Vous êtes sur la page 1sur 8

Research: Science and Education

edited by
Concepts in Biochemistry William M. Scovell
Bowling Green State University
Bowling Green, OH 43403

Cholesterol Biosynthesis: Lanosterol to Cholesterol


John M. Risley
Department of Chemistry, The University of North Carolina at Charlotte, Charlotte, NC 28223-0001;
jmrisley@email.uncc.edu

Cholesterol is an important molecule. It is important at the National Library of Medicine (13) yielded more than
biochemically because it maintains membrane fluidity, 10,000 citations, while a search of “cholesterol biosynthesis,
microdomain structure, and permeability, and it is a precursor review” resulted in more than 1000 citations. A significant
to steroid hormones, bile acids, vitamin D, and lipoproteins. proportion of the latter are reviews of drug effects on choles-
In addition to these well-recognized functions, cholesterol has terol biosynthesis. However, I am not aware of a recent review
been shown recently to have an essential role in mammalian of the pathway from lanosterol to cholesterol that describes
embryonic development (1). It is important medically because in detail the reactions along the line of ref 9. I did find an
it is correlated with a number of diseases, as, for example, excellent review published in 1997 that summarizes the cloning
atherosclerosis (2), cholestasis (broad and variable impairment of cDNAs or genes encoding the enzymes of sterol biosyn-
of bile secretion) (3), hypercholesterolemia (4 ), cholesterol thesis for fungi, mammals, and higher plants (14), including brief
gallstone disease (5), and Nieman–Pick type C disease (6 ), descriptions of the enzyme reactions for each of the pathways
on which major research efforts and monies are expended to ergosterol, cholesterol, stigmasterol, and campesterol. Ad-
annually. It is important commercially for the pharmaceutical vances have been made since the publication of the review
industry in terms of drug development, and for the food article. For example, the genes for C3 sterol dehydrogenase
industry in terms of “healthy foods” that are advertised either (C4 decarboxylase) (15) and 3-keto reductase (16 ) have been
to be “cholesterol-free” or to reduce cholesterol. For example, cloned, completing the elucidation of the pathway of ergos-
Zocor (simvastatin) has been advertised to the general public terol biosynthesis in the yeast Saccharomyces cerevisiae.
by Merck & Co. as “an effective medicine that along with With no recent review from which to teach the conversion
diet and exercise can significantly lower total cholesterol” (7). of lanosterol to cholesterol, I have used the outline presented
And, above the nutrition facts on a box of Kellogg’s Corn in 1985 (9), with supplements obtained in my readings from
Flakes the cereal is touted as a “cholesterol free food”, and the research literature describing subsequent information on the
on the front of a box of General Mills Cheerios it reads “As enzymes in this pathway. This paper presents the material
part of a heart-healthy diet, the soluble fiber in Cheerios May that I use in teaching this topic. It is divided into five sections:
Reduce Your Cholesterol!” Each cereal “Meets American Heart (i) general overview of the pathway, (ii) step-by-step synthesis,
Association food criteria for saturated fat and cholesterol for (iii) cholesterol biosynthesis in perspective, (iv) cholesterol
healthy people over age 2”. homeostasis, and (v) inherited diseases of cholesterol biosynthe-
While exogenous cholesterol is considered a major cause sis. I do not present a comprehensive review of the literature in
for many of the health problems, cholesterol biosynthesis in this area, but I have included recent relevant references from
vivo (endogenous cholesterol) is a target in reducing overall which a more comprehensive study may be pursued if desired.
cholesterol in humans (as, for example, by lovastatin, which
inhibits HMG-CoA reductase) (2). Considering the significance Overview of the Pathway
and interest in cholesterol, it is interesting that discussions
of cholesterol biosynthesis in general biochemistry textbooks The anabolic pathway for the conversion of lanosterol
are restricted to the synthesis of lanosterol from acetate, which to cholesterol is shown in Figure 1. The numbering scheme
is usually discussed in detail. The conversion of lanosterol to for lanosterol is shown in Figure 2.
cholesterol is most often simply indicated, without elaboration, The biosynthesis of cholesterol from lanosterol is a 19-
as a multistep process. The only textbook I know of that step process. It requires nine different enzymes; two enzymes
presents a pathway of lanosterol to cholesterol is that by Voet catalyze multiple steps and three are utilized two times in
and Voet (8), after Rilling and Chayet in 1985 (9). The the pathway. Three methyl groups in lanosterol (Fig. 2: C30,
Biochemical Pathways chart distributed by Roche Molecular C31, C32) are oxidized and removed as formic acid and two
Biochemicals (10) gives an abbreviated pathway showing carbon dioxide molecules. Two mechanisms are utilized in
major intermediates, but in general without elaboration; the these reactions: a cytochrome P450 and formation of a β-
chart is used by ExPASy with links to a few of the enzymes ketoacid. The other reactions in the pathway are five reduction
(11). Map 100 in the Kyoto Encyclopedia of Genes and Genomes reactions using NAD(P)H, one dehydrogenation (desaturation)
(KEGG ) (12) outlines the pathway, naming major inter- reaction using NADH/O2, and one isomerization. Two mecha-
mediates with links to a few of the enzymes, but without nisms are utilized in the reduction reactions: two ketoreductions
elaboration. This situation is unsatisfactory for numerous and three olefinic reductions involving carbocation intermedi-
students, especially those more medically oriented. ates. A radical mechanism is utilized in the dehydrogenation
The research literature on cholesterol biosynthesis is vo- (desaturation) reaction, and formation of a carbocation inter-
luminous. A search of “cholesterol biosynthesis” on PubMed mediate is utilized in the isomerization reaction.

JChemEd.chem.wisc.edu • Vol. 79 No. 3 March 2002 • Journal of Chemical Education 377


Research: Science and Education

Figure 1. The pathway for the 19-step conversion of lanosterol to cholesterol.

378 Journal of Chemical Education • Vol. 79 No. 3 March 2002 • JChemEd.chem.wisc.edu


Research: Science and Education

Step-by-Step Synthesis 22 24
27
21
18 20 23 25
In the following discussion of the 19-step reaction, I give
the various names used in the literature for the enzyme, sub- 19 11
12
13
17
16
26

strates, and products in each step. 9 14 15


1 8
10 α
Oxidative Demethylation of 1 to 5 β 3
2
5 32
7
6
The oxidative demethylation of lanosterol (4,4′,14α-tri- HO β
4
α
methyl-5α-cholesta-8(9),24-dien-3β-ol) (1) to cholestatriene 31 30
(4,4′-dimethyl-5α -cholesta-8(9),14,24-trien-3β -ol; 4,4′-
dimethylcholesta-8,14,24-trien-3β-ol) (5) is catalyzed by lanos- Figure 2. The numbering scheme for lanosterol.
terol 14α-methyl demethylase cytochrome P450 (lanosterol
14α-methyl demethylase; 14α-methyl demethylase; lanosterol
14α-demethylase) (17–22). The 14α-methyl group, which
protrudes into the α face of the sterol ring system, is the first
group to be removed; its removal is essential for membrane and
regulatory functions of the resulting sterols (14). The enzyme, Cl
a cytochrome P450, requires the C3-hydroxy group, 14α- N
methyl group, and ∆8(9) bond for binding and catalysis (22); it N CH2 CH O CH2 Cl
also requires three NADPH coenzymes and three O2 molecules. Cl
Upon substrate binding the heme shifts from a low-spin to a
high-spin state (22). miconazole
The enzyme catalyzes three successive monooxygenation Cl
reactions: (i) lanosterol (1) to 3β-hydroxylanost-8-en-32-ol
(lanost-8-en-3β,32-diol) (2), (ii) 3β-hydroxylanost-8-en-32-ol
(2) to 3β-hydroxylanost-8-en-32-al (3), and (iii) 3β-hydroxy- O N N COCH3
lanost-8-en-32-al (3) to 14α-formyloxy-lanost-8-en-3β-ol (4) O
N
to cholestatriene (5). The third oxidation has been proposed N CH2 C O
to proceed by peroxy attack, a Baeyer–Villiger rearrangement, Cl
and deformylation by a base rearrangement (17), or by peroxy ketoconazole
attack, homolytic cleavage, deformylation by fragmentation,
and disproportionation (20). (Deformylation by fragmentation
has been termed a radicalar mechanism [14].) The human cDNA Cl
clone has been reported (23) and a model of the Candida Figure 3. The structures for two inhibitors of lanosterol 14α-methyl
albicans enzyme has been proposed (24). The crystal structure demethylase cytochrome P450.
of the enzyme from Mycobacterium tuberculosis was recently
published (PDB ID 1E9X and 1EA1) (25). This is the first
enzyme in this pathway for which there is a crystal structure.
The enzyme is inhibited by compounds, such as
miconazole and ketoconazole (lipophilic imidazoles) (Fig. 3),
that coordinate through the imidazole nitrogen to the heme
iron of the cytochrome P450 and block binding of O2 (14,
25). The result is an accumulation of dihydrolanosterol (the
C24 reduction product of lanosterol).
Reduction of 5 to 6 HO 5
H* Hw
The reduction of cholestatriene (5) to 14-demethyl- HO
6
lanosterol (4,4′-dimethylzymosterol; 4,4′-dimethylcholesta-
Hw+
8,24-dien-3β-ol) (6) is catalyzed by ∆14-sterol reductase (sterol
∆14-reductase; steroid-14-reductase; ∆8,14-sterol ∆14-reductase) NADPH*

(26, 27 ). The enzyme requires NADPH and the mechanism


involves formation of a carbocation intermediate (Fig. 4). A
proton from the aqueous solution (or from a general acid in
the enzyme active site) protonates at C15 to form a carbo- + Hw
cation, which is stabilized by the active site of the enzyme HO
and by the adjacent double bond. The enzyme directs
transfer of a hydride from NADPH to C14 that completes
the reduction reaction. Figure 4. Reduction of the C14 double bond by ∆14-sterol reductase.

JChemEd.chem.wisc.edu • Vol. 79 No. 3 March 2002 • Journal of Chemical Education 379


Research: Science and Education

Oxidation of 6 to 9
The oxidation of 4,4′-dimethylzymosterol (6) to 4-
methylzymosterol-4-carboxylic acid (4β-methyl-4α-carboxy- NAD+
H
cholesta-8,24-dien-3β-ol) (9) is catalyzed by sterol-4α-methyl-
oxidase (C4 sterol methyl oxidase; C4 methyl sterol oxidase; O 9
H C O
4-methyl sterol oxidase; C4 methyl oxidase; C4-methyl- –
O
oxidase) (28, 29). The enzyme is specific for oxidation of the
C4α methyl group. It requires three NADH coenzymes and HO
three O2 molecules. The reducing equivalents from NADH are
acquired by NADH:cytochrome b5 reductase and cytochrome
b5. The enzyme catalyzes three successive (monooxygenation) CO2
oxidation reactions: (i) 4,4′-dimethylzymosterol (6) to 4-methyl-
4-hydroxymethylzymosterol (4β-methyl-4α-hydroxymethyl-
cholesta-8,24-dien-3β-ol) (7), (ii) 4-methyl-4-hydroxymethyl-
zymosterol (7) to 4-methylzymosterol-4-carboxaldehyde (4β- O
C O
methyl-4α-formyl-cholesta-8,24-dien-3β-ol) (8), and (iii) H O
O 10
4-methylzymosterol-4-carboxaldehyde (8) to 4-methyl- H
zymosterol-4-carboxylic acid (9). The human cDNA clone has
been mapped to chromosome 4q32-34, and histidine-rich Figure 5. Oxidative decarboxylation and enolization epimerization
clusters are suggested to be involved in the binding of oxo- of the 3β-hydroxycarboxylic acid by C3 sterol dehydrogenase (C4
diiron (Fe–O–Fe) (29). decarboxylase).

Oxidative Decarboxylation and Enolization


Epimerization of 9 to 10
The oxidative decarboxylation and enolization epimeriza-
tion of 4-methylzymosterol-4-carboxylic acid (9) to 4-methyl-
zymosterone (3-keto-4-methylzymosterol) (10) is catalyzed by
C3 sterol dehydrogenase (C4 decarboxylase) (4α-oic acid de-
carboxylase; C4 decarboxylase) (15). The enzyme requires an
NAD+ coenzyme. The enzyme catalyzes oxidation of 9 (Fig. 5) HO H 16
by transfer of the 3α hydrogen as a hydride to NAD+, leading H*
to a 3-ketocarboxylic acid intermediate that undergoes Hw+
decarboxylation to give an enol. The enzyme directs the epimer-
ization from the enol that moves the former 4β-methyl group
to the 4α-methyl position.
Hw
Reduction of 10 to 11 +
The reduction of 4-methylzymosterone (10) to 4α- HO H
methylzymosterol (4α-methyl-cholesta-8,24-dien-3β-ol) (11) is H* H*+
catalyzed by 3-keto reductase (3-ketosterol reductase; steroid-
3-ketoreductase) (16 ). The enzyme, which requires an HO 17
NAD(P)H coenzyme, catalyzes the 3-ketoreduction to convert
the 3-keto group to the β-hydroxy sterol.
Figure 6. The isomerization reaction catalyzed by ∆ 8-∆7 sterol
Oxidation of 11 to 14 isomerase.
The oxidation of 4 α -methylzymosterol (11) to
zymosterol-4α-carboxylic acid (4α-carboxy-cholesta-8,24-
dien-3β-ol) (14) is catalyzed by sterol-4α-methyl-oxidase, the
same enzyme that catalyzes the oxidation of 6 to 9. It cata-
lyzes three successive (monooxygenation) oxidation reactions: that catalyzes the oxidative decarboxylation and enolization
(i) 4 α -methylzymosterol (11) to 4 α -hydroxymethyl- epimerization of 9 to 10. The enzyme catalyzes oxidation of 14
zymosterol (4α -hydroxymethyl-cholesta-8,24-dien-3β-ol) to a 3-ketocarboxylic acid intermediate that undergoes decarbox-
(12), (ii) 4α-hydroxymethylzymosterol (12) to zymosterol- ylation to give an enol, which undergoes tautomerization.
4α-carboxaldehyde (4α-formyl-cholesta-8,24-dien-3β-ol)
(13), and (iii) zymosterol-4 α -carboxaldehyde (13) to Reduction of 15 to 16
zymosterol-4α-carboxylic acid (14). The reduction of zymosterone (15) to zymosterol (5α-
cholesta-8,24-dien-3 β -ol; 5 α -cholest-8-en-3 β -ol; ∆ 8,24-
Oxidative Decarboxylation of 14 to 15 cholestadien-3β-ol; ∆8-cholestenol) (16) is catalyzed by 3-keto
The oxidative decarboxylation of zymosterol-4α-carboxylic reductase, the same enzyme that catalyzes the reduction of
acid (14) to zymosterone (15) is catalyzed by C3 sterol de- 10 to 11. The enzyme catalyzes the 3-ketoreduction to convert
hydrogenase (C4 decarboxylase), which is the same enzyme the 3-keto group to the β-hydroxy sterol.

380 Journal of Chemical Education • Vol. 79 No. 3 March 2002 • JChemEd.chem.wisc.edu


Research: Science and Education

Isomerization of 16 to 17
The isomerization of zymosterol (16) to lanthosterol (5α-
cholesta-7,24-dien-3 β -ol; 5 α -cholest-7-en-3 β -ol; ∆ 7,24-
HO
cholestadien-3β-ol; cholest-7-en-3β-ol) (17) is catalyzed by ∆8-∆7
Lanthosterol sterol isomerase (sterol ∆8-isomerase; 3β-hydroxysteroid-∆8-
HO ∆7-isomerase; cholestenol ∆-isomerase; ∆7-cholestenol ∆7-∆8-
NADH
O2
Desmosterol isomerase, EC 5.3.3.5) (30–33). The enzyme is also an
NADPH
emopamil-binding protein and a sigma factor. The human
cDNA clone has been reported (30). The enzyme catalyzes
NAD(P)H
isomerization of the double bond (Fig. 6) by protonation at C9
by a proton from the aqueous solution (or from a general acid
in the enzyme active site) to form a carbocation intermediate,
which is stabilized by the active site of the enzyme. The enzyme
HO directs subsequent abstraction of the 7β H to give the product.
7-Dehydrodesmosterol
This isomerization is the only reversible reaction in the 19
steps of biosynthesis of cholesterol from lanosterol (33).
HO
Cholesterol
Figure 7. The position in the pathway of C24 reduction has most Reduction of 17 to 18
often been placed as the last step in the conversion of lanos-
terol to cholesterol. In this position, the pathway from lanthosterol
The reduction of lanthosterol (17) to lathosterol (5α-
to cholesterol is shown. cholest-7-en-3β-ol; ∆7-cholesten-3β-ol) (18) is catalyzed by sterol
∆24-reductase (24-reductase; 3β-hydroxysterol ∆24-reductase)
(34). The exact location in the pathway for the C24 reduction
is not known because the enzyme catalyzes the reduction of the
double bond at each step in the pathway (e.g., see conversion
H* of 1 to 5). The position of the C24 reduction has most often
18
HO been placed as the last step in the pathway so that lanthosterol
H H is converted first to 7-dehydrodesmosterol and then to
desmosterol before final reduction of C24 to cholesterol (Fig.
H
HO
7). Bae and Paik (34) reported that the relative rates of re-
H H duction by sterol ∆24-reductase were 1.0 for lanthosterol (17),
H* 0.34 for zymosterol (16), 0.31 for desmosterol, and 0.06 for
lanosterol (1); lanthosterol is more easily reduced than
desmosterol or zymosterol. On the basis of their proposal from
this study, I have placed the C24 reduction at this location
in the pathway. The enzyme requires an NAD(P)H coenzyme
and presumably catalyzes the reduction by a carbocation
19
HO intermediate.
*
H
Figure 8. Lathosterol oxidase catalyzes the formation of the C5 Dehydrogenation (Desaturation) of 18 to 19
double bond.
The dehydrogenation (desaturation) of lathosterol (18) to
7-dehydrocholesterol (cholesta-5,7-dien-3β-ol) (19) is catalyzed
by lathosterol oxidase (lathosterol 5-desaturase; ∆7-sterol ∆5-
reductase; ∆7-sterol 5-reductase; ∆7-sterol-C5(6)-desaturase;
sterol 5α,6α-desaturase; sterol ∆5 desaturase; ∆5-dehydrogenase;
19
Hw C5-sterol desaturase; 5α-cholest-7-en-3β-ol:O2 ∆5-oxido-
HO reductase, EC 1.3.3.2) (35, 36 ). The enzyme requires an
20
HO
NAD(P)H coenzyme (NADH is better than NADPH) and
Hw+ H*
an O2 molecule; the reducing equivalents from NADH are
acquired by NADH:cytochrome b5 reductase and cytochrome
NADPH* b5. The enzyme contains histidine-rich motifs that would
provide the ligands for a presumed catalytic Fe center (35).
It catalyzes the stereospecific dehydrogenation (desaturation)
Hw of the 5α and 6α hydrogen atoms in a cis abstraction. A study
indicated an asynchronous scission of the two C–H bonds
+ (Fig. 8) where the C6α H is removed first to form a radical
HO
intermediate, and then the C5α H is removed, rather than a
Figure 9. The last step in the pathway is reduction of the C7 double concerted desaturation mechanism (36 ). The human cDNA
bond by 7-dehydrocholesterol reductase. clone has been reported (37 ).

JChemEd.chem.wisc.edu • Vol. 79 No. 3 March 2002 • Journal of Chemical Education 381


Research: Science and Education

Reduction of 19 to 20 bond cleavage. Instead, a second enzyme catalyzes oxidation of


the β-hydroxyacid to a β-ketoacid, which undergoes decar-
The reduction of 7-dehydrocholesterol (19) to cholesterol
boxylation. Decarboxylation via a β-ketoacid is more familiar
(20) is catalyzed by 7-dehydrocholesterol reductase (∆7-sterol
in metabolic pathways, as, for example, in the citric acid cycle
reductase; 3β-hydroxysterol ∆7-reductase; 7-dehydrocholesterol
where isocitrate dehydrogenase catalyzes oxidative decarboxyla-
∆7-reductase; cholesterol:NADP + ∆7-oxidoreductase, EC
tion of isocitrate (via oxalosuccinate) to α-ketoglutarate, and
1.3.1.21) (38, 39). The enzyme requires an NADPH coenzyme.
in the oxidative phase of the pentose phosphate pathway
It catalyzes ∆7-reduction (Fig. 9) by protonation at C8β by a
where 6-phosphogluconate dehydrogenase catalyzes oxidative
proton from aqueous solution (or from a general acid in the
decarboxylation of 6-phosphogluconate (via 6-phospho-3-oxo-
enzyme active site) to form a carbocation intermediate that is
gluconate) to ribulose-5-phosphate. Therefore, the biosynthetic
stabilized by the active site of the enzyme and by the adjacent
pathway for cholesterol from lanosterol displays two signifi-
double bond. The enzyme directs hydride transfer from NADPH
cantly different mechanisms for carbon–carbon bond cleavage;
to C7α that completes the reduction reaction. The human
the mechanisms are related via the oxidation reactions using
cDNA clone has been reported (38) and mapped to chromo-
iron centers.
some 11q12-13 (40).
Finally, the dehydrogenation (desaturation) reaction (18 →
The 19-step conversion of lanosterol to cholesterol requires
19) is not catalyzed by a flavoenzyme, although these are wide-
nine different enzymes. It has been proposed to occur in both
spread in metabolic pathways (e.g., succinate dehydrogenase
the endoplasmic reticulum and the peroxisome compartments
in the citric acid cycle, fatty acyl-CoA dehydrogenase in the
of the cell (41). trans-1,4-Bis-(2-chlorobenzylamino-
β-oxidation of fatty acids, and dihydroorotate oxidase in py-
methyl)cyclohexane dihydrochloride (AY-9944) is an inhibitor
rimidine nucleotide biosynthesis). Instead, a reduced flavin
of ∆14-sterol reductase (conversion of 5 to 6), ∆8–∆7 sterol
(NAD[P]H) and O2 are involved, presumably with a non-
isomerase (conversion of 16 to 17), and 7-dehydrocholesterol
heme iron (oxo-diiron) center. These non-heme iron enzymes
reductase (conversion of 19 to 20) with application in hyper-
participate in the biosynthesis of unsaturated fatty acids as
lipidemia. trans-2-[4-(1,2-Diphenylbuten-1-yl)phenoxy]-
fatty acyl-CoA desaturases (e.g., ∆6-desaturase, ∆9-stearoyl-
N,N-dimethylethylamine (tamoxifin), a chemotherapeutic
CoA desaturase, ∆11-myristoyl-CoA desaturase, and ∆12-oleyl-
agent used in the treatment of breast cancer, is an inhibitor
CoA desaturase) (36 ). Thus, this type of dehydrogenation
of ∆8–∆7 sterol isomerase.
may be primarily associated with lipid metabolism.
Cholesterol Biosynthesis in Perspective
Cholesterol Homeostasis
The biosynthesis of cholesterol from lanosterol is not un-
like other metabolic pathways in that the enzymes catalyze a Much has been learned about cholesterol homeostasis,
series of oxidation and reduction reactions along with carbon– although a complete picture has not yet emerged. Central to
carbon bond cleavage, epimerization, isomerization, and de- cholesterol homeostasis are the sterol regulatory element-
hydrogenation. The five reduction reactions (5 → 6, 10 → 11, binding proteins (SREBPs) (particularly SREBP-2, but also
15 → 16, 17 → 18, 19 → 20) utilize reduced flavoenzymes SREBP-1c), which are under sterol regulation. SREBP-1a and
(NAD[P]H) that are common in metabolic pathways (42). The -1c primarily regulate fatty acid biosynthesis (47–49). When intra-
epimerization (part of 9 → 10) and isomerization (16 → 17) cellular cholesterol levels decline, SREBP-cleavage activating
reactions are also observed in other metabolic pathways (42). protein (SCAP) forms a complex with SREBP that mediates the
Removal of the three methyl groups from lanosterol interaction of SREBP with site-1-protease (S1P). A recent study
(C30, C31, C32) requires carbon–carbon bond cleavage. To indicated that modifications of the N-linked carbohydrate
accomplish this, the methyl groups are oxidized to carboxylic moieties on SCAP are important for this to occur (50).
acids and eliminated as a formic acid (4 → 5) and two carbon S1P is a membrane-bound subtilisin-related serine protease
dioxides (part of 9 → 10, 14 → 15). Three enzymes catalyze (51) that catalyzes the first proteolytic reaction on SREBP.
the demethylation reactions. A cytochrome P450 enzyme is used Site-2-protease (S2P), a member of the family of zinc metallo-
for one complete oxidative decarboxylation of C32 (1 → 5), proteases (51), catalyzes hydrolysis of the second proteolytic
and two enzymes are used in the oxidation (6 → 9, 11 → 14) reaction on SREBP, which releases a soluble transcription factor
and oxidative decarboxylation (9 → 10, 14 → 15) reactions (the amino-terminal fragment of SREBP). The soluble tran-
of C30 and C31. scription factor enters the nucleus, where (usually with a second
Oxidations catalyzed by cytochrome P450 enzymes are transcription factor such as Sp-1 or NF-1) it binds to direct
common (43–45), but oxidations ending in carbon–carbon repeat sterol response elements (SRE) in promotor regions,
bond cleavage appear unique to certain steroid-metabolizing activating transcription of sterol-regulated cholesterogenic
cytochrome P450 enzymes (43, 44), including lanosterol genes. Genes under this regulatory control include, at least,
14α-methyl demethylase cytochrome P450 here. Thus, this LDL receptor, HMG-CoA synthase, HMG-CoA reductase,
type of carbon–carbon bond cleavage is not widely seen in farnesyl diphosphate synthase, squalene synthase, lanosterol
metabolic pathways. On the other hand, the oxidation reactions synthase, and lanosterol 14α-methyl demethylase cytochrome
of C30 and C31 are thought to involve non-heme oxo-diiron P450. Activation of these genes restores intracellular choles-
(29). The mechanism for oxidation by these binuclear non- terol levels.
heme iron centers is believed to be formally equivalent to When intracellular cholesterol levels build up, the activa-
that of cytochrome P450 (36, 42, 46 ), and enzymes utilizing tion of SREBPs by proteolysis is suppressed, which decreases
non-heme oxo-diiron centers are being more widely identified gene transcription for cholesterol biosynthesis. Cholesterol
(36, 46 ). But the reaction here does not end in carbon–carbon itself, when added to cultured cells, only weakly suppresses

382 Journal of Chemical Education • Vol. 79 No. 3 March 2002 • JChemEd.chem.wisc.edu


Research: Science and Education

cholesterol biosynthesis. However, when 25-hydroxycholes- malformation of the face, limb abnormalities, ambiguous
terol (cholest-5-ene-3β,25-diol), an oxysterol (52) synthesized genitalia, and mental retardation. There are two phenotypes:
by cholesterol 25-hydroxylase (53), is added to cultured cells type I (mild) and type II (severe). The incidence of this auto-
there is a rapid decrease in the activation of SREBPs and a somal recessive disorder is estimated to be 1 in 20,000 births.
consequent rapid decrease in cholesterol biosynthesis. If 25- It may be the second most common autosomal recessive dis-
hydroxycholesterol is the regulatory agent, the recent study order in the white population of North Americans (after cystic
indicates that the sterol blocks modification of the N-linked fibrosis). Approximately three-quarters of an issue of the
carbohydrate moieties of SCAP, which precludes an interaction American Journal of Medical Genetics was recently devoted to
between S1P and SREBP (50). SLOS (64 ). Recognition of cholesterol’s essential role in
Oxysterols (formed by hydroxylation of the side chain mammalian embryonic development (1) established SLOS
of cholesterol) are modulators of numerous processes (52). as the prototypical developmental disorder (62).
They are important in cholesterol homeostasis in peripheral
tissues such as brain and kidney where excess cholesterol is Acknowledgments
converted to oxysterols that are more hydrophilic than cho-
lesterol. These are transported to the liver for conversion to I wish to thank James C. Crosthwaite, Thomas W.
bile acids, which is under control of nuclear orphan receptors Mattingly Jr., and Joanna K. Krueger for a critical reading of
(54). Bile acids are synthesized by two biosynthetic pathways: the manuscript.
from cholesterol (the “classical” pathway) and from oxysterols
(a newly described pathway) (54). Literature Cited

Inherited Diseases of Cholesterol Biosynthesis 1. Farese, R. V. Jr.; Herz, J. Trends Genet. 1998, 14, 115.
2. Atherosclerosis and Coronary Artery Disease; Fuster, V.; Ross,
Deficiencies in four enzymes of cholesterol biosynthesis R.; Topol, E. J., Eds.; Lippincott-Raven: Philadelphia, 1996.
from lanosterol have been reported to give rise to diseases. 3. Chowdhury, J. R.; Walkoff, A. W.; Chowdhury, N. R.; Arias,
Mutations have been identified in the genes for three of these I. M. In The Metabolic and Molecular Bases of Inherited Dis-
enzymes. Mutations in C3 sterol dehydrogenase (C4 decarbox- ease, 7th ed.; Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle,
ylase) (conversion of 9 to 10, 14 to 15) give rise to CHILD D., Eds.; McGraw-Hill: New York, 1995; p 2161.
(congenital hemidysplasia, ichthyosis, and limb defects) 4. Goldstein, J. L.; Hobbs, H. H.; Brown, M. S. In The Meta-
syndrome (55) (MIM 308050 [56 ]). CHILD syndrome is bolic and Molecular Bases of Inherited Disease, 7th ed.; Scriver,
characterized by dry, thickened, scaly skin lesions (“fishskin”) C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D., Eds.; McGraw-
on one side of the body; limb defects; calcification points on Hill: New York, 1995; p 1981.
cartilage structures; and internal organ anomalies. 5. Way, L. W. In Cecil Textbook of Medicine, 16th ed.;
Mutations in ∆8–∆7 sterol isomerase (conversion of 16 Wyngaarden, J. B.; Smith, L. H., Eds.; Saunders: Philadel-
to 17) give rise to two forms of chondrodysplasia punctata phia, 1982; p 752.
(CDP) (57–59). CDP describes calcification points on car- 6. Pentchev, P. G.; Vanier, M. T.; Suzuki, K.; Patterson, M. C.
tilage structures and disorders of bone growth and structure In The Metabolic and Molecular Bases of Inherited Disease, 7th
associated with skeletal dwarfism. One form of CDP is ed.; Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D., Eds.;
Conradi–Hünermann syndrome (CDPX2) (MIM 302960 McGraw-Hill: New York, 1995; p 2625.
[56 ]), which arises by mutations in ∆8–∆7 sterol isomerase (58). 7. U.S. News and World Report 2001, 130 (10, Mar 12), 14.
It resembles CHILD syndrome, but the dry, thickened, scaly 8. Voet, D.; Voet, J. G. Biochemistry, 2nd ed.; Wiley: New York,
skin lesions are on both sides of the body in an asymmetric 1995; p 699.
pattern. Because CHILD syndrome and CDPX2 have similar 9. Rilling, H. C.; Chayet, L. T. In Sterols and Bile Acids;
clinical features, a patient diagnosed with CHILD syndrome was Danielsson, H.; Sjövall, J., Eds.; Elsevier: New York, 1985; p
reported to have a mutation in the ∆8–∆7 sterol isomerase gene 33.
(59). Thus CHILD syndrome and CDPX2 were proposed 10. Biochemical Pathways, 3rd ed.; Michal, G., Ed.; Boehringer
to result from allelic mutations in the ∆8–∆7 sterol isomerase Mannheim: Mannheim, Germany, 1993.
gene. However, because CHILD syndrome was shown to arise 11. ExPASy Molecular Biology Server; http://www.expasy.ch (ac-
by mutations in C3 sterol dehydrogenase (C4 decarboxylase) cessed Dec 2001).
(55), the mutation in ∆8–∆7 sterol isomerase (59) may be 12. Bioinformatics Center, Institute for Chemical Research, Kyoto
another form of CDP. Therefore, CHILD syndrome is distinct University. KEGG: Kyoto Encyclopedia of Genes and Genomes;
from Conradi–Hünermann (with unilateral skin lesions and http://www.genome.ad.jp/kegg/ (accessed Dec 2001).
bilateral skin lesions). 13. National Center for Biotechnology Information Home Page;
A deficiency in sterol ∆24-reductase (conversion of 17 http://www.ncbi.nlm.nih.gov (accessed Dec 2001).
to 18) activity leading to an accumulation of desmosterol has 14. Bach, T. J.; Benveniste, P. Prog. Lipid Res. 1997, 36, 197.
been proposed to give rise to desmosterolosis (60). This disease 15. Gachotte, D.; Barbuch, R.; Gaylor, J.; Nickel, E.; Bard, M.
is characterized by abnormally large cranial capacity, cleft Proc. Natl. Acad. Sci. USA 1998, 95, 13794.
palate, ambiguous genitalia, and short limbs. 16. Gachotte, D.; Sen, S. E.; Eckstein, J.; Barbuch, R.; Krieger,
Mutations of 7-dehydrocholesterol reductase (conversion M.; Ray, B. D.; Bard, M. Proc. Natl. Acad. Sci. USA 1999,
of 19 to 20) give rise to Smith–Lemli–Opitz syndrome (SLOS) 96, 12655.
(40, 61–63) (MIM 270400 [56 ]). SLOS is characterized by 17. Fischer, R. T.; Trzaskos, J. M.; Magolda, R. L.; Ko, S. S.; Brosz,
abnormal smallness of the head, poor growth, cleft palate, C. S.; Larsen, B. J. Biol Chem. 1991, 266, 6124.

JChemEd.chem.wisc.edu • Vol. 79 No. 3 March 2002 • Journal of Chemical Education 383


Research: Science and Education

18. Trzaskos, J. M.; Ko, S. S.; Magolda, R. L.; Favata, M. F.; Linck, L. M.; Connor, W. E.; Steiner, R. D.; Porter, F. D. Am.
Fischer, R. T.; Stam, S. H.; Johnson, P. R.; Gaylor, J. L. Bio- J. Hum. Genet. 1998, 63, 55.
chemistry 1995, 34, 9670. 41. Krisans, S. K. Ann. N.Y. Acad. Sci. 1996, 804, 142.
19. Trzaskos, J. M.; Fischer, R. T.; Ko, S. S.; Magolda, R. L.; Stam, 42. Kyte, J. Mechanism in Protein Chemistry; Garland: New York,
S.; Johnson, P.; Gaylor, J. L. Biochemistry 1995, 34, 9677. 1995.
20. Shyadehi, A. Z.; Lamb, D. C.; Kelly, S. L.; Kelly, D. E.; 43. Cytochrome P450: Structure, Mechanism and Biochemistry, 2nd
Schunck, W.-H.; Wright, J. N.; Corina, D.; Akhtar, M. J. Biol. ed.; Ortiz de Montellano, P. R., Ed.; Plenum: New York, 1995.
Chem. 1996, 271, 12445. 44. Sono, M.; Roach, M. P.; Coulter, E. D.; Dawson, J. H. Chem.
21. Lamb, D. C.; Kelly, D. E.; Schunck, W.-H.; Shyadehi, A. Z.; Rev. 1996, 96, 2841.
Akhtar, M.; Lowe, D. J.; Baldwin, B. C.; Kelly, S. L. J. Biol. 45. Newcomb, M.; Toy, P. H. Acc. Chem. Res. 2000, 33, 449.
Chem. 1997, 272, 5682. 46. Wallar, B. J.; Lipscomb, J. D. Chem. Rev. 1996, 96, 2625.
22. Bellamine, A.; Mangla, A. T.; Nes, W. D.; Waterman, M. R. 47. Brown, M. S.; Goldstein, J. L. Cell 1997, 89, 331.
Proc. Natl. Acad. Sci. USA 1999, 96, 8937. 48. Ridgway, N. D.; Byers, D. M.; Cook, H. W.; Storey, M. K.
23. Strömstedt, M.; Rozman, D.; Waterman, M. R. Arch. Biochem. Prog. Lipid Res. 1999, 38, 337.
Biophys. 1996, 329, 73. 49. Osborne, T. F. J. Biol. Chem. 2000, 275, 32379.
24. Ji, H.; Zhang, W.; Zhou, Y.; Zhang, M.; Zhu, J.; Song, Y.; 50. Nohturfft, A.; DeBose-Boyd, R. A.; Scheek, S.; Goldstein, J.
Lü, J.; Zhu, J. J. Med. Chem. 2000, 43, 2493. L.; Brown, M. S. Proc. Natl. Acad. Sci. USA 1999, 96, 11235.
25. Podust, L. M.; Poulos, T. L.; Waterman, M. R. Proc. Natl. 51. Brown, M. S.; Goldstein, J. L. Proc. Natl. Acad. Sci. USA 1999,
Acad. Sci. USA 2001, 98, 3068. 96, 11041.
26. Kim, C.-K.; Jeon, K.-I.; Lim, D.-M.; Johng, T.-N.; Trzasdos, J. 52. Schroepfer, G. J. Jr. Physiol. Rev. 2000, 80, 361.
M.; Gaylor, J. L.; Paik, Y.-K. Biochim. Biophys. Acta 1995, 53. Lund, E. G.; Kerr, T. A.; Sakai, J.; Li, W.-P.; Russell, D. W. J.
1259, 39. Biol. Chem. 1998, 273, 34316.
27. Silve, S.; Dupuy, P. H.; Ferrara, P.; Loison, G. Biochim. Biophys. 54. Russell, D. W. Cell 1999, 97, 539.
Acta 1998, 1392, 233. 55. König, A.; Happle, R.; Bornholdt, D.; Engel, H.; Grzeschik,
28. Bard, M.; Bruner, D. A.; Pierson, C. A.; Lees, N. D.; K.-H. Am. J. Med. Genet. 2000, 90, 339.
Biermann, B.; Frye, L.; Koegel, C.; Barbuch, R. Proc. Natl. 56. Online Mendelian Inheritance in Man (OMIM) Home Page;
Acad. Sci. USA 1996, 93, 186. http://www.ncbi.nlm.nih.gov/omim (accessed Dec 2001).
29. Li, L.; Kaplan, J. J. Biol. Chem. 1996, 271, 16927. 57. Kelley, R. I.; Wilcox, W. G.; Smith, M.; Kratz, L. E.; Moser,
30. Hanner, M.; Moebius, F. F.; Weber, F.; Grabner, M.; Striessnig, A.; Rimoin, D. S. Am. J. Med. Genet. 1999, 83, 213.
J.; Glossmann, H. J. Biol. Chem. 1995, 270, 7551. 58. Braverman, N.; Lin, P.; Moebius, F. F.; Obie, C.; Moser, A.;
31. Silve, S.; Dupuy, P. H.; Labit-Lebouteiller, C.; Kaghad, M.; Glossmann, H.; Wilcox, W. R.; Rimoin, D. L.; Smith, M.;
Chalon, P.; Rahier, A.; Taton, M.; Lupker, J.; Shire, D.; Loison, Kratz, L. Nat. Genet. 1999, 22, 291.
G. J. Biol. Chem. 1996, 271, 22434. 59. Grange, D. K.; Kratz, L. E.; Braverman, N. E.; Kelley, R. I.
32. Moebius, F. F.; Soellner, K. E. M.; Fiechtner, B.; Huck, C. Am. J. Med. Genet. 2000, 90, 328.
W.; Bonn, G.; Glossmann, H. Biochemistry 1999, 38, 1119. 60. FitzPatrick, D. R.; Keeling, J. W.; Evans, M. J.; Kan, A. E.;
33. Bae, S.-H.; Seong, J.; Paik, Y.-K. Biochem. J. 2001, 353, 689. Bell, J. E.; Porteous, M. E. M.; Mills, K.; Winter, R. M.;
34. Bae, S.-H.; Paik, Y.-K. Biochem. J. 1997, 326, 609. Clayton, P. T. Am. J. Med. Genet. 1998, 75, 145.
35. Taton, M.; Husselstein, T.; Benveniste, P.; Rahier, A. Biochem- 61. Fitzky, B. U.; Witsch-Baumgartner, M.; Erdel, M.; Lee, J. N.;
istry 2000, 39, 701. Paik, Y.-K.; Glossmann, H.; Utermann, G.; Moebius, F. F.
36. Rahier, A. Biochemistry 2001, 40, 256. Proc. Natl. Acad. Sci. USA 1998, 95, 8181.
37. Matsushima, M.; Inazawa, J.; Takahashi, E.; Suzumori, K.; 62. Waterham, H. R.; Wijburg, F. A.; Hennekam, R. C. M.;
Nakamura, Y. Cytogenet. Cell Genet. 1996, 74, 252. Vreken, P.; Poll-The, B. T.; Dorland, L.; Duran, M.; Jira, P.
38. Moebius, F. F.; Fitzky, B. U.; Lee, J. N.; Paik, Y.-K.; E.; Smeitink, J. A. M.; Wevers, R. A.; Wanders, R. J. A. Am.
Glossmann, H. Proc. Natl. Acad. Sci. USA 1998, 95, 1899. J. Hum. Genet. 1998, 63, 329.
39. Bae, S.-H.; Lee, J. N.; Fitzky, B. U.; Seong, J.; Paik, Y.-K. J. 63. Yu, H.; Tint, G. S.; Salen, G.; Patel, S. B. Am. J. Med. Genet.
Biol. Chem. 1999, 274, 14624. 2000, 90, 347.
40. Wassif, C. A.; Maslen, C.; Kachilele-Linjewile, S.; Lin, D.; 64. Am. J. Med. Genet. 1997, 68 (3).

384 Journal of Chemical Education • Vol. 79 No. 3 March 2002 • JChemEd.chem.wisc.edu

Vous aimerez peut-être aussi