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.
Biogenic polyamines, putrescine, spermidine, and spermine, are ubiquitous cellular cations and exert multiple biological functions. Polyamine analogues mimic biogenic polyamines at macromolecular level but are unable to substitute for natural polyamines and maintain cell proliferation, indicating biomedical applications. The mechanistic differences in DNA binding mode between natural and synthetic polyamines have not been explored. The aim of this study was to examine the interaction of calf thymus DNA with three polyamine analogues, 1,11-diamino-4,8-diazaundecane (333), 3,7,11,15-tetrazaheptadecane x 4 HCl (BE-333), and 3,7,11,15,19-pentazahenicosane x 5 HCl (BE-3333), using FTIR, UV-visible, and CD spectroscopy. Polyamine analogues bind with guanine and backbone PO2 group as major targets in DNA, whereas biogenic polyamines bind to major and minor grooves as well as to phosphate groups. Weaker interaction with DNA was observed for analogues with respect to biogenic polyamines, with K(333) = 1.90 (+/-0.5) x 10(4) M(-1), K(BE-333) = 6.4 (+/-1.7) x 10(4) M(-1), K(BE-3333) = 4.7 (+/-1.4) x 10(4) M(-1) compared to K(Spm) = 2.3 (+/-1.1) x 10(5) M(-1), K(Spd) = 1.4 (+/-0.6) x 10(5) M(-1), and K(Put) = 1.02 (+/-0.5) x 10(5) M(-1). A partial B- to A-DNA transition was also provoked by analogues. These data suggest distinct differences in the binding of natural and synthetic polyamines with DNA.  相似文献   

2.
Polyamine analogues show antitumor activity in experimental models, and their ability to alter activity of cytotoxic chemotherapeutic agents in breast cancer is well documented. Association of polyamines with nucleic acids and protein is included in their mechanism of action. The aim of this study was to examine the interaction of human serum albumin (HSA) with several polyamine analogues, such as 1,11-diamino-4,8-diazaundecane (333), 3,7,11,15-tetrazaheptadecane.4HCl (BE-333), and 3,7,11,15,19-pentazahenicosane.5HCl (BE-3333), in aqueous solution at physiological conditions using a constant protein concentration and various polyamine contents (microM to mM). FTIR, UV-visible, and CD spectroscopic methods were used to determine the polyamine binding mode and the effects of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind nonspecifically (H-bonding) via polypeptide polar groups with binding constants of K333 = 9.30 x 10(3) M(-1), KBE-333 = 5.63 x 10(2) M(-1), and KBE-3333 = 3.66 x 10(2) M(-1). The protein secondary structure showed major alterations with a reduction of alpha-helix from 55% (free protein) to 43-50% and an increase of beta-sheet from 17% (free protein) to 29-36% in the 333, BE-333, and BE-3333 complexes, indicating partial protein unfolding upon polyamine interaction. HSA structure was less perturbed by polyamine analogues compared to those of the biogenic polyamines.  相似文献   

3.
The bindings of biogenic polyamines spermine (spm), spermidine (spmd) and synthetic polyamines 3,7,11,15-tetrazaheptadecane·4HCl (BE-333) and 3,7,11,15,19-pentazahenicosane·5HCl (BE-3333) with β-lactoglobulin (β-LG) were determined in aqueous solution. FTIR, UV-vis, CD and fluorescence spectroscopic methods as well as molecular modeling were used to determine the polyamine binding sites and the effect of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind β-LG via both hydrophilic and hydrophobic contacts. Stronger polyamine-protein complexes formed with synthetic polyamines than biogenic polyamines, with overall binding constants of Kspm-β-LG = 3.2(±0.6) × 104 M−1, Kspmd-β-LG = 1.8(±0.5) × 104 M−1, KBE-333-β-LG = 5.8(±0.3) × 104 M−1 and KBE-3333-β-LG = 6.2(±0.05) × 104 M−1. Molecular modeling showed the participation of several amino acids in the polyamine complexes with the following order of polyamine-protein binding affinity: BE-3333 > BE-333 > spermine > spermidine, which correlates with their positively charged amino group content. Alteration of protein conformation was observed with a reduction of β-sheet from 57% (free protein) to 55-51%, and a major increase of turn structure from 13% (free protein) to ∼21% in the polyamine-β-LG complexes, indicating a partial protein unfolding.  相似文献   

4.
Major attention has been focused on dendrimer-DNA complexes because of their applications in gene delivery systems. Dendrimers are also used to transport miRNA and siRNA in vitro. We examine the interaction of tRNA with several dendrimers of different compositions, mPEG-PAMAM (G3), mPEG-PAMAM (G4), and PAMAM (G4) under physiological conditions using constant tRNA concentration and various dendrimer contents. FTIR, UV-visible, and CD spectroscopic methods as well as atomic force microscopy (AFM) were used to analyze the macromolecule binding mode, the binding constant, and the effects of dendrimer complexation on RNA stability, aggregation, particle formation, and conformation. Structural analysis showed that dendrimer-tRNA complexation occurred via RNA bases and the backbone phosphate group with both hydrophilic and hydrophobic contacts. The overall binding constants of K(mPEG-G3) = 7.6 (± 0.9) × 10(3) M(-1), K(mPEG-G4) = 1.5 (± 0.40) × 10(4) M(-1), and K(PAMAM-G4) = 5.3 (± 0.60) × 10(4) M(-1) show stronger polymer-RNA complexation by PAMAM-G4 than pegylated dendrimers. RNA remains in the A-family structure, whereas biopolymer aggregation and particle formation occurred at high polymer concentrations.  相似文献   

5.
We describe a method for the profiling of polyamines, N-acetylated polyamines and the polyamine analogues N1,N11-bis(ethyl)norspermine (BE-3-3-3) and 1,19-bis(ethylamino)-5,10,15-triazanonadecane (BE-4-4-4-4) in L1210 murine leukaemia cells by capillary gas chromatography with nitrogen-phosphorus detection. The method makes use of four internal standards. Prepurification comprises deproteinization, isolation with Sep-Pak silica at pH 9.0, conversion to heptafluorobutyryl derivatives and postderivatization organic fluid extraction. Within- and between-series precisions (given as C.V.s) for analysis of 1–2×106 cells were: putrescine 5.5 and 29.4%; spermidine 1.6 and 7.1%; and spermine 3.2 and 7.6%, respectively. Recoveries relative to the respective internal standards, were in the 70.6–104.7% range. Accuracy and precision of measurements of BE-4-4-4-4 can probably be improved by the introduction of a separate pentamine internal standard. We conclude that the method can be used for studying the effect of BE-3-3-3 and BE-4-4-4-4, and possibly their metabolites, on polyamine homeostasis (biosynthesis, retroconversion, transport, terminal catabolism) and polyamine function.  相似文献   

6.
Dendrimers are unique synthetic macromolecules of nanometer dimensions with a highly branched structure and globular shape. Among dendrimers, polyamidoamine (PAMAM) have received most attention as potential transfection agents for gene delivery, because these macromolecules bind DNA at physiological pH. The aim of this study was to examine the interaction of calf-thymus DNA with several dendrimers of different compositions, such as mPEG-PAMAM (G3), mPEG-PAMAM (G4), and PAMAM (G4) at physiological conditions, using constant DNA concentration and various dendrimer contents. FTIR, UV-visible, and CD spectroscopic methods, as well as atomic force microscopy (AFM), were used to analyze the macromolecule binding mode, the binding constant, and the effects of dendrimer complexation on DNA stability, aggregation, condensation, and conformation. Structural analysis showed a strong dendrimer-DNA interaction via major and minor grooves and the backbone phosphate group with overall binding constants of K(mPEG-G3) = 1.5 (±0.5) × 10(3) M(-1), K(mPEG-G4) = 3.4 (±0.80) × 10(3) M(-1), and K(PAMAM-G4) = 8.2 (±0.90) × 10(4) M(-1). The order of stability of polymer-DNA complexation is PAMAM-G4 > mPEG-G4 > mPEG-G3. Both hydrophilic and hydrophobic interactions were observed for dendrimer-DNA complexes. DNA remained in the B-family structure, while biopolymer particle formation and condensation occurred at high dendrimer concentrations.  相似文献   

7.
Exposure of rat L6 cells in culture to exogenous polyamines led to a very large increase in the activity of spermidine/spermine N1-acetyltransferase. Spermine was more potent than spermidine in bringing about this increase, but in both cases the elevated acetyltransferase activity increased the cellular conversion of spermidine into putrescine. The N1-acetyltransferase turned over very rapidly in the L6 cells, with a half-life of 9 min after spermidine and 18 min after spermine. A wide variety of synthetic polyamine analogues also brought about a substantial induction of spermidine/spermine N1-acetyltransferase activity. These included sym-norspermidine, sym-norspermine, sym-homospermidine, N4-substituted spermidine derivatives, 1,3,6-triaminohexane, 1,4,7-triaminoheptane and deoxyspergualin, which were comparable with spermidine in their potency, and N1N8-bis(ethyl)spermidine, N1N9-bis(ethyl)homospermidine, methylglyoxal bis(guanylhydrazone), ethylglyoxal bis(guanylhydrazone) and 1,1'-[(methylethanediylidene)dinitrilo]bis(3-amino-guanidine ), which were even more active than spermidine. It is suggested that these polyamine analogues may bring about a decrease in cellular polyamines not only by inhibiting biosynthesis but by stimulating the degradation of spermidine into putrescine.  相似文献   

8.
In vitro interactions between polyamines and pectic substances   总被引:1,自引:0,他引:1  
Putrescine, spermidine and spermine induce a decrease in the pH value of 1 mM polygalacturonic acid or pectin solutions; spermidine and spermine also cause the precipitation of the polymers. The association constants between polyamines and polygalacturonic acid were in the order of 10(5) for putrescine and spermidine, and 10(6) for spermine. The number of galacturonic units per binding sites are proportional to the number of positive charges on the polyamine molecule. Low affinity binding sites appear at high polyamine concentrations. Calcium ions seem to compete weakly with spermine by lowering the association constant 4- to 6-fold. Two natural pectins tested, showed that methylation of the carboxylic groups influences only the number of galacturonic units per site but not the association constant.  相似文献   

9.
We have generated mouse embryonic stem cells with targeted disruption of spermidine/spermine N(1)-acetyltransferase (SSAT) gene. The targeted cells did not contain any inducible SSAT activity, and the SSAT protein was not present. The SSAT-deficient cells proliferated normally and appeared to maintain otherwise similar polyamine pools as did the wild-type cells, with the possible exception of constantly elevated (about 30%) cellular spermidine. As expected, the mutated cells were significantly more resistant toward the growth-inhibitory action of polyamine analogues, such as N(1),N(11)-diethylnorspermine. However, this resistance was not directly attributable to cellular depletion of the higher polyamines spermidine and spermine, as the analogue depleted the polyamine pools almost equally effectively in both wild-type and SSAT-deficient cells. Tracer experiments with [C(14)]-labeled spermidine revealed that SSAT activity is essential for the back-conversion of spermidine to putrescine as radioactive N(1)-acetylspermidine and putrescine were readily detectable in N(1),N(11)-diethylnorspermine-exposed wild-type cells but not in SSAT-deficient cells. Similar experiments with [C(14)]spermine indicated that the latter polyamine was converted to spermidine in both cell lines and, unexpectedly, more effectively in the targeted cells than in the parental cells. This back-conversion was only partly inhibited by MDL72527, an inhibitor of polyamine oxidase. These results indicated that SSAT does not play a major role in the maintenance of polyamine homeostasis, and the toxicity exerted by polyamine analogues is largely not based on SSAT-induced depletion of the natural polyamines. Moreover, embryonic stem cells appear to operate an SSAT-independent system for the back-conversion of spermine to spermidine.  相似文献   

10.
The presence of several high affinity binding sites on human serum albumin (HSA) makes it a possible target for many organic and inorganic molecules. Organic polyamines are widely distributed in living cells and their biological roles have been associated with their physical and chemical interactions with proteins, nucleic acids, and lipids. This study is designed to examine the effects of spermine, spermidine, putrescine, and cobalt [Co(III)]-hexamine cations on the solution structure of HSA using Fourier transform IR, UV-visible, and circular dichroism (CD) spectroscopic methods. The spectroscopic results show that polyamine cations are located along the polypeptide chains with no specific interaction. The order of perturbations is associated with the number of positive charges of the polyamine cation: spermine > Co(III)-hexamine > spermidine > putrescine. The overall binding constants are 1.7 x 10(4), 1.1 x 10(4), 5.4 x 10(3), and 3.9 x 10(3)M(-1), respectively. The protein conformation is altered (IR and CD data) with reductions of alpha helices from 60 to 55% for free HSA to 50-40% and with increases of beta structures from 22 to 15% for free HSA to 33-23% in the presence of polyamine cations.  相似文献   

11.
12.
Aminooxy analogues of spermidine, 1-aminooxy-3-N-[3-aminopropyl]- aminopropane (AP-APA) and N-[2-aminooxyethyl]-1,4-diaminobutane (AOE-PU), were tested as substrates or inhibitors of the enzymes involved in methionine and polyamine metabolism. Both compounds were good competitive inhibitors and poor substrates of spermine synthase, good substrates of cytosolic polyamine acetyltransferase, inactivators of S-adenosylmethionine decarboxylase and inhibitors of ornithine decarboxylase. AP-APA and AOE-PU showed K1-values of 1.5 and 186 microM as inhibitors of purified spermine synthase, and Km-values of 1.4 and 2.1 mM as substrates of the crude hepatic polyamine acetyltransferase activity. AP-APA was more potent than AOE-PU in crude enzyme preparations. Neither drug had any significant effect at 1 mM concentration on the activities of spermidine synthase, methionine adenosyltransferase, S-adenosylhomocysteine hydrolase, and methylthioadenosine phosphorylase. The results suggest that compounds of this type are valuable tools in unraveling the physiology of polyamines.  相似文献   

13.
To estimate the polyamine distribution in bovine lymphocytes and rat liver, the binding constants (K) for DNA, RNA, phospholipid, and ATP were determined under the conditions of 10 mM Tris-HCl, pH 7.5, 2 mM Mg2+, and 150 mM K+. The binding constants of spermine for calf thymus DNA, Escherichia coli 16 S rRNA, phospholipid in rat liver microsomes and ATP were 1.15 x 10(2), 6.69 x 10(2), 2.22 x 10(2), and 5.95 x 10(2) M-1, respectively. From these binding constants and experimentally determined cellular concentrations of macromolecules, ATP, and polyamines, spermine distribution in the cells was estimated. In bovine lymphocytes, the mols of spermine bound to DNA, RNA, phospholipid, and ATP were 0.79, 3.7, 0.23, and 4.3 per 100 mol of phosphate of macromolecules or ATP, respectively. In rat liver, they were 0.19, 1.0, 0.05, and 0.97/100 mol of phosphate of macromolecules or ATP, respectively. The binding constants of spermidine for macromolecules and ATP were smaller than those of spermine, but a similar tendency was observed with spermidine distribution among macromolecules and ATP in the above two cells. The amount of polyamine bound to DNA and phospholipid was significantly lower than that to RNA. When either the Mg2+ or K+ concentration increased, the amount of free spermine and that bound to RNA and ATP increased, but the amount of spermine bound to DNA and phospholipid decreased. The results indicate that most polyamines exist as a polyamine-RNA complex in cells. Under the conditions that globin synthesis is stimulated by spermine in a rabbit reticulocyte cell-free system, the amount of spermine bound to RNA was very close to the value estimated in the cells.  相似文献   

14.
Abstract: Spermine and other polyamines both stimulate and inhibit N -methyl- d -aspartate receptor function, probably by interacting with two separate sites. To characterize these two actions, the effect of spermine on the binding kinetics of the channel blocker [3H]dizocilpine was studied in the presence of glutamate and glycine. Low concentrations (10 µ M ) of spermine increased the association and dissociation rates without modifying equilibrium binding, indicating that spermine increases the accessibility of [3H]dizocilpine to the channel by interacting with a high-affinity, stimulatory site. At higher concentrations (1 m M ), spermine markedly decreased equilibrium [3H]-dizocilpine binding by decreasing both affinity and B max, indicating that spermine allosterically inhibits binding by interacting with a second, low-affinity site. The presumed polyamine antagonists arcaine, diethylenetriamine, and 1,10-diaminodecane completely inhibited equilibrium [3H]dizocilpine binding, probably by interacting with the inhibitory polyamine site or other sites, but not with the stimulatory polyamine site. Low concentrations (10 µ M ) of ifenprodil completely reversed the increase in association rate produced by spermine, whereas higher concentrations (IC50 = 123 µ M ) inhibited equilibrium binding, indicating that ifenprodil is both a potent antagonist of the stimulatory site and a low-affinity ligand of the inhibitory site. The polyamine agonists spermine, spermidine, and neomycin interacted with the inhibitory site, but produced only partial inhibition of equilibrium [3H]dizocilpine binding.  相似文献   

15.
The mechanism of inward rectification was examined in cell-attached and inside-out membrane patches from Xenopus oocytes expressing the cloned strong inward rectifier HRK1. Little or no outward current was measured in cell-attached patches. Inward currents reach their maximal value in two steps: an instantaneous phase followed by a time-dependent "activation" phase, requiring at least two exponentials to fit the time- dependent phase. After an activating pulse, the quasi-steady state current-voltage (I-V) relationship could be fit with a single Boltzmann equation (apparent gating charge, Z = 2.0 +/- 0.1, n = 3). Strong rectification and time-dependent activation were initially maintained after patch excision into high [K+] (K-INT) solution containing 1 mM EDTA, but disappeared gradually, until only a partial, slow inactivation of outward current remained. Biochemical characterization (Lopatin, A. N., E. N. Makhina, and C. G. Nichols, 1994. Nature. 372:366-396.) suggests that the active factors are naturally occurring polyamines (putrescine, spermidine, and spermine). Each polyamine causes reversible, steeply voltage-dependent rectification of HRK1 channels. Both the blocking affinity and the voltage sensitivity increased as the charge on the polyamine increased. The sum two Boltzmann functions is required to fit the spermine and spermidine steady state block. Putrescine unblock, like Mg2+ unblock, is almost instantaneous, whereas the spermine and spermidine unblocks are time dependent. Spermine and spermidine unblocks (current activation) can each be fit with single exponential functions. Time constants of unblock change e-fold every 15.0 +/- 0.7 mV (n = 3) and 33.3 +/- 6.4 mV (n = 5) for spermine and spermidine, respectively, matching the voltage sensitivity of the two time constants required to fit the activation phase in cell-attached patches. It is concluded that inward rectification in intact cells can be entirely accounted for by channel block. Putrescine and Mg2+ ions can account for instantaneous rectification; spermine and spermidine provide a slower rectification corresponding to so-called intrinsic gating of inward rectifier K channels. The structure of spermine and spermidine leads us to suggest a specific model in which the pore of the inward rectifier channel is plugged by polyamines that enter deeply into the pore and bind at sites within the membrane field. We propose a model that takes into account the linear structure of the natural polyamines and electrostatic repulsion between two molecules inside the pore. Experimentally observed instantaneous and steady state rectification of HRK1 channels as well as the time-dependent behavior of HRK1 currents are then well fit with the same set of parameters for all tested voltages and concentrations of spermine and spermidine.  相似文献   

16.
Naturally occurring polyamines putrescine, cadaverine, spermidine, and spermine are analogues of the species-specific long-chain polyamines found in diatoms. Scanning electron microscopy and energy-dispersive spectroscopy show that the reactions of a soluble Ti(IV) precursor with spermidine and spermine, but not putrescine or cadaverine, produce nanostructured irregular polyhedra of titanium oxide. At 25 degrees C, the average size of the particles formed with spermidine is 400 +/- 150 nm, and with spermine, 140 +/- 50 nm. Although the particles are X-ray amorphous at room temperature, annealing studies reveal that the particles adopt crystallinity at higher temperatures characteristic of anatase (TiO2). The major portion of the biopolyamines is not coprecipitated with the solid but is left in solution. Kinetic measurements reveal an initial fast step followed by two slower phases of reaction. At 25 degrees C, k(1obs) and k(2obs) for the reaction with spermidine are 5 x 10(-3) s(-1) and 3.6 x 10(-4) s(-1), respectively, and for spermine, 4.8 x 10(-3) s(-1) and 4.2 x 10(-4) s(-1), respectively. Taken together, the data suggest spermidine and spermine are biocatalysts for the precipitation of nanostructured titanium oxide.  相似文献   

17.
Recent studies suggest that overexpression of the polyamine-acetylating enzyme spermidine/spermine N(1)-acetyltransferase (SSAT) significantly increases metabolic flux through the polyamine pathway. The concept derives from the observation that SSAT-induced acetylation of polyamines gives rise to a compensatory increase in biosynthesis and presumably to increased flow through the pathway. Despite the strength of this deduction, the existence of heightened polyamine flux has not yet been experimentally demonstrated. Here, we use the artificial polyamine precursor 4-fluoro-ornithine to measure polyamine flux by tracking fluorine unit permeation of polyamine pools in human prostate carcinoma LNCaP cells. Conditional overexpression of SSAT was accompanied by a massive increase in intracellular and extracellular acetylated spermidine and by a 6-20-fold increase in biosynthetic enzyme activities. In the presence of 300 microM 4-fluoro-ornithine, SSAT overexpression led to the sequential appearance of fluorinated putrescine, spermidine, acetylated spermidine, and spermine. As fluorinated polyamines increased, endogenous polyamines decreased, so that the total polyamine pool size remained relatively constant. At 24 h, 56% of the spermine pool in the induced SSAT cells was fluorine-labeled compared with only 12% in uninduced cells. Thus, SSAT induction increased metabolic flux by approximately 5-fold. Flux could be interrupted by inhibition of polyamine biosynthesis but not by inhibition of polyamine oxidation. Overall, the findings are consistent with a paradigm whereby flux is initiated by SSAT acetylation of spermine and particularly spermidine followed by a marked increase in key biosynthetic enzymes. The latter sustains the flux cycle by providing a constant supply of polyamines for subsequent acetylation by SSAT. The broader metabolic implications of this futile metabolic cycling are discussed in detail.  相似文献   

18.
DL-alpha-Difluoromethylornithine (DFMO) causes polyamines of the AIDS-associated opportunistic pathogen Pneumocystis carinii to diminish 15 times more rapidly than mammalian host cells. The proposed mechanism was that, unlike mammalian cells, P. carinii is unable to regulate polyamine catabolism when synthesis is blocked. To test this, the responses of the polyamine catabolic enzymes spermidine/spermine acetyltransferase (SSAT) and polyamine oxidase (PAO) were determined using a new high-performance liquid chromatography assay to measure the products of these enzymes. The specific activities in untreated Pneumocystis carinii were 1.78 +/- 0.5 pmol min(-1) mg protein(-1) for SSAT, similar to mammalian cells, and 6.42 +/- 0.8 pmol min(-1) mg protein(-1) for PAO, 19% of that of mammalian cells. DFMO treatment for 12 h caused reductions of only 11 and 4% in SSAT and PAO, respectively, despite polyamine reductions of 94, 96, and 90% for putrescine, spermidine, and spermine, respectively. The P. carinii SSAT K(m) value of 25 microM spermidine is 20% of that of mammalian cells, and the PAO K(m) value of 14 nM N(1)-acetylspermidine is 0.01% of that of mammalian cells. Acetylated polyamines continue to be lost from P. carinii even when exposed to DFMO. Collectively, these results support the hypothesis that P. carinii is unable to regulate polyamine catabolism.  相似文献   

19.
Biogenic polyamines putrescine, spermidine, and spermine are essential molecules for proliferation in all living organisms. Direct interaction of polyamines with nucleic acids has been proposed in the past based on a series of experimental evidences, such as precipitation, thermal denaturation, or protection. However, binding between polyamines and nucleic acids is not clearly explained. Several interaction models have also been proposed, although they do not always agree with one another. In the present work, we make use of the Raman spectroscopy to extend our knowledge about polyamine-DNA interaction. Raman spectra of highly polymerized calf-thymus DNA at different polyamine concentrations, ranging from 1 to 50 mM, have been studied for putrescine, spermidine, and spermine. Both natural and heavy water were used as solvents. Difference Raman spectra have been computed by subtracting the sum of the separated component spectra from the experimental spectra of the complexes. The analysis of the Raman data has supported the existence of structural specificities in the interactions, at least under our experimental conditions. These specificities lead to preferential bindings through the DNA minor groove for putrescine and spermidine, whereas spermine binds by the major groove. On the other hand, spermine and spermidine present interstrand interactions, whereas putrescine presents intrastrand interactions in addition to exo-groove interactions by phosphate moieties.  相似文献   

20.
The effects of alpha-difluoromethylornithine, a specific inhibitor of ornithine decarboxylase, on cell growth rate, polyamine content and the content of decarboxylated S-adenosylmethionine in SV-3T3 transformed mouse fibroblasts were studied. DL-alpha-Difluoromethylornithine at 1 mM or higher concentrations decreased the growth rate by over 90% after 2 or more days of exposure, but the cells remained viable, although quiescent for at least 9 days. Addition of 10 microM-spermidine or -spermine or 50 microM-putrescine at any time throughout this period completely reversed the inhibition of growth. Treatment with alpha-difluoromethylornithine decreased putrescine and spermidine contents by more than 98% and that of spermine by 60%, but cells exposed to exogenous polyamines did not require complete replenishment of the polyamine pools to resume growth. In fact, a virtually normal growth rate was obtained in cells lacking putrescine, having 2% of normal spermidine content and 156% of normal spermine. These results suggest that the well-known increase in putrescine and spermidine in cells stimulated for growth is not essential for this to occur and that mammalian cells can utilize spermine as their only polyamine. A substantial reversal of the growth-inhibitory effect of alpha-difluoromethylornithine was produced by a number of polyamines not normally found in mammalian cells, including the spermidine analogues aminopropylcadaverine and sym-homospermidine, which were partially converted into their respective spermine analogues by addition of an aminopropyl group within the cell. The spermine analogue sym-norspermine was also effective, but the maximal growth rate produced by these unphysiological polyamines was only 60-70% of that produced by the normal polyamines. These results indicate that spermidine and spermine have the optimal length for activation of the cellular processes critically dependent on polyamines and should help in identifying these processes. Exposure to alpha-difluoromethylornithine leads to an enormous rise in the concentration of decarboxylated S-adenosylmethionine, which reached a peak at 530-fold after 3 days of exposure and steadily declined to 140-fold after 11 days. This increase was abolished by addition of exogenous polyamines, which rapidly decreased the activity of S-adenosylmethionine decarboxylase. The increase in decarboxylated S-adenosylmethionine is unlikely to be solely responsible for the decrease to the same extent by spermine, sym-norspermidine and sym-homospermidine, which produce 97%, 16% and 60% of the control growth rate, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)  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