首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
4.
5.
6.
Hyperhomocysteinemia has long been associated with atherosclerosis and thrombosis and is an independent risk factor for cardiovascular disease. Its causes include both genetic and environmental factors. Although homocysteine is produced in every cell as an intermediate of the methionine cycle, the liver contributes the major portion found in circulation, and fatty liver is a common finding in homocystinuric patients. To understand the spectrum of proteins and associated pathways affected by hyperhomocysteinemia, we analyzed the mouse liver proteome of gene-induced (cystathionine β-synthase (CBS)) and diet-induced (high methionine) hyperhomocysteinemic mice using two-dimensional difference gel electrophoresis and Ingenuity Pathway Analysis. Nine proteins were identified whose expression was significantly changed by 2-fold (p ≤ 0.05) as a result of genotype, 27 proteins were changed as a result of diet, and 14 proteins were changed in response to genotype and diet. Importantly, three enzymes of the methionine cycle were up-regulated. S-Adenosylhomocysteine hydrolase increased in response to genotype and/or diet, whereas glycine N-methyltransferase and betaine-homocysteine methyltransferase only increased in response to diet. The antioxidant proteins peroxiredoxins 1 and 2 increased in wild-type mice fed the high methionine diet but not in the CBS mutants, suggesting a dysregulation in the antioxidant capacity of those animals. Furthermore, thioredoxin 1 decreased in both wild-type and CBS mutants on the diet but not in the mutants fed a control diet. Several urea cycle proteins increased in both diet groups; however, arginase 1 decreased in the CBS+/− mice fed the control diet. Pathway analysis identified the retinoid X receptor signaling pathway as the top ranked network associated with the CBS+/− genotype, whereas xenobiotic metabolism and the NRF2-mediated oxidative stress response were associated with the high methionine diet. Our results show that hyperhomocysteinemia, whether caused by a genetic mutation or diet, alters the abundance of several liver proteins involved in homocysteine/methionine metabolism, the urea cycle, and antioxidant defense.Homocysteine (Hcy)1 is a thiol-containing amino acid that is produced in every cell of the body as an intermediate of the methionine cycle (Fig. 1, Reactions 1–5) (1). Once formed, the catabolism of homocysteine occurs via three enzymatic pathways. 1) Hcy is remethylated back to methionine using vitamin B12-dependent methionine synthase (Fig. 1, Reaction 4) and/or 2) betaine-homocysteine methyltransferase (BHMT) (Fig. 1, Reaction 5), and 3) Hcy is converted to cysteine via the transsulfuration pathway using CBS and γ-cystathionase (Fig. 1, Reactions 6 and 7). Under normal conditions ∼40–50% of the Hcy that is produced in the liver is remethylated, ∼40–50% is converted to cysteine, and a small amount is exported (13). However, when Hcy production is increased (i.e. increased dietary methionine/protein intake) or when Hcy catabolism is decreased (i.e. CBS deficiency or B vitamin deficiencies), excess Hcy is exported into the extracellular space, resulting in hyperhomocysteinemia (15).Open in a separate windowFig. 1.Homocysteine metabolism in liver and kidney. In classical homocystinuria, the initial enzyme of the transsulfuration pathway, CBS (Reaction 6), is deficient. MTHF, methylenetetrahydrofolate; THF, tetrahydrofolate; DHF, dihydrofolate; MeCbl, methylcobalamin; DMG, dimethylglycine; PLP, pyridoxal 5′-phosphate.Homocystinuria was first described in the 1960s by Carson et al. (6): they observed 10 pediatric patients with severely elevated levels of Hcy in the urine and hypermethioninemia. Normal concentrations of plasma total homocysteine (tHcy) range from 5 to 12 μm (7); however, in homocystinuria, tHcy levels can exceed 100 μm. Homocystinuric patients present with mental retardation, abnormal bone growth, fine hair, malar flush, and dislocation of the lens of the eye, and most die from premature cardiovascular disease (6, 8). Autopsy findings indicate widespread thromboembolism, arteriosclerosis, and fatty livers (6, 8). Mudd et al. (9, 10) identified the cause of homocystinuria as a defect in the enzyme cystathionine β-synthase. A recent study of newborn infants in Denmark estimated the birth prevalence for CBS heterozygosity to be about 1:20,000 (11).Plasma tHcy concentrations are also directly correlated with dietary methionine/protein intake (12, 13). Guttormsen et al. (13) demonstrated that a protein-rich meal affected tHcy for at least 8–24 h. When normal subjects were fed a low protein-containing breakfast (12–15 g), plasma methionine levels increased slightly after 2 h (22.5–27.5 μm), but tHcy levels did not change significantly. However, when these same subjects were fed a high protein meal (52 g), plasma methionine levels peaked after 4 h (38 μm), and tHcy rose steadily until a maximum level was reached 8 h postmeal (7.6 versus 8.5 μm) (13). Thus, the following questions can be raised. How does the hepatic proteome respond to a hyperhomocysteinemic diet, and are the changes that accompany such a diet the same as or different from those that may be observed in gene-induced hyperhomocysteinemia?Because hyperhomocysteinemia is a strong independent risk factor for cardiovascular, cerebrovascular, and peripheral vascular disease, most of the current research has focused on the mechanisms involved in Hcy-induced endothelial dysfunction (1424). The results of those studies have concluded that Hcy induces intracellular oxidative stress by generating ROS, which in turn lead to decreased bioavailable nitric oxide (NO), altered gene expression, increased endoplasmic reticulum stress, and activation of cholesterol biosynthesis. Also, several studies have examined the association between hyperhomocysteinemia and alcoholic liver disease, but few have looked at the effect of Hcy on the non-alcoholic liver even though fatty liver is a constant finding in homocystinuria (6, 8), and the liver is the major source of circulating Hcy (4, 5, 10). We hypothesize that 1) the liver proteome will respond to hyperhomocysteinemia by altering the expression of proteins involved in methionine/homocysteine metabolism and antioxidant defense and that 2) the set of proteins that are expressed when hyperhomocysteinemia is induced by CBS deficiency will differ from those expressed as a result of a high methionine diet. In the present study, we use a well established mouse model of CBS deficiency to study the early changes in the liver proteome that accompany hyperhomocysteinemia (25).  相似文献   

7.
Fe2+ is now shown to weaken binding between ferritin and mitochondrial aconitase messenger RNA noncoding regulatory structures ((iron-responsive element) (IRE)-RNAs) and the regulatory proteins (IRPs), which adds a direct role of iron to regulation that can complement the well known regulatory protein modification and degradative pathways related to iron-induced mRNA translation. We observe that the Kd value increases 17-fold in 5′-untranslated region IRE-RNA·repressor complexes; Fe2+, is studied in the absence of O2. Other metal ions, Mn2+ and Mg2+ have similar effects to Fe2+ but the required Mg2+ concentration is 100 times greater than for Fe2+ or Mn2+. Metal ions also weaken ethidium bromide binding to IRE-RNA with no effect on IRP fluorescence, using Mn2+ as an O2-resistant surrogate for Fe2+, indicating that metal ions bound IRE-RNA but not IRP. Fe2+ decreases IRP repressor complex stability of ferritin IRE-RNA 5–10 times compared with 2–5 times for mitochondrial aconitase IRE-RNA, over the same concentration range, suggesting that differences among IRE-RNA structures contribute to the differences in the iron responses observed in vivo. The results show the IRE-RNA·repressor complex literally responds to Fe2+, selectively for each IRE-mRNA.Iron (e.g. ferrous sulfate, ferric citrate, and hemin) added to animal cells changes translation rates of messenger RNAs encoding proteins of iron traffic and oxidative metabolism (14). To cross cell membranes, iron ions are transported by membrane proteins such as DMT1 or carried on proteins such as transferrin. Inside the cells, iron is mainly in heme, FeS clusters, non-heme iron cofactors of proteins, and iron oxide minerals coated by protein nanocages (ferritins). Iron in transit is thought to be Fe2+ in labile “pools” accessible to small molecular weight chelators, and/or bound loosely by chaperones.When iron concentrations in the cells increase, a group of mRNAs with three-dimensional, noncoding structures in the 5′-untranslated region (UTR)3 are derepressed (Fig. 1A), i.e. the fraction of the mRNAs in mRNA·repressor protein complexes, which inhibit ribosome binding, decreases and the fraction of the mRNAs in polyribosomes increases (57). The three-dimensional, noncoding mRNA structure, representing a family of related structures, is called the iron-responsive element, or IRE, and the repressors are called iron regulatory proteins (IRPs). Together they are one of the most extensively studied eukaryotic messenger RNA regulatory systems (14). In addition to large numbers of cell studies, structures of IRE-RNAs are known from solution NMR (812), and the RNA·protein complex from x-ray crystallography (13). Recent data indicate that demetallation of IRP1 and disruption of the [4Fe-4S] cluster that inhibits IRP1 binding to RNA will be enhanced by phosphorylation and low iron concentrations (1, 2, 1416). Such results can explain the increased IRP1 binding to IRE-mRNAs and increased translational repression when iron concentrations are abnormally low. However, the mechanism to explain dissociation of IRE-RNA·IRP complexes, thereby allowing ribosome assembly and increased proteosomal degradation of IRPs (1, 2, 14, 15) (Fig. 1A), when high iron concentrations are abnormally high, is currently unknown.Open in a separate windowFIGURE 1.IRE-RNA·IRP complexes and a model for depression by excess iron. A, a representative model of iron-induced translation of 5′-UTR IRE-RNAs. This figure is modified from Ref. 7. B, IRE-RNA sites influenced by metal binding related to the crystal structure of the ferritin-IRE-RNA·IRP complex from Ref. 13. The figure was created by T. Tosha using Discovery Studio 1.6 and Protein Data Bank file 2IPY. ■, hydrated Mg2+, determined by solution NMR; ▴, Cu1+-1.10-phenanthroline, determined by RNA cleavage in O2.Metal ion binding changes conformation and function of most RNA classes, e.g. rRNA (17), tRNA (18, 19), ribozymes (2023), riboswitches (24, 25), possibly hammerhead mRNAs in mammals (26), and proteins. Although the effects of metal ion binding on eukaryotic mRNAs have not been extensively studied, Mg2+ is known to cause changes in conformation, shown by changes in radical cleavage sites of IRE-RNA with 1,10-phenanthrolene-iron and proton shifts in the one-dimensional NMR spectrum (12, 27). The Mg2+ effects are observed at low magnesium concentrations (0.1–0.5 mm) and low molar stoichiometries (1:1 and 2:1 = Mg:RNA).We hypothesized that Fe2+ could directly change the binding of the IRE-mRNA to the iron regulatory protein for several reasons. First, other metal ions influence the IRE-RNA structure (12, 27). Second, in IRE-RNA/IRP cocrystals there are exposed RNA sites in the IRE-RNA/IRP complex that are accessible for interactions (13) (Fig. 1B). Third, regions in the IRE-RNA are hypersensitive to Fe2+-EDTA/ascorbate/H2O2, suggesting selective interactions with metals and/or solvent (28). We now report that Fe2+ weakens IRE-RNA/IRP binding, whereas Mg2+ requires 100 times the concentration and Mn2+ is comparable with Fe2+; the Fe2+ effect was diminished in mutant IRE-RNA and IRE-RNA selective in wild type sequences: ferritin IRE-RNA > mt-aconitase IRE-RNA.  相似文献   

8.
9.
10.
11.
12.
13.
Malic enzyme has a dimer of dimers quaternary structure in which the dimer interface associates more tightly than the tetramer interface. In addition, the enzyme has distinct active sites within each subunit. The mitochondrial NAD(P)+-dependent malic enzyme (m-NAD(P)-ME) isoform behaves cooperatively and allosterically and exhibits a quaternary structure in dimer-tetramer equilibrium. The cytosolic NADP+-dependent malic enzyme (c-NADP-ME) isoform is noncooperative and nonallosteric and exists as a stable tetramer. In this study, we analyze the essential factors governing the quaternary structure stability for human c-NADP-ME and m-NAD(P)-ME. Site-directed mutagenesis at the dimer and tetramer interfaces was employed to generate a series of dimers of c-NADP-ME and m-NAD(P)-ME. Size distribution analysis demonstrated that human c-NADP-ME exists mainly as a tetramer, whereas human m-NAD(P)-ME exists as a mixture of dimers and tetramers. Kinetic data indicated that the enzyme activity of c-NADP-ME is not affected by disruption of the interface. There are no significant differences in the kinetic properties between AB and AD dimers, and the dimeric form of c-NADP-ME is as active as tetramers. In contrast, disrupting the interface of m-NAD(P)-ME causes the enzyme to be less active than wild type and to become less cooperative for malate binding; the kcat values of mutants decreased with increasing Kd,24 values, indicating that the dissociation of subunits at the dimer or tetramer interfaces significantly affects the enzyme activity. The above results suggest that the tetramer is required for a fully functional m-NAD(P)-ME. Taken together, the analytical ultracentrifugation data and the kinetic analysis of these interface mutants demonstrate the differential role of tetramer organization for the c-NADP-ME and m-NAD(P)-ME isoforms. The regulatory mechanism of m-NAD(P)-ME is closely related to the tetramer formation of this isoform.Malic enzymes catalyze a reversible oxidative decarboxylation of l-malate to yield pyruvate and CO2 with reduction of NAD(P)+ to NAD(P)H. This reaction requires a divalent metal ion (Mg2+ or Mn2+) for catalysis (13). Malic enzymes are found in a broad spectrum of living organisms that share conserved amino acid sequences and structural topology; such shared characteristics reveal a crucial role for the biological functions of these enzymes (4, 5). In mammals, malic enzymes have been divided into three isoforms according to their cofactor specificity and subcellular localization as follows: cytosolic NADP+-dependent (c-NADP-ME),2 mitochondrial NADP+-dependent (m-NADP-ME), and mitochondrial NAD(P)+-dependent (m-NAD(P)-ME). The m-NAD(P)-ME isoform displays dual cofactor specificity; it can use both NAD+ and NADP+ as the coenzyme, but NAD+ is more favored in a physiological environment (68). Dissimilar to the other two isoforms, m-NAD(P)-ME binds malate cooperatively, and it can be allosterically activated by fumarate; the sigmoidal kinetics observed with cooperativity is abolished by fumarate (912). Mutagenesis and kinetic studies demonstrated that ATP is an active-site inhibitor, although it also binds to the exo sites in the tetramer interface (1315). Structural studies also revealed an allosteric binding site for fumarate residing at the dimer interface. Mutation in the binding site significantly affects the activating effects of fumarate (11, 16, 17).The c-NADP-ME and m-NADP-ME isoforms play an important role in lipogenesis by providing NADPH for the biosynthesis of long-chain fatty acids and steroids. Thus, c-NADP-ME together with acetyl-CoA carboxylase, fatty-acid synthase, and glucose-6-phosphate dehydrogenase are classified as lipogenic enzymes (2, 1821). The m-NAD(P)-ME isoform has attracted much attention because it is involved in glutaminolysis, which is an energy-producing pathway of tumor cells that utilizes glutamine and glutamate. Thus, m-NAD(P)-ME is considered to be a potential target in cancer therapy (2227).Various crystal structures of malic enzymes in complex with substrate, metal ion, coenzyme, regulator, and inhibitor are available in the Protein Data Bank (4, 11, 2832). The overall tertiary structures of these malic enzymes are similar, but there are still some differences that may be significant for catalysis and regulation of the enzyme. Malic enzyme exists as a dimer of dimers with a stronger association of the dimer interface than that of the tetramer interface (Fig. 1A). The dimer interface is formed by subunits A and B or C and D (Fig. 1B), whereas the tetramer interaction consists of contacts between subunits A and D or B and C (Fig. 1C). A hydrophobic interaction is the major driving force for subunit assembly, but hydrogen bonding and ionic interactions also contribute markedly. The homotetramer of the enzyme is composed of four identical monomers each with its own active site. In the structure of human m-NAD(P)-ME, aside from its well defined active site, there are two regulatory sites on the enzyme (Fig. 1A). One of these sites is located at the dimer interface and is occupied by fumarate (Fig. 1B), whereas the other site, which is referred to as the exo site, is located at the tetramer interface and is occupied by either an NAD or an ATP molecule (Fig. 1A). In the ME family, Ascaris suum and human m-NAD(P)-ME were found to be activated by fumarate (11, 1517, 31). However, the relationship between enzyme regulation and subunit-subunit interaction is still unclear.Open in a separate windowFIGURE 1.Dimer and tetramer interfaces of human m-NAD(P)-ME. A, dimer of dimers quaternary structure of human m-NAD(P)-ME (Protein Data Bank code 1PJ3). The active site, fumarate site, and exo site in each subunit are indicated. B, dimer interface between A and B subunits of m-NAD(P)-ME. C, tetramer interface between A and D subunits of m-NAD(P)-ME. The amino acid residues at the dimer interface, Gln-51, Glu-90, Asp-139, His-142, and Asp-568 and C terminus in the tetramer interface, are represented by ball-and-stick modeling. The amino acid residues 51 and 90 in human c-NADP-ME are His and Asp, respectively. This figure was generated with PyMOL (DeLano Scientific LLC, San Carlos, CA).Previous studies have shown that the quaternary structure stability of malic enzyme isoforms is diverse. At neutral pH, pigeon c-NADP-ME exists as a unique tetramer with a sedimentation coefficient of ∼10 S (3335), whereas human m-NAD(P)-ME exists as a mixture of tetramer and dimer of 9.5 S and 6.5 S, respectively (13, 35). Some mutations at the interface will affect the quaternary structure (3437). Although the crystal structure of human c-NADP-ME has not been resolved, it is believed that it is similar to pigeon c-NADP-ME.Here we analyze the essential factors governing quaternary structure stability for human c-NADP-ME and m-NAD(P)-ME. Site-directed mutagenesis at the dimer and tetramer interfaces was used to disrupt the tetramer organization to create a series of c-NADP-ME and m-NAD(P)-ME dimers. Combined with the analytical ultracentrifugation data and kinetic analysis of these interface mutants, we demonstrate the differential role of tetramer organization for the c-NADP-ME and m-NAD(P)-ME isoforms. The regulatory mechanism of m-NAD(P)-ME is highly associated with the tetramer formation of this isoform.  相似文献   

14.
Mammalian CD38 and its Aplysia homolog, ADP-ribosyl cyclase (cyclase), are two prominent enzymes that catalyze the synthesis and hydrolysis of cyclic ADP-ribose (cADPR), a Ca2+ messenger molecule responsible for regulating a wide range of cellular functions. Although both use NAD as a substrate, the cyclase produces cADPR, whereas CD38 produces mainly ADP-ribose (ADPR). To elucidate the catalytic differences and the mechanism of cyclizing NAD, the crystal structure of a stable complex of the cyclase with an NAD analog, ribosyl-2′F-2′deoxynicotinamide adenine dinucleotide (ribo-2′-F-NAD), was determined. The results show that the analog was a substrate of the cyclase and that during the reaction, the nicotinamide group was released and a stable intermediate was formed. The terminal ribosyl unit at one end of the intermediate formed a close linkage with the catalytic residue (Glu-179), whereas the adenine ring at the other end stacked closely with Phe-174, suggesting that the latter residue is likely to be responsible for folding the linear substrate so that the two ends can be cyclized. Mutating Phe-174 indeed reduced cADPR production but enhanced ADPR production, converting the cyclase to be more CD38-like. Changing the equivalent residue in CD38, Thr-221 to Phe, correspondingly enhanced cADPR production, and the double mutation, Thr-221 to Phe and Glu-146 to Ala, effectively converted CD38 to a cyclase. This study provides the first detailed evidence of the cyclization process and demonstrates the feasibility of engineering the reactivity of the enzymes by mutation, setting the stage for the development of tools to manipulate cADPR metabolism in vivo.Cyclic ADP-ribose is a novel cyclic nucleotide with Ca2+-mobilizing activity targeting the endoplasmic reticulum. Its activity was first described in sea urchin eggs (1, 2), and cADPR3 has since been established as a second messenger molecule responsible for regulating a wide range of physiological functions, from fission in the dinoflagellate (3) to social behavior in mice (Ref. 4 and reviewed in Refs. 5 and 6). The Aplysia ADP-ribosyl cyclase (cyclase) was the first protein identified that uses NAD, a linear substrate, and ligates its two ends to produce cADPR, with the release of the terminal moiety, nicotinamide (7). The cyclase is a soluble protein of 30 kDa and is present in large amounts in Aplysia ovotestis (7). It is also present in the neurons of the Aplysia buccal ganglion, where it is responsible for the synthesis of endogenous cADPR and the regulation of the evoked synaptic transmission (8). Recently, it is shown that depolarization of Aplysia neurons induces the translocation of the cyclase from the cytosol into the nucleus, providing a mechanism for fine tuning of nuclear Ca2+ signals in neurons (9).CD38 is the major mammalian homolog of the cyclase and is responsible for regulating a wide range of physiological functions. Deletion of the CD38 gene in mice produces multiple defects, including impairment of insulin secretion (10), neutrophil chemotaxis (11), and oxytocin release (4). Catalytically, CD38 is quite different from the cyclase. Although both use NAD as substrate, CD38 produces only a small amount of cADPR, whereas the major product is ADP-ribose (ADPR) instead (1215) (Fig. 1a). It can also use cADPR as a substrate and hydrolyze it to ADPR (1215). Ablation of the CD38 gene in mice, nevertheless, results in a large reduction in endogenous cADPR in many tissues (10, 11). CD38 is thus responsible for both the synthesis and the hydrolysis of cADPR in mammalian cells.Open in a separate windowFIGURE 1.Crystal structure of the complex of cyclase with ribo-2′F-NAD. a, chemical structure of the substrate ribo-2′F-NAD and the reactions catalyzed by CD38 and the cylase. b, crystal structure of the cyclase dimer with the intermediates at each of the active sites of the monomer. The color scheme for the secondary structures is: red, α-helix; yellow, β-sheet; gray, coil. The color scheme for the residues is: cyan, Tyr-81; beige, Phe-174; blue, Glu-179; magenta, Glu-98; purple, Phe-175. The intermediates are colored by their elements: green, carbon; red, oxygen; orange, phosphorus; blue, nitrogen; light green, fluorine. c, stereo view of the folded conformation with electron density from an omit FoFc map contoured at 2.7 σ and shown as blue wire mesh. Other color schemes are the same as in b. The average B-factor is 76 Å2 for the folded intermediate.In fact, both CD38 and the cyclase are multifunctional enzymes that can also use NADP as a substrate and, in the presence of nicotinic acid, produce nicotinic acid adenine dinucleotide phosphate (NAADP) via a base-exchange reaction (16). NAADP was first shown to have Ca2+-mobilizing activity in sea urchin eggs (17) and has since been established as another Ca2+ messenger molecule targeting yet another intracellular Ca2+ store, the lysosome, in a variety of cell types (1820).To elucidate the mechanism of cyclizing a long linear substrate such as NAD to a compact cyclic product, cADPR, here we present the crystal structure of a stable complex of the cyclase with a substrate analog of NAD. The structure clearly identified critical residues for the cyclization process, which were verified by site-directed mutagenesis. The results demonstrate that catalysis by CD38 or the cyclase is controlled by one or two critical residues and that mutating them can interconvert the reactivities of the two enzymes. This study sets the stage for engineering enzymes with specific activity toward cADPR for expression in cells, which should be valuable tools for manipulating the function and metabolism of this novel Ca2+ messenger.  相似文献   

15.
16.
Fibronectin (FN) is a large extracellular matrix glycoprotein important for development and wound healing in vertebrates. Recent work has focused on the ability of FN fragments and embryonic or tumorigenic splicing variants to stimulate fibroblast migration into collagen gels. This activity has been localized to specific sites and is not exhibited by full-length FN. Here we show that an N-terminal FN fragment, spanning the migration stimulation sites and including the first three type III FN domains, also lacks this activity. A screen for interdomain interactions by solution-state NMR spectroscopy revealed specific contacts between the Fn N terminus and two of the type III domains. A single amino acid substitution, R222A, disrupts the strongest interaction, between domains 4–5FnI and 3FnIII, and restores motogenic activity to the FN N-terminal fragment. Anastellin, which promotes fibril formation, destabilizes 3FnIII and disrupts the observed 4–5FnI-3FnIII interaction. We discuss these findings in the context of the control of cellular activity through exposure of masked sites.Fibronectin (FN),4 a large multidomain glycoprotein found in all vertebrates, plays a vital role in cell adhesion, tissue development, and wound healing (1). It exists in soluble form in plasma and tissue fluids but is also present in fibrillar networks as part of the extracellular matrix. The structures of many FN domains of all three types, FnI, FnII, and FnIII, are known, for example (24). Although interactions between domains that are close in primary sequence have been demonstrated (3, 5), studies of multidomain fragments generally assume a beads-on-string model (2). There is, however, much evidence for the presence of long range order in soluble FN as a number of functional sites, termed cryptic, are not active in the native molecule, until exposed through conformational change. These include self-association sites (58), sites of protein interactions (9), and sites that control cellular activity (10, 11). Low resolution studies of the FN dimer suggest a compact conformation under physiological conditions (1214); however, attempts to define large scale structure in FN by small angle scattering or electric birefringence (1517) have yielded contradictory results. Interpretation of domain stability changes in terms of interaction sites (18) has also not been straightforward (2), possibly because of domain stabilization through nearest-neighbor effects (19, 20).A FN splicing variant produced in fetal and cancer patient fibroblasts, termed migration stimulation factor (MSF), stimulates migration of adult skin fibroblasts into type I collagen gels (10, 21) and breast carcinoma cells using the Boyden chamber (22). MSF comprises FN domains 1FnI to 9FnI, a truncated 1FnIII, and a small C-terminal extension; a recombinant FN fragment corresponding to 1FnI-9FnI (Fn70kDa) displays the same activity (10). An overview of FN domain structure and nomenclature is presented in Fig. 1a. Further experiments sub-localized full motogenic activity to the gelatin binding domain of FN (GBD, domains 6FnI-9FnI) (23) and partial activity to a shorter fragment spanning domains 7–9FnI (24). Two IGD tripeptides of domains 7FnI and 9FnI were shown to be essential through residue substitutions and reconstitution of partial motogenic activity in synthetic peptides (10, 24, 25); however, similar IGD tripeptides outside the GBD, on domains 3FnI and 5FnI, appear to have little effect (10, 23). Full-length adult FN does not affect cell migration in similar assays (10, 23); thus motogenic activity sites are presumed to be masked in the conformation adopted by soluble FN, although they could be exposed by molecular rearrangement.Open in a separate windowFIGURE 1.Motogenic activity of FN fragments. a, schematic representation of the FN domain structure (top) and enlargement of the FN N terminus (bottom). Type I domains are shown as pentagons; type II domains as hexagons; and type III domains as ovals. b, comparison of motogenic activity versus protein concentration of wild-type Fn70kDa and Fn100kDa fragments. Error bars are derived from duplicate experiments, and a gray band denotes migration activity of media without additives. c, similar comparisons for mutant Fn100kDa fragments. d, analytical size exclusion chromatography of large FN fragments. The trace of UV absorbance at 280 nm versus elution volume shown here indicates a larger hydrodynamic radius for Fn100kDa R222A compared with the wild type, consistent with our model (Fig. 6a).Here we show that a recombinant fragment, closely matching a truncated form of FN identified in zebrafish (26), as well as amphibians, birds, and mammals (27), does not stimulate cell migration. This fragment is similar to MSF but includes the first three FnIII domains (1–3FnIII), suggesting that these domains are responsible for a conformational transition that masks the activity sites in this construct and probably in full-length FN. To identify the mechanism behind this transition, we performed structural studies by solution NMR spectroscopy and identified a specific long range interaction between domains 4–5FnI and 3FnIII as essential for this masking effect. Interestingly, this interaction does not involve direct contacts with the GBD but possibly represses motogenic activity through chain compaction, evident in analytical size exclusion assays. Intramolecular interactions thus provide a mechanism by which conformational rearrangement induced, for example, by tension or splicing variation can result in cellular activity differences.  相似文献   

17.
Most individuals infected with human immunodeficiency virus type 1 (HIV-1) initially harbor macrophage-tropic, non-syncytium-inducing (M-tropic, NSI) viruses that may evolve into T-cell-tropic, syncytium-inducing viruses (T-tropic, SI) after several years. The reasons for the more efficient transmission of M-tropic, NSI viruses and the slow evolution of T-tropic, SI viruses remain unclear, although they may be linked to expression of appropriate chemokine coreceptors for virus entry. We have examined plasma viral RNA levels and the extent of CD4+ T-cell depletion in SCID mice reconstituted with human peripheral blood leukocytes following infection with M-tropic, dual-tropic, or T-tropic HIV-1 isolates. The cell tropism was found to determine the course of viremia, with M-tropic viruses producing sustained high viral RNA levels and sparing some CD4+ T cells, dual-tropic viruses producing a transient and lower viral RNA spike and extremely rapid depletion of CD4+ T cells, and T-tropic viruses causing similarly lower viral RNA levels and rapid-intermediate rates of CD4+ T-cell depletion. A single amino acid change in the V3 region of gp120 was sufficient to cause one isolate to switch from M-tropic to dual-tropic and acquire the ability to rapidly deplete all CD4+ T cells.The envelope gene of human immunodeficiency virus type 1 (HIV-1) determines the cell tropism of the virus (11, 32, 47, 62), the use of chemokine receptors as cofactors for viral entry (4, 17), and the ability of the virus to induce syncytia in infected cells (55, 60). Cell tropism is closely linked to but probably not exclusively determined by the ability of different HIV-1 envelopes to bind CD4 and the CC or the CXC chemokine receptors and initiate viral fusion with the target cell. Macrophage-tropic (M-tropic) viruses infect primary cultures of macrophages and CD4+ T cells and use CCR5 as the preferred coreceptor (2, 5, 15, 23, 26, 31). T-cell-tropic (T-tropic) viruses can infect primary cultures of CD4+ T cells and established T-cell lines, but not primary macrophages. T-tropic viruses use CXCR4 as a coreceptor for viral entry (27). Dual-tropic viruses have both of these properties and can use either CCR5 or CXCR4 (and infrequently other chemokine receptors [25]) for viral entry (24, 37, 57). M-tropic viruses are most frequently transmitted during primary infection of humans and persist throughout the duration of the infection (63). Many, but not all, infected individuals show an evolution of virus cell tropism from M-tropic to dual-tropic and finally to T-tropic with increasing time after infection (21, 38, 57). Increases in replicative capacity of viruses from patients with long-term infection have also been noted (22), and the switch to the syncytium-inducing (SI) phenotype in T-tropic or dual-tropic isolates is associated with more rapid disease progression (10, 20, 60). Primary infection with dual-tropic or T-tropic HIV, although infrequent, often leads to rapid disease progression (16, 51). The viral and host factors that determine the higher transmission rate of M-tropic HIV-1 and the slow evolution of dual- or T-tropic variants remain to be elucidated (4).These observations suggest that infection with T-tropic, SI virus isolates in animal model systems with SCID mice grafted with human lymphoid cells or tissue should lead to a rapid course of disease (1, 8, 4446). While some studies in SCID mice grafted with fetal thymus and liver are in agreement with this concept (33, 34), our previous studies with the human peripheral blood leukocyte-SCID (hu-PBL-SCID) mouse model have shown that infection with M-tropic isolates (e.g., SF162) causes more rapid CD4+ T-cell depletion than infection with T-tropic, SI isolates (e.g., SF33), despite similar proviral copy numbers, and that this property mapped to envelope (28, 41, 43). However, the dual-tropic 89.6 isolate (19) caused extremely rapid CD4+ T-cell depletion in infected hu-PBL-SCID mice that was associated with an early and transient increase in HIV-1 plasma viral RNA (29). The relationship between cell tropism of the virus isolate and the pattern of disease in hu-PBL-SCID mice is thus uncertain. We have extended these studies by determining the kinetics of HIV-1 RNA levels in serial plasma samples of hu-PBL-SCID mice infected with primary patient isolates or laboratory stocks that differ in cell tropism and SI properties. The results showed significant differences in the kinetics of HIV-1 replication and CD4+ T-cell depletion that are determined by the cell tropism of the virus isolate.  相似文献   

18.
Epidermal growth factor (EGF)-like modules are defined in part by six cysteines joined by disulfides in a 1–3, 2–4, and 5–6 pattern. Thrombospondin-1 (TSP-1) is a multimodular glycoprotein with three EGF-like modules, E1, E2, and E3, arranged in tandem. These modules likely propagate conformational changes between surrounding C-terminal and N-terminal elements of TSP-1 and interact with other extracellular molecules. E1, E2, and their homologs in other TSPs are unique among EGF-like modules in having two residues rather than one between Cys-4 and Cys-5. In addition, E2 has a calcium-binding site and an unusually long loop between Cys-5 and Cys-6. The structure of E1, E2, or E3 expressed alone changed little upon heating as monitored by far-UV CD, whereas more marked changes occurred in E12, E23, and E123 tandem constructs. The individual modules denatured in differential scanning calorimetry experiments only at >85 °C. E12, E23, or E123 tandem constructs, however, had a transition in the range of 44–70 °C. The temperature of the transition was higher when calcium was present and higher with E123 than with E12 or E23. Isothermal titration calorimetry demonstrated KD values of binding of calcium to E2, E12, E23, or E123 at 25 °C of 11.5, 2.9, 2.2, or 0.3 μm, respectively. Monoclonal antibodies HB8432 and C6.7, which recognize epitopes in E2, bound to E12, E23, or E123 with greater affinity than to E2 alone. These results indicate that interactions among the modules of E123 influence the tertiary structure and calcium binding of E2.Thrombospondins (TSPs)2 are multimodule, calcium-binding extracellular glycoproteins with various functions (1). TSP-1, which was the first TSP to be discovered and remains the best characterized, and TSP-2 are trimers. Each subunit is composed of an N-terminal module, oligomerization domain, von Willebrand factor type C module, three properdin or TSP type 1 modules, and the C-terminal signature domain that includes three EGF-like modules (E123), 13 aspartate-rich calcium-binding repeats of the wire module, and a lectin-like module (24). The five mammalian TSPs fall into two groups, trimeric (TSP-1 and TSP-2) and pentameric (TSP-3, TSP-4, and TSP-5) (1). All have a signature domain, with the major difference being the presence of four rather than three EGF-like modules in the signature domain of pentameric TSPs.EGF-like modules exist in more than 300 human extracellular proteins and play important roles in biological processes such as blood clotting and cell-cell signaling (57). The modules are 30–50 residues long and characterized by six cysteine residues that form three disulfide bonds in the order 1–3, 2–4, and 5–6 (Fig. 1) (6, 7). The backbone structure of the EGF-like modules consists of two submodules, referred to as the major (N-terminal) and minor (C-terminal) submodules (6, 8, 9).Open in a separate windowFIGURE 1.Model of the structure of E123. The model is built based on the crystal structure of EGF modules in the TSP-2 signature domain (Protein Data Bank code 1YO8) using SYBYL 7.0. E1 is shown in red, E2 in pink, and E3 in purple. The cysteines are colored yellow; the backbones of the residues between the fourth and fifth Cys are in blue; Glu-609 recognized by HB8432 and C6.7 is shown in green; and the long loop in E2 between the fifth and sixth Cys is hot pink. Ca2+ bound to the binding site on E2 near the interface between E1 and E2 is depicted as a red ball.The crystal structure of the three EGF-like modules of TSP-2 has been solved as part of the TSP-2 signature domain in 2 mm calcium (Ca2+) (Fig. 1) (4). All have the 1–3, 2–4, and 5–6 disulfide pattern. There is one Ca2+-binding site in the second EGF-like module (E2), located near the interface between the first and second EGF-like modules (E1 and E2) (Fig. 1). There is only one residue between the fourth and fifth cysteines in most EGF-like modules (6). However, E1 and E2 of TSP-1 and TSP-2 and three of the four EGF-like modules (E1, E2, and E2′) of pentameric TSPs have two residues between the fourth and fifth Cys. This difference is potentially important because the N-terminal major submodule of the repeat containing the 1–3 and 2–4 disulfides and the C-terminal submodule with the 5–6 disulfide have the potential to undergo hinge-like motions around the residues between the fourth and fifth Cys (6, 8, 9). Having two rather than one residue between these two Cys increases the potential flexibility. In addition, E2 modules in all five TSPs contain an unusually long loop of 23 residues between the fifth and sixth Cys (Fig. 1). In the TSP-2 signature domain structure, residues from the long loop interact with repeat 12N of the wire module (4). E3, which has one residue between the fourth and fifth Cys, interacts with the wire and the lectin-like module (3, 4). A common polymorphism (N700S) in wire repeat 1C of human TSP-1 influences the stability of the EGF-like modules (10). This finding suggests that the interactions between the EGF-like modules and more C-terminal elements of the signature domain allow conformational changes in the more C-terminal elements to be propagated N-terminally.The EGF-like modules (E123) of TSP-1 denature in differential scanning calorimetry (DSC) with a melting temperature of ∼68 °C in 2 mm Ca2+ (10), although most EGF-like modules are stable to heating (7). We have investigated this transition in detail to learn its origins and the influence of Ca2+. The results indicate interactions among the modules of E123 that enhance Ca2+ binding and influence the tertiary structure of E2.  相似文献   

19.
In this study, we report that the purified wild-type FANCI (Fanconi anemia complementation group I) protein directly binds to a variety of DNA substrates. The DNA binding domain roughly encompasses residues 200–1000, as suggested by the truncation study. When co-expressed in insect cells, a small fraction of FANCI forms a stable complex with FANCD2 (Fanconi anemia complementation group D2). Intriguingly, the purified FANCI-FANCD2 complex preferentially binds to the branched DNA structures when compared with either FANCI or FANCD2 alone. Co-immunoprecipitation with purified proteins indicates that FANCI interacts with FANCD2 through its C-terminal amino acid 1001–1328 fragment. Although the C terminus of FANCI is dispensable for direct DNA binding, it seems to be involved in the regulation of DNA binding activity. This notion is further enhanced by two C-terminal point mutations, R1285Q and D1301A, which showed differentiated DNA binding activity. We also demonstrate that FANCI forms discrete nuclear foci in HeLa cells in the absence or presence of exogenous DNA damage. The FANCI foci are colocalized perfectly with FANCD2 and partially with proliferating cell nuclear antigen irrespective of mitomycin C treatment. An increased number of FANCI foci form and become resistant to Triton X extraction in response to mitomycin C treatment. Our data suggest that the FANCI-FANCD2 complex may participate in repair of damaged replication forks through its preferential recognition of branched structures.Fanconi anemia (FA)3 is a genetic disorder characterized by chromosome instability, predisposition to cancer, hypersensitivity to DNA cross-linking agents, developmental abnormalities, and bone marrow failure (19). There are at least 13 distinct FA complementation groups, each of which is associated with an identified gene (2, 9, 10). Eight of them are components of the FA core complex (FANC A, B, C, E, F, G, L, and M) that is epistatic to the monoubiquitination of both FANCI and FANCD2, a key event to initiate interstrand cross-link (ICL) repair (2, 9, 11). Downstream of or parallel to the FANCI and FANCD2 monoubiquitination are the proteins involved in double strand break repair and breast cancer susceptibility (i.e. FANCD1/BRCA2, FANCJ/BRIP1, and FANCN/PALB2) (2, 9).FANCI is the most recently identified FA gene (1113). FANCI protein is believed to form a FANCI-FANCD2 (ID) complex with FANCD2, because they co-immunoprecipitate with each other from cell lysates and their stabilities are interdependent of each other (9, 11, 13). FANCI and FANCD2 are paralogs to each other, since they share sequence homology and co-evolve in the same species (11). Both FANCI and FANCD2 can be phosphorylated by ATR/ATM (ataxia telangiectasia and Rad3-related/ataxia telangiectasia-mutated) kinases under genotoxic stress (11, 14, 15). The phosphorylation of FANCI seems to function as a molecular switch to turn on the FA repair pathway (16). The monoubiquitination of FANCD2 at lysine 561 plays a critical role in cellular resistance to DNA cross-linking agents and is required for FANCD2 to form damage-induced foci with BRCA1, BRCA2, RAD51, FANCJ, FANCN, and γ-H2AX on chromatin during S phase of the cell cycle (1725). In response to DNA damage or replication stress, FANCI is also monoubiquitinated at lysine 523 and recruited to the DNA repair nuclear foci (11, 13). The monoubiquitination of both FANCI and FANCD2 depends on the FA core complex (11, 13, 26), and the ubiquitination of FANCI relies on the FANCD2 monoubiquitination (2, 11). In an in vitro minimally reconstituted system, FANCI enhances FANCD2 monoubiquitination and increases its specificity toward the in vivo ubiquitination site (27).FANCI is a leucine-rich peptide (14.8% of leucine residues) with limited sequence information to indicate which processes it might be involved in. Besides the monoubiquitination site Lys523 and the putative nuclear localization signals (Fig. 1A), FANCI contains both ARM (armadillo) repeats and a conserved C-terminal EDGE motif as FANCD2 does (11, 28). The EDGE sequence in FANCD2 is not required for monoubiquitination but is required for mitomycin C (MMC) sensitivity (28). The ARM repeats form α-α superhelix folds and are involved in mediating protein-protein interactions (11, 29). In addition, FANCI, at its N terminus, contains a leucine zipper domain (aa 130–151) that could be involved in mediating protein-protein or protein-DNA interactions (Fig. 1A) (3033). FANCD2, the paralog of FANCI, was reported to bind to double strand DNA ends and Holliday junctions (34).Open in a separate windowFIGURE 1.Purified human FANCI binds to DNA promiscuously. A, schematic diagram of predicted FANCI motifs and mutagenesis strategy to define the DNA binding domain. The ranges of numbers indicate how FANCI was truncated (e.g. 801–1328 represents FANCI-(801–1328)). NLS, predicted nuclear localization signal (aa 779–795 and 1323–1328); K523, lysine 523, the monoubiquitination site. The leucine zipper (orange bars, aa 130–151), ARM repeats (green bars), and EDGE motif (blue bars) are indicated. Red bars with a slash indicate the point mutations shown on the left. B, SDS-PAGE of the purified proteins stained with Coomassie Brilliant Blue R-250. R1285Q and D1301A are two point mutants of FANCI. All FANCI variants are tagged by hexahistidine. FANCD2 is in its native form. Protein markers in kilodaltons are indicated. C, titration of WT-FANCI for the DNA binding activity. Diagrams of the DNA substrates are shown at the top of each set of reactions. *, 32P-labeled 5′-end. HJ, Holliday junction. Concentrations of FANCI were 0, 20, 40, 60, and 80 nm (ascending triangles). The substrate concentration was 1 nm. Protein-DNA complex is indicated by an arrow. D, supershift assay. 1 nm of ssDNA was incubated with PBS (lane 1), 80 nm FANCI alone (lane 2), and 80 nm FANCI preincubated with a specific FANCI antibody (lane 3) in the condition described under “Experimental Procedures.”In order to delineate the function of FANCI protein, we purified the recombinant FANCI from the baculovirus expression system. In this study, we report the DNA binding activity of FANCI. Unlike FANCD2, FANCI binds to different DNA structures, including single-stranded DNA (ssDNA), double-stranded DNA (dsDNA), 5′-tailed, 3′-tailed, splayed arm, 5′-flap, 3′-flap, static fork, and Holliday junction with preference toward branched structures in the presence of FANCD2. Our data suggest that the dynamic DNA binding activity of FANCI and the preferential recognition of branched structures by the ID complex are likely to be the mechanisms to initiate downstream repair events.  相似文献   

20.
Bisretinoid adducts accumulate as lipofuscin in retinal pigment epithelial (RPE) cells of the eye and are implicated in the pathology of inherited and age-related macular degeneration. Characterization of the bisretinoids A2E and the all-trans-retinal dimer series has shown that these pigments form from reactions in photoreceptor cell outer segments that involve all-trans-retinal, the product of photoisomerization of the visual chromophore 11-cis-retinal. Here we have identified two related but previously unknown RPE lipofuscin compounds. By high performance liquid chromatography-elec tro spray ionization-tandem mass spectrometry, we determined that the first of these compounds is a phosphatidyl-dihydropyridine bisretinoid; to indicate this structure and its formation from two vitamin A-aldehyde (A2), we will refer to it as A2-dihydropyridine-phosphatidyleth a nol amine (A2-DHP-PE). The second pigment, A2-dihydropyridine-eth a nol amine, forms from phosphate hydrolysis of A2-DHP-PE. The structure of A2-DHP-PE was corroborated by Fourier transform infrared spectroscopy, and density functional theory confirmed the presence of a dihydropyridine ring. This lipofuscin pigment is a fluorescent compound with absorbance maxima at ∼490 and 330 nm, and it was identified in human, mouse, and bovine eyes. We found that A2-DHP-PE forms in reaction mixtures of all-trans-retinal and phosphatidyleth a nol amine, and in mouse eyecups we observed an age-related accumulation. As compared with wild-type mice, A2-DHP-PE is more abundant in mice with a null mutation in Abca4 (ATP-binding cassette transporter 4), the gene causative for recessive Stargardt macular degeneration. Efforts to clarify the composition of RPE lipofuscin are important because these compounds are targets of gene-based and drug therapies that aim to alleviate ABCA4-related retinal disease.Throughout the life of an individual, retinal pigment epithelial (RPE)2 cells of the eye accumulate bisretinoid adducts that comprise the lipofuscin of these cells (13). The compounds form as a byproduct of light-mediated isomerization of the visual chromophore 11-cis-retinal. Accordingly, conditions that reduce the production of all-trans-retinal (atRAL) from 11-cis-retinal photoisomerization, such as reduced serum vitamin A (46), variants, or mutations in the visual cycle protein RPE65 (79) and inhibitors of RPE65 and 11-cis retinol dehydrogenase (1013), substantially reduce the formation of this material.Up to the present time, at least 17 constituents of RPE lipofuscin have been identified chromatographically and characterized structurally; added to these are biosynthetic intermediates such as N-retinylidene-phosphatidylethanolamine (NRPE), A2PE and dihydropyridinium-A2PE (see Fig. 1, A and B) (1419). The first RPE lipofuscin constituent to be described was A2E (see Fig. 1A, inset). The pyridinium bisretinoid (1416, 20, 21) structure of A2E (C42H58NO; molecular weight, 592) was confirmed by extensive nuclear magnetic resonance studies (14) and by total synthesis (22). A2E formation begins in photoreceptor outer segments when atRAL, instead of being reduced to all-trans-retinol, reacts with phosphatidylethanolamine (PE) in a 2:1 ratio. Although the double bonds along the side arms of A2E are all in the trans (E) position, Z-isomers of A2E have double bonds at the C-13/14 (isoA2E), C-9/9′-10/10′, and C-11/11′-12/12′ positions, and all are detectable in human and mouse RPE (16). These pigments exhibit absorbances in both the UV and visible regions of the spectrum (A2E: λmax, 439 and 338 nm; iso-A2E: λmax, 428 and 337 nm).Open in a separate windowFIGURE 1.Some bisretinoid compounds associated with RPE lipofuscin formation. Structures, molecular weight (Mw), UV-visible absorbance (nm), and electronic transition assignments (↔). Phosphate hydrolysis (dashed lines) of A2PE (A) and A2-DHP-PE (C) generates A2E and A2-DHP-E (insets in A and C). A2PE (A), dihydropyridinium-A2PE (B), A2-DHP-PE (C). The molecular weights are based on dipalmitic acid as the fatty acid constituent.Another bisretinoid compound of RPE lipofuscin also absorbs in the short wavelength region of the visible spectrum (17, 18, 23). This pigment, all-trans-retinal dimer (atRAL dimer; λmax, 432 and 290 nm) forms from the condensation of two atRAL and is present in RPE lipofuscin as Schiff base conjugates with either PE or ethanolamine (atRAL dimer-PE and atRAL dimer-E, respectively) or as unconjugated atRAL dimer. The pigments atRAL dimer-PE and atRAL dimer-E absorb in the visible range at about 510 nm, a “red” shift relative to atRAL dimer that is attributable to protonation of the Schiff base linkage. Although A2E is a pyridinium salt containing a quaternary amine nitrogen that does not deprotonate or reprotonate (24), the protonation state of the Schiff base linkage in atRAL dimer-PE and atRAL dimer-E is pH-dependent (18).Other known constituents of RPE lipofuscin are generated by photooxidation. By mass spectrometry, the photooxidation products of A2E and atRAL dimer present as a series of peaks differing by increments of mass 16 beginning with M+ 592 (A2E) or M+ 552 (atRAL dimer) (18, 25). The moieties generated by the addition of oxygens at CC bonds of these bisretinoid compounds include endoperoxides, furanoid oxides, and epoxides (2527). These oxidized products are more polar than the parent compound, and mono- and bis-oxidized forms of A2E and atRAL dimer have been detected in RPE from human eyes and in eyecups from mice with null mutations in Abca4/ (18, 25), the gene responsible for recessive Stargardt macular degeneration. It is also notable that unconjugated atRAL dimer is a more efficient generator of singlet oxygen than is A2E and is also a more efficient quencher of singlet oxygen (18).Insight into the composition of RPE lipofuscin and the biosynthetic pathways by which these compounds form aids in an understanding of retinal diseases characterized by lipofuscin overload, particularly those associated with mutations in ABCA4 (ATP-binding cassette transporter 4) of photoreceptor cells (13). We report that a previously unrecognized bisretinoid molecule absorbing with maxima at 490 and 331 nm is detected at elevated levels in Abca4/ mice, a model of recessive Stargardt macular degeneration. This compound is also present in human RPE. By high performance liquid chromatography-electrospray ionization-tandem mass spectrometry (HPLC-ESI-MS/MS), with corroboration by Fourier transform infrared spectroscopy (FTIR), we determined that this molecule is a bisretinoid presenting with a noncharged dihydropyridine core (Fig. 1C). We propose a biosynthetic pathway by which this pigment may form and demonstrate that enzymatic removal of the phosphatidic acid portion of the molecule generates a second novel component of RPE lipofuscin.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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