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.
Light converts rhodopsin, the prototypical G protein-coupled receptor, into a form capable of activating G proteins. Recent work has shown that the light-activated state of different rhodopsins can possess different molecular properties, especially different abilities to activate G protein. For example, bovine rhodopsin is ∼20-fold more effective at activating G protein than parapinopsin, a non-visual rhodopsin, although these rhodopsins share relatively high sequence similarity. Here we have investigated possible structural aspects that might underlie this difference. Using a site-directed fluorescence labeling approach, we attached the fluorescent probe bimane to cysteine residues introduced in the cytoplasmic ends of transmembrane helices V and VI in both rhodopsins. The fluorescence spectra of these probes as well as their accessibility to aqueous quenching agents changed dramatically upon photoactivation in bovine rhodopsin but only moderately so in parapinopsin. We also compared the relative movement of helices V and VI upon photoactivation of both rhodopsins by introducing a bimane label and the bimane-quenching residue tryptophan into helices VI and V, respectively. Both receptors showed movement in this region upon activation, although the movement appears much greater in bovine rhodopsin than in parapinopsin. Together, these data suggest that a larger conformational change in helices V and VI of bovine rhodopsin explains why it has greater G protein activation ability than other rhodopsins. The different amplitude of the helix movement may also be responsible for functional diversity of G protein-coupled receptors.Rhodopsin, the photosensitive G protein-coupled receptor (GPCR),3 is responsible for transmitting a light signal into an intracellular signaling cascade through activation of G protein in visual and non-visual photoreceptor cells. Rhodopsin consists of a protein moiety (opsin, comprising seven transmembrane α-helical segments) combined with a chromophore (11-cis retinal) that acts as the light-sensitive ligand. Photoisomerization of the 11-cis retinal to the all-trans form induces structural changes in the protein moiety that then enable it to couple with and activate the G protein.The crystal structure of inactive bovine rhodopsin has been extensively investigated (13). Recently, a crystal structure of inactive invertebrate squid rhodopsin was also solved (4), and crystal structures of the inactive form of β-adrenergic receptors and A2 adenosine receptor have been reported (57). Remarkably, all of these crystal structures exhibit a very similar arrangement for the seven transmembrane helices (4, 8). Together, these facts suggest that the architecture for the inactive form is conserved among rhodopsin-like GPCRs.The structural features of an activated GPCR are much less defined. Thus, a variety of biochemical and biophysical methods, including cross-linking methods (9, 10) and site-directed spin and fluorescence labeling methods (1013), have been employed to identify the dynamic and structural changes involved in forming the activated state. The data from these studies consistently suggest that some kind of movement of helix VI is involved in the formation of the active state of the rhodopsins. In particular, the cytoplasmic end of helix VI has been proposed to rotate and/or tilt toward helix V (1013). Remarkably, the recent crystal structures of bovine opsin are consistent with the widely accepted helix motion model. Both the structures of opsin (the ligand-free form of rhodopsin that has partial G protein activation ability) and a complex of opsin with a peptide derived from the G protein C terminus show a movement of helix VI toward helix V, compared with the dark state rhodopsin structure (14, 15). Studies of β-adrenergic and muscarinic receptors also show that agonist binding promotes movement of helix VI toward helix V in these receptors (16, 17). Because the region between the cytoplasmic ends of helices V and VI in various GPCRs is a main site of interaction with G proteins (18), it is possible that movement of helices V and VI leads to formation of a conformation capable of interacting with G protein (19).Together, these studies imply that the active state conformation of GPCRs may be similar. However, a detailed comparison of the active-state conformation for two different GPCRs has never been precisely undertaken in the same laboratory using the same methods.In this context we have been investigating rhodopsins with different functional properties to determine whether their active states have different conformations. Our goal was to determine whether any functional or structural differences in the active states of these GPCRs could be detected under the exact same experimental conditions.Previously, we have found that several rhodopsins, such as an invertebrate rhodopsin and a vertebrate non-visual rhodopsin parapinopsin (20, 21), can be activated not only by light but also by exogenous all-trans retinal acting as a full agonist (22). This is in contrast to vertebrate visual rhodopsins, including bovine rhodopsin, which cannot fully form the active state by direct binding of all-trans retinal (23), although all-trans retinal can fully activate some rhodopsin mutants (24). Other invertebrate rhodopsin (25) and the circadian photoreceptor melanopsin (26) can also bind all-trans retinal directly.Interestingly, the active form of the all-trans retinal-activated rhodopsins exhibit some striking differences in their spectroscopic and biochemical properties compared with vertebrate visual rhodopsins (27). In particular, the efficiency of bovine rhodopsin for activating G protein is ∼20∼50-fold higher than that of parapinopsin and invertebrate rhodopsin. This difference could be related to the difference in position of a specific amino acid residue counterion that is essential for rhodopsin to absorb visible light, namely one at position 113 or 181 (28).4 Further biochemical analyses using chimeric mutants combining rhodopsins with lower and higher G protein activation abilities suggested that the difference in G protein activation ability was because of a structural difference in transmembrane helices in the active states but not because of difference in amino acid sequence of G protein interaction site (29) (Fig. 1, A–C). In addition, the active states of parapinopsin and the invertebrate rhodopsin are thermally stable and can be reconverted to the inactive state by subsequent light absorption, showing photo-regenerable or bistable nature (21, 28), unlike the active state of bovine rhodopsin, which is thermally unstable and cannot revert to the inactive state by subsequent light absorption (30).Open in a separate windowFIGURE 1.Molecular properties and sites of fluorescent probe attachment for bovine rhodopsin and parapinopsin. A, sequence alignment of bovine rhodopsin and parapinopsin. Amino acid residues to which cysteine and fluorescence label were introduced are marked with red. The amino acid residues identical and similar between bovine rhodopsin and parapinopsin are shown with white characters with black and gray background, respectively. Bovine rhodopsin and parapinopsin show 41% sequence identity and 61% similarity. In this paper the residue number of parapinopsin is described by using the bovine rhodopsin numbering system. B and C, comparison of G protein activation ability of rhodopsin and parapinopsin wild type (WT) proteins and loop-replaced mutants. In these mutants the second and/or third cytoplasmic loop was swapped between the two receptors. ParaL2 and ParaL3 indicate mutants of bovine rhodopsin in which second and third loops were replaced with the corresponding loop of parapinopsin, respectively. RhoL2 and RhoL3 indicate mutants of parapinopsin in which the second and third loops were replaced with the corresponding loops of bovine rhodopsin, respectively. ParaL2L3 and RhoL2L3 are mutants of bovine rhodopsin and parapinopsin in which both the second and third loops were swapped, respectively. See Terakita et al. (29) for more details. Data are presented as the means ± S.E. of three separate experiments except for mutants RhoL3, RhoL2L3, and ParaL2L3 (n = 2). D, model of bovine rhodopsin. Amino acid residues which were mutated to cysteine to enable attachment of the fluorescent probe bimane or mutated to tryptophan are indicated. Positions 226, 227, 244, 250, and 251 in the crystal structure of the dark state of bovine rhodopsin (PDB code 1GZM) are shown. E, reaction of the mBBr label with a sulfhydryl group. The mutants labeled with mBBr are named by the number of the residue and the suffix B1. F, reaction of the PDT-bimane with a sulfhydryl group. The mutants labeled with PDT-bimane are named by the number of the residue and the suffix B2. The disulfide linkage between the label and protein can be cleaved using Tris(2-carboxyethyl)phosphine (32).In this study we used site-directed fluorescence labeling (13, 31) to compare the structural features of active states of bovine rhodopsin with lamprey parapinopsin, a UV-sensitive non-visual pigment in the pineal organs (21). Parapinopsin shows relatively high sequence similarity (∼60%) to bovine rhodopsin, yet it has a greatly reduced ability to activate G protein (see Fig. 1, A–C) (21, 28). Using established protocols, we introduced cysteine residues into the cytoplasmic ends of helices V and VI, the region proposed to rearrange upon activation in GPCRs (11, 12, 14, 18). We then site-specifically labeled these cysteines with the small, non-polar fluorescent probe, bimane, and used the spectral properties of these bimane probes to act as reporter groups for environmental changes around their site of attachment upon formation of the photoactivated state for both rhodopsins.In addition, we measured changes in the relative proximity of the cytoplasmic ends of helix VI to helix V in both rhodopsin and parapinopsin using the tryptophan-induced-quenching of bimane (TrIQ-bimane) fluorescence method (31, 32). TrIQ-bimane measures the efficiency of intramolecular fluorescence quenching of bimane caused by tryptophan (Trp), which occurs in a distance-dependent manner. The goal of this study was to determine whether the helices in both receptors moved in the same way during formation of the active state. Our results show that whereas movement of helix VI relative to helix V occurs during formation of the active state for both parapinopsin and bovine rhodopsin, the “amplitude” of the movement is markedly different between the two rhodopsins.  相似文献   

2.
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   

3.
G protein-coupled receptor (GPCR) kinases (GRKs) phosphorylate activated GPCRs and initiate their desensitization. Many prior studies suggest that activated GPCRs dock to an allosteric site on the GRKs and thereby stimulate kinase activity. The extreme N-terminal region of GRKs is clearly involved in this process, but its role is not understood. Using our recent structure of bovine GRK1 as a guide, we generated mutants of solvent-exposed residues in the GRK1 kinase domain that are conserved among GRKs but not in the extended protein kinase A, G, and C family and evaluated their catalytic activity. Mutation of select residues in strands β1 and β3 of the kinase small lobe, αD of the kinase large lobe, and the protein kinase A, G, and C kinase C-tail greatly impaired receptor phosphorylation. The most dramatic effect was observed for mutation of an invariant arginine on the β1-strand (∼1000-fold decrease in kcat/Km). These residues form a continuous surface that is uniquely available in GRKs for protein-protein interactions. Surprisingly, these mutants, as well as a 19-amino acid N-terminal truncation of GRK1, also show decreased catalytic efficiency for peptide substrates, although to a lesser extent than for receptor phosphorylation. Our data suggest that the N-terminal region and the newly identified surface interact and stabilize the closed, active conformation of the kinase domain. Receptor binding is proposed to promote this interaction, thereby enhancing GRK activity.G protein-coupled receptor kinases (GRKs)2 are members of the protein kinase A (PKA), G, and C (AGC) family that phosphorylate Ser/Thr residues in the cytoplasmic loops and C termini of activated G protein-coupled receptors (GPCRs) (1). Receptor phosphorylation facilitates the binding of arrestin, which uncouples heterotrimeric G proteins, promotes receptor internalization, and activates arrestin-dependent signaling pathways (2, 3). Although playing a beneficial role in receptor desensitization, GRKs are implicated in a range of human diseases, including retinal degeneration, hypertension, heart failure, rheumatoid arthritis, opiate addiction, and various cancers (2, 4). Seven GRKs have been identified in mammals. They can be divided into the following three subgroups based on their sequence homology: GRK1 (GRK1 and GRK7), GRK2 (GRK2 and GRK3), and GRK4 (GRK4, GRK5, and GRK6). The primary structures of the three GRK subgroups are similar, consisting of tandem regulator of G protein signaling homology (RH) and kinase domains. Less conserved sequences involved in phospholipid membrane attachment are found at their C termini (Fig. 1A).Open in a separate windowFIGURE 1.GRK surface residues potentially important for GPCR phosphorylation. A, domain architecture of bGRK1. B, sequence alignment of regions from GRKs that were targeted in this study with other AGC kinases. Colored boxes map these regions back to the domain structure shown in A. Regions of the core kinase domain that contain residues conserved in the GRK subfamily, but not in the extended AGC kinase family, are highlighted in brown. Conserved residues of the AGC kinase C-tail are highlighted in green. Positions investigated in this study are indicated with asterisks. Only one PDB accession code for each kinase of known structure is shown in parentheses. Residue numbers correspond to those of bGRK1. The PXXP, turn, and hydrophobic (HF) motifs (highlighted in gray) are characteristic features found in most AGC kinase C-tails (22). Yeast, Saccharomyces cerevisiae; M.tb, Mycobacterium tuberculosis. C, ribbon diagram of the bGRK1535-H6·ATP complex. The model is a hybrid that contains all the ordered elements from the two unique chains resolved in the crystal structure (PDB accession number 3C4W), such that the observed N terminus and a nearly complete AST region of the kinase C-tail are displayed in a single model. The RH domain is colored gray, and the β-strands and α-helices of the core kinase domain are dark and light brown, respectively. The hinge region joining the kinase small and large lobes (between β5 and αD) is colored yellow. The N-terminal region and the AGC kinase C-tail are shown in green and the AST loop in cyan. The ATP molecule bound in the active site is shown as a stick model, and the two associated Mg2+ ions are colored black. D, conservation scores of GRKs mapped onto the surface of bGRK1. The area boxed in C is shown. The conservation scores were calculated by comparing the sequence conservation of residues from the core kinase domain of GRKs with those of the entire AGC family (see text). Residues are colored using a gradient, from magenta (more conserved in GRKs than the AGC kinases) to white (as conserved in GRKs as in AGC kinases) and to yellow (more variable in GRKs than in AGC kinases). The RH domain, which was not included in this calculation, is colored gray. Highest conservation among GRKs cluster at two sites, “site 1” and “site 2.” Site 1 corresponds to a region of the small lobe left exposed by the shorter AST loop found in GRKs relative to other AGC kinases. The AST region of protein kinase B (PDB accession number 1O6K) is superimposed for comparison (blue ribbon).All eukaryotic protein kinases, including GRKs, contain a core catalytic domain of ∼250 amino acids that can be divided into two subdomains, called the small (or N) and large (or C) lobes. The small lobe consists of a five-stranded β-sheet (β1–5) and a conserved helix, αC, whereas the large lobe is mostly α-helical. The active site is located at their interface, with the nucleotide-binding pocket formed primarily by the small lobe and the phosphoacceptor-binding site primarily by the large lobe. In their active conformation, kinases position the hydroxyl group of the phosphoacceptor substrate in the proper orientation with respect to the γ-phosphate of ATP via a network of interactions formed by conserved structural elements from both lobes. Control of this network often underlies the molecular basis for allosteric regulation of protein kinase activity (59).In GRKs, this allosteric regulation appears to be mediated by interactions with activated GPCRs. Steady-state kinetics indicate that the Km values of receptor substrates are in the micromolar range, whereas those of peptide substrates, even those derived from receptors, are in the millimolar range (1013). Moreover, the catalytic efficiency for peptide phosphorylation by GRKs is much lower than that for receptor phosphorylation, and it can be enhanced in the presence of activated receptors (11, 12, 14). Thus, in addition to the peptide phosphoacceptor-binding site of the large lobe, an additional allosteric receptor-docking site appears to be required to promote catalytic activity in GRKs.The molecular basis for how GRKs interact with activated GPCRs is poorly understood. In vitro, GRKs show little specificity among GPCRs, requiring only that the receptor be in an activated conformation. For example, although GRK1 is the predominant kinase expressed in rod outer segments, GRK1, GRK2, and GRK5 all phosphorylate bovine rhodopsin in a light-dependent manner with comparable catalytic efficiencies (1517). Therefore, it seems likely that GRKs have a common molecular mechanism for the recognition of activated GPCRs. The region of GRKs most strongly linked to efficient receptor phosphorylation is the highly conserved N-terminal region, which is unique to the GRK subfamily and predicted to form an α-helix (Fig. 1B). Deletion of this region in GRK1, -2, or -5 abolishes receptor phosphorylation (1820). Additionally, the binding of antibodies (18) or of recoverin (21) to the GRK1 N-terminal region inhibits receptor phosphorylation. In GRK5, it has also been suggested that the N terminus plays a role in phospholipid interactions (20). Another region that is likely involved in the allosteric regulation of GRKs is the AGC kinase C-terminal tail (C-tail), which is an extension of the kinase core domain and often plays a regulatory role in AGC kinases (2224) (Fig. 1, B and C). The central segment of the C-tail, termed the active site tether (AST), contributes residues to the active site and is only well ordered in kinase domain structures that are in conformations resembling the active state.To date, crystal structures representing each GRK subgroup have been reported, i.e. bovine GRK1 (bGRK1), bovine GRK2 (bGRK2), and human GRK6 (hGRK6) (2529). Although most of these structures were co-crystallized in the presence of ATP or nucleotide analogs, none adopted the closed, active conformation exhibited by nucleotide-bound PKA (30), and the AST region of their AGC kinase C-tails were either partially or totally disordered. Similarly, the N-terminal region important for receptor phosphorylation was only observed in one crystal structure, namely that of one chain of the bGRK1·ATP complex. Thus, the regions believed to be most important for receptor interaction were largely disordered in these structures, leaving the molecular basis for how GPCRs interact with GRKs unclear. Because the kinase domains in these structures appear to be otherwise competent for catalysis, it is expected that activated GPCRs bind in a manner that promotes full kinase domain closure. Interactions with negatively charged lipids in the cell membrane are also expected to play a role in this transition (20, 31, 32).In this study, we used recently determined structures of bGRK1 as a template to identify surface residues of the kinase domain that are conserved in GRKs but not in the extended AGC family. Biochemical characterization of site-directed mutants of these residues in bGRK1 identified a continuous surface on the small lobe and the AGC kinase C-tail that is critical for GRK activation by GPCRs. The residue whose mutation showed the largest effect on receptor phosphorylation is nearly as important as the N-terminal region, and the analogous residue is also critical in the other two GRK subgroups, represented by bGRK2 and hGRK6. Comparison with other AGC kinases reveals that this surface is uniquely available for protein-protein interactions in the GRK subfamily. A model for activation that involves cooperative interactions between the N-terminal region and the kinase domain is presented.  相似文献   

4.
5.
6.
Retinoic acid (RA) causes HL-60 human myeloblastic leukemia cell myeloid differentiation that is dependent on MAPK signaling. The process is propelled by c-Cbl, which binds the CD38 receptor as part of a signaling complex generating MAPK signaling. Here we report that the capability of c-Cbl to do this is lost in the G306E tyrosine kinase-binding domain mutant. Unlike wild-type (WT) c-Cbl, the G306E mutant c-Cbl fails to propel RA-induced differentiation, and disrupts the normal association with CD38. The G306E mutant does, like WT c-Cbl, co-immunoprecipitate with Vav, Slp-76, and p38. But unlike WT c-Cbl, does not cause MAPK signaling. In contrast, the C381A Ring finger domain mutant functions like WT c-Cbl. It binds CD38 and is part of the same apparent c-Cbl/Slp-76/Vav/p38 signaling complex. The C381A mutant causes MAPK signaling and propels RA-induced differentiation. In addition to HL-60 cells and their WT or mutant c-Cbl stable transfectants, the c-Cbl/Vav/Slp-76 complex is also found in NB4 cells where c-Cbl was previously also found to bind CD38. The data are consistent with a model in which the G306E mutant c-Cbl forms a signaling complex that includes Slp-76, Vav, and p38; but does not drive MAPK signaling because it fails to bind the CD38 receptor. Without the G306E mutation the c-Cbl unites CD38 with the signaling complex and delivers a MAPK signal that drives RA-induced differentiation. The results demonstrate the importance of the Gly306 residue in the ability of c-Cbl to propel RA-induced differentiation.Retinoic acid (RA),2 a form of vitamin A, can cause activation of MAPK signaling leading to induced cell differentiation and G0 cell cycle arrest (12). Because of this, RA has been used therapeutically for the chemoprevention and treatment of cancer (3), notably acute promyelocytic leukemia, making its mechanism of action of significant interest. The HL-60 human myeloblastic leukemia cell line serves as a model for studying differentiation induction therapy and the mechanism of RA action (46). The HL-60 cell line was established from peripheral blood leukocytes of a patient originally diagnosed with acute promyelocytic leukemia (4), which was retrospectively re-evaluated to be acute myeloblastic leukemia (7). HL-60 cells undergo G0 cell cycle arrest and myeloid differentiation in response to RA or monocytic differentiation in response to 1,25-dihydroxyvitamin D3. Induced differentiation depends on hyperactive MAPK signaling (1, 8). c-Cbl contributes propulsion to this process (9, 10). It is part of a signaling complex connected to the CD38 receptor that activates the RAF/MEK/ERK MAPK signaling axis to drive differentiation and G0 cell cycle arrest (9).The 120-kDa product of the c-Cbl protooncogene is a prominent component of tyrosine kinase signaling cascades downstream of activated cell surface receptors, including the epidermal growth factor receptor, the B cell receptor, Fc receptor, and cytokine receptors (11, 12). Cbl proteins have a highly conserved N-terminal domain termed the tyrosine kinase-binding (TKB) domain, which binds to phosphotyrosines on activated receptor-tyrosine kinases and other signaling proteins (13, 14), a short linker region, and a Ring finger domain that binds to ubiquitin-conjugating enzymes (15, 16). The Ring finger domain of c-Cbl presents the most conserved region among Cbl family proteins, and it is implicated as an important element in the function of Cbl proteins (1719). The TKB domain contains a four-helix bundle, EF-hand calcium-binding domain, and a variant SH2 domain that binds to phosphotyrosine residues (13, 20). Several features of the Cbl/Zap-70 complex (13) suggest that the four-helix bundle, EF-hand, and SH2 domains together form an interactive structure that is crucial for phosphoprotein recognition. Studies by Thien et al. (21) indicated that the TKB mutation, G306E, fails to bind phosphotyrosine residues and is a loss of function mutation in the c-Cbl TKB domain.Cbl interacts with a number of intracellular signaling molecules including various receptors, adaptors, cytoskeletal proteins, ubiquitin, and structural proteins via its various domains (22, 23). One of the receptors is CD38. The human cell surface antigen CD38 can be found in lipid rafts and causes RAF and ERK activation (24, 25). Kontani et al. (26) suggested that CD38 causes tyrosine phosphorylation of target molecules, including the Cbl adaptor. Vav is another signal regulator that plays a role in several cellular functions, including cell proliferation and maturation, cytoskeletal reorganization, regulation of gene expression, and apoptosis (27, 28). Bertagnolo et al. (29) reported that Vav is a potential target of Syk and that the SH3-SH2-SH3 fragment of Vav can associate with Cbl and Slp76. Slp76 was first identified as a substrate of the protein-tyrosine kinases that are engaged by activated T-cell receptors. It lacks intrinsic enzymatic activity but contains multiple protein-binding domains. Slp76 can interact with Cbl (30, 31). p38 is a member of the MAPK signaling family and is associated with cellular stress responses and apoptosis (32, 33). Studies by Frey and colleagues (34) have shown that epidermal growth factor receptor ubiquitinylation and Cbl activation are dependent on p38 activity. It is not known if an interaction between c-Cbl and p38 is induced by RA to support RA-induced differentiation and G0 arrest.Although the biological importance of the adaptor function of the Cbl protein and E3 ligase activity is now apparent, the precise role of specific residues of Cbl in cell differentiation and cell growth, as well as in the interaction of c-Cbl with other cellular proteins still remains to be well elucidated. We have targeted two residues of c-Cbl located in the Ring finger domain and the TKB domain (see Fig. 1A). This study evaluates if these mutations affect the ability of c-Cbl to drive RA-induced differentiation by altering its interactions with partners. Binding to the CD38 receptor was determined. The interactions of c-Cbl and Vav, c-Cbl and Slp76, as well as c-Cbl and p38 were also investigated. The interaction studies were carried out in HL-60 cells as well as NB4 cells, which was derived from an APL patient and bears the t(15,17) promyelocytic leukemia-retinoic acid receptor α translocation (35, 36). The data are consistent with a model in which Cbl is in a complex with Vav, Slp76, and p38 and this subassembly MAPK signaling complex is recruited to CD38 by an interaction dependent on Cbl Gly306 to propel RA-induced differentiation through activation of MAPK signaling.Open in a separate windowFIGURE 1.c-Cbl TKB mutant failed to enhance CD11b expression and RA-induced cell differentiation compared with WT c-Cbl stably transfected cells (c-Cbl+). A, diagrammatic representation of the major domains of c-Cbl and mutations within the TKB and Ring finger domains used in this study. The c-Cbl protein consists of the TKB, the Ring finger domain (RING), proline-rich region, a PX(P/A)XXR motif (PR), and a ubiquitin-associated domain (UBA) at its C terminus. The TKB domain containing a four-helix bundle (4H), EF-hand calcium binding, and a variant SH2 domain is separated from the RING domain by a short linker (L) region. B, in stably transfected cells WT c-Cbl (c-Cbl+) or two c-Cbl mutants (G306E and C381A) were strongly overexpressed compared with c-Cbl in vector control cells. G306E and C381A were transfected into HL-60 cells. Western blots of c-Cbl (upper) and GAPDH (lower) expression in vector control, c-Cbl+, and G306E and C381C stable transfectant cells were untreated control (C) or RA treated (RA, 48 h). C, compared with vector controls, c-Cbl+ and C381A transfectants showed significantly enhanced expression of CD11b after RA treatment; however, the G306E mutant did not. Vector controls, WT c-Cbl, G306E, and C381A stable transfectants were treated with RA for the indicated times, and stained with allophycocyanin-conjugated anti-CD11b antibody. Bars are means ± S.E. of 3 repeats. D, RA-induced expression of the functional differentiation marker inducible oxidative metabolism was accelerated by WT c-Cbl and the C381A mutant, but not the G306E mutant. Vector control, c-Cbl+, G306E, and C381A stable transfectants were treated with RA for the indicated times, and the percentage of cells capable of inducible oxidation metabolism detected by 2′,7′-dichlorohydrofluorescein diacetate (DCF) fluorescence was analyzed by flow cytometry. E, representative DCF fluorescence histograms of untreated and (48 h) RA-treated vector control, WT c-Cbl, G306E, and C381A stable transfectants. The different letters indicate different time points. Asterisk indicates that CD11b or DCF expression levels from WT c-Cbl and C381A transfectants were significantly (p ≤ 0.05) different from vector controls and G306E transfectants. # denotes that C381A transfectants were significantly (p ≤ 0.05) different from c-Cbl+ transfectants.  相似文献   

7.
8.
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   

9.
The FAD-dependent choline oxidase has a flavin cofactor covalently attached to the protein via histidine 99 through an 8α-N(3)-histidyl linkage. The enzyme catalyzes the four-electron oxidation of choline to glycine betaine, forming betaine aldehyde as an enzyme-bound intermediate. The variant form of choline oxidase in which the histidine residue has been replaced with asparagine was used to investigate the contribution of the 8α-N(3)-histidyl linkage of FAD to the protein toward the reaction catalyzed by the enzyme. Decreases of 10-fold and 30-fold in the kcat/Km and kcat values were observed as compared with wild-type choline oxidase at pH 10 and 25 °C, with no significant effect on kcat/KO using choline as substrate. Both the kcat/Km and kcat values increased with increasing pH to limiting values at high pH consistent with the participation of an unprotonated group in the reductive half-reaction and the overall turnover of the enzyme. The pH independence of both D(kcat/Km) and Dkcat, with average values of 9.2 ± 3.3 and 7.4 ± 0.5, respectively, is consistent with absence of external forward and reverse commitments to catalysis, and the chemical step of CH bond cleavage being rate-limiting for both the reductive half-reaction and the overall enzyme turnover. The temperature dependence of the Dkred values suggests disruption of the preorganization in the asparagine variant enzyme. Altogether, the data presented in this study are consistent with the FAD-histidyl covalent linkage being important for the optimal positioning of the hydride ion donor and acceptor in the tunneling reaction catalyzed by choline oxidase.A number of enzymes, including dehydrogenases (13), monooxygenases (47), halogenases (811), and oxidases (7, 12, 13), employ flavin cofactors (FAD or FMN) for their catalytic processes. About a tenth of all flavoproteins have been shown to contain a covalently attached cofactor, which may be linked at the C8M position via histidyl, tyrosyl, or cysteinyl side chains or at the C6M position via a cysteinyl side chain (14). Glucooligosaccharide oxidase (15, 16), hexose oxidase (17), and berberine bridge enzyme (18, 19) are examples of flavoproteins (FAD as cofactor) with both linkages present in one flavin molecule. The covalent linkages in flavin-dependent enzymes have been shown to stabilize protein structure (2022), prevent loss of loosely bound flavin cofactors (23), modulate the redox potential of the flavin microenvironment (20, 2327), facilitate electron transfer reactions (28), and contribute to substrate binding as in the case of the cysteinyl linkage (20). However, no study has implicated a mechanistic role of the flavin covalent linkages in enzymatic reactions in which a hydride ion is transferred by quantum mechanical tunneling.The discovery of quantum mechanical tunneling in enzymatic reactions, in which hydrogen atoms, protons, and hydride ions are transferred, has attracted considerable interest in enzyme studies geared toward understanding the mechanisms underlying the several orders of magnitudes in the rate enhancements of protein-catalyzed reactions compared with non-enzymatic ones. Tunneling mechanisms have been shown in a wide array of cofactor-dependent enzymes, including flavoenzymes. Examples of flavoenzymes in which the tunneling mechanisms have been demonstrated include morphinone reductase (29, 30), pentaerythritol tetranitrate reductase (29), glucose oxidase (3133), and choline oxidase (34). Mechanistic data on Class 2 dihydroorotate dehydrogenases, also with a flavin cofactor (FMN) covalently linked to the protein moiety (35, 36), could only propose a mechanism that is either stepwise or concerted with significant quantum mechanical tunneling for the hydride transfer from C6 and the deprotonation at C5 in the oxidation of dihydroorotate to orotate (37). This leaves choline oxidase as the only characterized enzyme with a covalently attached flavin cofactor (12, 38), where the oxidation of its substrate occurs unequivocally by quantum mechanical tunneling.Choline oxidase from Arthrobacter globiformis catalyzes the two-step FAD-dependent oxidation of the primary alcohol substrate choline to glycine betaine with betaine aldehyde, which is predominantly bound to the enzyme and forms a gem-diol species, as intermediate (Scheme 1). Glycine betaine accumulates in the cytoplasm of plants and bacteria as a defensive mechanism against stress conditions, thus making genetic engineering of relevant plants of economic interest (3945), and the biosynthetic pathway for the osmolyte is a potential drug target in human microbial infections of clinical interest (4648). The first oxidation step catalyzed by choline oxidase involves the transfer of a hydride ion from a deprotonated choline to the protein-bound flavin followed by reaction of the anionic flavin hydroquinone with molecular oxygen to regenerate the oxidized FAD (for a recent review see Ref. 50). The gem-diol choline, i.e. hydrated betaine aldehyde, is the substrate for the second oxidation step (49), suggesting that the reaction may follow a similar mechanism. The isoalloxazine ring of the flavin cofactor, which is buried within the protein, is physically constrained through a covalent linkage via the C(8) methyl of the flavin and the N(3) atom of the histidine side chain at position 99 (Fig. 1) (12). Also contributing to the physical constrain are the proximity of Ile-103 to the pyrimidine ring and the interactions of the backbone atoms of residues His-99 through Ile-103 with the isoalloxazine ring. The rigid positioning of the isoalloxazine ring could only permit a solvent-excluded cavity of ∼125 Å3 adjacent to the re face of the FAD to accommodate a 93-Å3 choline molecule in the substrate binding domain (12). Mechanistic data thus far obtained on choline oxidase, coupled with the crystal structure of the wild-type enzyme resolved to 1.86 Å, are consistent with a quantum tunneling mechanism for the hydride ion transfer occurring within a highly preorganized enzyme-substrate complex (Scheme 2) (12, 34, 50). Exploitation of the tunneling mechanism requires minimal independent movement of the hydride ion donor and acceptor, with the only dynamic motions permitted being the ones that promote the hydride transfer reaction.Open in a separate windowSCHEME 1.Two-step, four-electron oxidation of choline catalyzed by choline oxidase.Open in a separate windowFIGURE 1.x-ray crystal structure of the active site of wild-type choline oxidase resolved to 1.86 Å (PDB 2jbv). Note the significant distortion of the flavin ring at the C(4a) atom, which is due to the presence of a C(4a) adduct (69).Open in a separate windowSCHEME 2.The hydride ion transfer reaction from the α-carbon of the activated choline alkoxide species to the N(5) atom of the isoalloxazine ring of the enzyme-bound flavin in choline oxidase.In the present study, the contribution of the physically constrained flavin isoalloxazine ring to the reaction catalyzed by choline oxidase has been investigated in a variant enzyme in which the histidine residue at position 99 was replaced with an asparagine. The results suggest that, although not being required per se, the covalent linkage in choline oxidase contributes to the hydride tunneling reaction by either preventing independent movement or contributing to the optimal positioning of the flavin acting as hydride ion acceptor with respect to the alkoxide species acting as a donor. However, the covalent linkage is not required for the reaction.  相似文献   

10.
11.
Cryptochromes and DNA photolyases are related flavoproteins with flavin adenine dinucleotide as the common cofactor. Whereas photolyases repair DNA lesions caused by UV radiation, cryptochromes generally lack repair activity but act as UV-A/blue light photoreceptors. Two distinct electron transfer (ET) pathways have been identified in DNA photolyases. One pathway uses within its catalytic cycle, light-driven electron transfer from FADH* to the DNA lesion and electron back-transfer to semireduced FADHo after photoproduct cleavage. This cyclic ET pathway seems to be unique for the photolyase subfamily. The second ET pathway mediates photoreduction of semireduced or fully oxidized FAD via a triad of aromatic residues that is conserved in photolyases and cryptochromes. The 5,10-methenyltetrahydrofolate (5,10-methenylTHF) antenna cofactor in members of the photolyase family is bleached upon light excitation. This process has been described as photodecomposition of 5,10-methenylTHF. We show that photobleaching of 5,10-methenylTHF in Arabidopsis cry3, a member of the cryptochrome DASH family, with repair activity for cyclobutane pyrimidine dimer lesions in single-stranded DNA and in Escherichia coli photolyase results from reduction of 5,10-methenylTHF to 5,10-methyleneTHF that requires the intact tryptophan triad. Thus, a third ET pathway exists in members of the photolyase family that remained undiscovered so far.DNA photolyases and cryptochromes (cry)2 form a large family of related flavoproteins with DNA repair activity and photoreceptor function, respectively. Members of this protein family were identified in all kingdoms of life and can be grouped in at least nine subclades (1). DNA photolyases repair cytotoxic and mutagenic DNA lesions that are formed during exposure of DNA to UV-B. These DNA lesions are cyclobutane pyrimidine dimers (CPDs) or pyrimidine-pyrimidone (6-4) photoproducts. According to their substrate specificity, DNA photolyases are designated as CPD photolyases or (6-4) photolyases (2). The repair of both types of DNA lesions by photolyase requires the catalytic fully reduced and anionic flavin cofactor FADH that, when photoexcited, injects an electron directly into the DNA lesion (1) as shown in Fig. 1A (electron transfer pathway 1). During extraction from the cell and purification under aerobic conditions the flavin cofactor is usually oxidized to the semireduced and eventually to the fully oxidized form. Reduction of these flavin species to FADH in vitro can be achieved by illumination of the enzyme in the presence of reducing agents such as dithiothreitol or β-mercaptoethanol. This process is named photoactivation (1). Photoactivation in vitro requires photoexcitation of the flavin and a triad of redox-active residues in the protein moiety that is highly conserved in DNA photolyases (3, 4) as shown in Fig. 1A (electron transfer pathway 2). These residues are generally tryptophans that allow transport of an electron from the protein surface to the U-shaped flavin cofactor, which is buried within the C-terminal α-helical domain (59). Whether the same mechanism is used by photolyase to photoreduce FAD in vivo is a matter of debate (10). Photoreduction of the flavin cofactor was also observed in cryptochrome blue/UV-A photoreceptors. However, instead of fully reduced flavin, semireduced flavin species (either anionic flavin semiquinone radical or neutral semiquinone radical) accumulate. This form of the photoreceptor is considered as the signaling state (1114).Open in a separate windowFIGURE 1.Electron transfer pathways in cry3 and structures of folates. A, indicated are the distances of the tryptophans in the tryptophan triad (Trp-356, -409, -432) of Trp-432 to FADH and of FADH to the 5,10-methenylTHF (MTHF) cofactor in cry3. Shown are also the two established routes of electrons from FADH to the DNA lesion (Route 1) and within the tryptophan triad to FAD (Route 2). The third electron transfer pathway from FADH to 5,10-methenylTHF (Route 3) is the subject of this study. B, chemical structures of folates and their molecular masses. Folypolyglutamate molecules have a pteridin and a p-aminobenzoate moiety linked with a glutamate chain with a variable number of glutamic acids. The various THF species differ in their oxidation state of the C1 unit that is attached at the N-5 or N-10 position or form a bridge between both.A recently discovered subclade of the DNA photolyase/cryptochrome family are DASH cryptochromes, which have members in plants, bacteria, and aquatic animals (6, 1517). Because DASH cryptochromes were found to lack repair activity for CPDs in double-stranded DNA, they were considered as cryptochrome-type photoreceptors (6, 16). However, it was recently shown that DASH cryptochromes repair CPDs in single-stranded DNA (18) and loop structures of double-stranded DNA (19) and, thus, belong to the CPD photolyase group. In contrast to conventional CPD photolyases, DASH cryptochromes are unable to flip the CPD lesion out of the DNA duplex (7).Besides the flavin cofactor that is essential for enzymatic activity, DNA photolyases and most likely all cryptochromes contain a second chromophore (1). Like the catalytic flavin, the second chromophore is non-covalently attached to the protein moiety. The majority of DNA photolyases and, as far as studied, the cryptochromes including the DASH-type like cry3 from Arabidopsis thaliana contain polyglutamated 5,10-methenyltetrahydrofolate (5,10-methenylTHF) as the second chromophore (1, 12, 17, 20, 21) (see Fig. 1B for folate structures). Several organisms like the cyanobacterium Anacystis nidulans (Synechococcus elongatus) produce deazariboflavins (7,8-didemethyl-8-hydroxy-5-deazariboflavin) and utilize them as second cofactor (22). In photolyases of thermophilic bacteria and Archaea of the genus Sulfolobus, FMN and FAD, respectively, were found as second cofactors (23, 24). The sole function of the second cofactors demonstrated at present is transfer of excitation energy to the catalytic flavin cofactor via a Förster-type mechanism. The crystal structures of DNA photolyases and DASH cryptochromes revealed that the second chromophores are located in a cleft between the N-terminal α/β domain and the C-terminal α domain (79). The centroid distances between the catalytic FAD and the second chomophore are in the range of 15–18 Å. The close distances and the angles between the transition dipole moments of the two cofactors are favorable for efficient energy transfer. Indeed, energy transfer efficiencies are about 70% for Escherichia coli photolyase (25), close to 100% for A. nidulans photolyase (26), and between 78% (dark-adapted) and 87% (light-adapted) for Arabidopsis cry3 (27). Although the second cofactors are not essential for catalysis (28, 29), they increase the efficiency of repair and possibly of photoactivation by having higher extinction coefficients than FADH in the near UV and blue region (30). The spectral overlap between 5,10-methenylTHF emission and the absorption of the different flavin redox states is on the order FADHo > FADox > FADH (31).Illumination in vitro of photolyase that contains fully oxidized or semireduced flavin results in light-induced absorbance changes. The decrease in absorption in the 450–470-nm region reflects a decrease in the amount of fully oxidized FAD concomitant with transient increase in absorption above 500 nm, which indicates the formation of a neutral semiquinone radical. Excitation of the 5,10-methenylTHF antenna chromophore at its absorption peak at 380 nm causes a likewise photoreduction of the catalytic FAD (1, 27, 28, 30, 31). However, irreversible bleaching of the 380-nm peak is observed under high irradiance UV-A or camera flash illumination (28, 30). This irreversible bleaching goes along with release of the folate cofactor from the protein moiety (30) and was named photodecomposition of 5,10-methenylTHF (28). However, the identity of the formed folate species remained unknown (30). In our previous spectroscopic characterization of Arabidopsis cry3, a similar bleaching of the 380-nm peak was observed (27).Here we show that a third electron transfer pathway exists in photolyase and DASH cryptochome, where the 5,10-methenylTHF cofactor is photoreduced to 5,10-methyleneTHF. Thus, bleaching at 380 nm does not simply reflect destruction but is a specific chemical conversion of the second chromophore.  相似文献   

12.
FTY720, a sphingosine analog, is in clinical trials as an immunomodulator. The biological effects of FTY720 are believed to occur after its metabolism to FTY720 phosphate. However, very little is known about whether FTY720 can interact with and modulate the activity of other enzymes of sphingolipid metabolism. We examined the ability of FTY720 to modulate de novo ceramide synthesis. In mammals, ceramide is synthesized by a family of six ceramide synthases, each of which utilizes a restricted subset of acyl-CoAs. We show that FTY720 inhibits ceramide synthase activity in vitro by noncompetitive inhibition toward acyl-CoA and uncompetitive inhibition toward sphinganine; surprisingly, the efficacy of inhibition depends on the acyl-CoA chain length. In cultured cells, FTY720 has a more complex effect, with ceramide synthesis inhibited at high (500 nm to 5 μm) but not low (<200 nm) sphinganine concentrations, consistent with FTY720 acting as an uncompetitive inhibitor toward sphinganine. Finally, electrospray ionization-tandem mass spectrometry demonstrated, unexpectedly, elevated levels of ceramide, sphingomyelin, and hexosylceramides after incubation with FTY720. Our data suggest a novel mechanism by which FTY720 might mediate some of its biological effects, which may be of mechanistic significance for understanding its mode of action.FTY720 (2-amino-(2-2-[4-octylphenyl]ethyl)propane 1,3-diol hydrochloride), also known as Fingolimod, is an immunosuppressant drug currently being tested in clinical trials for organ transplantation and autoimmune diseases such as multiple sclerosis (1). FTY720 is a structural analog of sphingosine, a key biosynthetic intermediate in sphingolipid (SL)2 metabolism (see Fig. 1). In vivo, FTY720 is rapidly phosphorylated by sphingosine kinase 2 (2, 3) to form FTY720 phosphate (FTY720-P), an analog of sphingosine 1-phosphate (S1P) (see Fig. 1A). FTY720-P binds to S1P receptors (S1PRs) (4, 5) and thereby induces a variety of phenomena such as T-lymphocyte migration from lymphoid organs (69); accordingly, FTY720 treatment results in lymphopenia as lymphocytes (especially T-cells) become sequestered inside lymphoid organs (1012). The ability of FTY720 to sequester lymphocytes has stimulated its use in treatment of allograft rejection and autoimmune diseases (13), and FTY720 is currently under phase III clinical trials for treatment of relapsing-remitting multiple sclerosis (14).Open in a separate windowFIGURE 1.SL structure and metabolism. A, structures of SLs and SL analogs used in this study. B, metabolic inter-relationships between SLs and the metabolism of FTY720. The enzymes are denoted in italics. LPP3, lipid phosphate phosphatase 3; LPP1α, lipid phosphate phosphatase 1α.Apart from the binding of FTY720-P to S1PRs, the ability of FTY720 to inhibit S1P lyase (15) (see Fig. 1B), and its inhibitory effect on cytosolic phospholipase A2 (16), whose activity can be modulated by ceramide 1-phosphate (17), little is known about whether FTY720 or FTY720-P can modulate the activity of other enzymes of SL metabolism. Because FTY720 is an analog of sphingosine, one of the two substrates of ceramide synthase (CerS) (see Fig. 1), we now examine whether FTY720 can modulate CerS activity. CerS utilizes fatty acyl-CoAs to N-acylate sphingoid long chain bases. Six CerS exist in mammals, each of which uses a restricted subset of acyl-CoAs (1823). We demonstrate that FTY720 inhibits CerS activity and that the extent of inhibition varies according to the acyl chain length of the acyl-CoA substrate. Surprisingly, FTY720 inhibits CerS activity toward acyl-CoA via noncompetitive inhibition and toward sphinganine via uncompetitive inhibition. Finally, the mode of interaction of FTY720 with CerS in cultured cells depends on the amount of available sphinganine. Together, we show that FTY720 modulates ceramide synthesis, which may be of relevance for understanding its biological effects in vivo and its role in immunomodulation.  相似文献   

13.
Apoptotic caspases, such as caspase-7, are stored as inactive protease zymogens, and when activated, lead to a fate-determining switch to induce cell death. We previously discovered small molecule thiol-containing inhibitors that when tethered revealed an allosteric site and trapped a conformation similar to the zymogen form of the enzyme. We noted three structural transitions that the compounds induced: (i) breaking of an interaction between Tyr-223 and Arg-187 in the allosteric site, which prevents proper ordering of the catalytic cysteine; (ii) pinning the L2′ loop over the allosteric site, which blocks critical interactions for proper ordering of the substrate-binding groove; and (iii) a hinge-like rotation at Gly-188 positioned after the catalytic Cys-186 and Arg-187. Here we report a systematic mutational analysis of these regions to dissect their functional importance to mediate the allosteric transition induced by these compounds. Mutating the hinge Gly-188 to the restrictive proline causes a massive ∼6000-fold reduction in catalytic efficiency. Mutations in the Arg-187–Tyr-223 couple have a far less dramatic effect (3–20-fold reductions). Interestingly, although the allosteric couple mutants still allow binding and allosteric inhibition, they partially relieve the mutual exclusivity of binding between inhibitors at the active and allosteric sites. These data highlight a small set of residues critical for mediating the transition from active to inactive zymogen-like states.Caspases are a family of dimeric cysteine proteases whose members control the ultimate steps for apoptosis (programmed cell death) or innate inflammation among others (for reviews, see Refs. 1 and 2). During apoptosis, the upstream initiator caspases (caspase-8 and -9) activate the downstream executioner caspases (caspase-3, -6, and-7) via zymogen maturation (3). The activated executioner caspases then cleave upwards of 500 key proteins (46) and DNA, leading to cell death. Due to their pivotal role in apoptosis, the caspases are involved both in embryonic development and in dysfunction in diseases including cancer and stroke (7). The 11 human caspases share a common active site cysteine-histidine dyad (8), and derive their name, cysteine aspartate proteases, from their exquisite specificity for cleaving substrate proteins after specific aspartate residues (913). Thus, it has been difficult to develop active site-directed inhibitors with significant specificity for one caspase over the others (14). Despite difficulties in obtaining specificity, there has been a long-standing correlation between efficacy of caspase inhibitors in vitro and their ability to inhibit caspases and apoptosis in vivo (for review, see Ref. 31). Thus, a clear understanding of in vitro inhibitor function is central to the ability control caspase function in vivo.Caspase-7 has been a paradigm for understanding the structure and dynamics of the executioner caspases (1521). The substrate-binding site is composed of four loops; L2, L3, and L4 are contributed from one-half of the caspase dimer, and L2′ is contributed from the other half of the caspase dimer (Fig. 1). These loops appear highly dynamic as they are only observed in x-ray structures when bound to substrate or substrate analogs in the catalytically competent conformation (1719, 22) (Fig. 1B).Open in a separate windowFIGURE 1.Allosteric site and dimeric structure in caspase-7. A, the surface of active site-bound caspase-7 shows a large open allosteric (yellow) site at the dimer interface. This cavity is distinct from the active sites, which are bound with the active site inhibitor DEVD (green sticks). B, large subunits of caspase-7 dimers (dark green and dark purple) contain the active site cysteine-histidine dyad. The small subunits (light green and light purple) contain the allosteric site cysteine 290. The conformation of the substrate-binding loops (L2, L2′, L3, and L4) in active caspase-7 (Protein Data Bank (PDB) number 1f1j) is depicted. The L2′ loop (spheres) from one-half of the dimer interacts with the L2 loop from the other half of the dimer. C, binding of allosteric inhibitors influences the conformation of the L2′ loop (spheres), which folds over the allosteric cavity (PDB number 1shj). Subunit rendering is as in panel A. Panels A, B, and C are in the same orientation.A potential alternative to active site inhibitors are allosteric inhibitors that have been seeded by the discovery of selective cysteine-tethered allosteric inhibitors for either apoptotic executioner caspase-3 or apoptotic executioner caspase-7 (23) as well as the inflammatory caspase-1 (24). These thiol-containing compounds bind to a putative allosteric site through disulfide bond formation with a thiol in the cavity at the dimer interface (Fig. 1A) (23, 24). X-ray structures of caspase-7 bound to allosteric inhibitors FICA3 and DICA (Fig. 2) show that these compounds trigger conformational rearrangements that stabilize the inactive zymogen-like conformation over the substrate-bound, active conformation. The ability of small molecules to hold mature caspase-7 in a conformation that mimics the naturally occurring, inactive zymogen state underscores the utility and biological relevance of the allosteric mechanism of inhibition. Several structural changes are evident between these allosterically inhibited and active states. (i) The allosteric inhibitors directly disrupt an interaction between Arg-187 (next to the catalytic Cys-186) and Tyr-223 that springs the Arg-187 into the active site (Fig. 3), (ii) this conformational change appears to be facilitated by a hinge-like movement about Gly-188, and (iii) the L2′ loop folds down to cover the allosteric inhibitor and assumes a zymogen-like conformation (Fig. 1C) (23).Open in a separate windowFIGURE 2.Structure of allosteric inhibitors DICA and FICA. DICA and FICA are hydrophobic small molecules that bind to an allosteric site at the dimer interface of caspase-7. Binding of DICA/FICA is mediated by a disulfide between the compound thiol and Cys-290 in caspase-7.Open in a separate windowFIGURE 3.Movement of L2′ blocking arm. The region of caspase-7 encompassing the allosteric couple Arg-187 and Tyr-223 is boxed. The inset shows the down orientation of Arg-187 and Tyr-223 in the active conformation with DEVD substrate mimic (orange spheres) in the active site. In the allosteric/zymogen conformation, Arg-187 and Tyr-223 are pushed up by DICA (blue spheres).Here, using mutational analysis and small molecule inhibitors, we assess the importance of these three structural units to modulate both the inhibition of the enzyme and the coupling between allosteric and active site labeling. Our data suggest that the hinge movement and pinning of the L2-L2′ are most critical for transitioning between the active and inactive forms of the enzyme.  相似文献   

14.
ARAP1 is a phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GTPase-activating protein (GAP) with five PH domains that regulates endocytic trafficking of the epidermal growth factor receptor (EGFR). Two tandem PH domains are immediately N-terminal of the Arf GAP domain, and one of these fits the consensus sequence for PtdIns(3,4,5)P3 binding. Here, we tested the hypothesis that PtdIns(3,4,5)P3-dependent recruitment mediated by the first PH domain of ARAP1 regulates the in vivo and in vitro function of ARAP1. We found that PH1 of ARAP1 specifically bound to PtdIns(3,4,5)P3, but with relatively low affinity (≈1.6 μm), and the PH domains did not mediate PtdIns(3,4,5)P3-dependent recruitment to membranes in cells. However, PtdIns(3,4,5)P3 binding to the PH domain stimulated GAP activity and was required for in vivo function of ARAP1 as a regulator of endocytic trafficking of the EGFR. Based on these results, we propose a variation on the model for the function of phosphoinositide-binding PH domains. In our model, ARAP1 is recruited to membranes independently of PtdIns(3,4,5)P3, the subsequent production of which triggers enzymatic activity.Pleckstrin homology (PH)2 domains are a common structural motif encoded by the human genome (1, 2). Approximately 10% of PH domains bind to phosphoinositides. These PH domains are thought to mediate phosphoinositide-dependent recruitment to membranes (13). Most PH domains likely have functions other than or in addition to phosphoinositide binding. For example, PH domains have been found to bind to protein and DNA (412). In addition, some PH domains have been found to be structurally and functionally integrated with adjacent domains (13, 14). A small fraction of PH domain-containing proteins (about 9% of the human proteins) have multiple PH domains arranged in tandem, which have been proposed to function as adaptors but have only been examined in one protein (15, 16). Arf GTPase-activating proteins (GAPs) of the ARAP family are phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GAPs with tandem PH domains (17, 18). The function of specific PH domains in regulating Arf GAP activity and for biologic activity has not been described.Arf GAPs are proteins that induce the hydrolysis of GTP bound to Arfs (1923). The Arf proteins are members of the Ras superfamily of GTP-binding proteins (2427). The six Arf proteins in mammals (five in humans) are divided into three classes based on primary sequence: Arf1, -2, and -3 are class 1, Arf4 and -5 are class 2, and Arf6 is class 3 (23, 24, 2729). Class 1 and class 3 Arf proteins have been studied more extensively than class 2. They have been found to regulate membrane traffic and the actin cytoskeleton.The Arf GAPs are a family of proteins with diverse domain structures (20, 21, 23, 30). ARAPs, the most structurally complex of the Arf GAPs, contain, in addition to an Arf GAP domain, the sterile α motif (SAM), five PH, Rho GAP, and Ras association domains (17, 18, 31, 32). The first and second and the third and fourth PH domains are tandem (Fig. 1). The first and third PH domains of the ARAPs fit the consensus for PtdIns(3,4,5)P3 binding (3335). ARAPs have been found to affect actin and membrane traffic (21, 23). ARAP3 regulates growth factor-induced ruffling of porcine aortic endothelial cells (31, 36, 37). The function is dependent on the Arf GAP and Rho GAP domains. ARAP2 regulates focal adhesions, an actin cytoskeletal structure (17). ARAP2 function requires Arf GAP activity and a Rho GAP domain capable of binding RhoA·GTP. ARAP1 has been found to have a role in membrane traffic (18). The protein associates with pre-early endosomes involved in the attenuation of EGFR signals. The function of the tandem PH domains in the ARAPs has not been examined.Open in a separate windowFIGURE 1.ARAP1 binding to phospholipids. A, schematic of the recombinant proteins used in this study. Domain abbreviations: Ank, ankyrin repeat; PLCδ-PH, PH domain of phospholipase C δ; RA, Ras association motif; RhoGAP, Rho GTPase-activating domain. B, ARAP1 phosphoinositide binding specificity. 500 nm PH1-Ank recombinant protein was incubated with sucrose-loaded LUVs formed by extrusion through a 1-μm pore filter. LUVs contained PtdIns alone or PtdIns with 2.5 μm PtdIns(3,4,5)P3, 2.5 μm PtdIns(3)P, 2.5 μm PtdIns(4)P, 2.5 μm PtdIns(5)P, 2.5 μm PtdIns(3,4)P2, 2.5 μm PtdIns(3,5)P2, or 2.5 μm PtdIns(4,5)P2 with a total phosphoinositide concentration of 50 μm and a total phospholipid concentration of 500 μm. Vesicles were precipitated by ultracentrifugation, and associated proteins were separated by SDS-PAGE. The amount of precipitated protein was determined by densitometry of the Coomassie Blue-stained gels with standards on each gel. C, PtdIns(3,4,5)P3-dependent binding of ARAP1 to LUVs. 1 μm PH1-Ank or ArfGAP-Ank recombinant protein was incubated with 1 mm sucrose-loaded LUVs formed by extrusion through a 1-μm pore size filter containing varying concentration of PtdIns(3,4,5)P3. Precipitation of LUVs and analysis of associated proteins were performed as described in B. The average ± S.E. of three independent experiments is presented.Here we investigated the role of the first two PH domains of ARAP1 for catalysis and in vivo function. The first PH domain fits the consensus sequence for PtdIns(3,4,5)P3 binding (3335). The second does not fit a phosphoinositide binding consensus but is immediately N-terminal to the GAP domain. We have previously reported that the PH domain that occurs immediately N-terminal of the Arf GAP domain of ASAP1 is critical for the catalytic function of the protein (38, 39). We tested the hypothesis that the two PH domains of ARAP1 function independently; one recruits ARAP1 to PtdIns(3,4,5)P3-rich membranes, and the other functions with the catalytic domain. As predicted, PH1 interacted specifically with PtdIns(3,4,5)P3, and PH2 did not. However, both PH domains contributed to catalysis independently of recruitment to membranes. None of the PH domains in ARAP1 mediated PtdIns(3,4,5)P3-dependent targeting to plasma membranes (PM). PtdIns(3,4,5)P3 stimulated GAP activity, and the ability to bind PtdIns(3,4,5)P3 was required for ARAP1 to regulate membrane traffic. We propose that ARAP1 is recruited independently of PtdIns(3,4,5)P3 to the PM where PtdIns(3,4,5)P3 subsequently regulates its GAP activity to control endocytic events.  相似文献   

15.
16.
17.
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   

18.
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   

19.
The molecular weight of hyaluronan is important for its rheological and biological function. The molecular mechanisms underlying chain termination and hence molecular weight control remain poorly understood, not only for hyaluronan synthases but also for other β-polysaccharide synthases, e.g. cellulose, chitin, and 1,3-betaglucan synthases. In this work, we manipulated metabolite concentrations in the hyaluronan pathway by overexpressing the five genes of the hyaluronan synthesis operon in Streptococcus equi subsp. zooepidemicus. Overexpression of genes involved in UDP-glucuronic acid biosynthesis decreased molecular weight, whereas overexpression of genes involved in UDP-N-acetylglucosamine biosynthesis increased molecular weight. The highest molecular mass observed was at 3.4 ± 0.1 MDa twice that observed in the wild-type strain, 1.8 ± 0.1 MDa. The data indicate that (a) high molecular weight is achieved when an appropriate balance of UDP-N-acetylglucosamine and UDP-glucuronic acid is achieved, (b) UDP-N-acetylglucosamine exerts the dominant effect on molecular weight, and (c) the wild-type strain has suboptimal levels of UDP-N-acetylglucosamine. Consistent herewith molecular weight correlated strongly (ρ = 0.84, p = 3 × 10−5) with the concentration of UDP-N-acetylglucosamine. Data presented in this paper represent the first model for hyaluronan molecular weight control based on the concentration of activated sugar precursors. These results can be used to engineer strains producing high molecular weight hyaluronan and may provide insight into similar polymerization mechanisms in other polysaccharides.Hyaluronan (HA)3 is a linear polymer of a repeating disaccharide, β1–3 d-N-acetylglucosamine (GlcNAc) β1–4 d-glucuronic acid (GlcUA) (1) (see Fig. 1). Ubiquitous in the extracellular matrix in vertebrates, HA is particularly abundant in cartilage, synovial fluid, dermis, and the vitreous humor of the eye, where it serves specialized functions. HA also plays a critical role during fertilization and embryogenesis. In many group A and C streptococci, HA forms a capsule that helps these microbes evade the host immune system (2). HA molecular weight is important for the physiochemical as well as biological properties of HA. High molecular weight is important for HA to exert its unique rheological properties (3), for mucoadherence (4, 5), and anti-inflammatory effects (6, 7), whereas low molecular weight is a potent signaling molecule (8).Open in a separate windowFIGURE 1.Biosynthetic pathway of HA in S. zooepidemicus.HA is produced by a processive synthase (9, 10) from the activated sugar precursors, UDP-glucuronic acid (UDP-GlcUA) and UDP-N-acetylglucosamine (UDP-GlcNAc) (see Fig. 1). In addition to the HA synthase (hasA), streptococcal has operons encode for one or more enzymes involved in biosynthesis of the activated sugars (11). The Streptococcus equi subsp. zooepidemicus (S. zooepidemicus) operon encodes for five genes: HA synthase (EC 2.4.1.212; hasA), UDP-glucose dehydrogenase (EC 1.1.1.22; hasB), UDP-glucose pyrophosphorylase (EC 2.7.7.9; hasC), a glmU paralog encoding for a dual function enzyme acetyltransferase and pyrophosphorylase activity (EC 2.3.1.4/EC 2.7.7.23; hasD), and a pgi paralog encoding for phosphoglucoisomerase (EC 5.3.1.9; hasE).Although the biosynthetic mechanism is well established, little is known about what controls HA molecular weight. This is true not only for HA, but also for the highly abundant β-polysaccharides: cellulose, chitin, and 1,3-betaglucan. Molecular weight is partly an intrinsic parameter of the HA synthase. Weigel and colleagues have demonstrated that, at least in vitro, mutation of conserved cysteine or polar residues in streptococcal HA synthases results in reduced molecular weight with limited effect on biosynthetic rate (1214). In a vertebrate HA synthase from Xenopus, the mutation of a serine or a cysteine residue yielded HA of higher, lower, or similar molecular weight depending on the amino acid substitution (15).We and others have demonstrated that in vivo molecular weight is also affected by culture parameters, e.g. temperature and aeration (1620). Although changed culture conditions affect the physiochemical environment of the HA synthase, a more likely explanation is that molecular weight is affected by the availability of activated sugar substrates (UDP-GlcUA and UDP-GlcNAc) as well as the concentration of possible effector molecules, such as free UDP (21). Although such a mechanism has been suggested for several processive synthases (22, 23), there has never been any direct evidence linking molecular weight to the concentration of a substrate.Experimental support for the hypothesis has been obtained for the type 3 polysaccharide of Streptococcus pneumoniae (24). Like HA, the type 3 polysaccharide in S. pneumoniae is synthesized by a processive synthase from alternating addition of activated sugars, in this case UDP-glucose (UDP-Glc) and UDP-GlcUA. Mutants with reduced UDP-glucose dehydrogenase (“hasB”) activity not only produce less polysaccharide, but also polysaccharide with lower molecular weight (24). Although the levels of UDP-GlcUA were below detection in all strains, this supports the idea that UDP-GlcUA concentration controls molecular weight. Moreover, it is consistent with previous in vitro studies showing that low levels of UDP-GlcUA cause chain termination and hence low molecular weight (25). It was proposed that the concentration of UDP-GlcUA is critical for the successful transition from oligosaccharide lipid to highly processive polysaccharide synthesis (26). A similar mechanism is not likely for HA biosynthesis, because there is no indication that the HA synthase needs a primer (27).In the present study, we manipulated metabolite concentrations in the HA pathway by overexpressing the five genes in the has operon of S. zooepidemicus (Fig. 1). Overexpression of these genes had a profound effect on HA molecular weight, which correlated with the levels of UDP-sugars and in particular, UDP-GlcNAc.  相似文献   

20.
Mammalian glutamate dehydrogenase (GDH) is a homohexameric enzyme that catalyzes the reversible oxidative deamination of l-glutamate to 2-oxoglutarate using NAD(P)+ as coenzyme. Unlike its counterparts from other animal kingdoms, mammalian GDH is regulated by a host of ligands. The recently discovered hyperinsulinism/hyperammonemia disorder showed that the loss of allosteric inhibition of GDH by GTP causes excessive secretion of insulin. Subsequent studies demonstrated that wild-type and hyperinsulinemia/hyperammonemia forms of GDH are inhibited by the green tea polyphenols, epigallocatechin gallate and epicatechin gallate. This was followed by high throughput studies that identified more stable inhibitors, including hexachlorophene, GW5074, and bithionol. Shown here are the structures of GDH complexed with these three compounds. Hexachlorophene forms a ring around the internal cavity in GDH through aromatic stacking interactions between the drug and GDH as well as between the drug molecules themselves. In contrast, GW5074 and bithionol both bind as pairs of stacked compounds at hexameric 2-fold axes between the dimers of subunits. The internal core of GDH contracts when the catalytic cleft closes during enzymatic turnover. None of the drugs cause conformational changes in the contact residues, but all bind to key interfaces involved in this contraction process. Therefore, it seems likely that the drugs inhibit enzymatic turnover by inhibiting this transition. Indeed, this expansion/contraction process may play a major role in the inter-subunit communication and allosteric regulation observed in GDH.Glutamate dehydrogenase (GDH)2 is found in all living organisms and catalyzes the reversible oxidative deamination of l-glutamate to 2-oxoglutarate using NAD(P)+ as coenzyme (1). In eukaryotic organisms, GDH resides within the inner mitochondrial matrix where it catabolizes glutamate to feed 2-oxoglutarate to the Krebs cycle. Although there is some debate as to the directionality of the reaction, the high Km for ammonium in the reductive amination reaction seems to prohibit the reverse reaction under normal conditions in most organisms (2). GDH from animals, but not other kingdoms (3), is allosterically regulated by a wide array of ligands (39). GTP (911), and with ∼100-fold lower affinity, ATP (3), is a potent inhibitor of the reaction and acts by increasing the binding affinity for the product, thereby slowing down enzymatic turnover (11). Hydrophobic compounds such as palmitoyl-CoA (12), steroid hormones (13), and steroid hormone analogs such as diethylstilbestrol (5) are also potent inhibitors. ADP is an activator of GDH (3, 6, 10, 11, 14) that acts in an opposite manner to GTP by facilitating product release. Leucine is a poor substrate for GDH but is also an allosteric activator for the enzyme (8). Its activation is akin to ADP but acts at site distinct from ADP (15).The crystal structures of the bacterial (1618) and animal forms (19, 20) of GDH have shown that the general architecture and the locations of the catalytically important residues have remained unchanged throughout evolution. The structure of GDH (Fig. 1) is essentially two trimers of subunits stacked directly on top of each other with each subunit being composed of at least three domains (1922). The bottom domain makes extensive contacts with a subunit from the other trimer. Resting on top of this domain is the “NAD binding domain” that has the conserved nucleotide binding motif. Animal GDH has a long protrusion, “antenna,” rising above the NAD binding domain that is not found in bacteria, plants, fungi, and the vast majority of protists. The antenna from each subunit lies immediately behind the adjacent, counterclockwise neighbor within the trimer. Because these intertwined antennae are only found in the forms of GDH that are allosterically regulated by numerous ligands, it is reasonable to speculate that it plays a role in regulation.Open in a separate windowFIGURE 1.Conformational transitions and locations of ligand binding sites in bovine glutamate dehydrogenase. A, a ribbon diagram of apo-bovine glutamate dehydrogenase with each of the identical subunits represented by different colors. The subunit arrangement is that of a trimer of dimers where anti-parallel β-strands form extensive interactions between the subunits stacked on top of each other. This pairing is represented by different shades of the same color. The conformational changes that during substrate binding are shown by the numbered arrows. As substrate binds, the NAD+ binding domain closes (1). The ascending helix of the antenna moves toward the pivot helix of the adjacent subunit (2). The short helix of the descending strand of the antenna becomes extended and distorted at the carboxyl end (3). Finally, the internal cavity of the helix compresses, bringing the three pairs closer together (4). B shows the structure of ADP (green spheres) bound to the apo-form of GDH and the location of Arg-463 (mauve spheres) that is involved in ADP activation (22). C shows the location of the inhibitor, GTP (mauve spheres), bound to the NADH (gray spheres), and glutamate (orange spheres) abortive complex. The green arrow notes the approximate location of one of the two sites (Lys-420) modified by 5′-FSBA (48). Comparing B and C, the closing of the catalytic cleft and the movement of the pivot helix is evident.From the structures GDH alone and complexed with various ligands, it is clear that GDH undergoes large conformational changes during each catalytic cycle (1922) (the locations of these changes are summarized in Fig. 1). Substrate binds to the deep recesses of the cleft between the coenzyme binding domain and the lower domain. Coenzyme binds along the coenzyme binding domain surface of the cleft. Upon binding, the coenzyme binding domain rotates by ∼18° to firmly close down upon the substrate and coenzyme (Fig. 1, arrow 1). As the catalytic cleft closes, the base of each of the long ascending helices in the antenna appears to rotate out in a counterclockwise manner to push against the “pivot” helix of the adjacent subunit (Fig. 1, arrow 2). There is a short helix in the descending loop of the antenna that becomes distended and shorter as the mouth closes in a manner akin to an extending spring (Fig. 1, arrow 3). The pivot helix rotates in a counterclockwise manner along the helical axes as well as rotating counterclockwise around the trimer 3-fold axis. Finally, the entire hexamer seems to compress as the mouth closes (Fig. 1, arrow 4). The three pairs of subunits that sit on top of each other move as a rigid units toward each other, compressing the cavity at the core of the hexamer. This last conformational change will be further examined in this work. Allosteric regulation is likely exacted by controlling some or all of these conformational changes.The reason for complex animal regulation came from studies that linked GDH regulation with insulin and ammonia homeostasis. The connection between GDH and insulin regulation was initially established using a non-metabolizable analog of leucine (7, 23), BCH (β-2-aminobicycle[2.2.1]-heptane-2-carboxylic acid). These studies demonstrated that activation of GDH was tightly correlated with increased glutaminolysis and release of insulin. In addition, it has also been noted that factors that regulate GDH also affect insulin secretion (24). The in vivo importance of GDH in glucose homeostasis was demonstrated by the discovery that a genetic hypoglycemic disorder, the hyperinsulinemia/hyperammonemia (HHS) syndrome, is caused by loss of GTP regulation of GDH (2527). Children with HHS have increased β-cell responsiveness to leucine and susceptibility to hypoglycemia following high protein meals (28). This is likely due to uncontrolled catabolism of amino acids yielding high ATP levels that stimulate insulin secretion as well as high serum ammonium levels. During glucose-stimulated insulin secretion in normal individuals, it has been proposed that the generation of high energy phosphates inhibits GDH and promotes conversion of glutamate to glutamine, which, alone or combined, might amplify the release of insulin (29, 30). Further support for this contention came from studies on the inhibitory effects of the polyphenolic compounds from green tea on BCH-stimulated insulin secretion (31). This not only lent support for the contention that GDH plays a significant role in insulin homeostasis, but also suggests that the HHS disorder might be directly controlled pharmaceutically. The role of GDH in insulin homeostasis is summarized in Fig. 2.Open in a separate windowFIGURE 2.Link between GDH and insulin homeostasis. This figure shows the role of GDH in BCH stimulated insulin secretion and how GDH inhibitors affect this process (29, 30). In energy-depleted β-cells, a BCH ramp stimulates insulin secretion. Here, the major energy source is glutaminolysis via phosphate-dependent glutaminase and GDH, because the concentration of GDH inhibitors (ATP/GTP) have been depleted and the phosphate-dependent glutaminase activator Pi (inorganic phosphate) has been increased. BCH stimulates glutamine utilization via GDH activation, thus providing the ATP signal necessary for insulin secretion. GDH inhibitors block this process by inhibiting GDH activity.To both find a more stable pharmaceutical agent to control HHS and to better understand the allosteric regulation of GDH, a high throughput screen was performed to identify new GDH inhibitors (32). Of the ∼30,000 compounds tested, ∼20 demonstrated significant activity. Three of the most active compounds, hexachlorophene, GW5074, and bithionol, were chosen for further analysis in this study. As shown here, all three compounds exhibit essentially non-competitive inhibition of the reaction and therefore do not compete with either substrate or coenzyme. Structural studies are presented here that demonstrate that six hexachlorophene (HCP) molecules bind to the inner core of the GDH hexamer, forming an internal ring via aromatic interactions. In contrast, bithionol and GW5074 bind as pairs between dimers of GDH subunits further away from the core of the enzyme. None of these compounds induce significant conformational changes in their immediate vicinity, and the mechanism of action is not clear from the location of their binding sites. However, detailed analysis of the various GDH complexes shows all of the drugs are binding to contact areas in the core of the hexamer that appear to be expanding and contracting during catalytic turnover. Therefore, inhibition is likely due interference with this “breathing” process.  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