Cholesterol Articles and Abstracts

For medical practitioners and the general public - Cholesterol Journal Article Catalog.

Cholesterol Journal Articles



Record 2481 to 2500
First Page Previous Page Next Page Last Page
Cholesterol induces fetal rat enterocyte death in culture
Gazzola, J., E. P. Silva, et al. (2004), Braz J Med Biol Res 37(7): 1087-94.
Abstract: The effect of cholesterol on fetal rat enterocytes and IEC-6 cells (line originated from normal rat small intestine) was examined. Both cells were cultured in the presence of 20 to 80 microM cholesterol for up to 72 h. Apoptosis was determined by flow cytometric analysis and fluorescence microscopy. The expression of HMG-CoA reductase and peroxisome proliferator-activated receptor gamma (PPARgamma) was measured by RT-PCR. The addition of 20 microM cholesterol reduced enterocyte proliferation as early as 6 h of culture. Reduction of enterocyte proliferation by 28 and 41% was observed after 24 h of culture in the presence and absence of 10% fetal calf serum, respectively, with the effect lasting up to 72 h. Treatment of IEC-6 cells with cholesterol for 24 h raised the proportion of cells with fragmented DNA by 9.7% at 40 microM and by 20.8% at 80 microM. When the culture period was extended to 48 h, the effect of cholesterol was still more pronounced, with the percent of cells with fragmented DNA reaching 53.5% for 40 microM and 84.3% for 80 microM. Chromatin condensation of IEC-6 cells was observed after treatment with cholesterol even at 20 microM. Cholesterol did not affect HMG-CoA reductase expression. A dose-dependent increase in PPARgamma expression in fetal rat enterocytes was observed. The expression of PPAR-gamma was raised by 7- and 40-fold, in the presence and absence of fetal calf serum, respectively, with cholesterol at 80 mM. The apoptotic effect of cholesterol on enterocytes was possibly due to an increase in PPARgamma expression.

Cholesterol inhibition, cancer, and chemotherapy
Buchwald, H. (1992), Lancet 339(8802): 1154-6.
Abstract: An important feature of malignant transformation is loss of the cholesterol feedback inhibition mechanism that regulates cholesterol synthesis. Cancer cells seem to require an increase in the concentrations of cholesterol and of cholesterol precursors. Therefore, a reasonable assumption is that prevention of tumour-cell growth can be achieved by restricting either cholesterol availability or cholesterol synthesis. In-vivo and cell-culture experiments have shown that lowering the plasma cholesterol concentration or intervening in the mevalonate pathway with 3-hydroxy-3-methylglutaryl (HMG) CoA reductase inhibitors decreases tumour growth. Currently prescribed doses of HMG-CoA reductase inhibitors given orally or continuously by an implantable infusion pump could achieve tumour therapeutic tissue concentrations of these agents. My hypothesis is that cholesterol inhibition can inhibit tumour cell growth, can act as an adjuvant to cancer chemotherapy, and, possibly, can prevent carcinogenesis.

Cholesterol inhibitory effects on human sperm-induced acrosome reaction
Motamed Khorasani, A., A. P. Cheung, et al. (2000), J Androl 21(4): 586-94.
Abstract: Progesterone (P4) is known to induce an acrosome reaction in mammalian sperm in vitro, whereas cholesterol is a major inhibitor of acrosome reaction. This study had three objectives: to study the in vitro effects of exogenous cholesterol on acrosome reactions in human sperm, to study the mechanism by which cholesterol affects P4-induced acrosome reaction and those induced by dibutyryl cyclic adenosine monophosphate (db-cAMP), and to study the status of the P4 surface receptor during capacitation and acrosome reaction and its relationship with cholesterol and different acrosome reaction inducers. Acrosome reaction was induced with exposure to 10 microg/mL of P4 for 30 minutes and 1 mM of db-cAMP for 30 minutes in motile sperm either in the presence or absence of 0.1-1 microg/mL of cholesterol for 30 minutes. The effects of a 30-minute exposure to 1 microg/mL of beta-sitosterol, a cholesterol plant analogue, as well as the effects of cholesterol on P4-induced acrosome reactions were compared. Fluorescein isothiocyanate-labeled albumin-progesterone conjugate (P4-FITC-BSA) was used as the probe in order to quantify the percentage of sperm in which the P, surface receptor was exposed. The results of this study indicate that cholesterol inhibited P4-induced acrosome reactions when added to the sperm during capacitation (long incubation) and when it was added with P4 during the induction of acrosome reactions (short incubation). Similarly, acrosome reaction that was induced by db-cAMP was also inhibited by cholesterol. Fifty percent of P4-induced acrosome reaction was inhibited by a cholesterol concentration of 0.2 microg/mL. Cholesterol's inhibition of induced acrosome reaction was independent of P4 concentration. Beta-sitosterol inhibited P4-induced acrosome reaction in a dose-dependent manner that was identical to that of cholesterol. We observed that increases in the P4 surface receptor exposure were time-dependent and receptors migrated toward the equatorial segment during the first 2 hours of capacitation. We also found that db-cAMP induced the appearance of the P4 surface receptor in the sperm plasma membrane and that cholesterol inhibited it. The results of this study suggest that cholesterol inhibits acrosome reaction in a noncompetitive manner by modifying the structure of the sperm plasma membrane, which prevents exposure of the P4 surface receptor for P4 binding.

Cholesterol inhibits glutamine metabolism in LLC WRC256 tumour cells but does not affect it in lymphocytes: possible implications for tumour cell proliferation
Lescano-De-Souza, A., Jr. and R. Curi (1999), Cell Biochem Funct 17(4): 223-8.
Abstract: The effect of cholesterol on proliferation and glutamine metabolism of lymphocytes and tumour cells was investigated. The addition of cholesterol to the culture medium did not cause a significant effect on 2-(14)C-thymidine incorporation in lymphocytes. In the presence of concanavalin A, lymphocyte proliferation was increased by cholesterol (from 0.013 up to 1.3 microm). At high concentrations (234 and 468 microm), however, a marked inhibition of lymphocyte proliferation occurred. The same inhibitory effect was observed in the presence of lipopolysaccharides. Cholesterol also caused a marked decrease of LLC WRC256 tumour cell growth at 117 and 234 microm. The same findings were obtained by the measurement of 2-(14)C-thymidine incorporation in these cells. The effect of cholesterol on phosphate-dependent glutaminase activity was then tested in cultured lymphocytes and LLC WRC256 tumour cells. Cholesterol at concentrations of 117 and 234 microm did not alter this enzyme activity in lymphocytes. However, this sterol, already at 26 microm, caused a 44 per cent reduction in glutaminase activity. Similar to the changes observed for glutaminase, cholesterol reduced glutamine oxidation in LLC WRC256 tumour cells, whereas no effect was observed on lymphocytes. Therefore, cholesterol might control lymphocyte and tumour cells proliferation by different mechanisms. The significance of these findings for the immune function in tumour-bearing patients remains to be investigated.

Cholesterol inhibits peroxidation of egg phosphatidylcholine in multilamellar liposomes
Gallova, J., P. Slosarcik, et al. (2002), Ceska Slov Farm 51(5): 240-3.
Abstract: The effect of cholesterol (CHOL) on the rate of autoperoxidation of chromatographically pure phosphatidylcholine from hen egg yolks (EYPC) in aerated multilamellar liposomes was examined at eight CHOL:EYPC molar ratios from 0.1 up to 1.0 at 45 degrees C. Autoperoxidation was measured as conjugated diene (CD) formation by UV-VIS spectrophotometry. In the presence of cholesterol, the initiation and propagation phases of autoperoxidation were prolonged, the autoperoxidation rate in the propagation phase was decreased, and the CD concentration in the termination phase was lowered. The inhibition of EYPC autoperoxidation by cholesterol can originate a) from the increased lateral separation of polyunsaturated EYPC acyl chains caused by intercalation of cholesterol between EYPC molecules, b) from the increased molecular packing of both the bilayer polar and hydrophobic regions due to the reduced bilayer hydration, c) from the decreased concentration of the molecular oxygen in the bilayer interior, and d) from the cholesterol epoxidation and hydroperoxidation induced by the oxidation products of EYPC polyunsaturated acyl chains.

Cholesterol inhibits spontaneous action potentials and calcium currents in guinea pig gallbladder smooth muscle
Jennings, L. J., Q. W. Xu, et al. (1999), Am J Physiol 277(5 Pt 1): G1017-26.
Abstract: Elevated cholesterol decreases agonist-induced contractility and enhances stone formation in the gallbladder. The current study was conducted to determine if and how the electrical properties and ionic conductances of gallbladder smooth muscle are altered by elevated cholesterol. Cholesterol was delivered as a complex with cyclodextrin, and effects were evaluated with intracellular recordings from intact gallbladder and whole cell patch-clamp recordings from isolated cells. Cholesterol significantly attenuated the spontaneous action potentials of intact tissue. Furthermore, calcium-dependent action potentials and calcium currents were reduced in the intact tissue and in isolated cells, respectively. However, neither membrane potential hyperpolarizations induced by the ATP-sensitive potassium channel opener, pinacidil, nor voltage-activated outward potassium currents were affected by cholesterol. Hyperpolarizations elicited by calcitonin gene-related peptide were reduced by cholesterol enrichment, indicating potential changes in receptor ligand binding and/or second messenger interactions. These data indicate that excess cholesterol can contribute to gallbladder stasis by affecting calcium channel activity, whereas potassium channels remained unaffected. In addition, cholesterol enrichment may also modulate receptor ligand behavior and/or second messenger interactions.

Cholesterol inhibits the lytic activity of melittin in erythrocytes
Raghuraman, H. and A. Chattopadhyay (2005), Chem Phys Lipids 134(2): 183-9.
Abstract: Although cell lysis by the hemolytic peptide, melittin, has been extensively studied, the role of specific lipids of the erythrocyte membrane on melittin-induced hemolysis remains unexplored. In this report, we have explored the modulatory role of cholesterol on the hemolytic activity of melittin by specifically depleting cholesterol from rat erythrocytes using methyl-beta-cyclodextrin (MbetaCD). Our results show that the hemolytic activity of melittin is increased by approximately 3-fold upon depletion of erythrocyte membrane cholesterol by approximately 55% without any appreciable loss of phospholipids. This result constitutes the first report demonstrating that the presence of cholesterol inhibits the lytic activity of melittin in its natural target membrane, i.e., the erythrocyte membrane. These results are relevant in understanding the role of cholesterol in the mechanism of action of melittin in the erythrocyte membrane.

Cholesterol inhibits the nuclear entry of estrogen receptor activation factor (E-RAF) and its dimerization with the nonactivated estrogen receptor (naER) in goat uterus
Thampan, R. V., A. Zafar, et al. (2000), J Cell Biochem 77(3): 382-95.
Abstract: An alternative form of estrogen receptor isolated from goat uterus, the nonactivated estrogen receptor (naER), has no DNA-binding function, although it is closely similar to the classical estrogen receptor (ER) in its hormone binding affinity and specificity. The naER dimerizes with a DNA binding protein, estrogen receptor activation factor (E-RAF). The heterodimer binds to the DNA. Assays carried out during the purification of E-RAF showed that an endogenous inhibitor that is heat stable and dialyzable bound to the E-RAF and prevented the formation of the heterodimer. The inhibitor has been isolated and purified. GC-MS analysis identifies this molecule to be cholesterol. Circular dichroism measurement has shown that the high-affinity binding of cholesterol to E-RAF results in subtle changes in the secondary and the tertiary structure of the protein. The E-RAF with altered conformation fails to dimerize with the naER. Instead of facilitating E-RAF entry into the nucleus, dimerization with the naER prevents it. Similarly, cholesterol binding blocks the nuclear entry of the protein, showing that E-RAF with altered conformation is incapable of interaction with the nuclear pore complex/membrane proteins. The naER-E-RAF heterodimer remains at the nuclear periphery, incapable of further transport. These results indicate the possibility that the dimerization between naER and the E-RAF takes place only within the nuclear compartment. The observation that cholesterol binding prevents nuclear entry of the E-RAF reflects the similarity of E-RAF with the sterol regulatory element (SRE) binding protein that enters the nucleus and binds to SRE only when the intracellular level of cholesterol remains low.

Cholesterol intake
Biro, G. and M. Antal (1992), Orv Hetil 133(42): 2731-2.

Cholesterol intake and plasma cholesterol: an update
DJ, P. N. (1997), J Am Coll Nutr 16(6): 530-4.
Abstract: The misperception that dietary cholesterol determines blood cholesterol is held by many consumers in spite of evidence to the contrary. Many studies reported over the past 2 years have shown that dietary cholesterol is not a significant factor in an individual's plasma cholesterol level or cardiovascular disease (CVD) risk. Reports from the Lipid Research Clinics Research Prevalence Study and the Framingham Heart Study have shown that dietary cholesterol is not related to either blood cholesterol or heart disease deaths. In a similar manner, 10 clinical trials (1994 to 1996) of the effects of dietary cholesterol on blood lipids and lipoproteins indicate that addition of an egg or two a day to a low-fat diet has little if any effect on blood cholesterol levels. This observation was noted in young men and women with normal cholesterol levels as well as older subjects with elevated plasms cholesterol concentrations. The consistency of the clinical and the epidemiological data demonstrating that dietary cholesterol has little effect on plasma cholesterol in most individuals raises a number of questions regarding the justification of population wide restrictions on dietary cholesterol intake and egg consumption.

Cholesterol intake and serum cholesterol level
Gotto, A. M. (1991), N Engl J Med 324(13): 912-3.

Cholesterol intake is associated with lecithin intake in Japanese people
Ishinaga, M., A. Ueda, et al. (2005), J Nutr 135(6): 1451-5.
Abstract: Many dietary recommendations for the prevention of heart disease and hyperlipidemia include restriction of cholesterol intake. However, limiting cholesterol intake might also affect the intake of other nutrients. The daily intakes of cholesterol, lecithin, total fatty acids, and SFAs by 388 Japanese subjects (ages 6-59 y) were analyzed directly using the duplicate portion sampling technique. Intakes were 266.1 +/- 146.5 mg/d, 1.6 +/- 0.9 g/d, 39.3 +/- 16.8 g/d, and 12.8 +/- 6.9 g/d, respectively. There was a strong positive correlation between cholesterol and lecithin intakes (r = 0.864, P < 0.001), and when food intake was adjusted to 1 kg/d, the correlation remained high (r = 0.881, P < 0.001). In contrast, the correlation between total fatty acid and lecithin intakes was lower (r = 0.423, P < 0.001), and when food intake was adjusted to 1 kg/d, the correlation coefficient remained stable (r = 0.448, P < 0.001). These results strongly indicate that limiting cholesterol intake decreases lecithin intake. Lecithin intake can be estimated from the following regression equation: lecithin intake (g/d) = 0.005 x cholesterol intake (mg/d) + 0.16 (R2= 0.747, P < 0.001). Furthermore, the intake of choline derived from lecithin can be estimated by the following equation: choline (mg/d) = 0.724 x cholesterol (mg/d) + 21.5.

Cholesterol interaction with free radicals produced from carbon tetrachloride or bromotrichloromethane by either catalytic decomposition or via liver microsomal activation
Fanelli, S. L., G. D. Castro, et al. (1995), Chem Biol Interact 98(3): 223-36.
Abstract: The reaction between cholesterol (Ch) and trichloromethyl or trichloromethyl peroxyl radicals was studied. The latter were generated from CCl4 either by benzoyl peroxide (BP) catalysis or via thermal activation or by liver microsomal NADPH-dependent biotransformation of CBrCl3. The structure of the products formed was elucidated by gas chromatography-mass spectrometry (GC/MS). Under aerobic conditions and using thermal activation of CCl4, the formation of 6 products was observed. Two (I and II) were dehydrated Ch derivatives (one also having a third double bond) (I). Another product was a delta(5)-3 ketone derivative of Ch (III). Two additional reaction products were determined as ketocholesterols (IV and V). One chloro Ch was also formed (VI). At low concentrations of BP, reaction was more extensive than under thermal activation, and the formation of peaks I to IV was also observed. When the reaction was conducted anaerobically and using thermal activation of CCl4 to generate radicals, only products I and II were formed in low yield. Under anaerobic conditions, but using catalyst, compounds I and III were produced plus two new isomeric ketocholesterol derivatives (VIII and IX) and also a compound having an extra hydroxyl group on the Ch structure (X). In order to check whether similar reactions are observable under biological experimental conditions, we used activation of CBrCl3 by liver microsomes. The incubation using only microsomes (without CBrCl3 or NADPH) showed two ketocholesterol peaks (A and B). In the presence of CBrCl3 we could detect peak B and hydroxycholesterol (C) and two others, ketocholesterols (D and E). D was the only peak showing close similarity (spectrum and retention time) to one of those observed in the chemical reaction system (V). The reaction of CBrCl3 in the presence of NADPH showed peaks B, C, D and E, in low abundance and a 7-ketocholesterol (F). If some of the reaction products reported here were formed during the intoxication with these haloalkanes, significant biological consequences might be expected.

Cholesterol interaction with recombinant human sterol carrier protein-2
Colles, S. M., J. K. Woodford, et al. (1995), Lipids 30(9): 795-803.
Abstract: The interaction of human recombinant sterol carrier protein-2 (SCP-2) with sterols was examined. Two independent ligand binding methods, Lipidex 1000 binding of 3Hcholesterol and a fluorescent dehydroergosterol binding assay, were used to determine the affinity of SCP-2 for sterols. Binding analysis indicated SCP-2 bound 3Hcholesterol and dehydroergosterol with a Kd of 0.3 and 1.7 microM, respectively, and suggested the presence of a single binding site. Phase fluorometry and circular dichroism were used to characterize the SCP-2 sterol binding site. Alterations in dehydroergosterol lifetime, SCP-2 tryptophan lifetime, and SCP-2 tryptophan quenching by acrylamide upon cholesterol binding demonstrated a shielding of the SCP-2 tryptophan from the aqueous solvent by bound sterol. Differential polarized phase fluorometry revealed decreased SCP-2 tryptophan rotational correlation time upon cholesterol binding. Circular dichroism of SCP-2 indicated that cholesterol elicited a small decrease in SCP-2 alpha helical content. The data suggest that SCP-2 binds sterols with affinity consistent with a lipid transfer protein that may act either as an aqueous carrier or at a membrane surface to enhance sterol desorption.

Cholesterol interaction with the daunorubicin binding site of P-glycoprotein
Wang, E., C. N. Casciano, et al. (2000), Biochem Biophys Res Commun 276(3): 909-16.
Abstract: The inherent complexities of cholesterol disposition and metabolism preclude a single transmembrane active transport avenue for this steroid-precursor, cell-membrane constituent. Yet the ABC (ATP binding cassette) transporters are inextricably linked to elements of cholesterol disposition. Recent observations have suggested that, under certain settings, the ABC transporter P-glycoprotein (P-gp) performs a direct role in cholesterol disposition. The gene product of MDR1 (multidrug resistance transporter), P-glycoprotein also confers protection against xenobiotics. Using a whole cell assay in which the retention of a marker substrate is evaluated and quantified, we studied the ability of cholesterol to inhibit directly the function of this transporter. In a NIH-G185 cell line presenting an overexpressed amount of the human transporter P-gp, cholesterol caused dramatic inhibition of daunorubicin transport with an IC(50) of about 8 microM yet had no effect on the parent cell line nor rhodamine 123 transport. Additionally, using the ATP-hydrolysis assay, we showed that cholesterol increases P-gp-mediated ATP hydrolysis by approximately 1.6-fold with a K(s) of 5 microM. Suggesting that cholesterol directly interacts with the substrate binding site of P-gp, these results are consistent with cholesterol being transported by MDR1 P-gp.

Cholesterol interactions with phospholipids in membranes
Ohvo-Rekila, H., B. Ramstedt, et al. (2002), Prog Lipid Res 41(1): 66-97.
Abstract: Mammalian cell membranes are composed of a complex array of glycerophospholipids and sphingolipids that vary in head-group and acyl-chain composition. In a given cell type, membrane phospholipids may amount to more than a thousand molecular species. The complexity of phospholipid and sphingolipid structures is most likely a consequence of their diverse roles in membrane dynamics, protein regulation, signal transduction and secretion. This review is mainly focused on two of the major classes of membrane phospholipids in eukaryotic organisms, sphingomyelins and phosphatidylcholines. These phospholipid classes constitute more than 50% of membrane phospholipids. Cholesterol is most likely to associate with these lipids in the membranes of the cells. We discuss the synthesis and distribution in the cell of these lipids, how they are believed to interact with each other, and what cellular consequences such interactions may have. We also include a discussion about findings in the recent literature regarding cholesterol/phospholipid interactions in model membrane systems. Finally, we look at the recent trends in computer and molecular dynamics simulations regarding phospholipid and cholesterol/phospholipid behavior in bilayer membranes.

Cholesterol interactions with tetracosenoic acid phospholipids in model cell membranes: role of the double-bond position
Ayanoglu, E., B. H. Chiche, et al. (1990), Biochemistry 29(14): 3466-71.
Abstract: The synthesis and thermotropic properties of 1,2-di-(9Z)-9-tetracosenoylphosphatidylcholine delta 9-PC(24:1,24:1), 1, 1,2-di-(5Z)-5-tetracosenoylphosphatidylcholine delta 5-PC(24:1,24:1), 2, and 1,2-di-(15Z)-15- tetracosenoylphosphatidylcholine delta 15-PC(24:1,24:1), 3 are reported. Liposomes prepared from these phospholipids differ from those of the natural sponge phospholipids, 1,2-di-(5Z,9Z)-5,9-hexacosadienoylphosphatidylcholine (4a) and the corresponding ethanolamine (4b), both of which virtually exclude cholesterol from their bilayers. The behavior of 1 and 2 is similar to that of 1,2-di-(6Z,9Z)-6,9-hexacosadienoylphosphatidylcholine (5), which exhibits a partial molecular interaction with cholesterol. In the case of 3, cholesterol appears to interact with the saturated acyl chain regions of this phospholipid in a manner similar to that of its interaction with DPPC acyl chains. This study delineates the effect of the double-bond location in long fatty acyl chains of phospholipids on their interactions with cholesterol.

Cholesterol interacts with all of the lipid in bilayer membranes. Implications for models
Singer, M. A. and L. Finegold (1990), Biophys J 57(1): 153-6.
Abstract: The interaction of cholesterol with lipid membranes has been studied by differential scanning calorimetry on liposomes, a technique which involves only the natural lipids, with no exogeneous probes. The influence of cholesterol at different molar percent concentrations c on the enthalpy delta H of the main gel to liquid crystal phase transition of saturated phosphatidylcholines of acyl chain length n = 12-20 was well represented by delta H = -9.43 + 1.01n - 0.268c kcal/mol. The linear dependence of delta H simultaneously upon chain length n and upon cholesterol concentration c shows clearly that cholesterol interacts with the deeper part of the lipids, as well as the superficial parts. This observation is not accommodated in any of the current models of cholesterol-lipid interactions.

Cholesterol interacts with lactosyl and maltosyl cerebrosides but not with glucosyl or galactosyl cerebrosides in mixed monolayers
Slotte, J. P., A. L. Ostman, et al. (1993), Biochemistry 32(31): 7886-92.
Abstract: Pure and mixed monolayers of mono- and dihexoside cerebrosides with cholesterol have been characterized at the air/water interface. Cholesterol oxidase was used as a reporter enzyme for the cholesterol-cerebroside interaction in the mixed monolayers. The cerebrosides either were derived from bovine brain extracts or were synthetic. The dihexoside cerebrosides were synthesized by coupling of the hepta-O-acetyl-alpha-lactosyl- or maltosylphosphoramidates with D-erythro-N-acylceramides in dichloromethane, in the presence of trimethylsilyl triflate and molecular sieves, followed by hydrolysis of the acetate-protecting groups. All of the bovine-brain-derived cerebrosides galactosyl cerebroside (GalCer, types I and II), glucosyl cerebroside (GlcCer), and lactosyl cerebroside (LacCer) had very condensed force-area isotherms (compressibility values of 3-5 x 10(-3) m/mN at 20 mN/m), as did the synthetic N-stearoylmaltosylceramide (N-18:0 MaltCer). Shorter-chain synthetic cerebrosides (N-8:0 LacCer and N-8:0 MaltCer) had more expanded isotherms, with compressibility values of 15-17 x 10(-3) m/mN. When cholesterol was included in mixed monolayers of monohexoside cerebroside, it did not induce significant condensation of packing (indicating that cholesterol did not increase the order of the acyl chains). However, with dihexoside cerebrosides, a cholesterol-induced condensing effect was observed, which amounted to a 11-19% reduction in the observed mean molecular area. When cholesterol oxidase was used to titrate the stoichiometry of cholesterol/cerebroside in mixed monolayers, at which pure cholesterol clusters appeared, it was observed that in monohexoside cerebroside monolayers cholesterol clusters were present even below a 1:1 molar stoichiometry.(ABSTRACT TRUNCATED AT 250 WORDS)

Cholesterol is accumulated by mycobacteria but its degradation is limited to non-pathogenic fast-growing mycobacteria
Av-Gay, Y. and R. Sobouti (2000), Can J Microbiol 46(9): 826-31.
Abstract: In this report we show that fast-growing non-pathogenic mycobacteria degrade cholesterol from liquid media, and are able to grow on cholesterol as a sole carbon source. In contrast, slow-growing mycobacteria, including pathogenic Mycobacterium tuberculosis and bacillus Calmette-Guerin (BCG), do not degrade and use cholesterol as a carbon source. Nevertheless, pathogenic mycobacteria are able to uptake, modify, and accumulate cholesterol from liquid growth media, and form a zone of clearance around a colony when plated on solid media containing cholesterol. These data suggest that cholesterol may have a role in mycobacterial infection other than its use as carbon source.


First Page Previous Page Next Page Last Page



Sitemap
Link | Link | Link | Link | Link | Link | Link | Link

Search the Dr Huxt site:

powered by FreeFind



Last Modified: 29 January 2006
http://www.huxt.com