首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The electronic structures of transition metal complexes of NO are controlled by the stereochemistry about the metal atom (stereochemical control of valence). The six-coordinate complex, trans-[CoNO(NCS)-(C6H4 [As(CH3)2]2)2]+, consists of [Co(III)-(N=O-)]2+ (angleCo-N-O = 135°), while the pentacoordinate trigonal bipyramidal complex, [CoNO(C6H4[As(CH3)2]2)2] 2+, is best formulated as [Co(I)-(NO+]2+ (angleCo-N-O = 179°). Evidence indicates that complexes of (NO)+ and N2 are electronically similar. Hence, the principles of stereochemical control of valence may be applied to metal complexes of N2. In a linearly coordinated Mn˜m(NN) complex, valence electrons can be transferred from the metal to the N2 ligand producing a bent, protonated, and/or metallated Mn+2-(N=N2-) complex. This reduction of N2 can be effected by the addition of an appropriate ligand to M or by a change in the coordination geometry about M. Stereo-chemical control of valence leads to the rejection of one of the previously proposed mechanisms for reduction by nitrogenase.  相似文献   

2.
Steady state phenylalanine and tyrosine turnover and the rate of conversion of phenylalanine to tyrosine in vivo were determined in 6 healthy postabsorptive adult volunteers. Continuous infusions of tracer amounts of L-[ring-2H5]phenylalanine were administered intravenously for 13–14 hr. After 9–10 hr, a priming dose followed by a continuous infusion of L-[1-13C]tyrosine was added and maintained, along with the [2H5]phenylalanine infusion, for 4 hr. Venous plasma samples were obtained before the initiation of each infusion and every 30 min during the course of the combined [2H5]phenylalanine and [13C]tyrosine infusion for determination of isotopic enrichments of [2H5]phenylalanine, [13C]tyrosine, and [2H4[tyrosine by gas chromatograph-mass spectrometric analysis of the N-trifluoroacetyl-, methyl ester derivatives of the amino acids. Calculated from the observed enrichments, free phenylalanine and tyrosine turnover rates were 36.1 ± 5.1 μmole · kg?1 · h?1 and 39.8 ± 3.5 μmole · kg?1 · h?1, respectively. Phenylalanine was converted to tyrosine at the rate of 5.83 ± 0.59 μmole · kg?1 · h?1, accounting for approximately 16% of either the phenylalanine or the tyrosine flux. The results indicate that the normal basal steady state phenylalanine hydroxylase activity in vivo in man is lower than that obtained from phenylalanine loading studies. This supports the existence of some type of substrate activation of the enzyme as reflected in the previously reported exponential relationship between phenylalanine concentration and phenylalanine hydroxylase activity in vitro. The use of continuous simultaneous infusions of tracer amounts of stable isotope-labeled phenylalanine and tyrosine provides a direct means for studying physiological regulation of phenylalanine hydroxylase activity in vivo.  相似文献   

3.
The all-ferric [Fe4S4]4+ cluster [Fe4S4{N(SiMe3)2}4] 1 and its one-electron reduced form [1]- serve as convenient precursors for the synthesis of 3∶1-site differentiated [Fe4S4] clusters and high-potential iron-sulfur protein (HiPIP) model clusters. The reaction of 1 with four equivalents (equiv) of the bulky thiol HSDmp (Dmp = 2,6-(mesityl)2C6H3, mesityl = 2,4,6-Me3C6H2) followed by treatment with tetrahydrofuran (THF) resulted in the isolation of [Fe4S4(SDmp)3(THF)3] 2. Cluster 2 contains an octahedral iron atom with three THF ligands, and its Fe(S)3(O)3 coordination environment is relevant to that in the active site of substrate-bound aconitase. An analogous reaction of [1]- with four equiv of HSDmp gave [Fe4S4(SDmp)4]- 3, which models the oxidized form of HiPIP. The THF ligands in 2 can be replaced by tetramethyl-imidazole (Me4Im) to give [Fe4S4(SDmp)3(Me4Im)] 4 modeling the [Fe4S4(Cys)3(His)] cluster in hydrogenases, and its one-electron reduced form [4]- was synthesized from the reaction of 3 with Me4Im. The reversible redox couple between 3 and [3]- was observed at E1/2 = -820 mV vs. Ag/Ag+, and the corresponding reversible couple for 4 and [4]- is positively shifted by +440 mV. The cyclic voltammogram of 3 also exhibited a reversible oxidation couple, which indicates generation of the all-ferric [Fe4S4]4+ cluster, [Fe4S4(SDmp)4].  相似文献   

4.
Resonance Raman spectra of the BR568, BR548, K625, and L550 intermediates of the bacteriorhodopsin photocycle have been obtained in 1H2O and 2H2O by using native purple membrane as well as purple membrane regenerated with 14,15-13C2 and 12,14-2H2 isotopic derivatives of retinal. These derivatives were selected to determine the contribution of the C14—C15 stretch to the normal modes in the 1100- to 1400-cm-1 fingerprint region and to characterize the coupling of the C14—C15 stretch with the NH rock. Normal mode calculations demonstrate that when the retinal Schiff base is in the C[unk]N cis configuration the C14—C15 stretch and the NH rock are strongly coupled, resulting in a large (≈50-cm-1) upshift of the C14—C15 stretch upon deuteration of the Schiff base nitrogen. In the C[unk]N trans geometry these vibrations are weakly coupled and only a slight (<5-cm-1) upshift of the C14—C15 stretch is predicted upon N-deuteration. In BR568, the insensitivity of the 1201-cm-1 C14—C15 stretch to N-deuteration demonstrates that its retinal C[unk]N configuration is trans. The C14—C15 stretch in BR548, however, shifts up from 1167 cm-1 in 1H2O to 1208 cm-1 in 2H2O, indicating that BR548 contains a C[unk]N cis chromophore. Thus, the conversion of BR568 to BR548 (dark adaptation) involves isomerization about the C[unk]N bond in addition to isomerization about the C13[unk]C14 bond. The insensitivity of the native, [14,15-13C2]-, and [12,14-2H2]K625 and L550 spectra to N-deuteration argues that these intermediates have a C[unk]N trans configuration. Thus, the primary photochemical step in bacteriorhodopsin (BR568 → K625) involves isomerization about the C13[unk]C14 bond alone. The significance of these results for the mechanism of proton-pumping by bacteriorhodopsin is discussed.  相似文献   

5.
Summary The antitumor activity of the four metallocene compounds decaphenylstannocene [5-(C6H5)5C5]2Sn(II), decabenzylstannocene [5-(C6H5CH2)5C5]2Sn(II), decaphenylgermanocene [5-(C6H5)5C5]2Ge(II), and decabenzylgermanocene [5-(C6H5CH2)5C5]2Ge(II), containing the main group IV elements tin or germanium as the central metal atom and two pentasubstituted cyclopentadienyl ring ligands in sandwich arrangement, were tested against Ehrlich ascites tumor in female CF1 mice. The complexes caused cure rates of 40% to 90% of the animals treated over rather broad dose ranges. With both germanocene complexes, no strong dose-activity relationship was manifest. The toxicity of all four metallocenes was low, the LD10 values of both stannocenes being 460 and 500 mg/kg, and those of both germanocenes higher than 700 mg/kg. Regarding the isolated pentasubstituted cyclopentadiene ligands (C6H5)5C5H and (C6H5CH2)5C5H, these also exhibited antitumor activity which was less pronounced than that of the metal-containing sandwich complexes. Decasubstituted stannocene and germanocene compounds represent a new type of non-platinum group metal antitumor agents structurally differing from known inorganic and organometallic cytostatics.  相似文献   

6.
n-Type Si has been shown to serve as a stable photoanode in a cell for the conversion of light to electricity. The other components of the cell are a Pt cathode and an electrolyte consisting of an ethanol solution of [n-Bu4N]ClO4 with a redox couple of ferricenium/ferrocene. Data from a two-compartment cell show that ferrocene is oxidized to ferricenium with 100 ± 2% current efficiency at the Si photoanode. Furthermore, prolonged irradiation of the Si in a one-compartment cell yields constant photocurrent and output characteristics. The maximum open-circuit photopotential is ~700 mV, and the short-circuit quantum yield for electron flow at low light intensity exceeds 0.5. Conversion of monochromatic 632.8-nm light to electricity with ~2% power efficiency at an output voltage of ~200 mV has been sustained. These results represent a stable n-type Si-based photoelectrochemical cell.  相似文献   

7.
To determine the contributions of galactose and fructose to glucose formation, 6 subjects (26 ± 2 years old; body mass index, 22.4 ± 0.2 kg/m2) (mean ± SE) were studied during fasting conditions. Three subjects received a primed constant intravenous infusion of [6,6-2H2]glucose for 3 hours followed by oral bolus ingestion of galactose labeled to 2% with [U-13C]galactose (0.72 g/kg); the other 3 subjects received a primed constant intravenous infusion of [6,6-2H2]glucose followed by either a bolus ingestion of fructose alone (0.72 g/kg) (labeled to 2% with [U-13C]fructose) or coingestion of fructose (labeled with [U-13C]fructose) (0.72 g/kg) and unlabeled glucose (0.72 g/kg). Four hours after ingestion, subjects received 1 mg of glucagon intravenously to stimulate glycogenolysis. When galactose was ingested alone, the area under the curve (AUC) of [13C6]glucose and [13C3]glucose was 7.28 ± 0.39 and 3.52 ± 0.05 mmol/L per 4 hours, respectively. When [U-13C]fructose was ingested with unlabeled fructose or unlabeled fructose plus glucose, no [13C6]glucose was detected in plasma. The AUC of [13C3]glucose after fructose and fructose plus glucose ingestion was 20.21 ± 2.41 and 6.25 ± 0.34 mmol/L per 4 hours, respectively. Comparing the AUC for the 13C3 vs 13C6 enrichments, 67% of oral galactose enters the systemic circulation via a direct route and 33% via an indirect route. In contrast, fructose only enters the systemic circulation via the indirect route. Finally, when ingested alone, fructose and galactose contribute little to glycogen synthesis. After the coingestion of fructose and glucose with the resultant insulin response from the glucose, fructose is a significant contributor to glycogen synthesis.  相似文献   

8.
Fifteen children, five with phenylketonuria (PKU), five with hyperphenylalaninaemia, and five phenotypically normal but at risk of being carriers for PKU, were given [ring2H5]phenylalanine orally in amounts ranging from 75 mg/kg to 10 mg/kg. Plama was assayed for [2H5]phenylalanine and [2H4]tyrosine at hourly intervals, the amino acids being measured as theN-acetyl, n-propyl esters by gas chromatography-mass spectroscopy. The results obtained were calculated as the log of the ratio [2H5]phenylalanine: [2H4]tyrosine in the plasma. The five patients with PKU had ratios of infinity because no [2H4]tyrosine was measured in their plasma during the experimental period. The patients with hyperphenylalaninaemia had log ratios over 2.00 throughout the assay period. Among the five normal children three are considered to be carriers for PKU as the logarithms of the [2H5]phenylalanine: [2H4]tyrosine ratios were 1.77, 1.73, and 1.33 and remained over 1.00 during the assay period. The other children had log ratios of 1.16 and 1.00 at the first hour which dropped below 1.00 subsequently, suggesting normal activity of phenylalanine hydroxylase.  相似文献   

9.
Summary In open-chest pigs during severe myocardial ischemia [K+]e, [Ca2+]e and [H+]e increase, [Na+]e increases transiently reaching control values after 30 min. Extracellular osmolality of the ischemic area increases due to an H2O-shift from the ECS to the ICS. The increase of [Na+]e and [Ca2+]e must be explained by the shrinkage of the ECS due to the H2O-shift. The increase of [Ca2+]e is additionally caused by the decrease of pHe. The increase of [K+]e is mainly caused by the release of K+ from the ICS. The changes of [K+]e and [K+]i cause a decrease of the membrane potential to a range in which slow response potentials and re-entry excitations can occur. The increase of [K+e therefore seems to be a major factor to cause early post-ischemic arrhythmias.Supported by Deutsche Forschungsgemeinschaft, SFB 68, A 7  相似文献   

10.
2H magnetic resonance spectroscopic imaging has been shown recently to be a viable technique for metabolic imaging in the clinic. We show here that 2H MR spectroscopy and spectroscopic imaging measurements of [2,3-2H2]malate production from [2,3-2H2]fumarate can be used to detect tumor cell death in vivo via the production of labeled malate. Production of [2,3-2H2]malate, following injection of [2,3-2H2]fumarate (1 g/kg) into tumor-bearing mice, was measured in a murine lymphoma (EL4) treated with etoposide, and in human breast (MDA-MB-231) and colorectal (Colo205) xenografts treated with a TRAILR2 agonist, using surface-coil localized 2H MR spectroscopy at 7 T. Malate production was also imaged in EL4 tumors using a fast 2H chemical shift imaging sequence. The malate/fumarate ratio increased from 0.016 ± 0.02 to 0.16 ± 0.14 in EL4 tumors 48 h after drug treatment (P = 0.0024, n = 3), and from 0.019 ± 0.03 to 0.25 ± 0.23 in MDA-MB-231 tumors (P = 0.0001, n = 5) and from 0.016 ± 0.04 to 0.28 ± 0.26 in Colo205 tumors (P = 0.0002, n = 5) 24 h after drug treatment. These increases were correlated with increased levels of cell death measured in excised tumor sections obtained immediately after imaging. 2H MR measurements of [2,3-2H2]malate production from [2,3-2H2]fumarate provide a potentially less expensive and more sensitive method for detecting cell death in vivo than 13C MR measurements of hyperpolarized [1,4-13C2]fumarate metabolism, which have been used previously for this purpose.

Currently, the response of solid tumors to treatment is assessed mainly on the basis of changes in tumor size [Response Evaluation Criteria in Solid Tumors (RECIST) (1)]. However, changes in size may take weeks to appear after the initiation of treatment and, in some cases, may not appear at all, for example, in the case of treatments that inhibit tumor growth but do not result in tumor regression (2, 3). Changes in metabolism can give an earlier indication of treatment response, for example, assessment of glycolytic activity using positron emission tomography (PET) measurements of 2-[18F]-fluoro-2-deoxy D-glucose uptake (FDG-PET). PET Response Criteria in Solid Tumors (PERCIST) was introduced as a potentially more sensitive method of assessing treatment response when compared to assessment based on changes in tumor size alone (4), particularly with therapies that stabilize disease. Imaging with hyperpolarized [1-13C]pyruvate, like FDG-PET, can also be used to detect drug target engagement as well as subsequent tumor cell death (5, 6). We have shown recently that imaging hyperpolarized [1-13C]pyruvate metabolism can be more sensitive than FDG-PET in detecting reductions in glycolytic flux associated with tumor cell death posttreatment (7).While metabolic imaging may indicate drug target engagement, and in some cases tumor cell death, there is a need for imaging methods that detect tumor cell death more directly posttreatment and that can give an indication of longer-term treatment outcomes (8). Fumarate is hydrated in the reaction catalyzed by the intracellular enzyme fumarase to produce malate. Previous 13C magnetic resonance spectroscopic imaging (MRSI) studies with hyperpolarized [1,4-13C2]fumarate in tumor models and in models of myocardial infarction and acute kidney necrosis (916) have demonstrated that the production of labeled malate can be used to image cell death in vivo. The increased production of malate was attributed to loss of the plasma membrane permeability barrier in necrotic cells and increased access of hyperpolarized [1,4-13C2]fumarate to fumarase. However, imaging with hyperpolarized 13C-labeled substrates is limited both by the transient nature of the hyperpolarization, which restricts its application to relatively fast metabolic processes, and the requirement for relatively large amounts of 13C-labeled compounds and access to clinical hyperpolarizers, which are expensive. The recent demonstration by De Feyter et al. that 2H MRSI can be used to image the metabolism of 2H-labeled substrates in vivo, including in human subjects (17), has provided a potentially lower-cost alternative for clinical metabolic imaging. The relatively low sensitivity of 2H detection is compensated by its very short T1, which means that signal can be acquired rapidly without saturation. The main limitation is the narrow frequency range, which requires the use of relatively high magnetic field strengths. Nevertheless, by collecting a series of rapidly acquired images, the technique is capable of generating quantitative images of metabolic flux (18). We show here that 2H magnetic resonance spectroscopy (MRS) and MRSI measurements of [2,3-2H2]fumarate conversion to 2H-labeled malate can be used to detect tumor cell death in vivo and that this is potentially a more sensitive method for detecting cell death than 13C MRSI with either hyperpolarized [1-13C]pyruvate or [1,4-13C2]fumarate.  相似文献   

11.
The endogenous gibberellins (GAs) were examined from young vegetative shoots of the dominant mutant, Dwarf-8, a GA-nonresponder, and normal maize; GA44, GA17, GA19, GA20, GA29, GA1, and GA8, members of the early-13-hydroxylation pathway, were identified from both kinds of shoots by full-scan mass spectra and Kovats retention indices. In addition, we report the identification of 3-epi-GA1, GA3, GA4, GA5, GA7, GA9, GA12, GA15, GA24, GA34, and GA53 by using the same criteria. [1,7,12,18-14C4]GA53 and -GA44, [17-2H2]GA19, and [17-13C,3H2]GA20, -GA29, -GA1, -GA8, and -GA5 were used as internal standards to determine the endogenous levels of these GAs by measurement of isotope dilution, using capillary gas chromatography and selected ion monitoring. Shoots of Dwarf-8 accumulate relatively high levels of GA20, GA1, and GA8. The accumulation of GA1 appears to be related to gene dosage. Since Dwarf-8 contains the same pattern of GAs as normals (including GA1 and GA3), the genetic control point probably lies after GA1 (and GA3). Thus Dwarf-8 may be a GA receptor mutant or a mutant that controls a product downstream from the binding of the bioactive GA to a receptor.  相似文献   

12.
The synthetic analog approach has been applied to a clarification of the active sites of 2Fe-2S* proteins. The compound (Et4N)2[FeS(SCH2)2C6H4]2, derived from o-xylyl-α,α′-dithiol, has been prepared and its structure has been determined by x-ray diffraction. The centrosymmetric anion contains two tetrahedrally coordinated ferric ions bridged by two sulfide ions and separated by 2.70 Å. Comparison of electronic, Mössbauer, and proton magnetic resonance spectra and magnetic susceptibility of the anion with the corresponding properties of the oxidized forms of the proteins reveals significant degrees of similarity. The anion also exhibits the essential redox capacity of the proteins. We conclude that [FeS(SCH2)2C6H4]22- possesses the same total oxidation level and electronic configuration as the active sites of the oxidized proteins, and that its structure provides a feasible representation of the minimal structure of the active site. [FeS(SCH2)2C6H4]22- is thus the first well-defined synthetic analog of the active sites of two-iron ferredoxins.  相似文献   

13.
Herein, we present dicarboxylate platinum(II) complexes of the general formula [Pt(mal)(DMSO)(L)] and [Pt(CBDC)(DMSO)(L)], where L is dbtp 5,7-ditertbutyl-1,2,4-triazolo[1,5-a]pyrimidine) or ibmtp (7-isobutyl-5-methyl-1,2,4- triazolo[1,5-a]pyrimidine), as prospective prodrugs. The platinum(II) complexes were synthesized in a one-pot reaction between cis-[PtCl2(DMSO)2], silver malonate or silver cyclobutane-1,1-dicarboxylate and triazolopyrimidines. All platinum(II) compounds were characterized by FT-IR, and 1H, 13C, 15N and 195Pt NMR; and their square planar geometries with one monodentate N(3)-bonded 5,7-disubstituted-1,2,4-triazolo[1,5-a]pyrimidine, one S-bonded molecule of dimethyl sulfoxide and one O,O-chelating malonato (1, 2) or O,O-chelating cyclobutane-1,1-dicarboxylato (3, 4) was determined. Additionally, [Pt(CBDC)(dbtp)(DMSO)] (3) exhibited (i) substantial in vitro cytotoxicity against the lung adenocarcinoma epithelial cell line (A549) (IC50 = 5.00 µM) and the cisplatin-resistant human ductal breast epithelial tumor cell line (T47D) (IC50 = 6.60 µM); and (ii) definitely exhibited low toxicity against normal murine embryonic fibroblast cells (BALB/3T3).  相似文献   

14.
Reactive oxygen species (ROS) and intracellular Ca2+ overload play key roles in myocardial ischemia-reperfusion (IR) injury but the relationships among ROS, Ca2+ overload and LV mechanical dysfunction remain unclear. We tested the hypothesis that H2O2 impairs LV function by causing Ca2+ overload by increasing late sodium current (INa), similar to Sea Anemone Toxin II (ATX-II). Diastolic and systolic Ca2+ concentrations (d[Ca2+]i and s[Ca2+]i) were measured by indo-1 fluorescence simultaneously with LV work in isolated working rat hearts. H2O2 (100 μM, 30 min) increased d[Ca2+]i and s[Ca2+]i. LV work increased transiently then declined to 32% of baseline before recovering to 70%. ATX-II (12 nM, 30 min) caused greater increases in d[Ca2+]i and s[Ca2+]i. LV work increased transiently before declining gradually to 17%. Ouabain (80 μM) exerted similar effects to ATX-II. Late INa inhibitors, lidocaine (10 μM) or R56865 (2 μM), reduced effects of ATX-II on [Ca2+]i and LV function, but did not alter effects of H2O2. The antioxidant, N-(2-mercaptopropionyl)glycine (MPG, 1 mM) prevented H2O2-induced LV dysfunction, but did not alter [Ca2+]i. Paradoxically, further increases in [Ca2+]i by ATX-II or ouabain, given 10 min after H2O2, improved function. The failure of late INa inhibitors to prevent H2O2-induced LV dysfunction, and the ability of MPG to prevent H2O2-induced LV dysfunction independent of changes in [Ca2+]i indicate that impaired contractility is not due to Ca2+ overload. The ability of further increases in [Ca2+]i to reverse H2O2-induced LV dysfunction suggests that Ca2+ desensitization is the predominant mechanism of ROS-induced contractile dysfunction.  相似文献   

15.
The influence of the substitution pattern in ferrocenyl α-thienyl thioketone used as a proligand in complexation reactions with Fe3(CO)12 was investigated. As a result, two new sulfur–iron complexes, considered [FeFe]-hydrogenase mimics, were obtained and characterized by spectroscopic techniques (1H, 13C{1H} NMR, IR, MS), as well as by elemental analysis and X-ray single crystal diffraction methods. The electrochemical properties of both complexes were studied and compared using cyclic voltammetry in the absence and in presence of acetic acid as a proton source. The performed measurements demonstrated that both complexes can catalyze the reduction of protons to molecular hydrogen H2. Moreover, the obtained results showed that the presence of the ferrocene moiety at the backbone of the linker of both complexes improved the stability of the reduced species.  相似文献   

16.
Reactive oxygen species (ROS) and intracellular Ca2+ overload play key roles in myocardial ischemia–reperfusion (IR) injury but the relationships among ROS, Ca2+ overload and LV mechanical dysfunction remain unclear. We tested the hypothesis that H2O2 impairs LV function by causing Ca2+ overload by increasing late sodium current (INa), similar to Sea Anemone Toxin II (ATX-II). Diastolic and systolic Ca2+ concentrations (d[Ca2+]i and s[Ca2+]i) were measured by indo-1 fluorescence simultaneously with LV work in isolated working rat hearts. H2O2 (100 μM, 30 min) increased d[Ca2+]i and s[Ca2+]i. LV work increased transiently then declined to 32% of baseline before recovering to 70%. ATX-II (12 nM, 30 min) caused greater increases in d[Ca2+]i and s[Ca2+]i. LV work increased transiently before declining gradually to 17%. Ouabain (80 μM) exerted similar effects to ATX-II. Late INa inhibitors, lidocaine (10 μM) or R56865 (2 μM), reduced effects of ATX-II on [Ca2+]i and LV function, but did not alter effects of H2O2. The antioxidant, N-(2-mercaptopropionyl)glycine (MPG, 1 mM) prevented H2O2-induced LV dysfunction, but did not alter [Ca2+]i. Paradoxically, further increases in [Ca2+]i by ATX-II or ouabain, given 10 min after H2O2, improved function. The failure of late INa inhibitors to prevent H2O2-induced LV dysfunction, and the ability of MPG to prevent H2O2-induced LV dysfunction independent of changes in [Ca2+]i indicate that impaired contractility is not due to Ca2+ overload. The ability of further increases in [Ca2+]i to reverse H2O2-induced LV dysfunction suggests that Ca2+ desensitization is the predominant mechanism of ROS-induced contractile dysfunction.  相似文献   

17.
Summary Bisphosphonates are compounds with a high affinity for bone and other calcified tissues. They inhibit tumor-induced bone destruction and the associated hypercalcemia by hindering the activity of the osteoclasts. Owing to a long biological half-life of bisphosphonates in the bone, a treatment using a prophylactic regimen seems possible. This paper summarizes preclinical studies with the bisphosphonate 3-amino-1-hydroxypropylidene-1, 1-diphosphonic acid and two methyl derivatives; S-N,N-dimethylamino-1-hydroxypropylidene-1,1-diphosphonic acid and 4-N,N-dimetyhlamino-1-hydroxybutylidene-1, 1-diphosphonic acid with respect to their bone-protecting activity in therapy as well as in prophylaxis. To find substances that are useful for the treatment of primary tumor, as well as bone metastasis, we synthesized and testedcis-diammine[nitrilotris(methylphosphonato) (2-)-O 1 ,N1]platinum(II) andcis-diammine{[bis-(phosphonatomethyl)amino]acetato(2-)-O1, N1} platinum(II), which contain both an osteotropic and an antineoplastic moiety. Experiments were carried out: (a) in the intratibial transplanted Walker carcinosarcoma 256B of the rat, which mimics osteolytic bone metastasis, and (b) in the transplantable osteosarcoma of the rat, which shows a histology and metastatic pattern similar to that found in man. These investigations indicate that it is possible to effect adjuvant therapy of bone metastases by combination of two compounds with different properties into one structure without losing the therapeutic characteristics of the parent compounds. They thus provide evidence that it may be possible to design compounds well suited for the therapeutic or prophylactic treatment of bone-related malignancies.Abbreviations NMR nuclear magnetic resonance - WCS Walker carcinosarcoma 256B - APD 3-amino-1-hydroxypropylidene-1,1-diphosphonic acid - 3Me2APD theN-dimethyl derivative of APD - 4Me2ABD A-N, N-dimethylamino-1-hydroxybutylidene-1,1-diphosphonic acid - AMDP cis-diammine[nitrilotris(methylphosphonato)(2-)-O 1,N 1]platinum(II) - DBP cis-diammine{[bis(phosphonatomethyl)amino]acetato(2-)-O 1,N 1}platmum(II)  相似文献   

18.
Ca2+ signaling regulates cell function. This is subject to modulation by H+ ions that are universal end-products of metabolism. Due to slow diffusion and common buffers, changes in cytoplasmic [Ca2+] ([Ca2+]i) or [H+] ([H+]i) can become compartmentalized, leading potentially to complex spatial Ca2+/H+ coupling. This was studied by fluorescence imaging of cardiac myocytes. An increase in [H+]i, produced by superfusion of acetate (salt of membrane-permeant weak acid), evoked a [Ca2+]i rise, independent of sarcolemmal Ca2+ influx or release from mitochondria, sarcoplasmic reticulum, or acidic stores. Photolytic H+ uncaging from 2-nitrobenzaldehyde also raised [Ca2+]i, and the yield was reduced following inhibition of glycolysis or mitochondrial respiration. H+ uncaging into buffer mixtures in vitro demonstrated that Ca2+ unloading from proteins, histidyl dipeptides (HDPs; e.g., carnosine), and ATP can underlie the H+-evoked [Ca2+]i rise. Raising [H+]i tonically at one end of a myocyte evoked a local [Ca2+]i rise in the acidic microdomain, which did not dissipate. The result is consistent with uphill Ca2+ transport into the acidic zone via Ca2+/H+ exchange on diffusible HDPs and ATP molecules, energized by the [H+]i gradient. Ca2+ recruitment to a localized acid microdomain was greatly reduced during intracellular Mg2+ overload or by ATP depletion, maneuvers that reduce the Ca2+-carrying capacity of HDPs. Cytoplasmic HDPs and ATP underlie spatial Ca2+/H+ coupling in the cardiac myocyte by providing ion exchange and transport on common buffer sites. Given the abundance of cellular HDPs and ATP, spatial Ca2+/H+ coupling is likely to be of general importance in cell signaling.Most cells are exquisitely responsive to calcium (Ca2+) (1) and hydrogen (H+) ions (i.e., pH) (2). In cardiac myocytes, Ca2+ ions trigger contraction and control growth and development (3), whereas H+ ions, which are generated or consumed metabolically, are potent modulators of essentially all biological processes (4). By acting on Ca2+-handling proteins directly or via other molecules, H+ ions exert both inhibitory and excitatory effects on Ca2+ signaling. For example, in the ventricular myocyte, H+ ions can reduce Ca2+ release from sarcoplasmic reticulum (SR) stores, through inhibition of the SR Ca2+ ATPase (SERCA) pump and ryanodine receptor (RyR) Ca2+ channels (5, 6). In contrast, H+ ions can enhance SR Ca2+ release by stimulating sarcolemmal Na+/H+ exchange (NHE), which raises intracellular [Na+] and reduces the driving force for Ca2+ extrusion on Na+/Ca2+ exchange (NCX), leading to cellular retention of Ca2+ (7, 8). Ca2+ signaling is thus subservient to pH.Cytoplasmic Ca2+ and H+ ions bind avidly to buffer molecules, such that <1% of all Ca2+ ions and <0.001% of all H+ ions are free. Some of these buffers bind H+ and Ca2+ ions competitively, and this has been proposed to be one mechanism underlying cytoplasmic Ca2+/H+ coupling (9). Reversible binding to buffers greatly reduces the effective mobility of Ca2+ and H+ ions in cytoplasm (10, 11) and can allow for highly compartmentalized ionic microdomains, and hence a spatially heterogeneous regulation of cell function. In cardiac myocytes under resting (diastolic) conditions, the cytoplasm-averaged concentration of free [Ca2+] ([Ca2+]i) and [H+] ([H+]i) ions is kept near 10−7 M by membrane transporter proteins. Thus, [H+]i is regulated by the balance of flux among acid-extruding and acid-loading transporter proteins at the sarcolemma [e.g., NHE and Cl/OH (CHE) exchangers, respectively] (4). Similarly, the activity of SERCA and NCX proteins returns [Ca2+]i to its diastolic level after evoked signaling events (3, 12). Despite these regulatory mechanisms, cytoplasmic gradients of [H+]i and [Ca2+]i do occur in myocytes and are an important part of their physiology. Gradients arise from local differences in transmembrane fluxes that alter [H+]i or [Ca2+]i. For example, spatial [H+]i gradients are produced when NHE transporters, expressed mainly at the intercalated disk region, are activated (4, 13) or when membrane-permeant weak acids, such as CO2, are presented locally (14). Similarly, release of Ca2+ through a cluster of RyR channels in the SR produces [Ca2+]i nonuniformity in the form of Ca2+ sparks (15). Given the propensity of cytoplasm to develop ionic gradients, it is important to understand their underlying mechanism and functional role.The present work demonstrates a distinct form of spatial interaction between Ca2+ and H+ ions. We show that cytoplasmic [H+] gradients can produce stable [Ca2+]i gradients, and vice versa, and that this interaction is mediated by low-molecular-weight (mobile) buffers with affinity for both ions. We demonstrate that the diffusive counterflux of H+ and Ca2+ bound to these buffers comprises a cytoplasmic Ca2+/H+ exchanger. This acts like a “pump” without a membrane, which can, for instance, recruit Ca2+ to acidic cellular microdomains. Cytoplasmic Ca2+/H+ exchange adds a spatial paradigm to our understanding of Ca2+ and H+ ion signaling.  相似文献   

19.
Summary Intact cultured fibroblasts from patients with deficiency of long-chain 3-hydroxyacyl-CoA dehydrogenase release3H2O from [9,10-3H]myristic acid and [9,10-3H]palmitic acid more slowly than normal. The ratio of activity (palmitate/myristate) is also low and the expression (rate with palmitate)2/(rate with myristate) gives good differentiation between affected and unaffected cells. In some patients who have shown hydroxydicarboxylic aciduria when unwell there is reduced3H2O production from [9,10-3H]myristic and [9,10-3H]palmitic acids by intact cultured fibroblasts but normal 3-hydroxyacyl-CoA dehydrogenase activities in disrupted cells. The palmitate/myristate ratio is higher than in long-chain 3-hydroxyacyl-CoA dehydrogenase deficiency. The basic defect in these patients is still unknown but it is suggested that caution be used over the administration of medium-chain triglyceride.  相似文献   

20.
We present a model of photosynthetic water oxidation that utilizes the property of higher-valent Mn ions in two different environments and the characteristic function of redox-active ligands to explain all known aspects of electron transfer from H2O to Z, the electron donor to P680, the photosystem II reaction center chlorophyll a. There are two major features of this model. (i) The four functional Mn atoms are divided into two groups of two Mn each: [Mn] complexes in a hydrophobic cavity in the intrinsic 34-kDa protein; and (Mn) complexes on the hydrophilic surface of the extrinsic 33-kDa protein. The oxidation of H2O is carried out by two [Mn] complexes, and the protons are transferred from a [Mn] complex to a (Mn) complex along the hydrogen bond between their respective ligand H2O molecules. (ii) Each of the two [Mn] ions binds one redox-active ligand (RAL), such as a quinone (alternatively, an aromatic amino acid residue). Electron transfer occurs from the reduced RAL to the oxidized Z. When the experimental data concerning atomic structure of the water-oxidizing center (WOC), electron transfer between the WOC and Z, the electronic structure of the WOC, the proton-release pattern, and the effect of Cl- are compared with the predictions of the model, satisfactory qualitative and, in many instances, quantitative agreements are obtained. In particular, this model clarifies the origin of the observed absorption-difference spectra, which have the same pattern in all S-state transitions, and of the effect of Cl--depletion on the S states.  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