by numerous biochemical studiesand the phenotype ofAbca4–/– mice, the main function of this transporter is to acceleratethe clearance ofall-trans-retinal from ROS/COS.401,518−520
Human ABCA4 is an integral membraneprotein with 2273 residues that form two homologous but nonidenticalparts. Each carries six membrane-spanning helices that constitutea TMD, a soluble cytoplasmic domain (CD) that hosts a canonical nucleotide-bindingsite (NBD) with Walker A and Walker B motifs characteristic of ATP-processingenzymes, and an exocytoplasmic (intradiscal) domain (ECD) (Figure33).76 The overall topologicalmodel for ABCA4 is supported by multiple glycosylation sites identifiedin both ECD domains, whereas CD1 hosts multiple phosphorylation sites.76,521,522 The structure of native bovineABCA4 has been determined by negative stain electron microscopy toa resolution of 18 Å, revealing the overall shape of the molecule.523 On the basis of the accepted model of ATP-driventransport across a lipid membrane, binding of a molecule to be transportedincreases the affinity of NBDs for ATP. The nucleotide then inducesa conformational change of NBDs that come in close contact to forma dimer with the two nucleotide molecules positioned at its interface.This movement is translated onto a TMD that induces the translocationof the substrate molecule across the lipid membrane. Such conformationalchanges have indeed been observed by electron microscopy and hydrogen–deuteriumexchange.523 Subsequent hydrolysis of ATPand dissociation of ADP prompt the separation of NBDs, returning thetransporter to its initial state and completing the cycle. In theabsence of substrate, ABC transporters undergo cycles of slow ATPhydrolysis by individual NBDs, resulting in a basal ATPase activity.Although numerous lipids stimulate ATP hydrolysis by ABCA4,all-trans-retinal or its phosphatidylethanolamine (PE) conjugate,N-retinylidene-PE (N-ret-PE) are the preferredsubstrates.524,525 The topology of this proteinin combination with the logic of the visual cycle suggest that ABCA4acts as an importer, an assumption recently confirmed experimentally.525 Interestingly, ABCA4 still remains the onlyknown example of an importer among eukaryotic ABC transporters.
Figure 33.
Structureand function of ABCA4. (A) Two-dimensional topology diagramof ABCA4. Positions of the Walker A motifs are indicated by blue dashedlines within the CDs. Glycosylation sites are marked with red stars,and an intramolecular disulfide bridge is indicated by S–S.ECD, exocytoplasmic domain; CD, cytoplasmic domain. (B) Electron microscopicstructure of ABCA4 and its dimensions relative to a ROS disk rim.TMDs, transmembrane domains. (C) Structural differences in ABCA4 inthe absence and presence of ATP. (D) Role of ABCA4 in the visual cycleand pathology of elevatedall-trans-retinal. ABCA4flips theall-trans-retinal–PE complex, aproduct ofall-trans-retinal (red line structure)condensation with PE (black line structure with blue sphere indicatingthe headgroup), to the outer leaflet of the disk membrane, allowingdissociation of the complex and subsequent reduction ofall-trans-retinal toall-trans-retinol, which then reentersthe visual cycle (left). Retinal pathology is observed in mice lackingABCA4 and RDH8 activities due to accumulation ofall-trans-retinal and its lipid adducts (right).
The importer activity of ABCA4 has consequences for vitaminA metabolismin photoreceptors (Figure33C). Decay of photoactivatedrhodopsin results in hydrolysis of the opsin-retinylidene Schiff basebond followed by subsequent release ofall-trans-retinalinto the disk membrane.84 The next stepin the visual cycle is reduction of newly liberatedall-trans-retinal toall-trans-retinol byall-trans-RDHs (mainly RDH8 in mouse retina), which associate with the cytoplasmicside of the disk bilayer.64,400 The relatively highhydrophobicity ofall-trans-retinal allows its rapidpartition between the inner and outer leaflets of disk membranes,ensuring its accessibility to RDHs.526,527 However,the high chemical reactivity of aldehydes also leads to spontaneousformation ofall-trans-retinal adducts with primaryand secondary amines. In the biological membrane environment,all-trans-retinal reacts predominantly with PE to formN-ret-PE that, similar to phospholipids, cannot freely flipbetween the inner and outer leaflets of the lipid bilayer. In a testtube, the formation of N-ret-PE in a mixture of chloroform/methanol(2:1) occurs with a bimolecular rate constant of about 3.75 ×10–2 M–1 s–1 while the rate of its hydrolysis is about 7.9 × 10–6 s–1.528 This reversible reaction is highly dependent on the presenceof water. The time scale for this processin vivo remains to be elucidated. Upon a 45% bleach of rhodopsin in wild-typemice, about 24% of releasedall-trans-retinal wasdetected in the form of a conjugate with PE.529 Thus, the main role of ABCA4 could be aiding the movement ofN-ret-PE to the cytoplasmic side of disk membranes. Becausethe Schiff base bond inN-ret-PE is susceptible tohydrolysis, the resultingall-trans-retinal eventuallycan be metabolized toall-trans-retinol. Taking intoaccount the partition of retinals into the lipid bilayer and a relativelyslow rate of ATP hydrolysis by ABCA4 equivalent to three enzymaticcycles per second, one can assume that only a fraction of the totalall-trans-retinal is processed by ABCA4.530 Nevertheless, the activity of ABCA4 turns out to be essentialfor loweringall-trans-retinal concentrations belowthe threshold that could cause photoreceptor toxicity.401
8.3. Condensation Reactionsof Retinal
Delayed clearance ofall-trans-retinal from thephotoreceptors after light exposure can have dramatic pathophysiologicalconsequences. In cultured ARPE-19 cells derived from human RPE cells,exposure to 10 μMall-trans-retinal causedprofound cytotoxicity after less than 1 min of incubation by inducinga Ca2+-driven apoptotic cell death pathway.318 Because rhodopsin concentration in ROS reaches4 mM,257 bleaching of just 0.5% of thetotal amount could potentially generate toxic levels ofall-trans-retinal if this retinoid is not efficiently removed from the retina.The mechanism ofall-trans-retinal toxicity involvesthe generation of superoxide radical, singlet oxygen, and peroxideswhen irradiated with UVA or blue light.279 In addition,all-trans-retinal can stimulate increasedlevels of reactive oxygen species in a NADPH oxidase-dependent manner.531,532 Unless quickly dissipated, such reactive oxygen species can causeoxidative damage to lipids and proteins that compromise photoreceptorstructure and function.533 Yet anotherconsequence of elevatedall-trans-retinal concentrationsis the accelerated formation and accumulation of bis-retinoid lipofuscinchromophores within retina and RPE.314,401,520,534,535 Although formation of the Schiff base adduct ofall-trans-retinal with PE is fully reversible, reaction ofN-ret-PE with a second molecule ofall-trans-retinalcan initiate a cascade of irreversible nonenzymatic conversions thatlead to the production of fluorescent diretinal compounds, includingpyridinium bisretinoid (A2E) (Figure34) andretinaldehyde dimer (RALdi) (Figure35). Thecommon precursor in the biosynthesis of these fluorophores is protonatedN-ret-PE, that undergoes spontaneous tautomerizationvia an H-shift [1,6] generating phosphatidyl analogs ofenamine.536,537 The subsequent reaction witha second molecule ofall-trans-retinal can occurthrough amine condensation followed by 6π-electrocyclizationto generate a reduced form of A2E, phosphatidyl-dihydropyridine bisretinoid(A2PE-H2) (Figure34). Alternatively,the nucleophilic carbon 20 of theN-ret-PE tautomercan react with carbon 13 ofall-trans-retinalvia [1,4] conjugate addition. Consequent Mannich condensationthen leads to the RALdi precursor (Figure35).534 The strong kinetic isotope effectobserved in RALdi synthesis as compared to that found for A2E supportsthis reaction mechanism537,538 Aromatization of thedihydropyridine in A2PE-H2 eliminates two hydrogens toyield phosphatidyl-pyridinium bisretinoid (A2PE) whereas the RALdiprecursor spontaneously rearranges to eliminate the amino group ofPE (Figures34 and35, respectively). In the final stage of A2E biosynthesis, its immediateprecursor, A2PE, is either hydrolytically cleaved by enzymatic actionof phosphodiesterase, which can occur in ROS before internalizationby the RPE, or undergoes nonenzymatic acid-catalyzed hydrolysis insideRPE phagosomes.314 As a result of incompletelysosomal degradation, these byproducts of retinal metabolism accumulatein the RPE as residual bodies called lipofuscin. These bodies providea useful fluorescent marker for lipofuscin quantification in the livingeye, which can serve as long-lasting evidence of excessive past accumulationofall-trans-retinal.539−541 Indeed, the primaryfluorescent components of lipofuscin areall-trans-retinal conjugates such as A2E and RALdi.542 A2E was shown to be further metabolized by horseradish peroxidase,543 but there is yet no evidence that this processoccursin vivo. Lipofuscin granules contain severalchromophores absorbing UV, blue, and green light whereas A2E actsas an acceptor of energy from their photoexcited states and dissipatesthat energy mainly by thermal deactivation and partly by emittingfluorescence.544 Considering that lipofuscingranules contain potent photosensitizers, quenching of excited statesof lipofuscin photosensitizers by A2E may play a protective antioxidantrole. Yet, A2E also exhibits some minor photoreactivity and severalexperimentsin vitro demonstrated that A2E and itsoxidation products can be cytotoxic to RPE cells and trigger complementactivation and inflammatory signaling.317 It remains to be established whether or not these deleterious propertiesof A2E are physiologically relevant.
Figure 34.
Mechanism of A2E formation fromall-trans-retinaland phosphatidylethanolamine (R-NH2).
Figure 35.
Mechanism ofall-trans-retinal dimer formationfromall-trans-retinal and phosphatidylethanolamine(R-NH2).
The connection betweendelayedall-trans-retinalclearance, lipofuscin buildup, and retinal degeneration is exemplifiedby Stargardt disease, a human condition in which juvenile maculardegeneration is caused by mutations in both alleles of theABCA4 gene.487,545 Consequently, retinal samplescollected from these patients revealed elevated levels of N-retinylidene-PEand overaccumulation of A2E in the RPE.529 Because the cytotoxicity of retinal conjugates has been widely accepted,it is believed that the precipitating cause of retinal degenerationin Stargardt patients is the deterioration of RPE cells responsiblefor maintenance of photoreceptors.546 However,recent studies ofAbca4–/–Rdh8–/– double knockout mice that closely recapitulate human retinalpathological conditions indicate thatall-trans-retinalitself may play a decisive role in light-induced photoreceptor degeneration.318,320,531
9. Aberrationsin the Retinoid Cycle and HumanRetinal Diseases
Studies of mutations in the retinoid cyclegenes can teach us aboutthe structure–function of key visual proteins and enzymes aswell as the cell biology of rod photoreceptor cells, one of our mostmetabolically active neurons. Most of the genes encoding retinoidcycle enzymes are associated with retinal disease (Figure36).64,76,77,285,547−552 For example, point mutations in the rod opsin gene are the mostcommon cause of autosomal dominant retinitis pigmentosa (RP). Themost frequent mutation and the first identified as causing blindnessis P23H,553 accounting for 10% of humancases of autosomal dominant RP. Only a few mutations, including severetruncation of the opsin gene554 and thec.448G > A (p.E150K) mutation,555,556 are inheritedin anautosomal recessive manner. RDH12, LRAT, and RPE65, when inactivatedby mutations, cause juvenial forms of blindness called LCA. For RPE65the severity of disease has been associated with the degree of residualenzymatic activity.274,281,557 Mutations in the transporter ABCA4 cause accumulation of condensationproducts ofall-trans-retinal.520 Here, lack of a functional transporter causes Stargardtdisease,545,558,559 a juvenial form of macular degeneration, whereas some other changeswere demonstrated to be associated with AMD.560,561 Mutations in other genes, such as those encoding RDH5 or IRBP, areassociated with slowly progressive retinal degeneration.562−565
Figure 36.
Retinal diseases caused by defects in visual cycle enzymes. Therapeuticagents used in the treatment of these conditions are indicated.
This relationship between geneticmutations and retinal diseasealso provides information about the physiological role of these proteinsin the chemistry of the retinoid cycle, as revealed by many animalmodel studies. Such generated animal models are also required fortesting possible remedies for these blinding diseases.566 In some cases the treatment is highly pertinentbecause of the significant number of patients involved and the socioeconomiccost of their associated blindness. In other cases, the small numberof affected individuals decreases the likelihood of commercially developinga possible treatment. But animal studies are still vital for thoseafflicted and also may lead to breakthroughs applicable to more commonretinal diseases.
Natural and genetically altered animal modelshave been used toinvestigate the effects of retinoid supplementation in treatment.Thus, blockage in production of 11-cis-retinal wasovercome with oral delivery of another photosensitive but chemicallymore stable retinoid, 9-cis-retinal (Figure36)567−571 (reviewed in ref (76)). Retinylamine and other primary amines were employed to bufferthe toxic effect ofall-trans-retinal either by inhibitingthe visual cycle or by chemically trapping an excess ofall-trans-retinal in the form of a Schiff base when it could not be effectivelycleared from photoreceptor cells by reduction toall-trans-retinol (Figure36).318,572,573 The toxicity ofall-trans-retinal could be a major contributor to retinal degeneration inseveral human diseases, including Stargardt disease.318,401,570 Moreover, there are also experimentaltherapies that utilize carotenoids, retinoid precursors that comprisean integral part of human macula.574 Thetherapeutic landscape for inherited and acquired retinal diseasesis rapidly evolving. Just a few years ago, it was difficult to imaginestrategies for treating these diseases with practical pharmacologicalapproaches. But today, many solid scientific findings provide hopethat these chronic diseases will become manageable.
10. Final Conclusions: The Retinoid Cycle and WholeBody Retinoid Metabolism
Vitamin A must be adequately distributedwithin the body to maintainthe biological function of retinoids in the peripheral tissues andthe production of visual chromophore in the eye. Transport of vitaminA is facilitated by RBP4, and its cellular receptor, STRA6, functionallycouples with LRATvia CRBP. Recent studies by vonLintig and co-workers suggest that ocular vitamin A uptake is favoredover other peripheral tissues in vitamin A deficient states.492 In contrast to other cell types, an overdoseofall-trans-retinol does not resultin excessive accumulation of retinyl esters in the RPE.567,575 Although the pivotal role of STRA6 and LRAT in vitamin A uptakeis widely recognized, the mechanisms that govern “buffering”of vitamin A within the eye still remain unknown. This also appliesto the potential role of light in the regulation of retinoid metabolism,especially the rate of the retinoid (visual) cycle. Despite earlywork showing that retinal G protein-coupled receptor, expressed inthe RPE, provides light-dependent modulation ofall-trans-retinyl ester synthesis and degradation as well as influences RPE65activity in mice,576 the molecular mechanismof light-induced stimulation of retinoid isomerization remains anintriguing biological mystery.
A classic unsolved problem ishow retinyl esters traffic from lipidstorage droplets in the retina to the ER and RPE65. It is still unknownif retinyl palmitate is transported intact or needs to undergo hydrolysisfollowed by re-esterification by LRAT in the ER close to RPE65 tosustain robust retinoid isomerization. Yet another aspect relatedto the organization of vitamin A metabolism in the eye is the putativeclose interaction of proteins involved in this process. Despite severalreports indicating interaction of particular proteins, the isomerizationcomplex containing RPE65, RDH5, LRAT, or CRALBP has not been purifiedor reconstitutedin vitro. Determining the moleculardetails ofall-trans-retinol processing proteinsat the atomic level through structural biology represents a complementaryapproach to biochemical studies. Until recently, progress in thisfield has been marked only by NMR or crystal structures of solubleretinoid-binding proteins such as CRBPs, IRBP, and CRALBP.507,577,578 The structure of RPE65452,455 definitely provided an incentive for obtaining structures of othermembrane associated proteins including LRAT and RDHs.
Furtherstructural studies of RPE65 and other CCOs with the goalof obtaining high resolution complexes with retinoids and relatedcompounds will help resolve lingering issues concerning the mechanismof the retinoid isomerization reaction. Spectroscopic approaches,which are becoming more feasible with the advent of improved expressionand purification methods, will also be of great utility in unravelingthe fine details of the bioinorganic chemistry of this enzyme family.
Analysis of the isomeric composition of retinoids in rod- and cone-dominantretinas reveals striking differences.579 Significant amounts of 11-cis-retinyl esters and11-cis-retinol in ground squirrel and chicken retinasnot observed in rod-dominant mice, rat, or cow are associated withalternatives to RPE65 and LRAT enzymatic activities responsible forformation of these retinoids.302,305,430 These findings suggest the existence of an unconventional cone-specificvisual chromophore regeneration pathway. Though supported by biochemicaldata, this pathway remains only a concept without molecular identificationof its protein components. Thus, efforts to clone or purify enzymesinvolved in cone regeneration currently represent an exciting challenge.Newly generated 11-cis-retinal is needed to reformlight-sensitive visual pigments. Thus, another challenging questionin studies of the rhodopsin (and other visual pigments) cycle is todetermine how the chromophore enters and exits the binding pocket.
In short, recent decades have witnessed an improved understandingof the retinoid cycle and retinoid metabolism in general. Indeed,it would be improper to treat the retinoid cycle as a separate entitybecause it is so dependent on the metabolism of the whole organism(Figure37). The coming years will certainlybring further molecular characterization of these processes from approachesinvolving combinations of human molecular genetics, characterizationof newly generated animal models, advanced imaging techniques, andstructural studies.
Figure 37.
Key proteins involved in the transport of retinol fromthe liverto target tissues. Retinol travels in the circulation bound to RBP.In turn, RBP complexes with a transthyretin (TTR) tetramer, whichprevents filtration of RBP across the glomeruli of the kidney. Holo-RBPcan dissociate from the TTR tetramer and bind to the retinol membranetransporter, STRA6.all-trans-Retinol is picked upon the cytoplasmic side of STRA6 by CRBP, which shuttles the retinoidto the ER of the RPE. There it is esterified by LRAT to formall-trans-retinyl-esters, which are either used as substratesfor visual chromophore production or stored in lipid bodies knownas retinosomes: atROL,all-trans-retinol; atRE,all-trans-retinyl ester.
Acknowledgments
We are gratefulto Dr. Johannes von Lintig for helpful discussionsand to Drs. Leslie T. Webster, Jr., and Malgorzata Rozanowska forcritical comments on the manuscript. We thank Grazyna Palczewska forimages used in Figure17, Debarshi Mustafifor images used in Figures13 and14A–C, and Dr. Sanae Sakami for the EM imageof the photoreceptor–RPE interface (Figure14D). This research was supported, in whole or in part, by EY009339(K.P.) and EY021126 (K.P.) grants from the National Institutes ofHealth. K.P. is John H. Hord Professor of Pharmacology.
Glossary
Symbols and Abbreviations
- A2E
pyridinium bisretinoid
- A2PE
phosphatidyl-pyridiniumbisretinoid
- A2PE-H2
phosphatidyl-dihydropyridinebisretinoid
- Abca4
ATP-binding cassette, subfamilyA (ABC1), transporter member 4
- Acetyl-CoA
acetyl-coenzyme A
- ACO
apocarotenoid 15,15′-oxygenase
- AMD
age-related maculardegeneration
- ARAT
acyl-CoA:retinol acyltransferase
- CCO
carotenoid cleavage oxygenase
- CD
cytoplasmic domain
- COS
cone outer segment(s)
- CRALBP
cellular retinaldehyde-bindingprotein
- CRBP
cellularretinol-binding protein
- CrtI
carotene desaturase
- CrtIso
carotenoid isomerase
- Des-1
dihydroceramide desaturase-1
- DGAT1
acyl-CoA:diacylglycerol acyltransferase1
- DMAPP
dimethylallyldiphosphate
- ECD
exocytoplasmic domain
- ER
endoplasmic reticulum
- GGPP
geranylgeranyl diphosphate
- HPLC
high performance liquid chromatography
- HRASLS
H-ras-like tumorsuppressors
- IDI
IPP isomerase
- IPP
isopentenyl diphosphate
- IRBP
interphotoreceptor-binding protein
- LCA
Leber Congenital Amaurosis
- LRAT
lecitihin:retinol acyltransferase
- Lyso-PC
1-hydroxy-2-acyl-sn-glycero-3-phosphocholine
- MEP
methylerythritol 4-phosphate
- NBD
nucleotide-binding domain
- PE
phosphatidylethanolamine
- PPARγ
peroxisomeproliferator-activatedreceptor γ
- RALdi
retinaldehyde dimer
- RBP4
serum retinol-binding protein 4
- RDH
retinol dehydrogenase
- RetSat
retinol saturase
- ROS
rod outer segment(s)
- RPE65
retinoid isomerase
- RPE
retinal pigmentepithelium
- SDR
short-chaindehydrogenases/reductase
- SR-BI
scavenger receptor class-B type-1
- STRA6
stimulated by retinoic acid 6
- TMD
transmembrane domain
- TPM
two-photon microscopy
- TTR
transthyretin
Biographies
Philip D. Kiser was bornin 1980 in Olney, IL. He received a Pharm.D.degree from the St. Louis College of Pharmacy in 2005. During pharmacyschool he performed research in the laboratories of Dr. Eric Barkerat Purdue University and Dr. Thomas Baranski at Washington University.He then entered graduate school in the Department of Pharmacologyat Case Western Reserve University (CWRU), where he earned a Ph.D.in molecular pharmacology in 2010 under the guidance of Prof. KrzysztofPalczewski. He then joined the Department of Pharmacology as an instructorin 2011, following a brief postdoctoral appointment. His primary researchinterests pertain to the biochemistry and structural biology of retinoidand carotenoid metabolism. He is also involved in teaching pharmacologyto undergraduate, graduate, and professional students at CWRU. Inhis free time, he enjoys playing and listening to music and fixingthings around the house with his three sons.
Marcin Golczakobtained his Master’s degree in Biotechnologyfrom the Wroclaw University of Technology, Poland, in 1999. He continuedresearch, working on Ca2+-binding proteins, at the NenckiInstitute of Experimental Biology, Polish Academy of Science in Warsaw,Poland, where he received his Ph.D. in 2003. During his subsequentpostdoctoral research in Prof. Krzysztof Palczewski’s laboratoryat the University of Washington, Seattle, WA. and Case Western ReserveUniversity, Cleveland, OH, Dr. Golczak has focused on vitamin A metabolism,especially the enzymatic pathway called the retinoid (visual) cycle,that leads to regeneration of the visual chromophore required forsight. Currently, he holds an Instructor position in the Departmentof Pharmacology at Case Western Reserve University. His main interestsare the elucidation of molecular mechanisms of vitamin A homeostasisin the eye and the development of small molecule-based therapeuticstrategies against light-induced retinal degeneration.
KrzysztofPalczewski, Ph.D., has been Chair and John H. Hord Professorof Pharmacology at Case Western Reserve University since 2005. Beforethat he was professor of Ophthalmology, Pharmacology and Chemistryat the University of Washington, Seattle. Dr. Palczewski’snotable scientific achievements span the fields of vertebrate vision,structural biology, and pharmacology, including the translation ofsome of his basic discoveries into treatments for human blinding diseases.He is author/coauthor of scientific publications involving topicssuch as the following: structures of rhodopsin, retinoid isomeraseRPE65, and other retinal proteins; transmembrane signaling; the “retinoid”or visual cycle; animal models of human blinding diseases; and retinalimaging techniques. Palczewski is recipient of several prestigiousawards, including the following: the Jules and Doris Stein Researchto Prevent Blindness Professor award; the Cogan Award from the Associationfor Research in Vision and Ophthalmology; the Knight’s Crossof the Order of Merit of the Republic of Poland award; The Roger HJohnson Macular Degeneration Award; and The Friedenwald Award. Healso is one of four recipients of the 2012 Award from Foundation forPolish Science, the highest recognition available to Polish scientistsliving outside that country. Professor Palczewski also serves on theeditorial boards of several well recognized biochemical journals andNational Institute of Health scientific panels appropriate to hisfield.
Author Contributions
* Contributed equally
M.G.andK.P. are inventors of one or more of patents (7,706,863; 8,324,270;7,951,841; 8,338,394; and pending 20120295895) submitted by Universityof Washington and Case Western Reserve University and briefly discussedin the paper. K.P. is CSO of Polgenix Inc. and Visum Inc., developingfurther some of these technologies.
Funding Statement
National Institutes of Health, United States
References
- Wolf G.J. Nutr.2001, 131, 1647. [DOI] [PubMed] [Google Scholar]
- Littré E.Oevres complètesde Hippocrate; J. B. Baillière: Paris, 1861; p 150. [Google Scholar]
- McCollum E. V.; Davis M.Nutr.Rev.1973, 31, 280. [DOI] [PubMed] [Google Scholar]
- Moore T.Biochem. J.1930, 24, 692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karrer P.; Morf R.; Schoepp K.Helv. Chim. Acta1931, 14, 1431. [Google Scholar]
- Wald G.Nature1968, 219, 800. [DOI] [PubMed] [Google Scholar]
- Olson J. A.Biochim. Biophys.Acta1960, 37, 166. [DOI] [PubMed] [Google Scholar]
- Olson J. A.Am. J. Clin. Nutr.1961, 9(4)Pt 21. [DOI] [PubMed] [Google Scholar]
- Olson J. A.J. Biol. Chem.1961, 236, 349. [PubMed] [Google Scholar]
- Zachman R. D.; Olson J. A.J. Biol. Chem.1963, 238, 541. [PubMed] [Google Scholar]
- Dunagin P. E. Jr.; Zachman R. D.; Olson J. A.Biochim. Biophys. Acta1964, 90, 432. [DOI] [PubMed] [Google Scholar]
- Olson J. A.J. Lipid Res.1964, 5, 281. [PubMed] [Google Scholar]
- Olson J. A.J. Lipid Res.1964, 5, 402. [PubMed] [Google Scholar]
- Zachman R. D.; Olson J. A.Nature1964, 201, 1222. [DOI] [PubMed] [Google Scholar]
- Dunagin P. E. Jr.; Meadows E. H. Jr.; Olson J. A.Science1965, 148, 86. [DOI] [PubMed] [Google Scholar]
- Olson J. A.; Hayaishi O.Proc. Natl. Acad. Sci. U.S.A.1965, 54, 1364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zachman R.D.; Olson J. A.J. Lipid Res.1965, 6, 27. [PubMed] [Google Scholar]
- Nath K.; Olson J. A.J. Nutr.1967, 93, 461. [DOI] [PubMed] [Google Scholar]
- Olson J. A.Pharmacol. Rev.1967, 19, 559. [PubMed] [Google Scholar]
- Olson J. A.Vitam. Horm.1968, 26, 1. [DOI] [PubMed] [Google Scholar]
- Goodman D. S.; Huang H. S.Science1965, 149, 879. [DOI] [PubMed] [Google Scholar]
- Huang H. S.; Goodman D. S.J. Biol. Chem.1965, 240, 2839. [PubMed] [Google Scholar]
- Goodman D. S.; Huang H. S.; Shiratori T.J. Biol. Chem.1966, 241, 1929. [PubMed] [Google Scholar]
- Kanai M.; Raz A.; Goodman D. S.J. Clin. Invest.1968, 47, 2025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fidge N. H.; Smith F. R.; Goodman D. S.Biochem. J.1969, 114, 689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goodman D. S.Am. J. Clin. Nutr.1969, 22, 963. [DOI] [PubMed] [Google Scholar]
- Smith F. R.; Goodman D. S.J. Clin. Invest.1971, 50, 2426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith J. E.; Goodman D. S.J. Clin. Invest.1971, 50, 2159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goodman D. S.Biochem. Soc.Symp.1972, 287. [PubMed] [Google Scholar]
- Muto Y.; Goodman D. S.J. Biol. Chem.1972, 247, 2533. [PubMed] [Google Scholar]
- Smith F. R.; Goodman D. S.; Arroyave G.; Viteri F.Am. J. Clin. Nutr.1973, 26, 982. [DOI] [PubMed] [Google Scholar]
- Smith J. E.; Muto Y.; Milch P. O.; Goodman D. S.J. Biol. Chem.1973, 248, 1544. [PubMed] [Google Scholar]
- Chytil F.; Ong D. E.Annu. Rev. Nutr.1987, 7, 321. [DOI] [PubMed] [Google Scholar]
- Ong D. E.; Takase S.; Chytil F.Ann. N. Y. Acad. Sci.1987, 513, 172. [DOI] [PubMed] [Google Scholar]
- Porter S. B.; Ong D. E.; Chytil F.Int. J. Vitam. Nutr. Res.1986, 56, 11. [PubMed] [Google Scholar]
- Ong D. E.; Chytil F.Vitam. Horm.1983, 40, 105. [DOI] [PubMed] [Google Scholar]
- Ong D. E.; Chytil F.Ann. N. Y. Acad. Sci.1981, 359, 415. [DOI] [PubMed] [Google Scholar]
- Chytil F.; Ong D. E.Fed. Proc.1979, 38, 2510. [PubMed] [Google Scholar]
- Ong D. E.; Chytil F.J. Biol. Chem.1979, 254, 8733. [PubMed] [Google Scholar]
- Chytil F.; Ong D. E.Vitam. Horm.1978, 36, 1. [DOI] [PubMed] [Google Scholar]
- Ong D. E.; Chytil F.Nature1975, 255, 74. [DOI] [PubMed] [Google Scholar]
- Hagen E.; Myhre A. M.; Smeland S.; Halvorsen B.; Norum K. R.; Blomhoff R.J. Nutr. Biochem.1999, 10, 345. [DOI] [PubMed] [Google Scholar]
- Senoo H.; Wake K.; Wold H. L.; Higashi N.; Imai K.; Moskaug J. J.; Kojima N.; Miura M.; Sato T.; Sato M.; Roos N.; Berg T.; Norum K. R.; Blomhoff R.Comp. Hepatol.2004, 3Suppl 1S18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Higashi N.; Imai K.; Sato M.; Sato T.; Kojima N.; Miura M.; Wold H. L.; Moskaug J. J.; Berg T.; Norum K. R.; Roos N.; Wake K.; Blomhoff R.; Senoo H.Comp.Hepatol.2004, 3Suppl 1S16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nagy N. E.; Holven K. B.; Roos N.; Senoo H.; Kojima N.; Norum K. R.; Blomhoff R.J. Lipid Res.1997, 38, 645. [PubMed] [Google Scholar]
- Green M. H.; Green J. B.; Berg T.; Norum K. R.; Blomhoff R.Am. J. Physiol.1993, 264, G509. [DOI] [PubMed] [Google Scholar]
- Norum K. R.; Blomhoff R.Am. J. Clin. Nutr.1992, 56, 735. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Green M. H.; Norum K. R.Annu. Rev. Nutr.1992, 12, 37. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Senoo H.; Smeland S.; Bjerknes T.; Norum K. R.J. Nutr. Sci.Vitaminol. (Tokyo)1992, Spec No327. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Skrede B.; Blomhoff H. K.; Norum K. R.J. Nutr. Sci. Vitaminol.(Tokyo)1992, Spec No473. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Green M. H.; Green J. B.; Berg T.; Norum K. R.Physiol. Rev.1991, 71, 951. [DOI] [PubMed] [Google Scholar]
- Norum K. R.; Blomhoff R.Tidsskr. Nor. Laegeforen.1991, 111, 1078. [PubMed] [Google Scholar]
- Blomhoff R.; Green M. H.; Berg T.; Norum K. R.Science1990, 250, 399. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Skrede B.; Norum K. R.J. Intern. Med.1990, 228, 207. [DOI] [PubMed] [Google Scholar]
- Botilsrud M.; Holmberg I.; Wathne K. O.; Blomhoff H. K.; Norum K. R.; Blomhoff R.Scand. J. Clin. Lab. Invest.1990, 50, 309. [DOI] [PubMed] [Google Scholar]
- Green M. H.; Green J. B.; Berg T.; Norum K. R.; Blomhoff R.J. Nutr.1988, 118, 1331. [DOI] [PubMed] [Google Scholar]
- Blomhoff R.; Helgerud P.; Dueland S.; Berg T.; Pedersen J. I.; Norum K. R.; Drevon C. A.Biochim. Biophys. Acta1984, 772, 109. [DOI] [PubMed] [Google Scholar]
- Tseng S. C.; Hirst L. W.; Farazdaghi M.; Green W. R.Invest. Ophthalmol. Vis.Sci.1987, 28, 538. [PubMed] [Google Scholar]
- McBee J. K.; Palczewski K.; Baehr W.; Pepperberg D. R.Prog. RetinalEye Res.2001, 20, 469. [DOI] [PubMed] [Google Scholar]
- Okada T.; Ernst O. P.; Palczewski K.; Hofmann K. P.Trends Biochem. Sci.2001, 26, 318. [DOI] [PubMed] [Google Scholar]
- Filipek S.; Stenkamp R. E.; Teller D. C.; Palczewski K.Annu. Rev. Physiol.2003, 65, 851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ridge K. D.; Abdulaev N. G.; Sousa M.; Palczewski K.Trends Biochem. Sci.2003, 28, 479. [DOI] [PubMed] [Google Scholar]
- Moise A. R.; von Lintig J.; Palczewski K.Trends Plant Sci.2005, 10, 178. [DOI] [PubMed] [Google Scholar]
- Travis G. H.; Golczak M.; Moise A. R.; Palczewski K.Annu. Rev. Pharmacol.Toxicol.2007, 47, 469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewski K.Trends Pharmacol.Sci.2010, 31, 284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewski K.J. Biol. Chem.2012, 287, 1612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewski K.J. Biol. Chem.2012, 287, 1610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parker R. O.; Crouch R. K.Exp. Eye Res.2010, 91, 788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang P. H.; Kono M.; Koutalos Y.; Ablonczy Z.; Crouch R. K.Prog. RetinalEye Res.2013, 32, 48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arshavsky V. Y.; Lamb T. D.; Pugh E. N. Jr.Annu. Rev. Physiol.2002, 64, 153. [DOI] [PubMed] [Google Scholar]
- Lamb T. D.; Pugh E. N. Jr.Prog.Retinal Eye Res.2004, 23, 307. [DOI] [PubMed] [Google Scholar]
- Rando R. R.Biochemistry1991, 30, 595. [DOI] [PubMed] [Google Scholar]
- Pepperberg D. R.; Okajima T. L.; Wiggert B.; Ripps H.; Crouch R. K.; Chader G. J.Mol. Neurobiol.1993, 7, 61. [DOI] [PubMed] [Google Scholar]
- Wolf G.Nutr. Rev.2002, 60, 62. [DOI] [PubMed] [Google Scholar]
- Takimoto N.; Kusakabe T.; Tsuda M.Photochem. Photobiol.2007, 83, 242. [DOI] [PubMed] [Google Scholar]
- Tsybovsky Y.; Molday R. S.; Palczewski K.Adv. Exp. Med. Biol.2010, 703, 105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollock N. L.; Callaghan R.FEBS J.2011, 278, 3204. [DOI] [PubMed] [Google Scholar]
- Redmond T. M.Exp. Eye Res.2009, 88, 846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hofmann K. P.; Scheerer P.; Hildebrand P. W.; Choe H. W.; Park J. H.; Heck M.; Ernst O. P.Trends Biochem. Sci.2009, 34, 540. [DOI] [PubMed] [Google Scholar]
- vonLintig J.; Vogt K.J. Nutr.2004, 134, 251S. [DOI] [PubMed] [Google Scholar]
- vonLintig J.; Wyss A.Arch. Biochem. Biophys.2001, 385, 47. [DOI] [PubMed] [Google Scholar]
- vonLintig J.; Hessel S.; Isken A.; Kiefer C.; Lampert J. M.; Voolstra O.; Vogt K.Biochim. Biophys. Acta2005, 1740, 122. [DOI] [PubMed] [Google Scholar]
- vonLintig J.; Kiser P. D.; Golczak M.; Palczewski K.Trends Biochem. Sci.2010, 35, 400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewski K.Annu. Rev. Biochem.2006, 75, 743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J. S.; Kefalov V. J.Prog. Retinal Eye Res.2011, 30, 115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hill G. E.; Johnson J. D.Am. Nat.2012, 180, E127. [DOI] [PubMed] [Google Scholar]
- Walter M. H.; Strack D.Nat. Prod. Rep.2011, 28, 663. [DOI] [PubMed] [Google Scholar]
- Ruzicka L.Experientia1953, 9, 357. [DOI] [PubMed] [Google Scholar]
- Cane D. E.IsoprenoidBiosynthesis: Overview. In Isoprenoid IncludingCarotenoids and Steroids, 1st ed.; Cane D. E., Ed.; Elsevier: Oxford, 1999; Vol. 1. [Google Scholar]
- Lange B. M.; Rujan T.; Martin W.; Croteau R.Proc. Natl. Acad. Sci. U.S.A.2000, 97, 13172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rohmer M.Nat. Prod. Rep.1999, 16, 565. [DOI] [PubMed] [Google Scholar]
- Schwender J.; Seemann M.; Lichtenthaler H. K.; Rohmer M.Biochem. J.1996, 316Pt 173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flesch G.; Rohmer M.Eur. J. Biochem.1988, 175, 405. [DOI] [PubMed] [Google Scholar]
- Rohmer M.; Seemann M.; Horbach S.; BringerMeyer S.; Sahm H.J.Am. Chem. Soc.1996, 118, 2564. [Google Scholar]
- Takahashi S.; Kuzuyama T.; Watanabe H.; Seto H.Proc.Natl. Acad. Sci. U.S.A.1998, 95, 9879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuzuyama T.; Takahashi S.; Watanabe H.; Seto H.Tetrahedron Lett.1998, 39, 4509. [Google Scholar]
- Grawert T.; Groll M.; Rohdich F.; Bacher A.; Eisenreich W.Cell. Mol. Life Sci.2011, 68, 3797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunter W. N.J. Biol. Chem.2007, 282, 21573. [DOI] [PubMed] [Google Scholar]
- Odom A. R.PLoS Pathog.2011, 7, e1002323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farre G.; Sanahuja G.; Naqvi S.; Bai C.; Capell T.; Zhu C. F.; Christou P.Plant Sci.2010, 179, 28. [DOI] [PubMed] [Google Scholar]
- Kasahara H.; Hanada A.; Kuzuyama T.; Takagi M.; Kamiya Y.; Yamaguchi S.J. Biol. Chem.2002, 277, 45188. [DOI] [PubMed] [Google Scholar]
- Hemmerlin A.; Hoeffler J. F.; Meyer O.; Tritsch D.; Kagan I. A.; Grosdemange-Billiard C.; Rohmer M.; Bach T. J.J. Biol. Chem.2003, 278, 26666. [DOI] [PubMed] [Google Scholar]
- Rodriguez-Concepcion M.; Campos N.; Maria Lois L.; Maldonado C.; Hoeffler J. F.; Grosdemange-Billiard C.; Rohmer M.; Boronat A.FEBS Lett.2000, 473, 328. [DOI] [PubMed] [Google Scholar]
- Hahn F. M.; Hurlburt A. P.; Poulter C. D.J. Bacteriol.1999, 181, 4499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaneda K.; Kuzuyama T.; Takagi M.; Hayakawa Y.; Seto H.Proc.Natl. Acad. Sci. U.S.A.2001, 98, 932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berthelot K.; Estevez Y.; Deffieux A.; Peruch F.Biochimie2012, 94, 1621. [DOI] [PubMed] [Google Scholar]
- Muehlbacher M.; Poulter C. D.Biochemistry1988, 27, 7315. [DOI] [PubMed] [Google Scholar]
- Toteva M. M.; Richard J. P.Bioorg. Chem.1997, 25, 239. [Google Scholar]
- Wouters J.; Oudjama Y.; Barkley S. J.; Tricot C.; Stalon V.; Droogmans L.; Poulter C. D.J. Biol. Chem.2003, 278, 11903. [DOI] [PubMed] [Google Scholar]
- Durbecq V.; Sainz G.; Oudjama Y.; Clantin B.; Bompard-Gilles C.; Tricot C.; Caillet J.; Stalon V.; Droogmans L.; Villeret V.EMBO J.2001, 20, 1530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cornfort J. W.; Clifford K.; Mallaby R.; Phillips G. T.Proc. R. Soc.London, B: Biol. Sci.1972, 182, 277. [Google Scholar]
- Hemmi H.; Ikeda Y.; Yamashita S.; Nakayama T.; Nishino T.Biochem. Biophys. Res. Commun.2004, 322, 905. [DOI] [PubMed] [Google Scholar]
- Sharma N. K.; Pan J. J.; Poulter C. D.Biochemistry2010, 49, 6228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heaps N. A.; Poulter C. D.J. Am. Chem. Soc.2011, 133, 19017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang K. C.; Ohnuma S.Biochim. Biophys. Acta2000, 1529, 33. [DOI] [PubMed] [Google Scholar]
- Laskovics F. M.; Poulter C. D.Biochemistry1981, 20, 1893. [DOI] [PubMed] [Google Scholar]
- Tarshis L. C.; Proteau P. J.; Kellogg B. A.; Sacchettini J. C.; Poulter C. D.Proc. Natl. Acad. Sci. U.S.A.1996, 93, 15018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poulter C.D.; Muscio O. J.; Goodfellow R. J.Biochemistry1974, 13, 1530. [DOI] [PubMed] [Google Scholar]
- Altman L. J.; Ash L.; Kowerski R. C.; Epstein W. W.; Larsen B. R.; Rilling H. C.; Muscio F.; Gregonis D. E.J. Am. Chem. Soc.1972, 94, 3257. [DOI] [PubMed] [Google Scholar]
- Davies B. H.; Taylor R. F.Pure Appl. Chem.1976, 47, 211. [Google Scholar]
- Dogbo O.; Laferriere A.; D’Harlingue A.; Camara B.Proc. Natl. Acad. Sci. U.S.A.1988, 85, 7054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthews P. D.; Luo R.; Wurtzel E. T.J. Exp. Bot.2003, 54, 2215. [DOI] [PubMed] [Google Scholar]
- Chen Y.; Li F.; Wurtzel E. T.Plant Physiol.2010, 153, 66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu Q.; Ghisla S.; Hirschberg J.; Mann V.; Beyer P.J. Biol. Chem.2011, 286, 8666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moise A. R.; Kuksa V.; Imanishi Y.; Palczewski K.J. Biol. Chem.2004, 279, 50230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cunningham F. X. Jr.; Sun Z.; Chamovitz D.; Hirschberg J.; Gantt E.Plant Cell1994, 6, 1107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cunningham F. X.; Chamovitz D.; Misawa N.; Gantt E.; Hirschberg J.FEBS Lett.1993, 328, 130. [DOI] [PubMed] [Google Scholar]
- Hornero-Mendez D.; Britton G.FEBS Lett.2002, 515, 133. [DOI] [PubMed] [Google Scholar]
- Bouvier F.; d’Harlingue A.; Camara B.Arch. Biochem. Biophys.1997, 346, 53. [DOI] [PubMed] [Google Scholar]
- Paik J.; Vogel S.; Quadro L.; Piantedosi R.; Gottesman M.; Lai K.; Hamberger L.; Vieira Mde M.; Blaner W. S.J. Nutr.2004, 134, 276S. [DOI] [PubMed] [Google Scholar]
- During A.; Harrison E. H.J. Lipid Res.2007, 48, 2283. [DOI] [PubMed] [Google Scholar]
- van Bennekum A.; Werder M.; Thuahnai S. T.; Han C. H.; Duong P.; Williams D. L.; Wettstein P.; Schulthess G.; Phillips M. C.; Hauser H.Biochemistry2005, 44, 4517. [DOI] [PubMed] [Google Scholar]
- von Lintig J.; Vogt K.J. Biol. Chem.2000, 275, 11915. [DOI] [PubMed] [Google Scholar]
- Hessel S.; Eichinger A.; Isken A.; Amengual J.; Hunzelmann S.; Hoeller U.; Elste V.; Hunziker W.; Goralczyk R.; Oberhauser V.; von Lintig J.; Wyss A.J.Biol. Chem.2007, 282, 33553. [DOI] [PubMed] [Google Scholar]
- Chambon P.FASEB J.1996, 10, 940. [PubMed] [Google Scholar]
- Kumar S.; Sandell L. L.; Trainor P. A.; Koentgen F.; Duester G.Biochim. Biophys.Acta2012, 1821, 198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Salyers K. L.; Cullum M. E.; Zile M. H.Biochim. Biophys.Acta1993, 1152, 328. [DOI] [PubMed] [Google Scholar]
- Ross A. C.; Zolfaghari R.Annu. Rev. Nutr.2011, 31, 65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewski K.; Kumasaka T.; Hori T.; Behnke C. A.; Motoshima H.; Fox B. A.; Le Trong I.; Teller D. C.; Okada T.; Stenkamp R. E.; Yamamoto M.; Miyano M.Science2000, 289, 739. [DOI] [PubMed] [Google Scholar]
- Nathans J.Biochemistry1990, 29, 9746. [DOI] [PubMed] [Google Scholar]
- Nathans J.; Merbs S. L.; Sung C. H.; Weitz C. J.; Wang Y.Annu.Rev. Genet.1992, 26, 403. [DOI] [PubMed] [Google Scholar]
- Bernstein P. S.; Rando R. R.Biochemistry1986, 25, 6473. [DOI] [PubMed] [Google Scholar]
- Rattner A.; Smallwood P. M.; Nathans J.J. Biol. Chem.2000, 275, 11034. [DOI] [PubMed] [Google Scholar]
- Haeseleer F.; Huang J.; Lebioda L.; Saari J. C.; Palczewski K.J. Biol. Chem.1998, 273, 21790. [DOI] [PubMed] [Google Scholar]
- Ruiz A.; Winston A.; Lim Y. H.; Gilbert B. A.; Rando R. R.; Bok D.J. Biol. Chem.1999, 274, 3834. [DOI] [PubMed] [Google Scholar]
- MacDonald P. N.; Ong D. E.Biochem. Biophys. Res. Commun.1988, 156, 157. [DOI] [PubMed] [Google Scholar]
- Batten M. L.; Imanishi Y.; Maeda T.; Tu D. C.; Moise A. R.; Bronson D.; Possin D.; Van Gelder R. N.; Baehr W.; Palczewski K.J. Biol. Chem.2004, 279, 10422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moiseyev G.; Crouch R. K.; Goletz P.; Oatis J. Jr.; Redmond T. M.; Ma J. X.Biochemistry2003, 42, 2229. [DOI] [PubMed] [Google Scholar]
- Buck J.; Derguini F.; Levi E.; Nakanishi K.; Hammerling U.Science1991, 254, 1654. [DOI] [PubMed] [Google Scholar]
- Buck J.; Grun F.; Derguini F.; Chen Y.; Kimura S.; Noy N.; Hammerling U.J. Exp. Med.1993, 178, 675. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Derguini F.; Nakanishi K.; Hammerling U.; Buck J.Biochemistry1994, 33, 623. [DOI] [PubMed] [Google Scholar]
- Frot-Coutaz J. P.; Silverman-Jones C. S.; De Luca L. M.J. Lipid Res.1976, 17, 220. [PubMed] [Google Scholar]
- Masushige S.; Schreiber J. B.; Wolf G.J.Lipid Res.1978, 19, 619. [PubMed] [Google Scholar]
- Rosso G. C.; Masushige S.; Quill H.; Wolf G.Proc. Natl. Acad.Sci. U.S.A.1977, 74, 3762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kahan J.J.Chromatogr.1967, 30, 506. [DOI] [PubMed] [Google Scholar]
- Targan S. R.; Merrill S.; Schwabe A. D.Clin. Chem. (Washington,DC, U.S.)1969, 15, 479. [PubMed] [Google Scholar]
- Vecchi J.; Vesely J.; Oesterhelt G.J. Chromatogr.1973, 83, 447. [DOI] [PubMed] [Google Scholar]
- Rotmans J. P.; Kropf A.Vision Res.1975, 15, 1301. [DOI] [PubMed] [Google Scholar]
- Frolik C. A.; Tavela T. E.; Peck G. L.; Sporn M. B.Anal. Biochem.1978, 86, 743. [DOI] [PubMed] [Google Scholar]
- Kane M. A.; Folias A. E.; Napoli J. L.Anal. Biochem.2008, 378, 71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kane M. A.; Folias A. E.; Wang C.; Napoli J. L.Anal. Chem.2008, 80, 1702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Hooser J. P.; Garwin G. G.; Saari J. C.Methods Enzymol.2000, 316, 565. [DOI] [PubMed] [Google Scholar]
- Golczak M.; Bereta G.; Maeda A.; Palczewski K.Methods Mol. Biol.2010, 652, 229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosenheim O.; Drummond J. C.Biochem. J.1925, 19, 753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carr F. H.; Price E. A.Biochem. J.1926, 20, 497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blatz P. E.; Estrada A.Anal. Chem.1972, 44, 570. [DOI] [PubMed] [Google Scholar]
- Schwieter U.; Englert G.; Rigassi N.; Vetter W.Pure Appl. Chem.1969, 20, 365. [DOI] [PubMed] [Google Scholar]
- Kildahl-Andersen G.; Naess S. N.; Aslaksen P. B.; Anthonsena T.; Liaaen-Jensen S.Org. Biomol. Chem.2007, 5, 3027. [DOI] [PubMed] [Google Scholar]
- Pienta N. J.; Kessler R. J.J. Am. Chem. Soc.1992, 114, 2419. [Google Scholar]
- Bobrowski K.; Das P. K.J. Am. Chem. Soc.1982, 104, 1704. [Google Scholar]
- McBee J. K.; Kuksa V.; Alvarez R.; de Lera A. R.; Prezhdo O.; Haeseleer F.; Sokal I.; Palczewski K.Biochemistry2000, 39, 11370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pakhomova S.; Kobayashi M.; Buck J.; Newcomer M. E.Nat. Struct. Biol.2001, 8, 447. [DOI] [PubMed] [Google Scholar]
- Honig B.; Ebrey T. G.Annu. Rev. Biophys. Bioeng.1974, 3, 151. [DOI] [PubMed] [Google Scholar]
- Warshel A.; Huler E.Chem.Phys.1974, 6, 463. [Google Scholar]
- Warshel A.; Huler E.; Rabinovi D.; Shakked Z.J. Mol. Struct.1974, 23, 175. [Google Scholar]
- Rowan R. 3rd; Warshel A.; Sykes B. D.; Karplus M.Biochemistry1974, 13, 970. [DOI] [PubMed] [Google Scholar]
- Hubbard R.J. Biol. Chem.1966, 241, 1814. [PubMed] [Google Scholar]
- Hepperle S. S.; Li Q.; East A. L.J. Phys. Chem. A2005, 109, 10975. [DOI] [PubMed] [Google Scholar]
- Benson S. W.; Golden D. M.; Egger K. W.J. Chem. Phys.1965, 42, 4265. [Google Scholar]
- Futterman S.; Rollins M. H.J. Biol. Chem.1973, 248, 7773. [PubMed] [Google Scholar]
- Rando R. R.; Chang A.J.Am. Chem. Soc.1983, 105, 2879. [Google Scholar]
- Hubbard R.; Wald G.J. Gen. Physiol.1952, 36, 269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kropf A.; Hubbard R.Photochem. Photobiol.1970, 12, 249. [DOI] [PubMed] [Google Scholar]
- Brown P. K.; Wald G.J.Biol. Chem.1956, 222, 865. [PubMed] [Google Scholar]
- Fisher M. M.; Weiss K.Photochem. Photobiol.1974, 20, 423. [Google Scholar]
- Bensasson R.; Land E. J.New J. Chem.1978, 2, 503. [Google Scholar]
- Bensasson R.; Land E. J.; Truscott T. G.Photochem. Photobiol.1973, 17, 53. [DOI] [PubMed] [Google Scholar]
- Rosenfel. T.; Alchalel A.; Ottoleng. M.J. Phys. Chem.1974, 78, 336. [Google Scholar]
- Land E. J.Photochem. Photobiol.1975, 22, 286. [DOI] [PubMed] [Google Scholar]
- Becker R. S.Photochem. Photobiol.1988, 48, 369. [DOI] [PubMed] [Google Scholar]
- Das P. K.; Becker R. S.J. Am. Chem. Soc.1979, 101, 6348. [Google Scholar]
- Deval P.; Singh A. K.J. Photochem. Photobiol., A: Chem.1988, 42, 329. [Google Scholar]
- Waddell W. H.; Crouch R.; Nakanishi K.; Turro N. J.J. Am. Chem. Soc.1976, 98, 4189. [DOI] [PubMed] [Google Scholar]
- Jensen N. H.; Wilbrandt R.; Bensasson R. V.J. Am. Chem. Soc.1989, 111, 7877. [Google Scholar]
- Raubach R. A.; Guzzo A. V.J. Phys. Chem.1973, 77, 889. [DOI] [PubMed] [Google Scholar]
- Waddell W. H.; Chihara K.J. Am. Chem. Soc.1981, 103, 7389. [Google Scholar]
- Ganapathy S.; Trehan A.; Liu R. S. H.J. Chem. Soc., Chem.Commun.1990, 199. [Google Scholar]
- Rosenfel. T.; Alchalel A.; Ottoleng. M.Photochem. Photobiol.1974, 20, 121. [DOI] [PubMed] [Google Scholar]
- Giuliano G.; Giliberto L.; Rosati C.Trends Plant Sci.2002, 7, 427. [DOI] [PubMed] [Google Scholar]
- Park H.; Kreunen S. S.; Cuttriss A. J.; DellaPenna D.; Pogson B. J.Plant Cell2002, 14, 321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Isaacson T.; Ronen G.; Zamir D.; Hirschberg J.Plant Cell2002, 14, 333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breitenbach J.; Vioque A.; Sandmann G.Z. Naturforsch.,C2001, 56, 915. [DOI] [PubMed] [Google Scholar]
- Masamoto K.; Wada H.; Kaneko T.; Takaichi S.Plant Cell Physiol.2001, 42, 1398. [DOI] [PubMed] [Google Scholar]
- Isaacson T.; Ohad I.; Beyer P.; Hirschberg J.Plant Physiol.2004, 136, 4246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moise A. R.; Isken A.; Dominguez M.; dLera A. R.; von Lintig J.; Palczewski K.Biochemistry2007, 46, 1811. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moise A. R.; Dominguez M.; Alvarez S.; Alvarez R.; Schupp M.; Cristancho A. G.; Kiser P. D.; de Lera A. R.; Lazar M. A.; Palczewski K.J. Am. Chem. Soc.2008, 130, 1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moise A. R.; Lobo G. P.; Erokwu B.; Wilson D. L.; Peck D.; Alvarez S.; Dominguez M.; Alvarez R.; Flask C. A.; de Lera A. R.; von Lintig J.; Palczewski K.FASEB J.2010, 24, 1261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schupp M.; Lefterova M. I.; Janke J.; Leitner K.; Cristancho A. G.; Mullican S. E.; Qatanani M.; Szwergold N.; Steger D. J.; Curtin J. C.; Kim R. J.; Suh M. J.; Albert M. R.; Engeli S.; Gudas L. J.; Lazar M. A.Proc. Natl. Acad.Sci. U.S.A.2009, 106, 1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pichaud F.; Desplan C.Curr. Opin. Genet. Dev.2002, 12, 430. [DOI] [PubMed] [Google Scholar]
- Pichaud F.; Desplan C.Nature2002, 416, 139. [DOI] [PubMed] [Google Scholar]
- Vopalensky P.; Pergner J.; Liegertova M.; Benito-Gutierrez E.; Arendt D.; Kozmik Z.Proc. Natl. Acad. Sci. U.S.A.2012, 109, 15383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arendt D.; Hausen H.; Purschke G.Philos. Trans. R. Soc. London,B: Biol. Sci.2009, 364, 2809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nilsson D. E.; Arendt D.Curr. Biol.2008, 18, R1096. [DOI] [PubMed] [Google Scholar]
- Loosli F.; Winkler S.; Burgtorf C.; Wurmbach E.; Ansorge W.; Henrich T.; Grabher C.; Arendt D.; Carl M.; Krone A.; Grzebisz E.; Wittbrodt J.Development2001, 128, 4035. [DOI] [PubMed] [Google Scholar]
- Arendt D.Int. J. Dev. Biol.2003, 47, 563. [PubMed] [Google Scholar]
- Darwin C.On the origin of speciesby means of natural selection; John Murray: London, 1859. [Google Scholar]
- Nagel G.; Szellas T.; Kateriya S.; Adeishvili N.; Hegemann P.; Bamberg E.Biochem. Soc. Trans.2005, 33, 863. [DOI] [PubMed] [Google Scholar]
- Zhai Y.; Heijne W. H.; Smith D. W.; Saier M. H. Jr.Biochim.Biophys. Acta2001, 1511, 206. [DOI] [PubMed] [Google Scholar]
- Pedros-Alio C.Int. Microbiol.2006, 9, 191. [PubMed] [Google Scholar]
- Kennis J. T.; Groot M. L.Curr. Opin. Struct. Biol.2007, 17, 623. [DOI] [PubMed] [Google Scholar]
- Rossle S.C.; Frank I.Front. Biosci.2009, 14, 4862. [DOI] [PubMed] [Google Scholar]
- Davies W. L.; Hankins M. W.; Foster R. G.Photochem. Photobiol. Sci.2010, 9, 1444. [DOI] [PubMed] [Google Scholar]
- Purschwitz J.; Muller S.; Kastner C.; Fischer R.Curr. Opin. Microbiol.2006, 9, 566. [DOI] [PubMed] [Google Scholar]
- Filipek S.; Teller D. C.; Palczewski K.; Stenkamp R.Annu. Rev. Biophys. Biomol. Struct.2003, 32, 375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Menon S. T.; Han M.; Sakmar T. P.Physiol. Rev.2001, 81, 1659. [DOI] [PubMed] [Google Scholar]
- Sakmar T.P.; Menon S. T.; Marin E. P.; Awad E. S.Annu. Rev. Biophys.Biomol. Struct.2002, 31, 443. [DOI] [PubMed] [Google Scholar]
- Okada T.; Palczewski K.Curr. Opin. Struct.Biol.2001, 11, 420. [DOI] [PubMed] [Google Scholar]
- Rao V. R.; Oprian D. D.Annu. Rev. Biophys. Biomol.Struct.1996, 25, 287. [DOI] [PubMed] [Google Scholar]
- Muller D. J.; Wu N.; Palczewski K.Pharmacol. Rev.2008, 60, 43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yau K. W.; Hardie R. C.Cell2009, 139, 246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo D. G.; Xue T.; Yau K. W.Proc. Natl. Acad.Sci.U.S.A.2008, 105, 9855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fain G. L.; Hardie R.; Laughlin S. B.Curr. Biol.2010, 20, R114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lamb T. D.Philos. Trans. R. Soc. London, B: Biol. Sci.2009, 364, 2911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Wang T.; Jiao Y.; von Lintig J.; Montell C.Curr. Biol.2010, 20, 93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mustafi D.; Engel A. H.; Palczewski K.Prog. Retinal EyeRes.2009, 28, 289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kefalov V. J.J. Biol. Chem.2012, 287, 1635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Butcher A. J.; Kong K. C.; Prihandoko R.; Tobin A. B.Handb. Exp. Pharmacol.2012, 79. [DOI] [PubMed] [Google Scholar]
- Nathans J.Sci. Am.1989, 260, 42. [DOI] [PubMed] [Google Scholar]
- Oprian D. D.Curr. Opin. Neurobiol.1992, 2, 428. [DOI] [PubMed] [Google Scholar]
- Stenkamp R. E.; Filipek S.; Driessen C. A.; Teller D. C.; Palczewski K.Biochim. Biophys.Acta2002, 1565, 168. [DOI] [PubMed] [Google Scholar]
- Swaroop A.; Kim D.; Forrest D.Nat. Rev. Neurosci.2010, 11, 563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodieck R. W.The firststeps inseeing; Sinauer Associates: Sunderland, MA, 1998. [Google Scholar]
- Quiring R.; Walldorf U.; Kloter U.; Gehring W. J.Science1994, 265, 785. [DOI] [PubMed] [Google Scholar]
- Halder G.; Callaerts P.; Gehring W. J.Science1995, 267, 1788. [DOI] [PubMed] [Google Scholar]
- Gehring W.; Rosbash M.J. Mol. Evol.2003, 57Suppl 1S286. [DOI] [PubMed] [Google Scholar]
- Baylor D. A.; Lamb T. D.; Yau K. W.J. Physiol.1979, 288, 613. [PMC free article] [PubMed] [Google Scholar]
- Baylor D. A.Invest. Ophthalmol. Vis. Sci.1987, 28, 34. [PubMed] [Google Scholar]
- Nakanishi K.; Baloghnair V.; Arnaboldi M.; Tsujimoto K.; Honig B.J.Am. Chem. Soc.1980, 102, 7945. [Google Scholar]
- Sakmar T. P.Science2012, 338, 1299. [DOI] [PubMed] [Google Scholar]
- Wang W.; Nossoni Z.; Berbasova T.; Watson C. T.; Yapici I.; Lee K. S.; Vasileiou C.; Geiger J. H.; Borhan B.Science2012, 338, 1340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szel A.; von Schantz M.; Rohlich P.; Farber D. B.; van Veen T.Invest. Ophthalmol. Vis. Sci.1993, 34, 3641. [PubMed] [Google Scholar]
- Jeon C. J.; Strettoi E.; Masland R. H.J. Neurosci.1998, 18, 8936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Daemen F. J.Biochim. Biophys. Acta1973, 300, 255. [DOI] [PubMed] [Google Scholar]
- Blaurock A. E.; Wilkins M. H.Nature1969, 223, 906. [DOI] [PubMed] [Google Scholar]
- Liang Y.; Fotiadis D.; Filipek S.; Saperstein D. A.; Palczewski K.; Engel A.J. Biol. Chem.2003, 278, 21655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mustafi D.; Kevany B. M.; Genoud C.; Okano K.; Cideciyan A. V.; Sumaroka A.; Roman A. J.; Jacobson S. G.; Engel A.; Adams M. D.; Palczewski K.FASEB J.2011, 25, 3157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nickell S.; Park P. S.; Baumeister W.; Palczewski K.J. Cell Biol.2007, 177, 917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilliam J. C.; Chang J. T.; Sandoval I. M.; Zhang Y.; Li T.; Pittler S. J.; Chiu W.; Wensel T. G.Cell2012, 151, 1029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liang Y.; Fotiadis D.; Maeda T.; Maeda A.; Modzelewska A.; Filipek S.; Saperstein D. A.; Engel A.; Palczewski K.J. Biol. Chem.2004, 279, 48189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nag T. C.; Wadhwa S.Micron2012, 43, 759. [DOI] [PubMed] [Google Scholar]
- Bok D.J.Cell Sci. Suppl.1993, 17, 189. [DOI] [PubMed] [Google Scholar]
- Booij J. C.; Baas D. C.; Beisekeeva J.; Gorgels T. G.; Bergen A. A.Prog. Retinal Eye Res.2010, 29, 1. [DOI] [PubMed] [Google Scholar]
- Gao H.; Hollyfield J. G.Invest. Ophthalmol. Vis. Sci.1992, 33, 1. [PubMed] [Google Scholar]
- Strauss O.Physiol. Rev.2005, 85, 845. [DOI] [PubMed] [Google Scholar]
- Bazan N. G.Trends Neurosci.2006, 29, 263. [DOI] [PubMed] [Google Scholar]
- Lange C.A.; Bainbridge J. W.Ophthalmologica2012, 227, 115. [DOI] [PubMed] [Google Scholar]
- Caprara C.; Grimm C.Prog. Retinal Eye Res.2012, 31, 89. [DOI] [PubMed] [Google Scholar]
- Kevany B. M.; Palczewski K.Physiology (Bethesda)2010, 25, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anderson D. H.; Fisher S. K.; Steinberg R. H.Invest. Ophthalmol.Vis. Sci.1978, 17, 117. [PubMed] [Google Scholar]
- Sparrow J. R.; Gregory-Roberts E.; Yamamoto K.; Blonska A.; Ghosh S. K.; Ueda K.; Zhou J.Prog.Retinal Eye Res.2012, 31, 121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pagon R. A.; Daiger S. P. In GeneReviews; Pagon R. A., Bird T. D., Dolan C. R., Stephens K., Adam M. P., Eds.; Universityof Washington: Seattle, WA, 1993. [Google Scholar]
- Weleber R. G.; Francis P. J.; Trzupek K. M. In GeneReviews; Pagon R. A., Bird T. D., Dolan C. R., Stephens K., Adam M. P., Eds.; Universityof Washington: Seattle, WA, 1993. [Google Scholar]
- Handa J.T.Mol. Aspects Med.2012, 33, 418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cideciyan A. V.Prog. Retinal Eye Res.2010, 29, 398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blomhoff R.; Blomhoff H. K.J. Neurobiol.2006, 66, 606. [DOI] [PubMed] [Google Scholar]
- Cottet S.; Schorderet D. F.Curr. Mol. Med.2009, 9, 375. [DOI] [PubMed] [Google Scholar]
- Sahni J. N.; Angi M.; Irigoyen C.; Semeraro F.; Romano M. R.; Parmeggiani F.Curr. Genomics2011, 12, 276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuksa V.; Imanishi Y.; Batten M.; Palczewski K.; Moise A. R.Vision Res.2003, 43, 2959. [DOI] [PubMed] [Google Scholar]
- Rozanowska M.; Sarna T.Photochem.Photobiol.2005, 81, 1305. [DOI] [PubMed] [Google Scholar]
- von Lintig J.Annu. Rev. Nutr.2010, 30, 35. [DOI] [PubMed] [Google Scholar]
- Cai X.; Conley S. M.; Naash M. I.Ophthalmic Genet.2009, 30, 57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imanishi Y.; Gerke V.; Palczewski K.J. Cell Biol.2004, 166, 447. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imanishi Y.; Lodowski K. H.; Koutalos Y.Biochemistry2007, 46, 9674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiser P. D.; Palczewski K.Prog. Retinal Eye Res.2010, 29, 428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saari J.C.Annu. Rev. Nutr.2012, 32, 125. [DOI] [PubMed] [Google Scholar]
- Thompson D. A.; Gal A.Prog. Retinal Eye Res.2003, 22, 683. [DOI] [PubMed] [Google Scholar]
- Kiser P. D.; Golczak M.; Maeda A.; Palczewski K.Biochim. Biophys.Acta2012, 1821, 137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kolesnikov A. V.; Tang P. H.; Parker R. O.; Crouch R. K.; Kefalov V. J.J. Neurosci.2011, 31, 7900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imanishi Y.; Palczewski K.Methods Mol. Biol.2010, 652, 247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imanishi Y.; Sun W.; Maeda T.; Maeda A.; Palczewski K.J. Biol. Chem.2008, 283, 25091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imanishi Y.; Batten M. L.; Piston D. W.; Baehr W.; Palczewski K.J. Cell Biol.2004, 164, 373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Q.; Chen C.; Koutalos Y.Biophys. J.2006, 91, 4678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koutalos Y.Methods Mol. Biol.2010, 652, 115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boll F.VisionRes.1977, 17, 1249. [DOI] [PubMed] [Google Scholar]
- Oyster C. W.The Human Eye: Structureand Function, 1st ed.; Sinauer: Sunderland, 1999; pp 612–613. [Google Scholar]
- Ripps H.FASEBJ.2008, 22, 4038. [DOI] [PubMed] [Google Scholar]
- Kuhne W.On the photochemistryof the retina; Cambridge University Press: 1878; pp 1–104. [Google Scholar]
- Dowling J. E.Nature1960, 188, 114. [DOI] [PubMed] [Google Scholar]
- Hecht S.; Haig C.; Chase A. M.J. Gen. Physiol.1937, 20, 831. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schnapf J. L.; Nunn B. J.; Meister M.; Baylor D. A.J. Physiol.1990, 427, 681. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Penn R. D.; Hagins W. A.Biophys. J.1972, 12, 1073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mata N. L.; Radu R. A.; Clemmons R. C.; Travis G. H.Neuron2002, 36, 69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berman E. R.; Horowitz J.; Segal N.; Fisher S.; Feeney-Burns L.Biochim. Biophys.Acta1980, 630, 36. [DOI] [PubMed] [Google Scholar]
- Rodriguez K. A.; Tsin A. T.Am. J. Physiol.1989, 256, R255. [DOI] [PubMed] [Google Scholar]
- Das S. R.; Bhardwaj N.; Kjeldbye H.; Gouras P.Biochem. J.1992, 285Pt 3907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bok D.; Ong D. E.; Chytil F.Invest. Ophthalmol.Vis. Sci.1984, 25, 877. [PubMed] [Google Scholar]
- Fleisch V. C.; Schonthaler H. B.; von Lintig J.; Neuhauss S. C.J. Neurosci.2008, 28, 8208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaylor J. J.; Yuan Q.; Cook J.; Sarfare S.; Makshanoff J.; Miu A.; Kim A.; Kim P.; Habib S.; Roybal C. N.; Xu T.; Nusinowitz S.; Travis G. H.Nat. Chem. Biol.2013, 9, 30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shanklin J.; Guy J. E.; Mishra G.; Lindqvist Y.J. Biol. Chem.2009, 284, 18559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ternes P.; Franke S.; Zahringer U.; Sperling P.; Heinz E.J. Biol. Chem.2002, 277, 25512. [DOI] [PubMed] [Google Scholar]
- Michel C.; van Echten-Deckert G.; Rother J.; Sandhoff K.; Wang E.; Merrill A. H. Jr.J. Biol. Chem.1997, 272, 22432. [DOI] [PubMed] [Google Scholar]
- Orban T.; Palczewska G.; Palczewski K.J. Biol. Chem.2011, 286, 17248. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palczewska G.; Maeda T.; Imanishi Y.; Sun W.; Chen Y.; Williams D. R.; Piston D. W.; Maeda A.; Palczewski K.Nat. Med.2010, 16, 1444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sparrow J. R.; Wu Y.; Kim C. Y.; Zhou J.J. Lipid Res.2010, 51, 247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sparrow J. R.; Yamamoto K.Adv. Exp. Med. Biol.2012, 723, 761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunter J. J.; Morgan J. I.; Merigan W. H.; Sliney D. H.; Sparrow J. R.; Williams D. R.Prog. Retinal Eye Res.2012, 31, 28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sparrow J.R.; Cai B.; Fishkin N.; Jang Y. P.; Krane S.; Vollmer H. R.; Zhou J.; Nakanishi K.Adv. Exp. Med. Biol.2003, 533, 205. [DOI] [PubMed] [Google Scholar]
- Maeda A.; Maeda T.; Golczak M.; Chou S.; Desai A.; Hoppel C. L.; Matsuyama S.; Palczewski K.J. Biol. Chem.2009, 284, 15173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shiose S.; Chen Y.; Okano K.; Roy S.; Kohno H.; Tang J.; Pearlman E.; Maeda T.; Palczewski K.; Maeda A.J. Biol. Chem.2011, 286, 15543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda T.; Golczak M.; Maeda A.Photochem. Photobiol.2012, 88, 1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farrens D. L.Photochem. Photobiol.Sci.2010, 9, 1466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park P. S.; Lodowski D. T.; Palczewski K.Annu. Rev. Pharmacol.Toxicol.2008, 48, 107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ridge K. D.; Palczewski K.J. Biol. Chem.2007, 282, 9297. [DOI] [PubMed] [Google Scholar]
- Vogel R.; Sakmar T. P.; Sheves M.; Siebert F.Photochem. Photobiol.2007, 83, 286. [DOI] [PubMed] [Google Scholar]
- Liang X.; Nazarian A.; Erdjument-Bromage H.; Bornmann W.; Tempst P.; Resh M. D.J. Biol. Chem.2001, 276, 30987. [DOI] [PubMed] [Google Scholar]
- Olsson J. E.; Gordon J. W.; Pawlyk B. S.; Roof D.; Hayes A.; Molday R. S.; Mukai S.; Cowley G. S.; Berson E. L.; Dryja T. P.Neuron1992, 9, 815. [DOI] [PubMed] [Google Scholar]
- Humphries M. M.; Rancourt D.; Farrar G. J.; Kenna P.; Hazel M.; Bush R. A.; Sieving P. A.; Sheils D. M.; McNally N.; Creighton P.; Erven A.; Boros A.; Gulya K.; Capecchi M. R.; Humphries P.Nat. Genet.1997, 15, 216. [DOI] [PubMed] [Google Scholar]
- Molday R. S.Invest. Ophthalmol.Vis. Sci.1998, 39, 2491. [PubMed] [Google Scholar]
- Corless J. M.; Worniallo E.; Schneider T. G.Exp. Eye Res.1995, 61, 335. [DOI] [PubMed] [Google Scholar]
- Fotiadis D.; Liang Y.; Filipek S.; Saperstein D. A.; Engel A.; Palczewski K.Nature2003, 421, 127. [DOI] [PubMed] [Google Scholar]
- Vahedi-Faridi A.; Jastrzebska B.; Palczewski K.; Engel A.J.Electron Microsc. (Tokyo)2013, 62, 95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jastrzebska B.; Orban T.; Golczak M.; Engel A.; Palczewski K.FASEB J.2013, 27, 1572. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orban T.; Jastrzebska B.; Gupta S.; Wang B.; Miyagi M.; Chance M. R.; Palczewski K.Structure2012, 20, 826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jastrzebska B.; Ringler P.; Lodowski D. T.; Moiseenkova-Bell V.; Golczak M.; Muller S. A.; Palczewski K.; Engel A.J.Struct. Biol.2011, 176, 387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jastrzebska B.; Maeda T.; Zhu L.; Fotiadis D.; Filipek S.; Engel A.; Stenkamp R. E.; Palczewski K.J. Biol. Chem.2004, 279, 54663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fotiadis D.; Jastrzebska B.; Philippsen A.; Muller D. J.; Palczewski K.; Engel A.Curr.Opin. Struct. Biol.2006, 16, 252. [DOI] [PubMed] [Google Scholar]
- Jastrzebska B.; Tsybovsky Y.; Palczewski K.Biochem. J.2010, 428, 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jastrzebska B.; Goc A.; Golczak M.; Palczewski K.Biochemistry2009, 48, 5159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wald G.; Brown P. K.J. Gen. Physiol.1957, 40, 627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Angel T. E.; Chance M. R.; Palczewski K.Proc. Natl. Acad.Sci. U.S.A.2009, 106, 8555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Angel T. E.; Gupta S.; Jastrzebska B.; Palczewski K.; Chance M. R.Proc. Natl. Acad. Sci. U.S.A.2009, 106, 14367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orban T.; Gupta S.; Palczewski K.; Chance M. R.Biochemistry2010, 49, 827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choe H. W.; Kim Y. J.; Park J. H.; Morizumi T.; Pai E. F.; Krauss N.; Hofmann K. P.; Scheerer P.; Ernst O. P.Nature2011, 471, 651. [DOI] [PubMed] [Google Scholar]
- Salom D.; Lodowski D. T.; Stenkamp R. E.; Le Trong I.; Golczak M.; Jastrzebska B.; Harris T.; Ballesteros J. A.; Palczewski K.Proc. Natl. Acad. Sci. U.S.A.2006, 103, 16123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jastrzebska B.; Ringler P.; Palczewski K.; Engel A.J.Struct. Biol.2013, 182, 164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Salon J. A.; Lodowski D. T.; Palczewski K.Pharmacol. Rev.2011, 63, 901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith S. O.Biochem. Soc. Trans.2012, 40, 389. [DOI] [PubMed] [Google Scholar]
- Eilers M.; Goncalves J. A.; Ahuja S.; Kirkup C.; Hirshfeld A.; Simmerling C.; Reeves P. J.; Sheves M.; Smith S. O.J. Phys. Chem.B2012, 116, 10477. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taddese B.; Simpson L. M.; Wall I. D.; Blaney F. E.; Kidley N. J.; Clark H. S.; Smith R. E.; Upton G. J.; Gouldson P. R.; Psaroudakis G.; Bywater R. P.; Reynolds C. A.Biochem. Soc.Trans.2012, 40, 394. [DOI] [PubMed] [Google Scholar]
- Bourne H. R.; Meng E. C.Science2000, 289, 733. [DOI] [PubMed] [Google Scholar]
- Meng E. C.; Bourne H. R.Trends Pharmacol. Sci.2001, 22, 587. [DOI] [PubMed] [Google Scholar]
- Sakmar T.P.Curr. Opin. Cell Biol.2002, 14, 189. [DOI] [PubMed] [Google Scholar]
- Stenkamp R. E.; Teller D. C.; Palczewski K.ChemBioChem2002, 3, 963. [DOI] [PubMed] [Google Scholar]
- Okada T.; Sugihara M.; Bondar A. N.; Elstner M.; Entel P.; Buss V.J. Mol. Biol.2004, 342, 571. [DOI] [PubMed] [Google Scholar]
- Teller D. C.; Okada T.; Behnke C. A.; Palczewski K.; Stenkamp R. E.Biochemistry2001, 40, 7761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kandori H.; Shichida Y.; Yoshizawa T.Biochemistry2001, 66, 1197. [DOI] [PubMed] [Google Scholar]
- Surya A.; Stadel J. M.; Knox B. E.Trends Pharmacol. Sci.1998, 19, 243. [DOI] [PubMed] [Google Scholar]
- Furutani Y.; Kandori H.; Shichida Y.Biochemistry2003, 42, 8494. [DOI] [PubMed] [Google Scholar]
- Nakamichi H.; Okada T.Angew. Chem., Int. Ed.2006, 45, 4270. [DOI] [PubMed] [Google Scholar]
- Schreiber M.; Sugihara M.; Okada T.; Buss V.Angew. Chem., Int.Ed.2006, 45, 4274. [DOI] [PubMed] [Google Scholar]
- Yan E. C.; Kazmi M. A.; Ganim Z.; Hou J. M.; Pan D.; Chang B. S.; Sakmar T. P.; Mathies R. A.Proc. Natl. Acad. Sci.U.S.A.2003, 100, 9262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ludeke S.; Beck M.; Yan E. C.; Sakmar T. P.; Siebert F.; Vogel R.J. Mol. Biol.2005, 353, 345. [DOI] [PubMed] [Google Scholar]
- Kapoor N.; Menon S. T.; Chauhan R.; Sachdev P.; Sakmar T. P.J. Mol. Biol.2009, 393, 882. [DOI] [PubMed] [Google Scholar]
- Maeda T.; Imanishi Y.; Palczewski K.Prog. Retinal EyeRes.2003, 22, 417. [DOI] [PubMed] [Google Scholar]
- Palczewski K.; Benovic J. L.Trends Biochem. Sci.1991, 16, 387. [DOI] [PubMed] [Google Scholar]
- Polans A.; Baehr W.; Palczewski K.Trends Neurosci.1996, 19, 547. [DOI] [PubMed] [Google Scholar]
- Vogel R.; Siebert F.; Zhang X. Y.; Fan G.; Sheves M.Biochemistry2004, 43, 9457. [DOI] [PubMed] [Google Scholar]
- Schadel S. A.; Heck M.; Maretzki D.; Filipek S.; Teller D. C.; Palczewski K.; Hofmann K. P.J. Biol. Chem.2003, 278, 24896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heck M.; Schadel S. A.; Maretzki D.; Hofmann K. P.Vision Res.2003, 43, 3003. [DOI] [PubMed] [Google Scholar]
- Scheerer P.; Park J. H.; Hildebrand P. W.; Kim Y. J.; Krauss N.; Choe H. W.; Hofmann K. P.; Ernst O. P.Nature2008, 455, 497. [DOI] [PubMed] [Google Scholar]
- Makino C. L.; Riley C. K.; Looney J.; Crouch R. K.; Okada T.Biophys.J.2010, 99, 2366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hildebrand P. W.; Scheerer P.; Park J. H.; Choe H. W.; Piechnick R.; Ernst O. P.; Hofmann K. P.; Heck M.PLoSOne2009, 4, e4382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Porte S.; Xavier Ruiz F.; Gimenez J.; Molist I.; Alvarez S.; Dominguez M.; Alvarez R.; deLera A. R.; Pares X.; Farres J.Chem. Biol. Interact.2013, 202, 186. [DOI] [PubMed] [Google Scholar]
- Duester G.Eur. J. Biochem.2000, 267, 4315. [DOI] [PubMed] [Google Scholar]
- Lesk A. M.Curr. Opin. Struct. Biol.1995, 5, 775. [DOI] [PubMed] [Google Scholar]
- Howard E. I.; Sanishvili R.; Cachau R. E.; Mitschler A.; Chevrier B.; Barth P.; Lamour V.; Van Zandt M.; Sibley E.; Bon C.; Moras D.; Schneider T. R.; Joachimiak A.; Podjarny A.Proteins2004, 55, 792. [DOI] [PubMed] [Google Scholar]
- Danielsson O.; Atrian S.; Luque T.; Hjelmqvist L.; Gonzalez-Duarte R.; Jornvall H.Proc. Natl. Acad. Sci. U.S.A.1994, 91, 4980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang Y. H.; Chuang L. Y.; Hwang C. C.J. Biol. Chem.2007, 282, 34306. [DOI] [PubMed] [Google Scholar]
- Ensor C. M.; Tai H. H.Biochem. Biophys. Res. Commun.1991, 176, 840. [DOI] [PubMed] [Google Scholar]
- Obeid J.; White P. C.Biochem. Biophys. Res. Commun.1992, 188, 222. [DOI] [PubMed] [Google Scholar]
- Kavanagh K. L.; Jornvall H.; Persson B.; Oppermann U.Cell. Mol. Life Sci.2008, 65, 3895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oppermann U. C.; Persson B.; Jornvall H.Eur. J. Biochem.1997, 249, 355. [DOI] [PubMed] [Google Scholar]
- Filling C.; Berndt K. D.; Benach J.; Knapp S.; Prozorovski T.; Nordling E.; Ladenstein R.; Jornvall H.; Oppermann U.J. Biol. Chem.2002, 277, 25677. [DOI] [PubMed] [Google Scholar]
- Kratzer R.; Wilson D. K.; Nidetzky B.IUBMB Life2006, 58, 499. [DOI] [PubMed] [Google Scholar]
- Zhang M.; Hu P.; Napoli J. L.J. Biol. Chem.2004, 279, 51482. [DOI] [PubMed] [Google Scholar]
- Janecke A.R.; Thompson D. A.; Utermann G.; Becker C.; Hubner C. A.; Schmid E.; McHenry C. L.; Nair A. R.; Ruschendorf F.; Heckenlively J.; Wissinger B.; Nurnberg P.; Gal A.Nat. Genet.2004, 36, 850. [DOI] [PubMed] [Google Scholar]
- Lapshina E. A.; Belyaeva O. V.; Chumakova O. V.; Kedishvili N. Y.Biochemistry2003, 42, 776. [DOI] [PubMed] [Google Scholar]
- Belyaeva O. V.; Stetsenko A. V.; Nelson P.; Kedishvili N. Y.Biochemistry2003, 42, 14838. [DOI] [PubMed] [Google Scholar]
- Luo W.; Marsh-Armstrong N.; Rattner A.; Nathans J.J. Neurosci.2004, 24, 2623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liden M.; Eriksson U.J. Biol. Chem.2006, 281, 13001. [DOI] [PubMed] [Google Scholar]
- Biswas M. G.; Russell D. W.J. Biol. Chem.1997, 272, 15959. [DOI] [PubMed] [Google Scholar]
- Gough W.H.; VanOoteghem S.; Sint T.; Kedishvili N. Y.J. Biol. Chem.1998, 273, 19778. [DOI] [PubMed] [Google Scholar]
- Su J.; Chai X.; Kahn B.; Napoli J. L.J. Biol. Chem.1998, 273, 17910. [DOI] [PubMed] [Google Scholar]
- Haeseleer F.; Jang G. F.; Imanishi Y.; Driessen C. A.; Matsumura M.; Nelson P. S.; Palczewski K.J. Biol. Chem.2002, 277, 45537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Merrill D. K.; Guynn R. W.Brain Res.1981, 221, 307. [DOI] [PubMed] [Google Scholar]
- Rhinn M.; Dolle P.Development2012, 139, 843. [DOI] [PubMed] [Google Scholar]
- Maeda A.; Maeda T.; Imanishi Y.; Sun W.; Jastrzebska B.; Hatala D. A.; Winkens H. J.; Hofmann K. P.; Janssen J. J.; Baehr W.; Driessen C. A.; Palczewski K.J. Biol. Chem.2006, 281, 37697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda A.; Maeda T.; Imanishi Y.; Kuksa V.; Alekseev A.; Bronson J. D.; Zhang H.; Zhu L.; Sun W.; Saperstein D. A.; Rieke F.; Baehr W.; Palczewski K.J. Biol. Chem.2005, 280, 18822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim T. S.; Maeda A.; Maeda T.; Heinlein C.; Kedishvili N.; Palczewski K.; Nelson P. S.J. Biol. Chem.2005, 280, 8694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda A.; Maeda T.; Sun W.; Zhang H.; Baehr W.; Palczewski K.Proc. Natl. Acad.Sci. U.S.A.2007, 104, 19565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda A.; Maeda T.; Golczak M.; Palczewski K.J. Biol. Chem.2008, 283, 26684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- O’Byrne S. M.; Wongsiriroj N.; Libien J.; Vogel S.; Goldberg I. J.; Baehr W.; Palczewski K.; Blaner W. S.J. Biol. Chem.2005, 280, 35647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu L.; Gudas L. J.J. Biol. Chem.2005, 280, 40226. [DOI] [PubMed] [Google Scholar]
- Wongsiriroj N.; Piantedosi R.; Palczewski K.; Goldberg I. J.; Johnston T. P.; Li E.; Blaner W. S.J. Biol. Chem.2008, 283, 13510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin M.; Li S.; Moghrabi W. N.; Sun H.; Travis G. H.Cell2005, 122, 449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moise A. R.; Golczak M.; Imanishi Y.; Palczewski K.J. Biol. Chem.2007, 282, 2081. [DOI] [PubMed] [Google Scholar]
- Bussieres S.; Cantin L.; Desbat B.; Salesse C.Langmuir2012, 28, 3516. [DOI] [PubMed] [Google Scholar]
- Mateja A.; Szlachcic A.; Downing M. E.; Dobosz M.; Mariappan M.; Hegde R. S.; Keenan R. J.Nature2009, 461, 361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anantharaman V.; Aravind L.Genome Biol.2003, 4, R11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albalat R.Mol. Biol. Evol.2012, 29, 1461. [DOI] [PubMed] [Google Scholar]
- Golczak M.; Kiser P. D.; Sears A. E.; Lodowski D. T.; Blaner W. S.; Palczewski K.J. Biol. Chem.2012, 287, 23790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pang X. Y.; Cao J.; Addington L.; Lovell S.; Battaile K. P.; Zhang N.; Rao J. L.; Dennis E. A.; Moise A. R.J. Biol. Chem.2012, 287, 35260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu Q.; Rawlings N. D.; Chiu H. J.; Jaroszewski L.; Klock H. E.; Knuth M. W.; Miller M. D.; Elsliger M. A.; Deacon A. M.; Godzik A.; Lesley S. A.; Wilson I. A.PLoS One2011, 6, e22013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bok D.; Ruiz A.; Yaron O.; Jahng W. J.; Ray A.; Xue L.; Rando R. R.Biochemistry2003, 42, 6090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Golczak M.; Palczewski K.J. Biol. Chem.2010, 285, 29217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi Y. Q.; Hubacek I.; Rando R. R.Biochemistry1993, 32, 1257. [DOI] [PubMed] [Google Scholar]
- Jahng W. J.; Cheung E.; Rando R. R.Biochemistry2002, 41, 6311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xue L.; Rando R. R.Biochemistry2004, 43, 6120. [DOI] [PubMed] [Google Scholar]
- Jin X. H.; Okamoto Y.; Morishita J.; Tsuboi K.; Tonai T.; Ueda N.J. Biol. Chem.2007, 282, 3614. [DOI] [PubMed] [Google Scholar]
- Shinohara N.; Uyama T.; Jin X. H.; Tsuboi K.; Tonai T.; Houchi H.; Ueda N.J. Lipid Res.2011, 52, 1927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Uyama T.; Jin X. H.; Tsuboi K.; Tonai T.; Ueda N.Biochim.Biophys. Acta2009, 1791, 1114. [DOI] [PubMed] [Google Scholar]
- Uyama T.; Morishita J.; Jin X. H.; Okamoto Y.; Tsuboi K.; Ueda N.J. Lipid Res.2009, 50, 685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ross A. C.J. Biol. Chem.1982, 257, 2453. [PubMed] [Google Scholar]
- Helgerud P.; Petersen L. B.; Norum K. R.J. Lipid Res.1982, 23, 609. [PubMed] [Google Scholar]
- Torma H.; Vahlquist A.J. Invest. Dermatol.1987, 88, 398. [DOI] [PubMed] [Google Scholar]
- Chaudhary L. R.; Nelson E. C.Biochim. Biophys. Acta1987, 917, 24. [DOI] [PubMed] [Google Scholar]
- Kaschula C. H.; Jin M. H.; Desmond-Smith N. S.; Travis G. H.Exp. Eye Res.2006, 82, 111. [DOI] [PubMed] [Google Scholar]
- Muniz A.; Villazana-Espinoza E.T.; Thackeray B.; Tsin A. T.Biochemistry2006, 45, 12265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Randolph R. K.; Winkler K. E.; Ross A. C.Arch. Biochem.Biophys.1991, 288, 500. [DOI] [PubMed] [Google Scholar]
- Mata N. L.; Ruiz A.; Radu R. A.; Bui T. V.; Travis G. H.Biochemistry2005, 44, 11715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J. S.; Estevez M. E.; Cornwall M. C.; Kefalov V. J.Nat. Neurosci.2009, 12, 295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yen C. L.; Brown C. H. t.; Monetti M.; Farese R. V. Jr.J. Lipid Res.2005, 46, 2388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orland M. D.; Anwar K.; Cromley D.; Chu C. H.; Chen L.; Billheimer J. T.; Hussain M. M.; Cheng D.Biochim. Biophys.Acta2005, 1737, 76. [DOI] [PubMed] [Google Scholar]
- Ables G. P.; Yang K. J.; Vogel S.; Hernandez-Ono A.; Yu S.; Yuen J. J.; Birtles S.; Buckett L. K.; Turnbull A. V.; Goldberg I. J.; Blaner W. S.; Huang L. S.; Ginsberg H. N.J. Lipid Res.2012, 53, 2364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rando R. R.Chem. Rev.2001, 101, 1881. [DOI] [PubMed] [Google Scholar]
- Hamel C. P.; Tsilou E.; Harris E.; Pfeffer B. A.; Hooks J. J.; Detrick B.; Redmond T. M.J. Neurosci. Res.1993, 34, 414. [DOI] [PubMed] [Google Scholar]
- Hamel C. P.; Tsilou E.; Pfeffer B. A.; Hooks J. J.; Detrick B.; Redmond T. M.J. Biol. Chem.1993, 268, 15751. [PubMed] [Google Scholar]
- Bavik C. O.; Levy F.; Hellman U.; Wernstedt C.; Eriksson U.J. Biol. Chem.1993, 268, 20540. [PubMed] [Google Scholar]
- Redmond T.M.; Yu S.; Lee E.; Bok D.; Hamasaki D.; Chen N.; Goletz P.; Ma J. X.; Crouch R. K.; Pfeifer K.Nat. Genet.1998, 20, 344. [DOI] [PubMed] [Google Scholar]
- Marlhens F.; Bareil C.; Griffoin J. M.; Zrenner E.; Amalric P.; Eliaou C.; Liu S. Y.; Harris E.; Redmond T. M.; Arnaud B.; Claustres M.; Hamel C. P.Nat. Genet.1997, 17, 139. [DOI] [PubMed] [Google Scholar]
- Gu S. M.; Thompson D. A.; Srikumari C. R.; Lorenz B.; Finckh U.; Nicoletti A.; Murthy K. R.; Rathmann M.; Kumaramanickavel G.; Denton M. J.; Gal A.Nat. Genet.1997, 17, 194. [DOI] [PubMed] [Google Scholar]
- Schwartz S. H.; Tan B. C.; Gage D. A.; Zeevaart J. A.; McCarty D. R.Science1997, 276, 1872. [DOI] [PubMed] [Google Scholar]
- Tan B. C.; Schwartz S. H.; Zeevaart J. A.; McCarty D. R.Proc. Natl. Acad.Sci. U.S.A.1997, 94, 12235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choo D. W.; Cheung E.; Rando R. R.FEBS Lett.1998, 440, 195. [DOI] [PubMed] [Google Scholar]
- Mata N. L.; Moghrabi W. N.; Lee J. S.; Bui T. V.; Radu R. A.; Horwitz J.; Travis G. H.J. Biol. Chem.2004, 279, 635. [DOI] [PubMed] [Google Scholar]
- Moiseyev G.; Chen Y.; Takahashi Y.; Wu B. X.; Ma J. X.Proc. Natl. Acad. Sci. U.S.A.2005, 102, 12413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Redmond T. M.; Poliakov E.; Yu S.; Tsai J. Y.; Lu Z.; Gentleman S.Proc. Natl. Acad. Sci. U.S.A.2005, 102, 13658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kloer D. P.; Schulz G. E.Cell. Mol. Life Sci.2006, 63, 2291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moiseyev G.; Takahashi Y.; Chen Y.; Gentleman S.; Redmond T. M.; Crouch R. K.; Ma J. X.J.Biol. Chem.2006, 281, 2835. [DOI] [PubMed] [Google Scholar]
- Poliakov E.; Gubin A. N.; Stearn O.; Li Y.; Campos M. M.; Gentleman S.; Rogozin I. B.; Redmond T. M.PLoS One2012, 7, e49975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kusakabe T. G.; Takimoto N.; Jin M.; Tsuda M.Philos. Trans. R.Soc. London, B: Biol. Sci.2009, 364, 2897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiser P.D.; Golczak M.; Lodowski D. T.; Chance M. R.; Palczewski K.Proc. Natl. Acad.Sci. U.S.A.2009, 106, 17325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kloer D. P.; Ruch S.; Al-Babili S.; Beyer P.; Schulz G. E.Science2005, 308, 267. [DOI] [PubMed] [Google Scholar]
- Messing S. A. J.; Gabelli S. B.; Echeverria I.; Vogel J. T.; Guan J. C.; Tan B. C.; Klee H. J.; McCarty D. R.; Amzel L. M.Plant Cell2010, 22, 2970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiser P. D.; Farquhar E. R.; Shi W.; Sui X.; Chance M. R.; Palczewski K.Proc. Natl. Acad. Sci. U.S.A.2012, 109, E2747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bavik C. O.; Eriksson U.; Allen R. A.; Peterson P. A.J. Biol. Chem.1991, 266, 14978. [PubMed] [Google Scholar]
- Sagara H.; Hirosawa K.Exp. Eye Res.1991, 53, 765. [DOI] [PubMed] [Google Scholar]
- Garron L. K.Trans. Am. Ophthalmol.Soc.1963, 61, 545. [PMC free article] [PubMed] [Google Scholar]
- Golczak M.; Kiser P. D.; Lodowski D. T.; Maeda A.; Palczewski K.J. Biol. Chem.2010, 285, 9667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hemati N.; Feathers K. L.; Chrispell J. D.; Reed D. M.; Carlson T. J.; Thompson D. A.Mol. Vis.2005, 11, 1151. [PubMed] [Google Scholar]
- Ma J.; Zhang J.; Othersen K. L.; Moiseyev G.; Ablonczy Z.; Redmond T. M.; Chen Y.; Crouch R. K.Invest. Ophthalmol. Vis.Sci.2001, 42, 1429. [PubMed] [Google Scholar]
- Takahashi Y.; Moiseyev G.; Ablonczy Z.; Chen Y.; Crouch R. K.; Ma J. X.J. Biol. Chem.2009, 284, 3211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Picot D.; Garavito R. M.FEBS Lett.1994, 346, 21. [DOI] [PubMed] [Google Scholar]
- Nikolaeva O.; Moiseyev G.; Rodgers K. K.; Ma J. X.Biochem. J.2011, 436, 591. [DOI] [PubMed] [Google Scholar]
- Bernstein P. S.; Law W. C.; Rando R. R.J. Biol. Chem.1987, 262, 16848. [PubMed] [Google Scholar]
- Bracey M. H.; Cravatt B. F.; Stevens R. C.FEBS Lett.2004, 567, 159. [DOI] [PubMed] [Google Scholar]
- Chander P.; Gentleman S.; Poliakov E.; Redmond T. M.J. Biol. Chem.2012, 287, 30552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Redmond T. M.; Poliakov E.; Kuo S.; Chander P.; Gentleman S.J. Biol. Chem.2010, 285, 1919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lancaster C. R.; Michel H.J. Mol. Biol.1999, 286, 883. [DOI] [PubMed] [Google Scholar]
- Ferreira K. N.; Iverson T. M.; Maghlaoui K.; Barber J.; Iwata S.Science2004, 303, 1831. [DOI] [PubMed] [Google Scholar]
- Gillmor S. A.; Villasenor A.; Fletterick R.; Sigal E.; Browner M. F.Nat. Struct. Biol.1997, 4, 1003. [DOI] [PubMed] [Google Scholar]
- Hanein S.; Perrault I.; Gerber S.; Tanguy G.; Barbet F.; Ducroq D.; Calvas P.; Dollfus H.; Hamel C.; Lopponen T.; Munier F.; Santos L.; Shalev S.; Zafeiriou D.; Dufier J. L.; Munnich A.; Rozet J. M.; Kaplan J.Hum. Mutat.2004, 23, 306. [DOI] [PubMed] [Google Scholar]
- Nikolaeva O.; Takahashi Y.; Moiseyev G.; Ma J. X.Biochem. Biophys.Res. Commun.2010, 391, 1757. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi Y.; Moiseyev G.; Chen Y.; Ma J. X.FEBSLett.2005, 579, 5414. [DOI] [PubMed] [Google Scholar]
- Simovich M. J.; Miller B.; Ezzeldin H.; Kirkland B. T.; McLeod G.; Fulmer C.; Nathans J.; Jacobson S. G.; Pittler S. J.Hum. Mutat.2001, 18, 164. [DOI] [PubMed] [Google Scholar]
- Stecher H.; Gelb M. H.; Saari J. C.; Palczewski K.J. Biol. Chem.1999, 274, 8577. [DOI] [PubMed] [Google Scholar]
- Law W. C.; Rando R. R.Biochemistry1988, 27, 4147. [DOI] [PubMed] [Google Scholar]
- Jang G. F.; McBee J. K.; Alekseev A. M.; Haeseleer F.; Palczewski K.J. Biol. Chem.2000, 275, 28128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deigner P.S.; Law W. C.; Canada F. J.; Rando R. R.Science1989, 244, 968. [DOI] [PubMed] [Google Scholar]
- Wendt K. U.; Poralla K.; Schulz G. E.Science1997, 277, 1811. [DOI] [PubMed] [Google Scholar]
- Poliakov E.; Parikh T.; Ayele M.; Kuo S.; Chander P.; Gentleman S.; Redmond T. M.Biochemistry2011, 50, 6739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mandal M. N.; Moiseyev G. P.; Elliott M. H.; Kasus-Jacobi A.; Li X.; Chen H.; Zheng L.; Nikolaeva O.; Floyd R. A.; Ma J. X.; Anderson R. E.J. Biol. Chem.2011, 286, 32491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Noy N.; Xu Z. J.Biochemistry1990, 29, 3888. [DOI] [PubMed] [Google Scholar]
- Noy N.; Xu Z. J.Biochemistry1990, 29, 3883. [DOI] [PubMed] [Google Scholar]
- Molday R. S.; Beharry S.; Ahn J.; Zhong M.Adv. Exp. Med. Biol.2006, 572, 465. [DOI] [PubMed] [Google Scholar]
- Molday R. S.; Zhang K.Prog. Lipid Res.2010, 49, 476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molday R. S.; Zhong M.; Quazi F.Biochim. Biophys. Acta2009, 1791, 573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawaguchi R.; Yu J.; Honda J.; Hu J.; Whitelegge J.; Ping P.; Wiita P.; Bok D.; Sun H.Science2007, 315, 820. [DOI] [PubMed] [Google Scholar]
- Isken A.; Golczak M.; Oberhauser V.; Hunzelmann S.; Driever W.; Imanishi Y.; Palczewski K.; von Lintig J.Cell Metab.2008, 7, 258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim Y. K.; Wassef L.; Hamberger L.; Piantedosi R.; Palczewski K.; Blaner W. S.; Quadro L.J. Biol. Chem.2008, 283, 5611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J.; Bok D.Methods Mol. Biol.2010, 652, 55. [DOI] [PubMed] [Google Scholar]
- Amengual J.; Golczak M.; Palczewski K.; von Lintig J.J. Biol. Chem.2012, 287, 24216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alapatt P.; Guo F.; Komanetsky S. M.; Wang S.; Cai J.; Sargsyan A.; RodriguezDiaz E.; Bacon B. T.; Aryal P.; Graham T. E.J. Biol. Chem.2013, 288, 1250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Noy N.Biochem.J.2000, 348Pt 3481. [PMC free article] [PubMed] [Google Scholar]
- Newcomer M. E.; Jones T. A.; Aqvist J.; Sundelin J.; Eriksson U.; Rask L.; Peterson P. A.EMBO J.1984, 3, 1451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Naylor H. M.; Newcomer M. E.Biochemistry1999, 38, 2647. [DOI] [PubMed] [Google Scholar]
- Cowan S. W.; Newcomer M. E.; Jones T. A.J. Mol. Biol.1993, 230, 1225. [DOI] [PubMed] [Google Scholar]
- Saari J. C.; Nawrot M.; Garwin G. G.; Kennedy M. J.; Hurley J. B.; Ghyselinck N. B.; Chambon P.Invest. Ophthalmol. Vis. Sci.2002, 43, 1730. [PubMed] [Google Scholar]
- Futterman S.; Saari J. C.; Blair S.J. Biol. Chem.1977, 252, 3267. [PubMed] [Google Scholar]
- Futterman S.; Saari J. C.Invest. Ophthalmol. Vis. Sci.1977, 16, 768. [PubMed] [Google Scholar]
- Bunt-Milam A. H.; Saari J. C.J. Cell Biol.1983, 97, 703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saari J. C.; Bredberg D. L.J. Biol. Chem.1987, 262, 7618. [PubMed] [Google Scholar]
- Crabb J. W.; Goldflam S.; Harris S. E.; Saari J. C.J. Biol. Chem.1988, 263, 18688. [PubMed] [Google Scholar]
- Sparkes R.S.; Heinzmann C.; Goldflam S.; Kojis T.; Saari J. C.; Mohandas T.; Klisak I.; Bateman J. B.; Crabb J. W.Genomics1992, 12, 58. [DOI] [PubMed] [Google Scholar]
- Punta M.; Coggill P. C.; Eberhardt R. Y.; Mistry J.; Tate J.; Boursnell C.; Pang N.; Forslund K.; Ceric G.; Clements J.; Heger A.; Holm L.; Sonnhammer E. L.; Eddy S. R.; Bateman A.; Finn R. D.Nucleic Acids Res.2012, 40, D290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sha B.; Phillips S. E.; Bankaitis V. A.; Luo M.Nature1998, 391, 506. [DOI] [PubMed] [Google Scholar]
- He X. Q.; Lobsiger J.; Stocker A.Proc. Natl. Acad. Sci. U.S.A.2009, 106, 18545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saari J. C.; Nawrot M.; Stenkamp R. E.; Teller D. C.; Garwin G. G.Mol. Vis.2009, 15, 844. [PMC free article] [PubMed] [Google Scholar]
- Saari J. C.; Crabb J. W.Exp. Eye Res.2005, 81, 245. [DOI] [PubMed] [Google Scholar]
- Shaw N. S.; Noy N.Exp. Eye Res.2001, 72, 183. [DOI] [PubMed] [Google Scholar]
- Gonzalez-Fernandez F.; Baer C. A.; Ghosh D.BMC Biochem.2007, 8, 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gonzalez-Fernandez F.; Ghosh D.Exp.Eye Res.2008, 86, 169. [DOI] [PubMed] [Google Scholar]
- Adler A. J.; Stafford W. F. 3rd; Slayter H. S.J. Biol. Chem.1987, 262, 13198. [PubMed] [Google Scholar]
- Loew A.; Gonzalez-Fernandez F.Structure2002, 10, 43. [DOI] [PubMed] [Google Scholar]
- Molday L. L.; Rabin A. R.; Molday R. S.Nat. Genet.2000, 25, 257. [DOI] [PubMed] [Google Scholar]
- Sun H.; Nathans J.Nat. Genet.1997, 17, 15. [DOI] [PubMed] [Google Scholar]
- Papermaster D. S.; Reilly P.; Schneider B. G.Vision Res.1982, 22, 1417. [DOI] [PubMed] [Google Scholar]
- Beharry S.; Zhong M.; Molday R. S.J. Biol. Chem.2004, 279, 53972. [DOI] [PubMed] [Google Scholar]
- Sun H.; Smallwood P. M.; Nathans J.Nat. Genet.2000, 26, 242. [DOI] [PubMed] [Google Scholar]
- Weng J.; Mata N. L.; Azarian S. M.; Tzekov R. T.; Birch D. G.; Travis G. H.Cell1999, 98, 13. [DOI] [PubMed] [Google Scholar]
- Tsybovsky Y.; Wang B.; Quazi F.; Molday R. S.; Palczewski K.Biochemistry2011, 50, 6855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bungert S.; Molday L. L.; Molday R. S.J. Biol. Chem.2001, 276, 23539. [DOI] [PubMed] [Google Scholar]
- Tsybovsky Y.; Orban T.; Molday R. S.; Taylor D.; Palczewski K.Structure2013, 21, 854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun H.; Molday R. S.; Nathans J.J. Biol. Chem.1999, 274, 8269. [DOI] [PubMed] [Google Scholar]
- Quazi F.; Lenevich S.; Molday R. S.Nat. Commun.2012, 3, 925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho M. T.; Massey J. B.; Pownall H. J.; Anderson R. E.; Hollyfield J. G.J. Biol. Chem.1989, 264, 928. [PubMed] [Google Scholar]
- Rando R. R.; Bangerter F. W.Biochem. Biophys. Res. Commun.1982, 104, 430. [DOI] [PubMed] [Google Scholar]
- Plack P. A.; Pritchard D. J.Biochem. J.1969, 115, 927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mata N. L.; Weng J.; Travis G. H.Proc. Natl. Acad. Sci.U.S.A.2000, 97, 7154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahn J.; Wong J. T.; Molday R. S.J. Biol. Chem.2000, 275, 20399. [DOI] [PubMed] [Google Scholar]
- Chen Y.; Okano K.; Maeda T.; Chauhan V.; Golczak M.; Maeda A.; Palczewski K.J. Biol. Chem.2012, 287, 5059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lochner J. E.; Badwey J. A.; Horn W.; Karnovsky M. L.Proc. Natl. Acad.Sci. U.S.A.1986, 83, 7673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carmody R. J.; McGowan A. J.; Cotter T. G.Exp. Cell Res.1999, 248, 520. [DOI] [PubMed] [Google Scholar]
- Fishkin N. E.; Sparrow J. R.; Allikmets R.; Nakanishi K.Proc. Natl. Acad. Sci. U.S.A.2005, 102, 7091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parish C.A.; Hashimoto M.; Nakanishi K.; Dillon J.; Sparrow J.Proc. Natl. Acad. Sci. U.S.A.1998, 95, 14609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ben-Shabat S.; Parish C. A.; Vollmer H. R.; Itagaki Y.; Fishkin N.; Nakanishi K.; Sparrow J. R.J. Biol. Chem.2002, 277, 7183. [DOI] [PubMed] [Google Scholar]
- Kaufman Y.; Ma L.; Washington I.J. Biol. Chem.2011, 286, 7958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma L.; Kaufman Y.; Zhang J.; Washington I.J. Biol. Chem.2011, 286, 7966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sparrow J. R.; Fishkin N.; Zhou J.; Cai B.; Jang Y. P.; Krane S.; Itagaki Y.; Nakanishi K.Vision Res.2003, 43, 2983. [DOI] [PubMed] [Google Scholar]
- Rozanowski B.; Burke J.; Sarna T.; Rozanowska M.Photochem. Photobiol.2008, 84, 658. [DOI] [PubMed] [Google Scholar]
- Rozanowski B.; Burke J. M.; Boulton M. E.; Sarna T.; Rozanowska M.Invest. Ophthalmol. Vis. Sci.2008, 49, 2838. [DOI] [PubMed] [Google Scholar]
- Ng K. P.; Gugiu B.; Renganathan K.; Davies M. W.; Gu X.; Crabb J. S.; Kim S. R.; Rozanowska M. B.; Bonilha V. L.; Rayborn M. E.; Salomon R. G.; Sparrow J. R.; Boulton M. E.; Hollyfield J. G.; Crabb J. W.Mol. Cell. Proteomics2008, 7, 1397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Y.; Zhou J.; Fishkin N.; Rittmann B. E.; Sparrow J. R.J. Am. Chem. Soc.2011, 133, 849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haralampus-Grynaviski N. M.; Lamb L. E.; Clancy C. M.; Skumatz C.; Burke J. M.; Sarna T.; Simon J. D.Proc. Natl. Acad.Sci. U.S.A.2003, 100, 3179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allikmets R.; Singh N.; Sun H.; Shroyer N. F.; Hutchinson A.; Chidambaram A.; Gerrard B.; Baird L.; Stauffer D.; Peiffer A.; Rattner A.; Smallwood P.; Li Y.; Anderson K. L.; Lewis R. A.; Nathans J.; Leppert M.; Dean M.; Lupski J. R.Nat. Genet.1997, 15, 236. [DOI] [PubMed] [Google Scholar]
- Sparrow J. R.; Boulton M.Exp. Eye Res.2005, 80, 595. [DOI] [PubMed] [Google Scholar]
- Rattner A.; Sun H.; Nathans J.Annu. Rev. Genet.1999, 33, 89. [DOI] [PubMed] [Google Scholar]
- Hartong D. T.; Berson E. L.; Dryja T. P.Lancet2006, 368, 1795. [DOI] [PubMed] [Google Scholar]
- Rivolta C.; Sharon D.; DeAngelis M. M.; Dryja T. P.Hum. Mol. Genet.2002, 11, 1219. [DOI] [PubMed] [Google Scholar]
- Dryja T. P.; Berson E. L.Invest. Ophthalmol. Vis. Sci.1995, 36, 1197. [PubMed] [Google Scholar]
- Dryja T. P.; Li T.Hum. Mol. Genet.1995, 4 Spec No, 1739. [DOI] [PubMed] [Google Scholar]
- Allikmets R.Am. J. Hum. Genet.2000, 67, 793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dryja T. P.; McGee T. L.; Reichel E.; Hahn L. B.; Cowley G. S.; Yandell D. W.; Sandberg M. A.; Berson E. L.Nature1990, 343, 364. [DOI] [PubMed] [Google Scholar]
- Rosenfeld P. J.; Cowley G. S.; McGee T. L.; Sandberg M. A.; Berson E. L.; Dryja T. P.Nat. Genet.1992, 1, 209. [DOI] [PubMed] [Google Scholar]
- Azam M.; Khan M. I.; Gal A.; Hussain A.; Shah S. T.; Khan M. S.; Sadeque A.; Bokhari H.; Collin R. W.; Orth U.; van Genderen M. M.; den Hollander A. I.; Cremers F. P.; Qamar R.Mol. Vis.2009, 15, 2526. [PMC free article] [PubMed] [Google Scholar]
- Kumaramanickavel G.; Maw M.; Denton M. J.; John S.; Srikumari C. R.; Orth U.; Oehlmann R.; Gal A.Nat. Genet.1994, 8, 10. [DOI] [PubMed] [Google Scholar]
- Caruso R. C.; Aleman T. S.; Cideciyan A. V.; Roman A. J.; Sumaroka A.; Mullins C. L.; Boye S. L.; Hauswirth W. W.; Jacobson S. G.Invest. Ophthalmol. Vis. Sci.2010, 51, 5304. [DOI] [PubMed] [Google Scholar]
- Cideciyan A. V.; Swider M.; Aleman T. S.; Tsybovsky Y.; Schwartz S. B.; Windsor E. A.; Roman A. J.; Sumaroka A.; Steinberg J. D.; Jacobson S. G.; Stone E. M.; Palczewski K.Hum. Mol. Genet.2009, 18, 931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allikmets R.Nat. Genet.1997, 17, 122. [DOI] [PubMed] [Google Scholar]
- Allikmets R.Eur. J. Ophthalmol.1999, 9, 255. [DOI] [PubMed] [Google Scholar]
- Allikmets R.Am. J. Hum. Genet.2000, 67, 487. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Driessen C. A.; Winkens H. J.; Hoffmann K.; Kuhlmann L. D.; Janssen B. P.; Van Vugt A. H.; Van Hooser J. P.; Wieringa B. E.; Deutman A. F.; Palczewski K.; Ruether K.; Janssen J. J.Mol. Cell. Biol.2000, 20, 4275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto H.; Simon A.; Eriksson U.; Harris E.; Berson E. L.; Dryja T. P.Nat. Genet.1999, 22, 188. [DOI] [PubMed] [Google Scholar]
- Dryja T. P.Am. J. Ophthalmol.2000, 130, 547. [DOI] [PubMed] [Google Scholar]
- denHollander A. I.; McGee T. L.; Ziviello C.; Banfi S.; Dryja T. P.; Gonzalez-Fernandez F.; Ghosh D.; Berson E. L.Invest. Ophthalmol.Vis. Sci.2009, 50, 1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baehr W.; Frederick J. M.Vision Res.2009, 49, 2636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Batten M. L.; Imanishi Y.; Tu D. C.; Doan T.; Zhu L.; Pang J. J.; Glushakova L.; Moise A. R.; Baehr W.; Van Gelder R. N.; Hauswirth W. W.; Rieke F.; Palczewski K.PLoS Med.2005, 2, 1177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Hooser J. P.; Liang Y.; Maeda T.; Kuksa V.; Jang G. F.; He Y. G.; Rieke F.; Fong H. K.; Detwiler P. B.; Palczewski K.J. Biol. Chem.2002, 277, 19173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Hooser J. P.; Aleman T. S.; He Y. G.; Cideciyan A. V.; Kuksa V.; Pittler S. J.; Stone E. M.; Jacobson S. G.; Palczewski K.Proc. Natl. Acad. Sci. U.S.A.2000, 97, 8623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda T.; Maeda A.; Leahy P.; Saperstein D. A.; Palczewski K.Invest. Ophthalmol. Vis. Sci.2009, 50, 322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda T.; Perusek L.; Amengual J.; Babino D.; Palczewski K.; von Lintig J.Mol. Pharmacol.2011, 80, 943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda A.; Golczak M.; Chen Y.; Okano K.; Kohno H.; Shiose S.; Ishikawa K.; Harte W.; Palczewska G.; Maeda T.; Palczewski K.Nat. Chem. Biol.2012, 8, 170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Golczak M.; Maeda A.; Bereta G.; Maeda T.; Kiser P. D.; Hunzelmann S.; von Lintig J.; Blaner W. S.; Palczewski K.J. Biol. Chem.2008, 283, 9543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kijlstra A.; Tian Y.; Kelly E. R.; Berendschot T. T.Prog. RetinalEye Res.2012, 31, 303. [DOI] [PubMed] [Google Scholar]
- Maeda T.; Dong Z.; Jin H.; Sawada O.; Gao S.; Utkhede D.; Monk W.; Palczewska G.; Palczewski K.Invest. Ophthalmol. Vis. Sci.2013, 54, 455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Radu R. A.; Hu J.; Peng J.; Bok D.; Mata N. L.; Travis G. H.J. Biol. Chem.2008, 283, 19730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Franzoni L.; Lucke C.; Perez C.; Cavazzini D.; Rademacher M.; Ludwig C.; Spisni A.; Rossi G. L.; Ruterjans H.J. Biol. Chem.2002, 277, 21983. [DOI] [PubMed] [Google Scholar]
- Lu J.; Lin C. L.; Tang C.; Ponder J. W.; Kao J. L.; Cistola D. P.; Li E.J. Mol. Biol.1999, 286, 1179. [DOI] [PubMed] [Google Scholar]
- Travis G. H.; Radu R. A.; Lee J.; Mata N. L.Invest. Ophthalmol.Vis. Sci.2002, 43, U1006. [Google Scholar]