DataPDF Available

Figures

Overview of physiological methanol biogenesis. This figure summarizes the data on methanol. The methyl group donor SAM is synthesized via the catalytic activity of methionine adenosyltransferase, which transfers the adenosyl group of ATP to methionine (step 1). S-adenosyl homocysteine is formed after SAM transfers a methyl group to a methyl acceptor (step 2) such as DNA; thus methanol is involved in gene regulation (step 11). Methyl esters such as carboxyl methyl esters are unstable and are readily hydrolyzed in neutral and basic pH conditions or by methylesterase to produce methanol (step 3). Other sources of methanol include the human diet, which supplies the methanol-generating pectin/PME complex via fruits and vegetables (steps 4 and 5), aspartame as a synthetic nonnutritive sweetener (step 6) and alcoholic beverages (step 7). The human gut microbiota is a putative methanol source (step 8) and takes part in the generation of human endogenous ethanol (step 9). We suggest that endogenous and dietary methanol may be involved in the regulation of genes involved in the metabolic clearance of methanol (step 10). The first stage of the oxidative metabolism of methanol is executed by the catalase-H 2 O 2 system (step 12), cytochrome P450 (CYP2E1)-mediated oxidation (step 13) and, mainly, the alcohol dehydrogenase I (ADH1) class of enzymes (step 14). Although ADH1 converts methanol into toxic formaldehyde, physiological ethanol in the bloodstream substantively prevents all formaldehyde production from endogenous and dietary methanol in humans (step 15).
… 
Content may be subject to copyright.
METABOLIC METHANOL: MOLECULAR PATHWAYS
AND PHYSIOLOGICAL ROLES
Yuri L. Dorokhov, Anastasia V. Shindyapina, Ekaterina V. Sheshukova,
and Tatiana V. Komarova
A. N. Belozersky Institute of Physico-Chemical Biology, Moscow State University, Moscow, Russia; and N. I.
Vavilov Institute of General Genetics, Russian Academy of Science, Moscow, Russia
LDorokhov YL, Shindyapina AV, Sheshukova EV, Komarova TV. Metabolic Meth-
anol: Molecular Pathways and Physiological Roles. Physiol Rev 95: 603–644,
2015; doi:10.1152/physrev.00034.2014.—Methanol has been historically con-
sidered an exogenous product that leads only to pathological changes in the human
body when consumed. However, in normal, healthy individuals, methanol and its
short-lived oxidized product, formaldehyde, are naturally occurring compounds whose functions
and origins have received limited attention. There are several sources of human physiological
methanol. Fruits, vegetables, and alcoholic beverages are likely the main sources of exogenous
methanol in the healthy human body. Metabolic methanol may occur as a result of fermentation
by gut bacteria and metabolic processes involving S-adenosyl methionine. Regardless of its
source, low levels of methanol in the body are maintained by physiological and metabolic
clearance mechanisms. Although human blood contains small amounts of methanol and
formaldehyde, the content of these molecules increases sharply after receiving even methanol-
free ethanol, indicating an endogenous source of the metabolic methanol present at low levels
in the blood regulated by a cluster of genes. Recent studies of the pathogenesis of neurological
disorders indicate metabolic formaldehyde as a putative causative agent. The detection of
increased formaldehyde content in the blood of both neurological patients and the elderly
indicates the important role of genetic and biochemical mechanisms of maintaining low levels
of methanol and formaldehyde.
I. INTRODUCTION 603
II. TERRESTRIAL METHANOL 605
III. EXOGENOUS SOURCES OF... 607
IV. HUMAN GUT MICROBIOTA AS A... 611
V. METHANOL METABOLISM IN... 615
VI. METHANOL/FORMALDEHYDE... 624
VII. GENES INVOLVED IN ENDOGENOUS... 628
VIII. CONCLUDING COMMENTS 630
I. INTRODUCTION
Robert Boyle first described wood spirits, or methanol, as
the “sowrish spirit” of boxwood pyrolysis in 1661 (44),
and the function of methanol in plant and animal life has
since been unclear. In higher plants, cell wall (CW) pectin
methylesterase (PME) produces methanol by pectin de-
methylation (248). Terrestrial atmospheric methanol
emission comes from volcanoes, H
2
and CO
2
generation
within seafloor hydrothermal systems, and biomass com-
bustion, but PME-mediated emission from plants is most
likely the largest source of methanol in the atmosphere
(512). Methanol accumulates in the intercellular air
space or the liquid pool when the stomata close at night,
and a large quantity of methanol is released when the
stomata open in the morning (193). Gaseous methanol
was traditionally considered to be a biochemical “waste
product.” However, the effects of PME-generated plant
methanol (“emitters”) on plant defensive reactions (“re-
ceivers”) and plant-animal communication have recently
been shown (111, 112).
In humans, methanol is considered to be a poison because
alcohol dehydrogenase 1b (ADH1b) mainly metabolizes
methanol into toxic formaldehyde (63). Methanol itself is
not toxic to animal cells; however, formaldehyde is respon-
sible for carcinogenesis and age-related damage to neurons
in the brain (472).
Until recently, it was believed that the trace amounts of
methanol and formaldehyde in the blood of healthy people
came only from the consumption of fake or low-quality
alcoholic beverages. However, recent data have indicated
that methanol and short-lived formaldehyde are actually
naturally occurring compounds in normal, healthy human
individuals (413). There are several sources of physiological
methanol in humans (FIGURE 1). Fruits, vegetables, and al-
coholic beverages are likely to be the main sources of exog-
enous methanol in the healthy humans. More than 50 years
ago (126), two other sources were suggested: anaerobic
Physiol Rev 95: 603–644, 2015
doi:10.1152/physrev.00034.2014
6030031-9333/15 Copyright © 2015 the American Physiological Society
on April 2, 2015Downloaded from
fermentation by gut bacteria (38, 208, 209, 418) and the
transformation of S-adenosyl methionine (SAM) to metha-
nol by certain metabolic processes (16). Although human
blood contains small amounts of methanol and formalde-
hyde, their contents are sharply increased after receiving
even methanol-free ethanol (413). This indicates the exis-
tence of an endogenous source of low levels of metabolic
methanol in the blood, the regulation of which is controlled
by a cluster of genes (246, 413).
Interest in the literature is mainly focused on the mechanism
of ethanol metabolism (63, 347). Ethanol consumption, in
small amounts, can have a beneficial effect (147), but etha-
nol abuse leads to pathological changes in the human body
(355). In the past, methanol was mainly considered an ex-
ogenous product; the consumption of methanol was be-
lieved only to lead to pathological changes in the human
body (453). Researchers did not pay attention to either the
role or the origin of endogenous methanol and formalde-
hyde in humans. Recent studies of the pathogenesis of neu-
rological disorders indicate that metabolic formaldehyde is
a putative causative agent of human pathology (464, 466).
The detection of increased formaldehyde content in the
blood of not only neurological patients but also elderly
people (463) indicates the important role of genetic and
biochemical mechanisms in maintaining low levels of meth-
anol and formaldehyde.
This review will cover the wide spectrum of metabolic
methanol phenomena, both physiological and patholog-
ical, including the origin of human metabolic methanol,
the processes of methanol conversion and clearance, and
the roles of methanol and formaldehyde in human pa-
thology. Ultimately, an imbalance between the inflow
and outflow of formaldehyde may result in the onset of
pathologies.
Regulation of genes
involved in methanol
metabolic clearance
Methanol
Catalase-
H2O2 system
ADHI
Ethanol
Alcoholic
beverages
Aspartame as
a sweetener
Gut
microbiota
Fruits &
vegetables
Pectin/PME
complex
Methionine
+ATP
SAM
methyl
acceptor
1
2
3
4
5
6
7
8
9
10
11
14
15
Formaldehyde
Formic acid
CO2 + H2O
CYP2E1
12 13
FIGURE 1. Overview of physiological methanol biogenesis. This figure summarizes the data on methanol.
The methyl group donor SAM is synthesized via the catalytic activity of methionine adenosyltransferase,
which transfers the adenosyl group of ATP to methionine (step 1). S-adenosyl homocysteine is formed after
SAM transfers a methyl group to a methyl acceptor (step 2) such as DNA; thus methanol is involved in gene
regulation (step 11). Methyl esters such as carboxyl methyl esters are unstable and are readily hydrolyzed
in neutral and basic pH conditions or by methylesterase to produce methanol (step 3). Other sources of
methanol include the human diet, which supplies the methanol-generating pectin/PME complex via fruits
and vegetables (steps 4 and 5), aspartame as a synthetic nonnutritive sweetener (step 6) and alcoholic
beverages (step 7). The human gut microbiota is a putative methanol source (step 8) and takes part in the
generation of human endogenous ethanol (step 9). We suggest that endogenous and dietary methanol
may be involved in the regulation of genes involved in the metabolic clearance of methanol (step 10). The
first stage of the oxidative metabolism of methanol is executed by the catalase-H
2
O
2
system (step 12),
cytochrome P450 (CYP2E1)-mediated oxidation (step 13) and, mainly, the alcohol dehydrogenase I
(ADH1) class of enzymes (step 14). Although ADH1 converts methanol into toxic formaldehyde, physio-
logical ethanol in the bloodstream substantively prevents all formaldehyde production from endogenous
and dietary methanol in humans (step 15).
DOROKHOV ET AL.
604 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
II. TERRESTRIAL METHANOL
A. Methanol in the Earth’s Atmosphere
Methanol is an ubiquitous, biogeochemically active com-
pound and a significant component of the volatile organic
carbon in the atmosphere (177, 199, 239, 356, 390, 519).
The atmosphere contains 4 Tg (teragrams, 10
12
g) of
methanol (177). Terrestrial atmospheric methanol emis-
sions come from different sources, including volcanoes and
H
2
and CO
2
generation within seafloor hydrothermal sys-
tems (512). Another source of atmospheric methanol is bio-
mass combustion (131, 315), in which the wood pyrolysis
of plant fibers (i.e., cellulose and lignin) induces methanol
emission (149, 300, 515). Wood pyrolysis, i.e., the decom-
position of wood at elevated temperatures in the absence of
oxygen, was the method that allowed Robert Boyle to ob-
tain the “sowrish spirit” (44) from boxwood; much later,
this became known mainly as methanol and acetone. How-
ever, volatile organic compound emissions from plants are
most likely the largest sources of methanol in the atmo-
sphere (193, 407). After CO
2
and isoprene, gaseous meth-
anol is one of the most abundant carbon-containing volatile
organic compounds in the atmosphere (156, 407).
B. Plant-Made Methanol
Plant-emitted gaseous methanol is an abundant, volatile
organic compound that was considered for a long time to be
a waste product of plant metabolism. Now, the diverse
biological effects of methanol have been discovered and
demonstrated. The main source of plant methanol release is
the above-ground parts of the plant (353). Plant methanol
accumulates in the leaf intercellular spaces in either a dis-
solved or a gaseous state when the stomata are closed at
night. When the leaf stomata open in the morning, a signif-
icant amount of methanol is released (193, 343). Plant
methanol release occurs with varying intensity depending
on the stage of plant development (343), time of day (193),
and other conditions. The quantity of the leaf-emitted
methanol varies from a minimum of 0.38
g·g fresh weight
(FW)
1
·h
1
at night to a maximum of 7
g·g FW
1
·h
1
in
the morning, while the average value is 3
g·g FW
1
·h
1
.
It should be noted that under certain types of stress, meth-
anol emission is dramatically increased to up to 100
g·g
FW
1
·h
1
. In addition, damaged plants emit significant
amounts of methanol (48, 230). For example, in alfalfa
fields, high levels of gaseous methanol were detected after
mowing, and methanol emission levels continued to rise
over the next 3 days (505).
Plant tissues have been shown to metabolize methanol
(115). The majority of endogenous methanol reaches the
leaf surface and evaporates, and a minor amount is nonen-
zymatically oxidized to formaldehyde, which could later be
involved in the synthesis of serine, methionine, and phos-
phatidylcholine. In addition, methanol could be enzymati-
cally oxidized to CO
2
and then directed to the Calvin cycle
(151). Methanol metabolism in plants can be accompanied
by significant increases in biomass; in some C3 plants, this is
often accompanied by an increased photosynthetic effi-
ciency and developmental rate (151, 346). Moreover, plant-
generated methanol could be involved in leaf growth during
plant development (247). Small amounts of methanol emit-
ted through the stomata are oxidized to carbon dioxide by
methylotrophic bacteria either directly on the leaves or later
in the soil (244). In general, methanol is a rather stable
substance under normal conditions with a half-life of 10
days (199).
Plants produce methanol in the CW pectin de-methyl-ester-
ification reaction (FIGURE 2). Pectin is a matrix-forming
component of the plant CW. Cellulose and xyloglucan fi-
brils are immersed in the pectin matrix (90). Pectin is mainly
formed from homogalacturonan blocks of
-1,4-linked ga-
lacturonic acids, but rhamnogalacturonan I, rhamnogalac-
turonan II, and xylogalacturonan are the other major com-
ponents of the pectin matrix (165, 369). Pectin homogalac-
turonan is a compound that is synthesized by plant cells and
secreted into the CW in a highly methyl-esterified form
(FIGURE 2).
During the growth and development of cells or CW-me-
chanical damage pectin in the CW can be de-methyl-esteri-
fied, resulting in the production of methanol and the accu-
mulation of negatively charged galacturonic acid residues
(FIGURE 2). Due to the interaction of Ca
2
with the charged
carboxyl groups, the rigidity of the CW can be increased if
the de-methylated residues are arranged in blocks. On the
other hand, in the case of the random de-methyl-esterifica-
tion of pectin, CW softening can occur, i.e., homogalac-
turonan becomes available for hydrolysis by pectin-degrad-
ing enzymes of the CW, resulting in the initiation of pectin
cleavage (511). Both PME-associated modulations of the
CW and the degree of pectin methylation play essential
roles in mediating the plasticity of the CW, cellular adhe-
sion, ionic composition, and pH (371). All of these aspects
are very important in the growth and development of
plants, pollination, and fruit ripening, as well as in the re-
sponses of plant cells to stress factors.
While they are conserved proteins, plant PMEs comprise a
multigene family (371). Some PME genes are ubiquitously
expressed, while others are specifically expressed during
fruit ripening, microsporogenesis, and germination of the
pollen grain, or stem elongation (304). Most higher plant
PMEs are synthesized as a precursor protein that contains
an NH
2
terminus pre-pro-sequence, which is necessary for
the correct processing to form a mature enzyme and deliver
it to the CW. The pre-sequence, which may be presented
with a signal peptide and/or a transmembrane domain, is
METABOLIC METHANOL
605Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
cleaved at the stage of precursor delivery through the endo-
plasmic reticulum to the CW (290, 303), and the pro-se-
quence is removed at the next step of maturation (114,
513). Only group 2 PMEs possess the pro-sequence, which
is necessary for the correct folding of the enzyme and is
believed to function as an intramolecular chaperone (318,
371). In addition to the importance of PMEs during pro-
cesses of plant growth and development, PMEs play a sig-
nificant role in the protection of plants against external
stresses. PMEs are involved in modifications of the CW,
which is a natural barrier that protects and separates the cell
from the environment. Aside from the direct effects of
PMEs on the CW, the enzymes also act indirectly through
different compounds, including the methanol released via
pectin de-esterification (488a, 112, 371). Pectin with a
higher degree of methylation provides better resistance to
fungal and bacterial pectolytic enzymes. Thus plants that
are transgenic for the PME inhibitor (PMEI) gene exhibit
increased resistance to pathogenic fungi and bacteria com-
pared with wild-type plants (282, 283, 488). On the other
hand, PME is activated in response to the invasion of patho-
genic fungi and bacteria, and increased PME activity occurs
as a part of pattern-triggered immunity (31). CW integrity
damage by pathogenic pectolytic enzymes results in the ac-
cumulation of oligogalacturonide fragments, which play a
role in damage-associated molecular patterns (140). Oli-
gogalacturonide de-methylation by PMEs leads to the en-
hancement of their activity as elicitors and, thus, the im-
provement of plant defense responses to the invasion of
pathogenic fungi and bacteria (31, 383). With regard to
viral pathogens, PMEs have been shown to participate in
the virus cell-to-cell movement (74), which is necessary for
successful viral infection. PME is known to interact with the
movement protein of tobacco mosaic virus (TMV) (75,
113), and this interaction is required for the intercellular
transport of TMV. The important role of PME in the
spreading of viruses through the plant was confirmed by the
fact that Arabidopsis transgenic for the PMEI gene exhibits
decreased intercellular transport of TMV and lower suscep-
tibility to the virus (284). On the other hand, PMEs were
also shown to enhance virus-induced gene silencing (113)
and to indirectly interfere with nucleocytoplasmic transport
(245).
One of the indirect ways by which PME affects host-patho-
gen relationships is via gaseous methanol that is formed
during the pectin demethylation reaction in response to in-
jury. The effects of PME-generated methanol from plant
emitters on the defensive response of neighboring plant re-
ceivers were recently studied (111). It was shown that an
increase in methanol emission from PME-transgenic or
nontransgenic wounded plants restrains the growth of the
pathogenic bacteria Ralstonia solanacearum in neighboring
receiver plants. However, at the same time, plants exposed
to gaseous methanol became more sensitive to the viral
infection, likely due to the general activation of intercellular
transport. Such antibacterial resistance and virus suscepti-
bility are accompanied by increased activities of the genes
responsible for stress control and intercellular communica-
tion in receiver plants. These results led to the conclusion
that methanol is a signaling molecule that is involved in the
intra- and interplant communication (111, 247, 248). The
first portion of methanol emitted by damaged leaves is likely
produced by PME that is preexisting in the CW; this allows
for rapid methanol release into the atmosphere (250, 488a).
However, simultaneous de novo PME synthesis is induced,
which allows a high level of methanol emission to be main-
tained (111). Based on the available data, we can conclude
that methanol, which is released into the air by damaged
plants or plants attacked by herbivorous insects, serves as
O
O
OOCH
3
C
OH
COO
HO
HO HO
O
O
PME
MeOH
PME
O
OO
O
CH3
C
OH
COO
HO
HO HO
O
O
O
O
O
OO
C
OH
COO
HO
HO HO
O
O
O
OO
O
C
OH
COO
HO
HO HO
O
O
O
FIGURE 2. Demethylesterification of pectin ho-
mogalacturonan by PME with methanol formation.
DOROKHOV ET AL.
606 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
an alarm to help neighboring plants or adjacent leaves pre-
pare for defense.
C. Roles of Methanol and Ethanol in Plant-
Animal Communication
The role of methanol in plant-herbivore relationships re-
mains rather controversial. As was mentioned above, plants
release methanol during their growth and development, as
well as in response to wounding. Gaseous methanol plays a
role in signaling in the plant community and is also likely to
play a role in plant-animal communication. Methanol is
toxic to some insects (106, 288) and mediates defense reac-
tions against some herbivores (22). On the other hand, gas-
eous methanol attracts certain insects such as bark beetles
(Hylurgops palliatus, Tomicus piniperda, and Trypoden-
dron domesticum), while long-chain alcohols do not act as
attractants (56). Methanol emission occurs in response to
Manduca sexta larvae feeding on Nicotiana plants and ap-
pears to play a beneficial role for that insect: the treatment
of plants with the same amount of methanol as is released
during larvae feeding led to decreased plant defense reac-
tions and increased larvae attacking performance (488a).
Thus methanol release may induce metabolic changes that
influence the susceptibility of plants to herbivores. Further-
more, laboratory mice prefer the odor of methanol to that
of other plant-emitted volatiles or ethanol. Methanol emit-
ted by wounded plants was shown to be an attractant: mice
in a two-choice Y-maze chose methanol over other plant
volatiles. Moreover, inhalation of vapors released by
wounded plants led to increased blood methanol levels in
mice and changes in the expression profiles of the so-called
methanol-responsive genes in the brain (112, 246). This
discovery led to the conclusion that methanol emitted from
damaged plants should be regarded not only as a signal of
plant-to-plant communication but also as a plant-to-animal
cross-kingdom signaling molecule that modulates both be-
havior and mammalian gene expression (112).
As for ethanol, another alcohol, its increased emission
mainly accompanies fruit ripening due to the fermentation
of sugars by yeasts. Ethanol, among other volatiles, is an
important component of the scent of ripe fruit. Moreover,
an optimal ethanol concentration is a significant indicator
of the quality of fresh citrus fruit (331, 351). However, it
seems that increased ethanol levels in fruit adversely af-
fected the taste (80). Dominy (108) analyzed the ethanol
content of some Asian fruits and showed that small
amounts of ethanol (from 0.005 to 0.48%) were detected in
fruits of all developmental stages; the ethanol levels posi-
tively correlated with the soluble sugar levels. In this case,
ethanol is regarded as an olfactory cue for frugivores to
locate food of good quality with high carbohydrate content.
Frugivores such as Egyptian fruit bats (Rousettus aegyptia-
cus) were shown to locate fruits by assessing the degree of
ripeness and quality based on the intensity of the scent of
alcohol (methanol and ethanol). However, the Egyptian
fruit bats avoided overripe fruits, or fruits with ethanol
contents exceeding 1% (401). Moreover, yellow-vented
bulbuls (Pycnonotus xanthopygus) were shown to decrease
their food intake by 36% if the ethanol concentrations in
fruits exceeded 3% (309). The frugivorous tropical butter-
fly Bicyclus anynana locates fruits by using ethanol odor
cues as long range signals that guide the butterflies in the
direction of food sources containing soluble sugars (101).
An interesting relationship has been observed between the
bertam palm (Eugeissona tristis) and its pollinator, the pen-
tailed tree shrew (Ptilocercus lowii). The palm exudes nec-
tar from its flowers when the petals are still closed. This
nectar contains a high ethanol concentration and is pro-
duced by the plant for periods of up to 46 days. This facil-
itates the development of yeasts living in the nectar. Thus
these flowers release a strong ethanol odor that attracts tree
shrews, which then act as pollinators (509).
Thus plant-associated methanol and ethanol both play roles
in plant-animal communication as either attractants or re-
pellents, depending on the particular case.
III. EXOGENOUS SOURCES OF
PHYSIOLOGICAL METHANOL IN
HUMANS
A. Methanol Content in Healthy Human
Blood and Breathing Air
Methanol and ethanol are integral components of life for
humans and mammals (299). Healthy human blood con-
tains small amounts of metabolic methanol (112, 246, 413)
and ethanol (269, 362). Gaseous methanol and ethanol are
also detected in the air exhaled by healthy people (126, 212,
213, 474). The methanol concentration in the blood is 400 –
1,000 times less than the toxic concentration (214, 251) and
has been estimated by different researchers to range from
0.20 0.035 to 5.37 0.08 mg/l (24, 25, 78, 84, 112, 385,
413). Methanol is also detected in the urine, saliva (263),
and breast milk (370). The methanol concentration in ve-
nous blood is, on average, 3 times higher than that in saliva
(413) and 1.3 times lower than that in urine (263).
B. Plant Food as a Source of Exogenous
Methanol
Fruits and vegetables are the main sources of methanol in
the human body. Methanol is known to accumulate in fruit
during ripening as a result of PME activity, which among
other CW degrading enzymes plays a significant role in fruit
softening and ripening (367, 384). Methanol is detected
among other volatiles in the headspace of cut fruits (133).
Freshly squeezed citrus juices contain an average of 2040
mg/l methanol, whereas methanol is detected in a lower
METABOLIC METHANOL
607Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
concentration of 6.2 mg/l in processed juices or juice con-
centrates (293).
The consumption of vegetables such as leafy salads results
in increased blood methanol contents, by 30% (112). The
ingestion of pectin, fruits with different degrees of ripeness,
or fruit juices leads to the rise of the methanol content of
exhaled air (126, 451, 474) and blood (278). This occurs
mainly due to the demethylation of pectin contained in
vegetables by PME, which is also present in the plant CW.
When citrus pectin in the presence of PME activity was
administered to mice, their blood methanol levels increased
by more than 15-fold compared with those administered
pectin without PME activity just 10 min after ingestion
(112). Pectin demethylation is also believed to occur in the
gut by the gut microbiota (418), as some strains of gut-
inhabiting bacteria possess pectolytic enzymes. However,
the administration of pectin containing no active PME in
the gastrointestinal tract of mice did not lead to significant
differences in the concentrations of methanol in their blood
compared with the control; thus it can be concluded that the
contribution of the intestinal microflora in methanol gener-
ation from pectin is rather low compared with the methanol
production from ingested plant food by PME (112).
Pectin is a water-soluble fiber that is resistant to hydrolysis
by gastrointestinal tract enzymes but can be fermented by
the microflora in the large intestine. This fermentation re-
sults in the formation of short-chain fatty acids, which are
then absorbed and metabolized (232). The positive effect of
apple pectin on the human gastrointestinal microflora was
demonstrated: the daily intake of two apples increases the
amount of Bifidobacterium and Lactobacillus, most strains
of which are capable of utilizing pectin (414). Moreover,
the daily consumption of whole apples, pomace, or unclari-
fied juice, i.e., products containing water-soluble fibers such
as pectin, was shown to lead to lower levels of low-density
lipoproteins in the blood (389, 456). This positive effect of
dietary pectin can be explained by the ability of pectin to
inhibit the reabsorption of cholesterol and bile acids in the
lower gastrointestinal tract due to the good gel-forming
properties of pectin, as well as by the ability of short-chain
fatty acids to decrease cholesterol biosynthesis by the liver
(207).
The beneficial effects of plant food are generally recognized,
and according to the World Health Organization’s (WHO)
recommendations, the minimal daily intake of plant food
should not be less than 400 g (250 g vegetables and 150 g
fruits). Fresh fruits and vegetables are particularly rich in
pectin complexed with enzymatically active PME. Despite
that PME is a rather thermostable enzyme, its activity sig-
nificantly decreases after boiling (112). On the other hand,
more intense pectin deesterification and methanol forma-
tion occurs after the moderate heating of plant food to
45–65°C (blanching) (5). Pectin, mainly from apples and
citrus fruits, is used as a dietary supplement to improve
digestion and also as a therapeutic and prophylactic agent
in several diseases, including atherosclerosis (52, 468) and
cardiovascular diseases (458), due to its role in slightly de-
creasing blood cholesterol levels (51, 160). Pectin is also
believed to lower the risk of cancer by decreasing tumor cell
proliferation (30, 477), inducing apoptosis (198), and sup-
pressing the metastatic ability of tumors (146). Notably,
citrus pectin consumption led to significantly increased
blood methanol levels in volunteers; this resulted in changes
in the expression profiles of genes in human white blood
cells, as demonstrated by microarray analysis (413). Pectin
can modulate detoxifying enzymes, stimulate the immune
system, modulate cholesterol synthesis, and act as an anti-
bacterial, antioxidant, or neuroprotective agent (261).
With consideration of the recommendations of nutritionists
to include a large proportion of plant food and vegetable
fiber in the diet, as well as the popularity of vegetarianism,
it can be concluded that fresh fruits and vegetables are two
of the main sources of exogenous methanol in human
blood.
C. Plant Food as a Source of Ethanol
Plant food, mainly ripe fruit, can also be a source of ethanol.
It is known that plants are able to synthesize ethanol mainly
via anaerobic fermentation by alcohol dehydrogenases
(ADHs) (67, 141, 237, 329). Transcription from ADH pro-
moters is induced in response to a lack of oxygen and to
cold stress in maize and Arabidopsis (107). When respira-
tion is impeded or inhibited, pyruvate is formed as a result
of glycolysis. Then, pyruvate is converted to lactate by lac-
tate dehydrogenase. Pyruvate decarboxylase is a key en-
zyme in the alternative pyruvate metabolic pathway, which
results in acetaldehyde formation followed by its reduction
to ethanol by ADH. Ethanol is much less toxic to plant cells
than lactate or acetaldehyde. This metabolic switch allows
plant cells to continuously regenerate ATP and NAD
, but
is less effective than normal conditions (437). Usually, the
Michaelis constant (K
m
) of ADH for acetaldehyde is less
than or equal to 1 mM, while the K
m
for ethanol is more
than 10-fold higher (33). Thus one of the main functions of
ADH in plant cells is the disposal of toxic acetaldehyde.
Another function of ADH is the synthesis of C6 and C9
volatiles that define fruit flavor (406), which attracts fruit-
eating, seed-spreading animals (frugivores). Thus the roles
of ADH in plants include its participation in ethanolic fer-
mentation in hypoxic or anoxic conditions and, in normal
aerobic conditions, in fruit and seed ripening to allow for
the production and emission of the characteristic scent cues
used by pollinators or frugivores. Nowadays, to prolong the
storage lengths of fruits and vegetables, they are kept in a
low-oxygen atmosphere (for example, with an increased
CO
2
content) and are covered with special waxes that also
induce hypoxia. These treatments lead to increased ethanol
DOROKHOV ET AL.
608 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
contents in the stored fruits and vegetables (11, 12, 80).
Notably, the methanol concentrations in such fruits and
vegetables also rise (11, 12).
Still, the main source of ethanol from fruits is the fermen-
tation of fruit sugars by yeast. The increased ethanol con-
tent in fruits during ripening is likely to be an evolutionary
adaptation because raised ethanol concentrations restrain
active bacterial propagation (118). Thus ethanol is ubiqui-
tous in ripe fruits, and its concentration ranges from 0.04 to
0.72% (7–125 mM) (117, 270). Thus frugivorous animals
consume significant amounts of ethanol, as they consume
ripe fruits.
D. Alcoholic Beverages as Sources of
Ethanol and Methanol
The amount of ethanol obtained by humans from fruits and
vegetables is at least two orders of magnitude less than that
from alcoholic beverages (beer, wine, distilled spirits),
which appear to be the main sources of exogenous ethanol
for people. The first evidence of artificial fermented-bever-
age production comes from China, nearly nine millennia
ago (312). Later, fermented beverages started to appear in
different cultures; today, they are a part of human life.
At present, the production of different alcoholic beverages
is under strict governmental control, and there are regula-
tions that define the maximum acceptable concentrations
(MAC) of congeners (methanol, formaldehyde, acetalde-
hyde, etc.) in produced alcoholic beverages. Methanol has
been detected in beer (6–27 mg/l of product) (153), wine,
and distilled spirits. On the basis of data available since
1949 on the methanol content in different wines, reviewed
by Gnekow and Ough (148), the highest concentrations of
methanol, up to 635 mg/l, were found in Spanish and Italian
red wines. The average methanol content of the analyzed
wines was 100 mg/l of the final product. The maximum
limits for the methanol content in wines allowed by Inter-
national Organization of Vine and Wine are 400 mg/l for
red wines and 250 mg/l for white and rosé wines (354).
Higher methanol concentrations are allowed in distilled
fruit spirits due to the nature of the raw material, which is
derived from pectin-rich fruits. As was mentioned above,
PME activity increases during fruit ripening; thus the meth-
anol content is elevated. Methanol released from pectin
after demethylation accumulates in ripe fruits and ends up
in the final product, distilled fruit spirits (34) Under labo-
ratory conditions, it was shown that fruit spirits made from
cherries and plums could contain from 18.3 to 25 g meth-
anol/l of the 100% vol ethanol (405). Undoubtedly, such
high methanol contents exceed the MACs. In the United
States, the allowed methanol concentration for distilled
fruit spirits is 6–7 g/l of 100% vol ethanol (34, 477a). Since
2008, there have been EU limits on methanol (per liter of
100% vol ethanol) in different distilled spirits, as follows:
0.1 g for vodka; up to 1.35 g for some fruit spirits; 1.5 g for
fruit marc spirits; and 2 g, the highest allowed methanol
content, for wine spirits and brandy (130a).
Thus the methanol concentration in alcoholic beverages is
strictly controlled by legislation, and manufacturers are ea-
ger to reduce this concentration in their beverages. Despite
this, shockingly, when an individual drinks wine or distilled
spirits even with very low or undetectable methanol con-
tents, the endogenous methanol level increases significantly
and is comparable to the increased ethanol concentration
that results from ingesting ethanol (413). Earlier, it was
shown that the ingestion of methanol-free whisky or grain
alcohol by a volunteer led to increased methanol content in
his breath (451). Healthy volunteers who drank a glass of
red wine (150 ml) containing 13.7% vol of ethanol and 33
mg/l of methanol showed nearly twofold increases in their
blood methanol content. Moreover, the orders of magni-
tude for methanol and ethanol concentrations were the
same or at least comparable 60–120 min after ingestion.
Furthermore, when 50–90 ml per person (1 ml of 40% vol
ethanol per 1 kg of weight) of methanol-free 40% vol eth-
anol was ingested by the same volunteers in another exper-
iment, the methanol contents in their blood were also of the
same order of magnitude as ethanol and were even compa-
rable 90 min after ingestion, 350400
M (413). Thus
the ingestion of alcoholic beverages always leads to in-
creased blood methanol levels, regardless of how low the
methanol content is in the beverage. This likely happens due
to the partial depletion of liver ADH by ethanol, while a
methanol level continues to rise via endogenous sources.
From these data, the level of endogenous methanol pro-
duction can be estimated to be at least 1.7 mg·kg
1
·h
1
(413). The above-mentioned methanol concentrations
observed after red wine or 40% vol ethanol consumption
or even 5-fold higher methanol contents could be toler-
ated without negative consequences; only a nearly 20-
fold higher concentration is regarded as dangerous (364).
Despite the fact that alcoholic beverages contain ethanol,
which “distracts” ADH from metabolizing methanol,
blood formaldehyde levels also rise significantly after
ethanol ingestion, but with slight retardation; they start
to increase 60 min after ingestion and reach concentra-
tions comparable to methanol and ethanol 60–90 min
after red wine or 90–120 min after ethanol consumption
(413). Thus the ingestion of alcoholic beverages, even
those of high quality, always leads to increased ethanol,
methanol, and formaldehyde contents in the blood.
Moreover, plant food, which is undoubtedly believed to
be healthy, also results in increased methanol and form-
aldehyde levels in the blood. Taking into account that
plant food was the main food source during human evo-
lution, such oscillations of methanol and formaldehyde
concentrations might have some beneficial effects in hu-
mans.
METABOLIC METHANOL
609Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
E. Aspartame
Aspartame, a synthetic nonnutritive sweetener, is a dipep-
tide composed of aspartic acid and phenylalanine methyl
ester. Compared with sucrose, aspartame is 200-fold
sweeter, which allows it to be consumed in much lower
doses to make foods or beverages sweet; for this reason,
aspartame is regarded as a low-calorie sweetener (298). As
aspartame is a methyl ester, its degradation could result in
methanol formation. The stability of aspartame depends on
the storage conditions: it becomes unstable after heating in
aqueous solutions and at neutral or alkaline pH. None of
the degradation products of aspartame is sweet. Thus as-
partame is usually used in dry form or in carbonated bev-
erages and is not recommended for heat-treated foods. Fur-
thermore, the appropriate storage conditions and the expi-
ration dates of foods and beverages with aspartame are very
important. Despite the fact that aspartame was approved as
a safe sweetener by the United States Food and Drug Ad-
ministration (FDA) in 1996 (136, 477b) and in Europe in
1994 (130), the debates around this compound have not
subsided, and research into the effects of aspartame on hu-
man health are in progress; the question of its safety re-
mains acute (423). The main safety concern regarding as-
partame is the formation of methanol as a result of the
inevitable degradation of phenylalanine methyl ester. As
methanol is essentially converted to formaldehyde in mam-
malian organisms, excessive aspartame consumption may
be hazardous because of its contribution to the formation of
formaldehyde (467).
Currently, the acceptable daily intake (ADI) of aspartame in
Europe is 40 mg/kg body weight (bw) and in the United
States is 50 mg/kg bw. The highest aspartame consumption
is observed in teenagers and diabetics; the maximum intake
was estimated at 13.6 mg·kg bw
1
·day
1
for teenage girls
consuming high amounts of soft drinks (9), which, never-
theless, is at least two times lower than the ADI. For chil-
dren and adults with diabetes, the daily consumption of
aspartame in the worst-case scenario also did not exceed the
ADI (298). When volunteers consumed 600 mg aspartame
eight times a day at 1-h intervals, there were no detectable
differences in their blood methanol levels (434). It is worth
mentioning that the average body weight of the volunteers
was 70.8 13.5 kg; thus the doses of aspartame used in the
experiment exceeded the present ADI. According to Ste-
gink’s (433) data, the ingestion of 34 mg/kg bw of aspar-
tame (which is very close to the ADI) did not lead to a
detectable increase of blood methanol levels, likely due to a
lack of sensitivity of the detection method. Only doses from
100 mg/kg bw (twice as high as the ADI) resulted in detect-
able rises of blood methanol levels up to a mean value of
12.7 mg/l (433). Notably, a comparable methanol concen-
tration of 11 mg/l in the blood is obtained from endoge-
nous sources after methanol-free 40% vol ethanol ingestion
(413). In another study, the ingestion of 7.5–8.5 mg/kg
aspartame (the FDA estimation for the average daily con-
sumption) resulted in a mean rise over endogenous values of
1.06 mg/l serum (which is 0.5 mg/l blood), but this in-
crease is of the same order of magnitude as the variations in
endogenous methanol levels and between-individuals dif-
ferences (95). Thus it can be concluded that moderate as-
partame consumption within the recommended doses leads
to a rise in the blood methanol content, similar to the con-
sequence of plant food, juice, or alcoholic beverage inges-
tion.
F. Food as an Exogenous Source of
Acetaldehyde and Formaldehyde
Aside from being a source of ethanol and methanol, food
may also be an exogenous source of acetaldehyde and form-
aldehyde. For example, the acetaldehyde content in apples
ranged from 0.4 to 2.3 mg/kg, in bananas ranged from 1.88
to 18.27 mg/kg, in fresh orange juice was 5.89 mg/kg, and
in fermented dairy products, such as fruit-flavored yogurt,
was up to 17 mg/kg. Daily acetaldehyde consumption with
food (excluding ethanol) is estimated to be 40
g/kg bw,
which is considerably lower than the exposure from ethanol
consumption or tobacco smoking (476). The average daily
exposure to acetaldehyde from alcoholic beverages was es-
timated at 0.112 mg/kg bw (257). Significant amounts of
acetaldehyde (up to 37 mg/l) were detected in different sam-
ples of beer, whiskey, and other distilled spirits (up to 10
mg/l) (440).
Some food could be also a source of exogenous formal-
dehyde. The formaldehyde contents in numerous prod-
ucts were analyzed with chromatographic methods. In
different fruits and vegetables, the formaldehyde concen-
trations ranged from 3 to 26 mg/kg (469). High levels of
formaldehyde were detected in fish of the Gadidae fam-
ily, ranging from 6.4 to 293 mg/kg (32); in milk (0.027
mg/kg in fresh milk and 0.164 mg/kg in processed milk)
(222) and milk products [in Italy, formaldehyde is used
as a bacteriostatic agent in cheese production (391)]; and
in coffee [3.44.5 mg/kg in commercially brewed and
10–16.3 mg/kg in instant coffees (171)]. Moreover, nat-
ural antioxidants that are O- and N-demethylated by
cytochrome P450 in humans (125) could be regarded as a
source of formaldehyde. The same applies to some drugs
such as those containing codeine (98, 490). According to
various assessments, the daily intake of direct formalde-
hyde is estimated to be in the range of 1.5–14 mg/day for
the average adult (98).
One significant non-food source of exogenous methanol
and formaldehyde for humans that cannot be ignored is
cigarette smoke, which has an estimated methanol content
of 180
g/cigarette and formaldehyde content of 45–73
mg/cigarette (157, 302).
DOROKHOV ET AL.
610 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
G. Food Modulating Activity of Alcohol-
Metabolizing Enzymes
Aside from the food- and beverage-related intake of alco-
hols (methanol and ethanol) and aldehydes, changes in the
contents of these compounds can also result from ADH and
aldehyde dehydrogenase (ALDH) activity modulation.
Agents that affect these enzymes could also be food compo-
nents. Plants that contain inhibitors or activators of ADH
and ALDH are used in traditional Chinese medicine to re-
lieve hangovers or to treat alcoholism. Recently, several of
these phytotherapy-based products also appeared in West-
ern societies (313). The most popular herb for treating al-
coholism is Kudzu (Pueraria lobata). It contains daidzein
(4=,7-dihydroxyisoflavone), puerarin (8-C-glucoside of
daidzein), and daidzin (7-glucoside of daidzein), which re-
versibly inhibits mitochondrial ALDH2 (235). The struc-
ture of the daidzin/ALDH2 complex was resolved at 2.4 Å
and showed that the daidzin isoflavone moiety is placed
very closely to the substrate-binding site and glycosyl to a
hydrophobic patch immediately outside the isoflavone-
binding pocket (291). Daidzein is a dietary phytoestrogen
that is also found in soy products. Soy products that are
popular in Asia contain another isoflavone, genistein,
which together with daidzin lowers ethanol and acetalde-
hyde concentrations in the blood due to the enhancement of
ethanol metabolism; this was shown in rats that ingested
soy milk and fermented soy milk together with ethanol
(225). In humans, daidzein is metabolized to daidzin by
bowel bacteria (313). High levels of serum daidzein, 240
280
M, are characteristic for the Japanese population and
are 20 times higher than those in the United Kingdom
population (330); this difference can be explained by di-
etary preferences.
Thus the risks of acetaldehyde and formaldehyde intoxica-
tion as a result of ethanol consumption or smoking are
higher for the Asian population.
Another plant, Hovenia dulcis, also known as the Japanese
raisin tree, is used in traditional Asian medicine for detox-
ification after ethanol consumption. Extracts from the fruits
of this plant significantly decreased the blood ethanol con-
centration by stimulating the activity of liver ADH, ALDH,
and glutathione S-transferase in mice and rats. Hovenod-
ulinol is believed to be the compound responsible for this
effect (194), but the exact mechanism of its action remains
unknown.
Other foods have also been shown to affect ethanol metab-
olism. In experiments in which mango flesh or mango peel
was administered to mice, followed by ethanol ingestion,
blood ethanol levels were lower by more than 50% com-
pared with the control group. This difference in ethanol
concentration likely occurred due to the activation of ADH
and cytosolic ALDH in the liver (238). The authors as-
sumed that the effect observed could be mediated by fruc-
tose and aspartate from the mango, which produce NAD
and thus stimulate ADH and ALDH.
Some phytophenols were also shown to modulate ethanol
metabolism. For example, vanillin, syringaldehyde, caffeic
acid, and ellagic acid appeared to strongly inhibit liver
ADH1 in mice. When these phytophenols were adminis-
tered to mice together with ethanol, they prevented the
elimination of blood ethanol through the ADH metabolic
pathway (170). This effect led to reduced acetaldehyde ac-
cumulation after ethanol ingestion. Interestingly, all of
these compounds were found in mature whisky, and their
concentrations rose with the time of maturation. These
compounds are also present in different foods: vanillin is
found in vanilla beans and as a flavoring agent in confec-
tions; caffeic acid is found in coffee beans, soy beans, argan
oil, and barley; and ellagic acid is present in blackberries,
cranberries, pecans, pomegranates, raspberries, strawber-
ries, walnuts, grapes, and peaches.
Thus methanol, ethanol, formaldehyde, and acetaldehyde
are natural components found in the human body. People
come into daily contact with all of these compounds via
food and inhaled air.
IV. HUMAN GUT MICROBIOTA AS A
SOURCE OF METHANOL
A. Overview of Methanol Metabolism in
Bacteria
In mammals, the intestinal flora is the most relevant candi-
date for the production of methanol followed by its oxida-
tion to formaldehyde, formic acid, and carbon dioxide. To
prove this, it was recently tested whether the intestinal mi-
crobes in rats generate methanol (246). Removal of the
bowel in rats resulted in a lessened methanol increase in the
blood following the administration of 4-methylpyrazole
into the liver compared with the control group (246). Cur-
rently, a large number of bacteria and several types of yeast
are known to be capable of growing on medium wherein
methanol is the sole carbon source (TABLE 1). Bacteria uti-
lize three classic mechanisms of include methanol metabo-
lism: the Calvin cycle, the ribulose monophosphate cycle,
and the serine cycle (387). First, the most energetically
costly pathway of methanol metabolism begins with the
oxidation of methanol to carbon dioxide, which is then
incorporated into the Calvin cycle as a carbon source. The
process of methanol utilization through the Calvin cycle
was detected for the first time in Micrococcus denitrificans
(91). Later, this process was observed in the bacteria Xan-
thobacter autotrophicus (21), Methylacidiphilum inferno-
rum (358), as well as in bacteria belonging to the NC10
phylum (129). The classical method of methanol oxidation
to carbon dioxide takes place in three stages: the oxidation
METABOLIC METHANOL
611Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
of methanol to formaldehyde, formaldehyde to formic acid,
and formic acid to carbon dioxide (491). In some facultative
methylotrophs, instead of three reactions for converting
methanol to carbon dioxide, formaldehyde is oxidized to
carbon dioxide via the dissimilatory hexulose phosphate
cycle (82). In this section, we focus on bacterial enzymes
that either use methanol as substrate or produce methanol
as a product.
Bacteria oxidize alcohols via a few different groups of en-
zymes. The first group includes mammalian-like ADHs
with different electron acceptors (i.e., NAD, NADP, heme).
ADHs from the aerobic bacterial intestinal flora actively
participate in ethanol metabolism. The intestinal flora in
piglets produces great amounts of acetaldehyde from etha-
nol in vivo (211). Bacteria contain a branch of enzymes that
metabolize ethanol, but ADHs oxidize most of it. Although
ADHs play an important role in methanol metabolism in
mammals, bacteria utilize methanol in alternative ways.
Bacterial NADP-dependent ADH (493), NAD-dependent
ADH (124, 348), quinohemoprotein ADH (70), and mem-
brane-associated quinohemoprotein ADH (412) have little
or no activity toward methanol and mostly oxidize C2–C6
alcohols. An exception is NAD-dependent ADH (EC
1.1.1.1) from Bacillus stearothermophilus, DSM 2334,
which is able to oxidize methanol with an efficacy compa-
rable to that of horse liver ADH (411).
Enzymes that use hydrogen peroxide to metabolize alcohols
(methanol and ethanol) comprise a second group of en-
zymes that catalyze the first step of alcohol oxidation in
bacteria. The first catalase enzyme (EC 1.11.1.6) is likely
the most ancient alcohol-oxidizing enzyme and can be
found in many organisms, from archaea to human. In bac-
teria, catalase is encoded by the katE gene, which is an
ortholog of the human CAT gene. The second catalase en-
zyme (EC 1.11.1.21) is encoded by the katG gene, which is
specifically found in bacteria and fungi. According to the
KEGG database, both genes are present in all phyla of bac-
teria and in more than 1,000 species, highlighting their
crucial role in bacterial alcohol metabolism. Along with
ADHs, bacterial catalases from the intestinal flora partici-
pate in ethanol oxidation and metabolize 26–32% of the
total ethanol in vitro (459). Although the data on methanol
oxidation by catalase in microorganisms are very limited, it
is known that methanol can act as a hydrogen donor in
catalase reaction in some bacteria and yeast (166, 395,
482b). Based on the number of intestinal bacteria that carry
the katG and katE genes (TABLE 1) and the experimental
evidence of methanol utilization by the catalase enzymes,
Table 1. Summary of the most abundant bacteria that carry genes encoding methanol-producing (PME, BioH, PMG) and methanol-
oxidizing enzymes (katG, katE) in stool samples
Bacteria Name
Number of Samples
Containing Bacteria
(Out of 5)
Average Number of
Clones per Sample PME BioH PGM katG katE
Eubacterium rectale 5 3,412
Escherichia coli 5 3,183 ⫹⫹⫹
Bacteroides vulgatus 5 2,293 ⫹⫹
Alistipes finegoldii 5 589
Bacteroides thetaiotaomicron 5 538 ⫹⫹
Parabacteroides distasonis 5 527
Bifidobacterium breve 5 402
Butyrivibrio fibrisolvens 5 156
Eubacterium eligens 5 143 ⫹⫹
Bacteroides fragilis 5 119
Butyrate-producing bacterium SS3/4 472
Ruminococcus albus 359⫹⫹
Clostridium phytofermentans 528
Geobacillus thermoglucosidasius 523 ⫹⫹
Alistipes shahii 518⫹⫹
Xanthomonas campestris 418⫹⫹
Klebsiella pneumoniae 315⫹⫹ ⫹⫹
Enterobacter cloacae 217⫹⫹⫹ ⫹
Bacillus cereus 34⫹⫹
Geobacillus thermodenitrificans 33 ⫹⫹⫹
For the analyses, we took the 16S RNA sequencing data from citizens of New Zealand, United States, Hungary,
Norway, and Hong Kong. The data were obtained from the MG-RAST database, loaded and performed by the
American Gut Project (http://metagenomics.anl.gov/linkin.cgi?project6594).
DOROKHOV ET AL.
612 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
we hypothesize that catalases may participate in methanol
oxidation by the intestinal microflora.
Bacteria mainly oxidize methanol to formaldehyde by the
following methanol dehydrogenases (MDH): cytochrome
c-dependent methanol dehydrogenase, which contains a
pyrroloquinoline quinine cofactor (PQQ-MDH); NAD-de-
pendent methanol dehydrogenase (NAD-MDH) and
NADPH-dependent methanol dehydrogenase (NADPH-
MDH). On the basis of the analysis of bacteria from human
stool samples, three types of bacteria that oxidize methanol
by MDHs were revealed: Methylobacterium extorquens
and Methylobacterium nodulans, using PQQ-MDH, and
Bacillus methanolicus, using NAD-MDH (10). However,
the MDH-containing bacteria are not typical residents of
the human microflora, and few clones of these bacteria were
detected in analyzed stool samples.
Formaldehyde that is produced in the methanol and meth-
ane oxidation reaction is either assimilated in the ribulose
monophosphate (RuMP) or the serine cycle or further oxi-
dized into carbon dioxide, which is then incorporated into
the Calvin cycle (FIGURE 3). In the RuMP cycle, formalde-
hyde and D-ribulose 5-phosphate are condensed by hexu-
lose-6-phosphate synthase to form hexulose 6-phosphate.
Bacillus subtilis (521) and Brevibacillus brevis (528) are
two examples of intestinal flora bacteria that assimilate
formaldehyde via the RuMP cycle. In the serine cycle, serine
hydroxymethyltransferase uses formaldehyde and tetrahy-
drofolate as cofactors to form serine from glycine. The ser-
ine pathway of formaldehyde assimilation was discovered
in Methylobacterium extorquens AM1 (76), a gram-nega-
tive Alphaproteobacteria that can be found in human intes-
tines.
There are two classes of reactions that result in methanol
production: methyl ester hydrolysis and redox reactions
(TABLE 2). Despite the versatility of methanol-producing
reactions, only a few are typical for bacterial metabolism
under normal conditions. On the other hand, some of the
methanol-producing reactions can be found only in a few
species of bacteria. For example, methane monooxygenases
(soluble, sMMO; membrane-bound form, pMMO) are en-
zymes specific to methanotrophs, a small group of bacteria
Methanol
Formaldehyde Formaldehyde
Methane
Formic acid
NAD(P)H + H+, H2O
O2, NAD(P)H + H+
FDMMD/ADH
RuMP
CYCLE
SERINE
CYCLE
GSH-FaDH
FAE
MtdB FolD MtdA
Formate dehydrogenases
sMMO, pMMO
Substrate: 4-MO
Enzyme: 4-MO esterase
Substrate: Pectin
Enzyme: PE
Substrate: PAM
Enzyme: BioH
Carbon dioxide
Methyl esters
CH2 = H4FCH2 = H4MPT
FIGURE 3. Methanol metabolism and production by bacteria. 4-MO, 4-methyl oxaloacetate; PAM, pimelyl-
[acp] methyl ester; FDM, formaldehyde dismutase; MDs, methanol dehydrogenases; ADHs, alcohol dehydro-
genases; sMMO/pMMO, soluble/membrane-bound methane monooxygenase; GSH-FaDH, glutathione-linked
formaldehyde oxidation; FAE, formaldehyde activating enzyme; FolD, 5,10-methylene-H
4
folate dehydroge-
nase/5,10-methenylH
4
folate cyclohydrolase.
METABOLIC METHANOL
613Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
able to grow on methane as the sole carbon and energy
source (163).
Methanol concentrations in the blood of volunteers of dif-
ferent ages, genders, diets, as well as smoking and alcohol
habits are very similar. Additionally, methanol accumulates
with similar kinetics in humans and mice after the admin-
istration of ADH inhibitors. Both parameters indicate that
endogenous sources of methanol work constitutively under
normal conditions. Here, we will review the reactions that
result in methanol production and 1) occur under normal
conditions, 2) are common among intestinal bacteria, and
3) work constitutively.
B. Bacterial Pectin Methylesterase
Pectin methylesterase (PME; EC 3.1.1.11) is generally asso-
ciated with plant physiology and CW structure. PME from
bacteria is of particular interest due to its role in the degra-
dation of the CW in plants (260) and its various industrial
applications (187). Additionally, bacterial PME is an essen-
tial part of pectin fermentation in humans because our en-
zymes cannot break down pectin (514). PME is one of the
most abundant enzymes participating in methanol produc-
tion by intestinal bacteria. Approximately 30 species of in-
testinal bacteria carry the PME gene, and some of them are
commonly present in the microflora (TABLE 1). However,
little is known about characteristics of the intestinal bacte-
ria that produce methanol via PME, and few works have
demonstrated the ability of the mammalian intestinal flora
to produce methanol as the result of pectin breakdown by
PME under aerobic and anaerobic conditions (38, 208,
209, 418). As a result of the digestion of dietary pectin
(from vegetables and fruit) and pectin capsules by the mi-
croflora and/or by pectin-associated PME, methanol levels
in human blood increased up to 1.5 times than that before
exposure (112). However, PME only catalyzes methanol
production in the presence of exogenous pectin; thus PME
activity alone cannot explain the basic, similar methanol
concentration found in the diverse group of volunteers and
the constitutive production of methanol in mammals.
C. Protein-Glutamate Methylesterase
Protein-glutamate methylesterase (PGM) is another com-
mon enzyme in the microflora (TABLE 1) that hydrolyzes
methyl esters of glutamate that bind to proteins. Methyl-
ation and demethylation are well-known mechanisms of
controlling expression, translation, and protein activities.
Depending on the type of bind between methyl radicals and
biopolymers, active demethylation can produce either
formaldehyde or methanol. The product of carbonic acid
(amino acids, glutamate, oxaloacetate, etc.) or methyl ester
demethylation is methanol, while DNA demethylation gen-
erates formaldehyde. PGM is of particular interest due to its
ability to produce detectable levels of methanol in vivo
(461), and, in contrast to PME, only endogenous molecules
are substrates of PGM for methanol production. The CheB
gene, which encodes PMG, is present in more than 50 spe-
cies of intestinal bacteria from stool samples, including
abundant species (TABLE 1). However, CheB expression
correlates with decreased chemotactic stimuli (43), and
PMG does not maintain permanent methanol concentra-
tions over time. PMG seems to participate in methanol for-
mation by intestinal bacteria, but it is unlikely that PMG
plays a crucial role in this process.
D. Pimeloyl-[Acyl-Carrier Protein]-Methyl-
Ester Hydrolase
Pimeloyl-[acyl-carrier protein]-methyl-ester hydrolase (BioH)
is an enzyme with hydrolysis activity that might also
participate in methanol production by microflora. BioH
participates in biotin (vitamin H) production in bacteria
and other organisms. The demethylation reaction termi-
nates the part of the biotin biosynthesis pathway that is
catalyzed by fatty acid synthesis enzymes (281) and results
in methanol and biotin precursor production. Vitamin H
serves in all living organisms as a covalently bound enzyme
cofactor that is essential for the introduction of carboxylic
acid groups to substrates via carboxylases. Mammals have
no enzymes capable of biotin synthesis and thus use micro-
flora- and food-derived biotin. Hence, bacteria containing
genes that encode components of the biotin biosynthesis
pathway (including BioH,TABLE 1) are present in the intes-
Table 2. List of methanol-producing enzymes, their substrate,
and the type of reaction they catalyze in bacteria
Ferment Name [EC
Number] Substrate
Reaction
Class
Pectinesterase
[3.1.1.11]
Methyl esters of
pectin
Hydrolysis
sMMO/pMMO
[1.14.13.25]
Methane, oxygen Redox
reaction
Formaldehyde
dismutase
[1.2.99.4]
Formaldehyde Redox
reaction
Protocatechuate 4,5-
dioxygenase
[1.13.11.8]
3-O-Methylgallate Redox
reaction
4-Methyloxaloacetate
esterase [3.1.1.44]
4-Methyloxaloacetate Hydrolysis
Pimeloyl-[acyl-carrier
protein]-methyl-ester
hydrolase (BioH)
[3.1.1.85]
Pimeloyl-[acyl-carrier
protein]-methyl
ester
Hydrolysis
Protein-L-glutamate-O4-
methyl-ester
acylhydrolase
[3.1.1.61]
Protein glutamate
methyl ester
Hydrolysis
sMMO, soluble methane monooxygenase; pMMO, particulate meth-
ane monooxygenase.
DOROKHOV ET AL.
614 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
tinal flora. BioH activity was originally demonstrated in
Escherichia coli, which is one of the major microflora bac-
teria. Thus BioH participates in an essential biosynthetic
pathway and is widespread in the microflora; hence, BioH is
the best candidate for methanol production by the intestinal
flora.
To summarize, there are a few strong candidates among
bacterial enzymes that might participate in methanol me-
tabolism and serve as the source of methanol metabolism in
mammals. Catalase seems to oxidize most of the methanol
in bacteria, while ADH and MDH are less important. PME
and PGM require exogenous sources to produce methanol,
so they cannot be an efficient endogenous source. Because
BioH participates in an essential biosynthetic pathway and
is widespread in the microflora, it is the best candidate for
methanol production by the intestinal flora. This analysis is
in agreement with the experimental data on the microflora
of rats (246) and gives new clues for further investigation of
methanol metabolism by intestinal bacteria in mammals.
V. METHANOL METABOLISM IN HUMANS
Along with dietary methanol and methanol produced by the
intestinal microflora, processes that involve SAM participa-
tion also contribute to the pool of physiological methanol
(265). SAM is a universal, endogenous methyl donor for
several reactions, including the methylation of proteins,
phospholipids, DNA, RNA, and other molecules involved
in basic epigenetic mechanisms (196). Genome-wide meth-
ylation analysis has identified DNA methylation profiles
specific to aging and longevity. Moreover, analysis of the
DNA methylation landscape revealed that DNA obtained
from a 103-yr-old donor was more unmethylated overall
than DNA from the same cell type isolated from a neonate
(181). Differentially methylated genes are strikingly en-
riched in loci associated with neurological disorders, psy-
chological disorders, and cancers (531). Protein carboxym-
ethylation involves the methylation of amino acid COOH
groups; this reaction is catalyzed by methyltransferases,
which produce carboxyl methyl esters that are readily hy-
drolyzed under neutral and basic pH conditions, or by
methyl esterase, which produces methanol (104). Protein
carboxymethylase is highly localized to the brain and pitu-
itary gland in several mammalian species (105).
In a healthy person, no matter from what sources methanol
is derived in the body, it is eventually displayed and kept at
a low physiological level via physiological and metabolic
clearance mechanisms.
A. Nonmetabolic (Physiological) Clearance of
Methanol
Because methanol metabolism involves the same enzymes
as ethanol metabolism, information about the nonmeta-
bolic clearance of ethanol may be relevant to our consider-
ation of methanol metabolism. Approximately 90% of eth-
anol is removed by oxidation, and 10% of ethanol is
excreted in breath, sweat, and urine (63, 347). Animal stud-
ies have reported that inhaled methanol is eliminated
mainly by metabolism (70–97% of absorbed dose), and
only a small fraction is eliminated as unchanged methanol
in the urine and expired air (3–4%) (42, 110, 188). The
determination of methanol and ethanol contents in the hu-
man breath after the consumption of alcoholic beverages
and various amounts of fruits revealed that methanol con-
centrations increased from a natural (physiological) level of
0.4 to 2 ppm a few hours after eating 0.5 kg fruit
(451). The same concentration was reached after drinking
of 100 ml brandy containing 24% ethanol and 0.19%
methanol. We estimated the levels of endogenous methanol
generation by considering the probable removal of metha-
nol via pulmonary and renal excretion (413). Assuming
renal methanol clearance in the average volunteer of 1.0
ml/min (231, 483), we believe that a clearance of 60 ml of
blood in 1 h occurs and has little effect on the total blood
methanol content. For the estimation of the pulmonary ex-
cretion of methanol, we used estimates (200) indicating that
each minute, 5.6 ml of blood is hypothetically completely
cleared of methanol, i.e., the effect of the pulmonary clear-
ance of plasma methanol is also negligible. Thus, following
exogenous uptake or endogenous production, unchanged
methanol either is excreted (direct excretion) in the urine or
exhaled breath or enters a metabolic pathway.
B. First Phase of Methanol Catabolism
The main way methanol is eliminated from the body is via
its oxidation to formaldehyde and then to formic acid,
which can then be either excreted in the urine or further
oxidized to carbon dioxide. The first phase of the oxidative
metabolism of methanol and ethanol is similar in insects
(162, 503, 504), animals, and humans (63, 168, 453). Eth-
anol and methanol are converted into formaldehyde and
acetaldehyde, respectively, via one of at least three separate
pathways (FIGURE 4). The first involves cytochrome P450
monooxygenases (CYP) such as CYP2E1 in humans, which
catabolizes 9% of exogenous alcohol, especially at high
concentrations (57, 87, 495). The second process, which
oxidizes up to 1% of exogenous alcohol, is greatly depen-
dent on catalase (64, 97). The third pathway, which oxi-
dizes up to 90% of alcohol, is catalyzed by cytosolic alcohol
dehydrogenase I (ADH I) via NAD
-dependent oxidation
(63, 168, 297). The contribution of each of these mecha-
nisms in the catabolism of alcohol varies by human organs.
In the liver, most alcohol catabolism occurs via ADH1b
(122), while in the brain, the first phase of methanol and
ethanol oxidative metabolism is mainly carried out by two
other mechanisms involving CYP2E1 and catalase-H
2
O
2
compound I (89, 143, 272). Although there are species dif-
ferences in methanol and formic acid pharmacokinetics in
METABOLIC METHANOL
615Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
mice, rabbits, and primates (442, 443), methanol is metab-
olized to formaldehyde primarily by ADH in humans. Ro-
dents are reported to rely on the peroxidative activity of
catalase (45, 64, 229).
1. ADH-mediated methanol catabolism
ADH is a zinc-containing enzyme family that consists of
two subunits of 40 kDa (345). The function of ADH in
the human body consists mainly of the oxidation of ethanol
and methanol produced in the body by microorganisms of
the gastrointestinal tract and taken in via fruit, vegetable,
and alcoholic beverage consumption. The enzyme is located
in the cytosolic fraction of cells, primarily in the liver, gas-
trointestinal tract, kidney, nasal mucosa, testes, and uterus,
and is absent in brain cells (143). Approximately 90% of
ingested ethanol and methanol is metabolized via hepatic
ADH1b (TABLE 3), which catalyzes an oxidative pathway
(65) according to the general scheme presented in FIGURE 4.
The control of ADH activity is complex and includes 1)
dissociation of the product NADH, which is the rate-limit-
ing step; 2) product inhibition by NADH and acetaldehyde
or formaldehyde; and 3) substrate inhibition by high con-
centrations of ethanol and methanol. Mammalian ADHs
(TABLE 3) catalyze the reversible oxidation of alcohols to
aldehydes, acting on a wide variety of substrates ranging
from methanol to long-chain alcohols and sterols (120).
Mammalian liver ADH1b is a well-characterized enzyme
that uses ethanol as a substrate. Relatively little is known
about the interaction between ADH1b and methanol, but
there is extensive structural homology between methanol
and ethanol. ADHs were isolated from the human liver by
Vallee et al. (489) 50 years ago; this resulted in partially
purified preparations. Subsequent efforts improved upon
the purity, which allowed the study of human liver ADH
enzymology (286). Later, its primary and tertiary struc-
tures, catalytic mechanisms, and enzymatic properties were
well studied (301, 376, 381). In vitro, ADH1b catalyzes the
transfer of a hydride ion from an alcohol substrate to the
NAD
cofactor, yielding the corresponding aldehyde and
the reduced cofactor NADH. ADH1b is also an excellent
catalyst of the reverse reaction. In enzymatic tests carried
out in vitro, methanol is oxidized to formaldehyde by
NAD
in the presence of ADH; then, formaldehyde can be
oxidized to formate by NAD
or reduced to methanol by
NADH. These reactions follow a rapid equilibrium random
mechanism. Among these three reactions, the reduction of
formaldehyde is the most rapid. The rate of formaldehyde
oxidation is faster than that of methanol oxidation (380).
Each of the two liver ADH1b subunits contains one cata-
lytic and one structural zinc ion (83). The catalytic zinc ion
is situated at the bottom of the cleft between the coenzyme
binding domain and the catalytic domain. It is accessible
through two channels, one that accommodates the coen-
zyme and another through which the substrate approaches
the catalytic site (382). When the ternary complex forms,
the catalytic domain rotates 10° to allow the residues
in the catalytic domain to move closer to the coenzyme do-
main, and the cleft between domains closes (83). There are
significant differences in the enzymatic behavior for meth-
anol and ethanol. In vitro, methanol is oxidized by live
ADH1b at a much slower rate. The rate of methanol oxi-
dation catalyzed by purified ADH1b is only 3% of the
rate of ethanol oxidation. Within the alcohol binding site of
liver ADH1b, there are two binding regions: a hydroxyl
binding region and a hydrophobic binding region (99, 100).
Methanol binds to a similar site as ethanol in ADH1b.
Pyrazole, an efficient liver ADH1b inhibitor, binds to the
enzyme-NAD
complex and blocks the substrate binding
site (378). The inhibition patterns and inhibition constants
of pyrazole are similar for both methanol and ethanol
(380). This confirms that the binding of methanol to the
catalytic zinc is similar to that of ethanol. On the other
A
Ethanol Acetaldehyde Acetate
NAD+NADH
ADH
O2 + NADPH NADP+ + H2O2
H2O2H2O
CYP2E1
Catalase
B
NAD+NADH
ADH
O2 + NADPH NADP+ + H2O2
H2O2H2O
CYP2E1
Catalase
FormaldehydeMethanol Formate
FIGURE 4. The first phase of ethanol (A) and methanol (B) catab-
olism to acetaldehyde and formaldehyde, respectively, by three en-
zymatic systems.
DOROKHOV ET AL.
616 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
hand, the second step, hydride transfer, which is the con-
version of the ternary complex, is the rate-limiting step for
methanol oxidation. Because of the lack of an appreciable
hydrophobic chain, methanol exhibits different kinetics
than ethanol (50, 122, 380). The hydrophobic interaction
in the active site plays a fundamental role in substrate bind-
ing. When a substrate binds to the enzyme, the interaction
between the hydrophobic binding region and the hydropho-
bic chain on the substrate stabilizes substrate binding.
Methanol, with a chain of only one carbon atom, cannot
bind to the hydrophobic region as ethanol can. Because the
methyl group is not held in the proper orientation for hy-
dride transfer, methanol exhibits both a weak binding con-
stant and a slow turnover rate. The binding of NAD
to the
enzyme does not exhibit a significant effect on the binding
of methanol, nor does methanol affect the binding of
NAD
(380). Pyrazole and its derivatives such as 4-meth-
ylpyrazole are potent inhibitors of ADH1b and block the
interactions of all class I ADH isozymes with ethanol (523)
and methanol (378). These compounds function as ligands
of zinc, which is essential for the catalytic activity of the
native forms of ADH. They form slowly dissociating ter-
nary complexes with ADH and NAD
and act as compet-
itive inhibitors with respect to ethanol and methanol.
4-Methylpyrazole, the most potent inhibitor, binds to the
enzyme-NAD
complex as an inner sphere ligand of the
catalytic zinc (416), thus blocking the substrate binding
site and the formation of a dead-end ternary complex with the
enzyme and NAD
(455). The strong sensitivity of class I
ADHs to pyrazole inhibition explains the powerful inhibi-
tion of ethanol metabolism in humans by these agents. The
K
i
(4-methylpyrazole) value of class I ADHs is 0.1
M
(63) with ethanol as a substrate, and the K
i
value for meth-
anol is similar, equal to 0.09
M (378). The
␲␲
ADH
isoform (TABLE 3), named for its pyrazole insensitivity, is
inhibited much less effectively by 4-methylpyrazole (K
i
2
mM), although the pyrazole derivatives 4-bromo-, 4-nitro-,
or 4-pentylpyrazole are more efficient with ethanol (K
i
val-
ues between 4 and 27
M) (40).
In addition to ethanol and methanol, ADH isoforms also
oxidize several metabolic alcohols with high catalytic effi-
ciency, including retinol,
=-hydroxy fatty acids, hydroxy-
steroids, and hydroxyl derivatives of dopamine and epineph-
rine metabolites (37, 382). Vitamin A and its derivatives
(retinoids) are essential components in vision. Animals are,
in general, unable to synthesize vitamin A de novo. Retinol,
retinal, and retinoic acid are C20 isoprenoids that are met-
abolically derived from the oxidative cleavage of C40 caro-
tenoids. In the presence of NAD
, the enzyme oxidizes the
alcohol retinol to the aldehyde retinal, which is necessary
for vision. ADH utilizes a mechanism for retinol-retinal
interconversion that is similar to that used for ethanol/
methanol oxidation and acetaldehyde/formaldehyde reduc-
tion (37, 382). In contrast to the methanol-formaldehyde,
ethanol-acetaldehyde, and other short-chain alcohol-alde-
hyde systems, the equilibrium constant for the retinol-reti-
nal interconversion is 200 times higher. The K
m
value (60
M) indicates that the interactions between the long carbon
chain of the retinoid and the hydrophobic pocket of the
enzyme provide a major driving force for the binding pro-
cess. The oxidation of these alcohols can be inhibited by
ethanol and methanol, and therefore, the role of primary
alcohol substrate competition is an important issue in the
effects of alcohol on humans.
2. Human ADH genes
The human genome contains seven genes encoding ADH
(122) (TABLE 3) (FIGURE 5). They form a cluster of genes
located in chromosome 4 (4q21–23), spanning 370 kb
(359, 360) and encoding multiple forms of ADH with dif-
Table 3. ADH genes and proteins
Gene Name Class
Amino Acid Differences
Between Alleles Protein Name K
m
(Ethanol), mM K
m
(Methanol), mM
ADH1A I4.0
a
to 4.2
d
150.0
b,c
ADHB*1 Arg48, Arg370
1
0.05
a,d
to 1.2
b
6.0
d
to 7.0
e
ADH1B*2 His48, Arg370
2
0.9
d
ND
ADH1B*3 Arg48, Cys370
3
34.0
d
ND
ADH1C*1 Arg272, Ile350
1
1.0
d
ND
ADH1C*2 Gln272, Val350
2
0.6
d
to 1.0
b
30.0
b
ADH1C*352Thr Thr352 ND ND ND
ADH4 II 34
d
No activity
b
ADH5 III
or FDH 1,000.0
f
No activity
b
ADH6 V ADH6
g
ND ND
ADH7 IV 30.0
f
ND
K
m
indicates the concentration of alcohol (ethanol or methanol) at which the enzyme works at 50% capacity.
ND, not determined; FDH, formaldehyde dehydrogenase.
a
Bosron et al., 1983 (41);
b
Wagner et al., 1983
(492);
c
For
␣␥
1 heterodimer;
d
Crabb et al., 1987 (92);
e
Pietruszko, 1975 (378);
f
Edenberg, 2007 (122);
g
Ostberg et al., 2013 (361).
METABOLIC METHANOL
617Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
ferent substrate specificities (63, 92). The human liver con-
tains seven ADH isoforms, which are divided into five
classes (TABLE 3).
Class I ADHs are encoded by three genes, ADH1,ADH2,
and ADH3, which encode these subunits:
(ADHIA),
1,
2,
3 (ADHIB), and
1 and
2 (ADHC). The ADH1 sub-
units share 94% sequence identity. These different units
and polymorphs can form homodimers or various het-
erodimers (e.g.,
␣␣
,
1
1,
␣␤
2,
1
2) (41, 393, 492). The
class I ADH isoforms play the most important role in alco-
hol oxidation. The study of ADH1-null mice showed that
ADH1 enzymes together are responsible for 70% of the
total ethanol-oxidizing capacity of liver, indicating that the
non-ADH1 pathway accounts for the remaining 30%
(169). The relative contributions of each of the ADH
isozymes to ethanol oxidation change with the hepatic eth-
anol concentration. The K
m
of human class I ADH for eth-
anol is very low (TABLE 3). Therefore, the enzyme becomes
saturated after the intake of just 28–30 ml of ethanol, and
ethanol is removed from the human body at a constant rate,
independent of the concentration (347). Hence, the rate of
disappearance of ethanol from the blood is virtually inde-
pendent of the prevailing blood-ethanol concentration
(zero-order kinetics) (92). All isozymes of class I ADH show
substrate inhibition ([ethanol] 20 mM) because an excess
of ethanol decreases the speed of NADH dissociation from
the enzyme (223).
The ADH
-subunits, encoded by the gene ADH1A, are
monomorphic, whereas the other two subunits were iden-
tified as isoforms that differ in ADH catalytic activity levels.
The ADH1B gene was shown to comprise two functionally
important polymorphisms (TABLE 3):1) one encoding
ADH1B*2, which occurs in human populations of North
Africa, Eurasia, and Oceania, and 2) one encoding
ADH1B*3, which is found only in African populations.
Experiments on isolated preparations of human liver ADH
showed that the arginine amino acid at position 48 partic-
ipates in the binding of the pyrophosphate group NAD
,
but the replacement of arginine for histidine leads to a lower
optimal pH and a 100-fold increase in the
2-ADH turn-
over (215, 308). The study of the role of this change on the
metabolism of ethanol in the human body did not provide
such an unambiguous picture. The first-pass study showed
that the hepatic ethanol clearance of ADH1B*2 individuals
was higher than that of the ADH1B*1 individuals (267).
Evidence of the high enzymatic activity of
2-ADH was
also obtained in the study of another ADH1B*2/*2
group (224); the blood acetaldehyde concentrations of
ADH1B*2/*2 individuals were higher than those of
ADH1B*1/*2 individuals only in the ALDH2*1/*2 group.
However, the blood ethanol concentrations of the
ADH1B*2/*2 group were higher than those of the
ADH1B*1/*2 group regardless of the ALDH2 genotype.
These findings are unexpected and are difficult to explain
from the viewpoint of ADH enzyme activity alone.
The Arg370Cys amino acid replacement in the ADH1B*3
allele, which almost never occurs in populations of Euro-
pean and Asiatic origins, also leads to an increased reaction
speed due to a changed efficiency of the interaction between
the
3
3
dimer and NAD
(55).
ADH1C also contains single nucleotide polymorphisms, of
which ADH1C1 and ADH1C*2 are the most studied. The
enzyme encoded by ADH1C*1 (
1-ADH) contains Arg at
position 272 and isoleucine (Ile) at position 350, whereas
that encoded by ADH1C*2 (
2-ADH) contains glutamine
(Gln) at position 272 and valine (Val) at position 350 (359,
360). The kinetic differences between
1-ADH and
2-
ADH are small (185) (TABLE 3).
Classes II, III, and IV enzymes are homodimeric forms of
the
,
, and
subunits, respectively. The ADH4 gene
encodes class II ADHs, including the
␲␲
ADH isoform,
named for its pyrazole insensitivity. The
␲␲
ADH isoform is
more labile than the other ADH isoforms and was first
detected in liver biopsies and fresh autopsy samples. Due to
its rather high K
m
for ethanol (TABLE 3), this enzyme may be
more important in the metabolism of high concentrations of
ethanol (92).
␲␲
ADH accounts for nearly 30% of the total
ethanol-oxidizing capacity of the liver (192) and exhibits a
more limited substrate specificity than do the other molec-
ular forms of the human liver. Methanol, glycerol, and eth-
ylene glycol, even at concentrations up to 100 mM, cannot
be oxidized by
␲␲
ADH (40).
The ADH5 gene encodes class III ADHs, including the
␹␹
ADH isoform, which is widely distributed in the tissues
ADH1C ADH1B ADH1A
ADH7 1C 1B 1A ADH6 ADH4 ADH5
Class 1 ADH
1B*1: R48, R370
1B*2: H48, R370
1B*3: R48, C370
1C*1: R272, I350
1C*2: Q272, V350
1C*352Thr: T352
FIGURE 5. Relative map of the SNP sites
and ADH alleles. Top panel represents the
whole ADH gene cluster, and bottom panel
shows the genes of the class 1 ADH cluster
with the different alleles and amino acid sub-
stitutions indicated. [Modified from Osier et
al. (360) and Edenberg (122), both with per-
mission from Elsevier.]
DOROKHOV ET AL.
618 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
(63).
␹␹
ADH is virtually not inhibited at all by 4-meth-
ylpyrazole (492), has a very high K
m
for ethanol, and
likely plays almost no role in ethanol and methanol elim-
ination (63).
3. Cytochrome P450s-mediated oxidation of
methanol and ethanol
The CYP enzymes are named CYP for cytochrome P450,
followed by an Arabic number denoting the family, a letter
designating the subfamily and, finally, an Arabic numeral
representing the individual gene in the subfamily (236, 341,
530). The P450s are arranged in families based on sequence
homologies (340, 342) and are very versatile catalysts that
activate dioxygen and insert a single oxygen atom into al-
most any organic compound imaginable (255). Cyto-
chrome P450s are a family of heme-containing monooxy-
genases that are involved in the oxidation of ethanol, meth-
anol, steroids, fatty acids, and numerous xenobiotics
(acetone, benzene, and other alcohols) ingested from the
environment (62, 249, 252, 295, 530). The heme iron of the
CYP enzymes binds two oxygen atoms. However, in con-
trast to the iron found in hemoglobin, this iron atom is
bound rather tightly to the anionic thiolate sulfur of cys-
teine. This binding mode gives heme the properties neces-
sary for splitting the dioxygen molecule into two atoms.
Because one oxygen atom forms a water molecule, the CYP
enzymes were also named mixed function oxidases. The
second oxygen atom is activated for introduction into the
substrate molecule; therefore, CYP enzymes have been in-
cluded in the monooxygenase enzyme class. Hence, the
splitting of the dioxygen molecule to two atoms results in
the formation of a hydroxylated product (6). Most P450s
are considered to operate according to a general scheme
(FIGURE 6) (26, 27, 86, 508). There are many P450 isoforms
encoded by more than 100 gene families (249, 252, 340,
342, 530, 538). P450 functions in conjunction with micro-
somal enzymes such as NADPH-cytochrome P450 reduc-
tase and cytochrome b
5
(342) (FIGURE 6).
CYP2E1 is the P450 enzyme with the highest activity for
oxidizing ethanol and methanol to acetaldehyde and form-
aldehyde, respectively. CYP2E1 accounts for 6% of the
total P450 content in the human liver and catalyzes the
metabolism of 2% of commercially available drugs (150,
538). The human gene encoding CYP2E1 is the only gene of
the CYP2E subfamily located at chromosome 10q26.3; it
contains 9 exons and comprises several polymorphisms
(236). Several different CYP2E1 polymorphisms have been
identified (266). Human CYP2E1 expression is undetect-
able in the fetal liver but is the highest expressed cyto-
chrome P450 in the adult liver, where it is present mainly in
the endoplasmic reticulum (microsomal fraction) (271,
272, 274, 275) but are also found in mitochondria (15, 18,
19, 242, 273). Aside from the liver, CYP2E1 is also ex-
pressed in the brain (137, 138, 164, 178, 216, 246, 349,
482) and lungs (227). CYP2E1 expression has been found
also in the nasal mucosa, kidney cortex, testes, ovaries, and
gastrointestinal tract at lower levels (273, 454) and in car-
diac tissue at somewhat higher levels (317, 532, 533).
CYP2E1 is regulated by several mechanisms, including
transcriptional and posttranscriptional processes (350,
530). CYP2E1 induction appears to occur via two steps: a
posttranslational mechanism at low ethanol concentrations
and an additional transcriptional mechanism at high etha-
nol concentrations (17, 397).
Fe3+ ROH
FeOH3+ R•
FeO3+ RH
Fe-OOH RH Fe2+-O2 RH
O2
O2
Fe2+ RH
H+
NADPH-P450 reductaseox
NADPH-P450 reductaseox
NADPH-P450 reductasered
NADPH-P450 reductasered
Fe3+
Fe2+-O2 RH
b5ox
b5red
H2O
–H2O
–ROH RH
H2O2
2e
1e
1e
Fe3+ RH
1
2
3
4
5
6
7
8
FIGURE 6. The general scheme for P450-
catalyzed oxidation reactions (26, 27). RH,
substrate; ROH, product. The reversibility of
some of the latter steps is unknown. The
outlets for the uncoupled reduced oxygen
products O
2
,H
2
O
2
, and H
2
O are shown. Fol-
lowing substrate binding (step 1), ferric
P450 receives 1 electron via NADPH-P450
reductase (step 2). The ferrous form of hem
binds O
2
(step 3) before undergoing a sec-
ond 1-electron reduction to begin O
2
activa-
tion (step 4). Although this second electron
originates from NADPH-P450 reductase,
the accessory protein cytochrome b
5
takes
part in the delivery of the electron to P450.
Insertion of the activated oxygen into the sub-
strate occurs via C-H bond cleavage (step 6)
followed by rapid oxygen rebinding to form
the product (step 7). Step 8 is the release of
the product from the active site of the en-
zyme.
METABOLIC METHANOL
619Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
Ethanol and methanol are both inducers and substrates of
CYP2E1 (61, 63, 226, 357, 398). Acetaminophen, caffeine,
chlorzoxazone, and an array of carcinogens such as ben-
zene, styrene, acrylonitrile, and nitrosamines are also me-
tabolized by CYP2E1. The K
m
of CYP2E1 for ethanol is 10
mM, which is 10-fold higher than the K
m
of ADH for eth-
anol. Therefore, liver ADH has a much higher capacity for
ethanol oxidation than the CYP2E1 system at low ethanol
concentrations, but CYP2E1 may be the basis of the meta-
bolic adaptation to high concentrations of ethanol that de-
velops upon chronic ethanol consumption (272). CYP2E1
is likely to provide 10% of the total ethanol-oxidizing
capacity of the liver during modest ethanol intake (62).
Alcohol oxidation increases at higher ethanol concentra-
tions, and much of this increase is due to the metabolism of
ethanol by CYP2E1. CYP2E1 levels are increased by
chronic ethanol administration by a mechanism that largely
involves protection of the enzyme against proteolysis.
CYP2E1 expression is also induced in diabetics, in the
fasted nutritional state, and in nonalcoholic steatohepatitis
(NASH). NASH is a progressive liver disorder that occurs in
patients with high hepatic CYP2E1 levels but without sig-
nificant ethanol consumption (506).
CYP2E1 is highly conserved within the human population
(271), suggesting significant physiological functions, in-
cluding nutritional support via the catalysis of fatty acid
=-1 and
=-2 hydroxylation reactions (2, 4, 259). Thus
CYP2E1 plays dual physiological roles, one in the conver-
sion of ketones into glucose and one in detoxification (271).
An important feature of ethanol and methanol oxidation
via CYP2E1 is the generation of reactive oxygen species
(ROS), which contribute to the damage of liver cells (57,
272, 516). CYP2E1 is an effective generator of ROS such as
the superoxide anion radical and hydrogen peroxide as a
result of the uncoupling of oxygen consumption and
NADPH oxidation. The significant levels of CYP2E1 within
the mitochondria could further contribute to ROS deleteri-
ous effects (18). Moreover, metabolism by CYP2E1 results
in a significant release of free radicals that, in turn, diminish
reduced glutathione levels (18) and other oxidative stress
defense systems that play major pathogenic roles in alco-
holic (60, 242, 272) and nonalcoholic liver disease (14, 66).
CYP2E1 is a causative player in alcoholic liver disease, as
well as in NASH, likely through the enhancement of hepatic
lipid peroxidation (14, 60, 242).
4. Catalase-H
2
O
2
system
The ability of peroxisomal catalase to oxidize short-chain
alcohols such as methanol and ethanol was discovered long
ago (64, 81, 234, 274). With the use of hydrogen peroxide
(H
2
O
2
) as a cosubstrate, the heme-containing enzyme cat-
alase forms the catalase-H
2
O
2
system (also known as com-
pound I), which is also the major pathway of formaldehyde
oxidation. The role of this system in the oxidation of alco-
hols is small in the liver (63) but significant in the brain
(499, 536, 537), which lacks ADH activity (143). Ethanol
in the brain is oxidized into acetaldehyde by the action of
the catalase-H
2
O
2
system (7, 8, 81, 145, 203, 228) and
cytochrome P4502E1 (CYP2E1) (537). It is estimated that
the catalase-H
2
O
2
system in the brain is responsible for
60–70% of the acetaldehyde generated in this organ, while
CYP2E1 accounts for 10–20% (537). Inhibitors of cata-
lase, aminotriazole, and sodium azide depress the oxidation
of ethanol and methanol to acetaldehyde and formalde-
hyde, respectively, in the brain (64, 537). Acetaldehyde de-
rived from the catalase-dependent oxidation of ethanol in
the brain has been suggested to play a role in the mediation
of many effects of ethanol in the brain, including behav-
ioral, neurochemical, and neurotoxic actions, and plays a
crucial role in the development of alcoholism (89, 334).
Participation of the catalase-H
2
O
2
system in the oxidation
of ethanol and methanol in the brain was confirmed by
experiments with alpha lipoic acid (ALA, 1,2-dithiolane-3-
pentanoic acid). ALA is considered to be an H
2
O
2
scaveng-
ing agent and thus an inhibitor of the catalase-H
2
O
2
com-
plex; treatment with ALA prevents the formation of ac-
etaldehyde in the brain and, therefore, prevents its
neurochemical and neurotoxic actions (264). Indeed, on the
one hand, ALA reduced ethanol self-administration in rats
(368) and mice (264), and on the other hand, ALA treat-
ment prevented methanol exposure-induced oxidative dam-
age of tissues of the rat nervous system (388). However,
these results can also be explained by another property of
ALA, namely, the ability of ALA to activate ALDH2 (116,
171a, 285, 310). The ALA-mediated decreases of formal-
dehyde and acetaldehyde in animal tissues can be induced
by both interfering with their formation and accelerating
their oxidation by ALDH2. It could be suggested that the
beneficial effects of ALA on ethanol metabolism are related
to the capacity of ALA to control both reactions.
C. Second Phase of Methanol Catabolism
Formaldehyde is a naturally occurring biological com-
pound that is present in all tissues, cells, and biological
fluids (174). The concentration of endogenous formalde-
hyde in the blood of rats (176), monkeys (58), and humans
(176, 413, 463) does not exceed 0.1 mM, and that in the rat
and human brain is 0.2–0.4 mM (464, 466). Most formal-
dehyde is produced as the oxidation product of methanol.
However, another endogenous source of formaldehyde is
the oxidative deamination of methylamine derived primar-
ily from creatine (526) by semicarbazide-sensitive amine
oxidases (SSAOs) (201, 436, 527), which generate formal-
dehyde together with ammonia and hydrogen peroxide,
as follows (233, 363, 525, 527): CH
3
NH
2
O
2
H2O
SSAO
HCHO H
2
O
2
NH
3
.
In mammals, among the SSAOs, vascular adhesion protein
1 is one of the most extensively studied members of this
group of enzymes (119, 202, 524). It is interesting that some
DOROKHOV ET AL.
620 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
monoamine oxidase inhibitors such as phenelzine with an-
tidepressant properties (20) are able to protect neurons and
astrocytes against formaldehyde (424).
Formaldehyde is also part of the one-carbon pool, which is
utilized for the biosynthesis purines, thymidine, and several
amino acids, which are incorporated into DNA, RNA, and
proteins during macromolecular synthesis (220). Besides
the oxidation of methanol, formaldehyde is formed by the
conversion of serine and glycine in the presence of tetrahy-
drofolate (35). Formaldehyde is also generated during chro-
matin structure remodeling in the reactions of removing
methyl groups from lysine residues in histones that are cat-
alyzed by lysine-specific demethylase 1 and JmjC domain-
containing histone demethylases (79, 189, 287, 470, 497).
Aside from endogenous sources, formaldehyde can also be
produced and released from different exogenous products
that naturally contain formaldehyde, including food prod-
ucts such as coffee, codfish, meat, poultry, and maple syrup
(95b, 98, 298, 391). Drugs also may be sources of formal-
dehyde. For example, Selegiline (Anipryl, L-deprenyl, El-
depryl, Emsam, Zelapar), which is used to treat Parkinson’s
disease, gives off formaldehyde as a byproduct as a result of
the metabolic transformation circuit (219).
Efficient systems of formaldehyde oxidation exist in mam-
mals. Despite the multiple endogenous and exogenous
sources of formaldehyde, a low physiological level of form-
aldehyde in bodily fluids and tissues is maintained by the
continuous action of cellular formaldehyde-metabolizing
enzymes (FIGURE 7).
The oxidation of formaldehyde to formate occurs by at least
three separate pathways with the participation of P450
monooxygenases, mitochondrial aldehyde dehydrogenase
2 (ALDH2), and the gene encoding ADH5,
␹␹
ADH,
(TABLE 3), also called ADH3 or formaldehyde dehydroge-
nase (FDH) (473) (FIGURE 7). Liver ADH1b plays the main
role in the first phase of methanol oxidation but has a neg-
ligible effect on the reduction of formaldehyde to methanol
because of its higher K
m
value for formaldehyde (30 mM)
(419).
Similarly to other aldehydes, formaldehyde can be oxidized
by P450s (27, 183, 255, 289, 452), as shown in FIGURE 6.
Interestingly, transgenic Petunia hybrida plants harboring
the CYP2E1 gene have improved resistance to formalde-
hyde (533).
In humans, two members of the divergent ALDH superfam-
ily, a cytosolic (ALDH1A1) and a mitochondrial (ALDH2)
enzyme (204, 420, 427, 435, 484, 502), can directly metab-
olize formaldehyde (142, 297, 452, 453, 473, 485). Mito-
chondrial ALDH2, which is especially important for the
oxidation of high concentrations of formaldehyde, has a
high K
m
for formaldehyde (0.2–0.5 mM) compared with
that of FDH (see below) (0.01 mM) (59, 175, 336, 432).
The intracellular concentration of free formaldehyde is
likely too low to result in significant oxidation by ALDH2
(175).
Human FDH or ADH5 gene encoding (FIGURE 6)
␹␹
ADH
belongs to the family of medium-chain zinc-containing
ALDH1A1
ALDH2
Glutathione
Glutathione
S-hydroxymethyl glutathione
S-formylglutathione
FDH
NAD+
NADH
NAD+
NADH
NAD+
NADH
S-formylglutathione
hydrolase
MethanolEndogenous
sources
Exogenous
sources
Formaldehyde
Formate
CYP450s
CO2 + H2O
FIGURE 7. Formaldehyde metabolism. Overview
of the pathways of the conversion of formaldehyde
into water and CO
2
. For details, see text.
METABOLIC METHANOL
621Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
ADHs (TABLE 3). In the literature, different ADH nomen-
clatures have been used in early reports, in which ADH3
was referred to as glutathione-dependent FDH,
␹␹
ADH or
class III ADH (218, 432). We will use glutathione-depen-
dent FDH in our review because this name more precisely
reflects its formaldehyde oxidizing function.
In contrast to other mammalian ADH isoenzymes that are
active toward ethanol and methanol, FDH shows very poor
activity toward ethanol and methanol and prefers long-
chain alcohol substrates such as
=-hydroxy fatty acids,
including 12-hydroxydodecanoic acid and retinol (vitamin
A) (243, 268, 320, 332, 432). Compared with the other
ADH classes, FDH has two additional activities: a glutathi-
one-dependent formaldehyde oxidizing activity and an S-
nitrosoglutathione (GSNO) reductase activity (186).
Human FDH is the most efficient formaldehyde-metaboliz-
ing enzyme among all of the enzymes elucidated so far (432)
and was first purified from human liver and partially char-
acterized by Uotila and Koivusalo (479, 480). FDH is the
most conserved class of all of the ADHs (186). Like other
human ADH isoenzymes, FDH exists as a dimer and con-
tains two covalently bound zinc ions per subunit (520).
FDH is considered to be the main scavenger of exogenous
formaldehyde and is ubiquitously expressed with relatively
little inter-tissue variation in mammals, in contrast to other
ADHs. FDH is primarily a cytosolic enzyme, although was
revealed in nucleus where it likely protects DNA from form-
aldehyde-mediated damage (139, 195).
FDH is a glutathione-dependent
-ADH and oxidizes form-
aldehyde to formate in a two-step process (FIGURE 7). The
polarized carbonyl group formaldehyde is a compound
with high reactivity with thiols, spontaneously forming S-
hydroxymethyl glutathione after the interaction of formal-
dehyde with glutathione (256). Thus, in the first step of
FDH-mediated formaldehyde oxidation, S-hydroxymethyl
glutathione is formed and is subsequently used as an FDH
substrate to generate S-formylglutathione (168, 297, 431,
457). The conjugate S-formylglutathione is then hydrolyzed
byaS-formylglutathione transferase to generate formate
and glutathione (167, 168, 297, 430, 431, 452, 457, 473).
The formate generated by formaldehyde oxidation can
undergo further oxidization to carbon dioxide (FIGURE 7)
in metabolic pathways involving catalase (85, 419), 10-
formyltetrahydrofolate dehydrogenase, also known as
ALDH1L1, or ALDH1L2, its mitochondrial isoform
(419, 473).
Formaldehyde interacts with GSNO and functions as a
GSNO reductase in NO homeostasis (431, 432, 457).
GSNO depletion is associated with various diseases, includ-
ing asthma. Importantly, FDH-mediated S-(hydroxymeth-
yl)glutathione oxidation is accelerated in the presence of
GSNO, which is concurrently reduced by immediate cofac-
tor recycling (430, 432, 457).
D. Features of Methanol Metabolism in the
Brain and in Embryos
The metabolism of methanol and formaldehyde in brain
cells is important for the function of this vital organ (472).
Moreover, the elevation of brain formaldehyde levels is
likely to result in neurodegenerative diseases (386, 448,
450, 463, 464, 517, 518). Methanol and the products of its
metabolism in brain cells may be synthesized in situ or be
introduced from peripheral organs via the bloodstream by
overcoming the blood-brain barrier.
1. Production of methanol and formaldehyde by brain
cells
Cultured brain cells, including astrocytes and neurons, con-
tain mRNAs encoding SSAO and lysine-specific demethyl-
ase 1, as well as for the enzymes involved in formaldehyde
metabolism (472, 473). Synthesis of formaldehyde by brain
cells can occur via protein carboxymethylation (69, 265), as
well from the participation of SSAOs and in the reactions of
removing methyl groups from lysine residues in histones,
which are catalyzed by lysine-specific demethylase 1 and
JmjC domain-containing histone demethylases (79, 189,
287, 470, 497).
SAM might be transformed to methanol and S-adenosyl
homocysteine in the bovine pituitary gland and other
animal brain tissue (16, 422). SAM is a universal endog-
enous methyl donor and is a limiting factor in various
methylation reactions, including protein carboxymethy-
lation (135). Protein carboxymethylase is highly ex-
pressed in the brain and pituitary gland of several mam-
malian species (103, 105). The effects of SAM-induced
protein carboxymethylation on the formation of metha-
nol, formaldehyde, and formic acid in rat brain striatal
tissues were investigated; this study directly showed that
excessive SAM-dependent methylation increased levels
of methanol, formaldehyde, and formic acid in rat brain
striatal homogenates (265).
Formaldehyde can also be generated in the brain by the
reactions catalyzed by lysine-specific demethylase 1 and
JmjC domain-containing histone demethylases (79, 189,
288, 470, 472, 497).
In other metabolic pathways, SSAOs located in the outer
membranes of vascular smooth muscles and the endothe-
lium of the brain catalyze the deamination of methylamine
and generate formaldehyde together with ammonia and hy-
drogen peroxide (73, 233, 363, 527). It has been shown that
the incubation of methylamine in the presence of SSAO-rich
tissues, e.g., human brain meninges, results in cross-linked
DOROKHOV ET AL.
622 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
proteins. This cross-linkage can be completely blocked by a
selective SSAO inhibitor (155). Recently (386), the partici-
pation of brain SSAOs in formaldehyde generation was
shown in a well-established mouse model of Alzheimer’s
disease, the senescence accelerated mouse-prone 8 (SAMP8)
strain (466). Formaldehyde levels were elevated in the
brains of 3-mo-old SAMP8 mice (94.43 1.32
mol/g)
compared with those of age-matched senescence accelerated
resistant mouse 1 (SAMR1) mice (85.21 1.39
mol/g).
In contrast to serine hydroxymethyl transferase and
CYP450 2E1, SSAOs play a major role in the generation of
brain formaldehyde in SAMP8 mice. Levels of formalde-
hyde-producing SSAOs are higher in the brains of SAMP8
mice compared with SAMR1 mice, whereas the FDH
mRNA content was reduced in SAMP8 mice. Both the re-
duced ADH3 activity and SSAO mRNA/protein levels cor-
relate well with the observed formaldehyde accumulation in
the brains of SAMP8 mice; these data allowed the authors
to suggest a causal relationship between the two phenom-
ena (386).
2. Metabolism of methanol/formaldehyde introduced
into the brain through blood-brain barrier
One pertinent question is whether methanol and formalde-
hyde synthesized in the gastrointestinal tract and liver can
enter the blood and ultimately reach the brain by crossing
the blood-brain barrier (BBB) (472). The BBB ensures the
optimal control of homeostasis of the brain’s internal envi-
ronment. The anatomical structure of the BBB comprises
endothelial cells of arterioles, capillaries, veins, and epithe-
lial cell surfaces, with the presence of tight junctions be-
tween the endothelial cells of brain capillaries and oligoden-
drocytes. The endothelial cells enable the very selective
transport of substances from the blood to the brain, and
vice versa (326, 481). Methanol and ethanol from the blood
apparently reach the brain cells without hindrance (112,
159, 197). At the same time, peripherally produced acetal-
dehyde and formaldehyde penetrate the BBB to enter the
brain with difficulty. Although the physical-chemical prop-
erties of formaldehyde and acetaldehyde suggest that these
molecules would easily penetrate the BBB, the presence of
the “metabolic” barrier of ALDH and FDH hinders their
crossing of the BBB (89, 334, 410, 535).
3. The modes of brain methanol and formaldehyde
catabolism
ADH is responsible for the metabolism of ethanol and
methanol in the liver but has not been definitively demon-
strated to play a significant role in the brain. The dominat-
ing view is that ADH1 does not participate in methanol
metabolism in the brain (143, 217, 472). This conclusion is
strongly supported by the direct determination of the local-
ization of different classes of ADHs in the brain by North-
ern blot and enzyme activity analyses (143). This study did
not reveal any ADH1 activity in the brain, whereas class III
␹␹
ADH or FDH had high activity in the brains of adult
humans, mice, and rats. No ADH1 mRNAs were revealed
in the mouse brain following methanol inhalation (246).
Although ADH1 activity was not detected in the whole
brain homogenate, it was found to be expressed specifically
in granular and Purkinje cells of the cerebellum, indicating
that ADHs are likely to play a role in particular regions of
the brain (143). On the other hand, brain cell malignization
results in increased total ADH activity and, specifically,
increased class I ADH activity. The other tested classes of
ADH and ALDH enzymes did not show statistically signif-
icant differences in activity in cancer and normal cells (262).
The literature does not provide conclusive or reliable infor-
mation about phase I metabolism of methanol in the brain.
However, considering that ethanol is converted into acetal-
dehyde by the system that includes the participation of
CYP450s and the catalase-H
2
O
2
system, it is safe to assume
that these systems are also involved in methanol conversion
in the brain (178). Catalase and CYP450s represent the
major enzymes in the brain that catalyze ethanol oxidation.
CYP450s are abundantly expressed within the microsomes
of certain brain cells and are localized to particular brain
regions. There is evidence that catalase serves as the primary
ethanol-metabolizing enzyme in the brain (486) and ac-
counts for 60% of the ethanol oxidation in brain (537). The
studies of ethanol metabolism conducted by perfusing liv-
ing rat brains confirmed the participation of catalase in
acetaldehyde formation. Moreover, the addition of the cat-
alase inhibitor aminotriazole to the perfusing fluid in-
creased the ethanol level in the perfusate, i.e., ethanol elim-
ination was significantly decreased. The presence of
CYP450s in the brain has been established (178), but the
total level of CYP450s in the brain is much lower than that
in the liver, and these enzymes are concentrated in specific
regions and cell types, indicating their potentially consider-
able impact on local metabolism (138). The expression of
CYP2E1 and CYP2B6 within the brain can be substantially
increased in response to ethanol (137, 400). Brain CYP450s
are regulated by transcriptional, posttranscriptional, and
posttranslational mechanisms (138). Experiments on ani-
mals harboring genetic deficiencies in CYP2E1 indicated
that CYP2E1 is responsible for 20% of the ethanol me-
tabolism in the brain (537), but the mechanism responsible
for metabolizing the remaining 20% of ethanol remains
unclear (178, 375, 537).
Although no direct studies of the metabolic conversion of
formaldehyde in the mammalian brain have been con-
ducted, experiments with exogenous formaldehyde primar-
ily indicated its neurotoxicity (472). The extent of damage
depends on the dose of formaldehyde and the duration of
the exposure (13, 425, 426, 450). The conversion of form-
aldehyde into formic acid apparently occurs with the par-
ticipation of ALDHs (472). Fifteen Aldh isozymes and their
METABOLIC METHANOL
623Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
mRNAs were revealed in the mouse brain (3). After meth-
anol inhalation by mice, Aldh 1 and 2mRNA contents were
dramatically increased in the brain (246). The contribution
of ALDHs in the oxidation of formaldehyde in the human
brain is unknown. However, the mitochondrial ALDH2
activity toward formaldehyde in the liver was essentially
less in human subjects with mutant alleles (ALDH2*2)
compared with those with wild-type alleles (502).
The enzymatic degradation of formaldehyde in the brain is
also regulated by FDH/ADH3. In the human brain, FDH is
also likely involved in the oxidation of formaldehyde be-
cause its activity was detected in histological preparations
(328). The most intense FDH immunostaining was ob-
served in hippocampal pyramidal neurons, cerebellar Pur-
kinje cells, cerebral cortical neurons in layers IV and V, and
perivascular and subependymal astrocytes.
Studies conducted in adult rats and mice showed that Adh3
is the only ADH class present in the rodent brain (507). The
important role of this enzyme for mammalian life was
proven by studies of mice with a mutated Adh3 gene. Adh3-
knockout mice have a high sensitivity to formaldehyde and
a significantly reduced LD
50
value (96). Moreover,
Adh3/mice exhibited smaller litter sizes, smaller body
weights at 14 wk of age, and lower postnatal survival rates
compared with wild-type mice. Less than 20% of these mice
survive past postnatal day 6 (320). These features indicate
the inability of Adh3/mice to provide metabolic clear-
ance of endogenous formaldehyde. Reduced expression of
the Adh3 gene was also revealed in SAMP8 mice; along
with increased SSAO activity, high formaldehyde contents
in the brain and cognitive function disruption were ob-
served in these mice (386).
The metabolism of methanol in embryos is important for
mammalian procreation and has a great deal in common
with those in the central nervous system. In general, prena-
tal ethanol exposure has been found to be a risk factor for
fetal mortality, stillbirth, and infant and child mortality.
The elimination of ethanol from the fetus relies on the met-
abolic capacity of the mother (54). The main feature of
methanol metabolism in embryos is the lack of ADH1 ac-
tivity (507). Testing for Adh class mRNAs by in situ hybrid-
ization revealed the complete absence of Adh1 and Adh4
mRNAs from all brain structures at all prenatal stages in-
vestigated during the embryonic development of mice and
rats. Adh3 is the only ADH class investigated that was
found to be present in the rodent brain during embryogen-
esis. These data were in agreement with another in situ
hybridization study (143) that demonstrated only the pres-
ence of Adh3 in the brains of adult mice, rats, and humans.
The ubiquitous expression of Adh3 mRNA found in both
humans and mice from the earliest stages examined suggests
a housekeeping role for this gene. The important role of
Adh3 in the detoxification of formaldehyde also strength-
ens this suggested housekeeping function (432).
Thus the metabolic clearance of methanol and formalde-
hyde in the central nervous system and the embryo occurs
via two strategies. The first strategy aims to prevent the
oxidation of methanol, i.e., preventing in situ formaldehyde
formation. In accordance with this strategy, there is no
ADH1 activity in the brain and in embryos, thereby pre-
venting the creation of endogenous formaldehyde, i.e.,
formaldehyde synthesized in situ; furthermore, any metha-
nol introduced through the BBB is made nontoxic by phys-
iological clearance. The second strategy involves the oxida-
tion of any formaldehyde in the brain and in embryos that is
synthesized in situ or inserted into bloodstream by enzymes,
including FDH and ALDH2.
VI. METHANOL/FORMALDEHYDE AND
HUMAN PATHOLOGY
This section discuss the results of studies pointing to the
possible involvement of metabolic methanol and formalde-
hyde in human pathology. We believe that the mechanisms
of human poisoning by exogenous methanol, as well as the
practical therapies in cases of accidental methanol intake,
cannot be ignored.
A. Formaldehyde Neurotoxicity
Formaldehyde easily penetrates cell membranes, possesses
neurotoxic potential, and causes deficits in memory and
learning (386, 425, 466, 472). High levels of exogenous
formaldehyde without a concurrent elevation in the capac-
ity to clear formaldehyde raise the formaldehyde levels in
the body and lead to formaldehyde stress (171b, 472). Al-
though low concentrations of metabolic formaldehyde can-
not cause acute toxicity in brain cells, experiments on cul-
tured cells have provided some idea of how the metabolism
in the brain is changed by the putative impact of formalde-
hyde. It has been shown that the metabolism of cultured
astrocytes is altered after formaldehyde exposure (471). Al-
though the cells efficiently oxidized formaldehyde to for-
mate, likely by using ALDH2 and FDH, enhanced glyco-
lytic flux was observed due to the inhibition of mitochon-
drial respiration. In cultured cerebellar granule neurons
exposed to formaldehyde, a high rate of formaldehyde ox-
idation accompanied by significant increases in the cellular
and extracellular formate concentrations increased glucose
consumption, and lactate release strongly accelerated the
export of the antioxidant glutathione (473). Hydrogen sul-
fide (H
2
S) is considered the third gaseous neuromediator,
along with nitric oxide and carbon monoxide (327, 500,
501). In the mammalian brain, H
2
S is formed from the
amino acid cysteine by the action of cystathionine
-syn-
thase, the key enzyme in the transsulfuration pathway that
DOROKHOV ET AL.
624 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
processes homocysteine, which induces neurotoxicity (449)
and defects in the learning and memory of rats (277). In
Alzheimer’s disease brains, H
2
S was found to be decreased,
but homocysteine was increased (128). Formaldehyde has
also been shown to decrease endogenous H
2
S generation
and to induce neurotoxicity and intracellular ROS accumu-
lation in PC12 cells derived from rat adrenal medulla pheo-
chromocytoma (448). The interaction of formaldehyde
with H
2
S was the trigger of learning and memory dysfunc-
tions in rats (448).
B. Formaldehyde and Neurodegenerative
Diseases
Although the concentration of formaldehyde is low in the
blood, it is high in the brain, suggesting that formaldehyde
plays roles in brain function and dysfunction (441). There is
a direct correlation between elevated formaldehyde levels in
the brain and memory impairment (463, 464, 466); it is
currently unclear whether formaldehyde or a disruption in
gene expression is the cause of the pathogenesis of the men-
tal disease. Notably, brain formaldehyde concentrations
were significantly elevated to pathological levels (0.5
mM) in aged rats. It has been suggested that excess formal-
dehyde is associated with a global decline DNA methylation
(DNA demethylation) in the hippocampus during aging
(464). Intrahippocampal injection with excess formalde-
hyde induces memory deficits in adult rats. Furthermore,
intrahippocampal injection with excess formaldehyde (0.5
mM) not only obviously impaired the ability of spatial
learning in the adult rats but also damaged the memory
abilities of the rats (464, 465). In humans, a correlation was
observed between the degree of senile dementia and the
levels of urine formaldehyde in different patients (466).
Moreover, brain formaldehyde was increased in Alzhei-
mer’s disease patients (463, 466) and animal models (365,
386). The cortex and hippocampal formaldehyde levels of
Alzheimer’s disease patients were significantly higher than
those of age-matched controls or young people, reaching
0.5 mM (463). Interestingly, regularly drinking water can
decrease endogenous formaldehyde. Furthermore, this ac-
tivity can mitigate age-related cognitive impairment (460).
Highly reactive formaldehyde can be both a cause and a
consequence of pathological processes in humans. To un-
derstand the mechanism of its pathological effects on brain
cells, the chemical properties of formaldehyde must be
taken into account. The polarized carbonyl group of form-
aldehyde reacts with the amine group of lysine or arginine,
forming Schiff bases and producing irreversibly covalently
cross-linked complexes between proteins, as well as be-
tween proteins and single-stranded DNA (73, 155, 171b,
307, 432, 534). Formaldehyde, which is found at elevated
levels in Alzheimer’s disease (386, 439, 448, 450, 463, 464,
517, 518), may play important roles in
-amyloid (A
)
aggregation and cerebral amyloid angiopathy related to
Alzheimer’s disease pathology (319, 421, 428, 478, 525).
The potential implications of endogenous aldehydes in A
misfolding, oligomerization, and fibrillogenesis was con-
firmed by in vitro experiments in which formaldehyde was
capable not only of enhancing the rate of formation of A
oligomers and A
-APOE protein complexes but also of in-
creasing the size of the aggregates (72, 73, 394). In line with
this hypothesis, the overexpression of SSAOs, which were
colocalized with A
deposits, was detected in the cerebro-
vascular tissue of patients with Alzheimer’s disease (210,
478).
Another biomarker of Alzheimer’s disease is neuronal tau,
an important protein in the promotion and stabilization of
the microtubule system that is involved in cellular transport
and neuronal morphogenesis (429). Whereas tau normally
contains 2–3 mol of phosphates per mole, tau phosphory-
lation levels in Alzheimer’s disease brains are three- to four-
fold higher. Formaldehyde at low concentrations induces
tau polymerization into globular amyloid-like aggregates in
vitro and in vivo (191, 344). As shown in the study of mouse
neuroblastoma (N2a) cells treated with formaldehyde, tau
hyperphosphorylation and DNA damage occurred (296).
Tau became hyperphosphorylated not only in the cyto-
plasm but also in the nucleus of mouse brains (292).
Chronic methanol feeding led to pathological changes that
were related to Alzheimer’s disease development, including
tau hyperphosphorylation in the brains of mice (517) and
rhesus macaques, non-human primates (518).
Indirect evidence of the involvement of formaldehyde in the
pathogenesis of Alzheimer’s disease was obtained by inves-
tigating ALDH2 activity, a key enzyme that oxidizes alde-
hydes, including formaldehyde. The hypothesis that a de-
crease in ALDH2 activity contributes to Alzheimer’s disease
was supported by studies of transgenic mice with low
ALDH2 activity; these mice exhibited age-dependent neu-
rodegeneration and memory loss (352).
In humans, ALDH2 dysfunction is likely to be one of the
risk factors for Alzheimer’s disease. The ALDH2*2 allele is
a risk factor for late-onset Alzheimer’s disease. Approxi-
mately 0.6–1% of the Japanese population are estimated to
belong to the group with a combination of ALDH2*2 ge-
notypes, and nearly all of these people are expected to de-
velop Alzheimer’s disease (221, 352). A Chinese survey con-
firmed these data (498).
C. Formaldehyde and Hangover
Methanol and formaldehyde are known to participate in
the development of a hangover, known for unpleasant sen-
sations and uncomfortable “morning after” symptoms in
people following excessive ethanol intake (373, 377).
Hangovers after binge drinking are likely more common in
the young than in older aged persons (462). Among hang-
METABOLIC METHANOL
625Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
over causes, which include imbalances in the immune sys-
tem (373), effects of dehydration such as headaches (415)
and sleep disturbances, acetaldehyde accumulation, dys-
regulated cytokine pathways, and hormonal alterations
(487, 510), the metabolism of methanol and production of
formaldehyde are likely the most important (28, 446). The
first support for the contribution of methanol to hangovers
comes from data showing that brandies and whiskeys,
which are more frequently associated with the development
of a hangover, contain the highest methanol concentra-
tions. The first excellent study of methanol metabolism and
hangover found that methanol accumulated in the blood of
alcoholic subjects during a 10- to 15-day period of chronic
ethanol intake (299). The disappearance of blood methanol
lagged behind the linear disappearance of ethanol by 6–8
h, and complete clearance of blood methanol took several
days. Importantly, the accumulation and clearance patterns
of methanol and ethanol were similar in subjects who con-
sumed either whisky (bourbon) with a high methanol con-
tent or grain alcohol with low methanol content. The au-
thors suggested that methanol accumulates in the blood as a
result of the well-known competitive inhibition of ADH by
ethanol (263) and the presence of endogenous methanol
(413), which may contribute to the hangover severity (299).
An experimental study with healthy subjects who con-
sumed red wine containing 100 mg/l of methanol also
showed that elevated blood methanol levels persisted for
several hours after the ethanol was metabolized and that
this corresponded to the time course of hangover symptoms
(212, 213). The half-life of methanol in healthy men during
a hangover was estimated to be 142 min. This indicates that
elevated methanol concentrations in the blood persist for
12 h (213). The author suggested that methanol lingers
after ethanol levels drop because ethanol competitively in-
hibits methanol metabolism. Ethanol readministration
(“hair-of-the-dog” drinking) (377) fends off the hangover
effects that may be based on the ability of ethanol to block
methanol metabolism, thereby slowing the production of
formaldehyde and formic acid (213, 413).
D. Effect of Alcoholic Beverages on the
Cardiovascular System: U-Shape
The consumption of ethanol has an effect on the cardiovas-
cular system in humans and can cause coronary heart dis-
ease (CHD) (333, 355). A plethora of epidemiological evi-
dence has demonstrated a J- or U-shaped association be-
tween alcoholic beverages consumption and all-cause
mortality, as well as cardiovascular morbidity and mortal-
ity. On the other hand, moderate alcoholic beverages con-
sumption was inversely associated with CDH mortality
(144, 147, 241, 305, 335, 337, 366, 402). According to the
Dietary Guidelines for Americans (314), moderate alco-
holic beverages consumption is considered an intake of no
more than 1 drink per day for women and no more than 2
drinks per day for men, where 1 drink is equal to 12 g
ethanol. Moderate ethanol consumption also has a benefi-
cial effect in reducing the risk of vascular disorders of the
brain. Heavy alcoholic beverage consumption increases the
relative risk of any type of stroke, whereas light or moderate
ethanol consumption may be protective against ischemic
stroke (241, 366). The mechanism of this phenomenon is
not completely understood. Numerous hypotheses have
been proposed to explain the benefit of light-to-moderate
ethanol intake on the heart and brain, including an increase
of high-density lipoprotein cholesterol (47) and the promo-
tion of antioxidant effects with the participation of a pro-
tein kinase B/nuclear factor (erythroid-derived 2)-like 2-de-
pendent mechanism (77, 241, 494). A recent study also
suggested the involvement of the ADH1b gene in the U-
shaped curve. A survey of over 260,000 individuals showed
that carriers of the rs1229984 A-allele (ADH1B*2,TABLE
3) with a 100-fold increase in
2-ADH turnover (215, 308)
consumed less ethanol and had a reduced frequency of cor-
onary heart disease compared with noncarriers of this allele
(184). In the search for mechanisms to explain the U-shaped
relationship between ethanol consumption levels and coro-
nary heart disease, researchers have focused on methanol
and its oxidation product, formaldehyde. Recently, formal-
dehyde was hypothesized to participate in the process
(322). The proposed mechanism relies on the fact that the
common ADH1b enzyme carries out one-phase catabolism
of methanol and ethanol. Moderate ethanol consumption
competitively inhibits the conversion of methanol to form-
aldehyde, thus reducing the endogenous formaldehyde con-
tent in the organs. Another possible mechanism for the
beneficial effects of moderate ethanol intake was indicated
by the results of a study in rats showing that ALDH2 oxi-
dizes aldehydes, including formaldehyde, and may serve as
a potential endogenous neuroprotective target and a prom-
ising therapeutic strategy for the management of stroke
(438). Ethanol administration activated ALDH2 and en-
hanced the detoxification of aldehydes (161). The study of
ALA also indicated that ALDH2 participates in the formal-
dehyde and acetaldehyde detoxification process (116, 171a,
285, 310).
E. Methanol Fatalities and Antidote Therapy
Many cases of poisoning that result in fatalities are caused
by methanol ingestion. When methanol was discovered,
nothing was known about the toxicity of its metabolites in
humans; thus the use of methanol as a substitute for ethanol
became widespread in industry (453). Quickly thereafter,
the dangers of methanol intoxication through skin absorp-
tion and inhalation became apparent (311), as numerous
cases of industrial poisoning were reported (263). More-
over, as the odor and appearance of methanol are almost
the same as those of ethanol, these two alcohols were often
confused, and methanol was used instead of ethanol as a
beverage, thereby increasing the cases of poisonings and
fatalities.
DOROKHOV ET AL.
626 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
Today, methanol is more widely used in industry, but the
number of industrial poisonings are reduced due to the
existence of the necessary precautions and safety instruc-
tions. Nevertheless, methanol poisoning and fatal cases re-
main a problem. These cases are mainly due to the intended
or unintended ingestion of “surrogate alcohols,” namely,
non-beverage alcohol (e.g., industrial spirits, fire -liquids,
windshield washer liquid, antifreeze), moonshine, home-
made alcohol, samogon, etc. (258). Cases of homicide and
suicide with methanol-containing liquids have also been
reported (214). Fewer cases are related to methanol intox-
ication through percutaneous methanol absorption or va-
por inhalation (316a).
The minimal dose of methanol that could lead to a lethal
outcome without adequate treatment is currently consid-
ered to be 1 g/kg (396), and 10–15 ml may cause blind-
ness (316). According to the literature, the toxic effects of
methanol could be observed after the ingestion of 15 ml of
40% methanol (29) to 600 ml of pure methanol (253).
These significantly different results could be explained by
insufficient and incorrect data regarding the amount of al-
cohol consumed; usually, the ingested volume is self-re-
ported by the patients or is indirectly obtained. Moreover,
the important role of different factors associated with the
ingestion of methanol, including joint ethanol-methanol
consumption, for example, should not be underestimated.
Furthermore, the variability of toxic or lethal doses is de-
pendent on the general state of health, diet, and the amount
of tetrahydrofolate, which is known to take part in metha-
nol metabolism, of the organism (396, 453).
Upon ingestion or inhalation, methanol initially has a nar-
cotic effect as ethyl alcohol, followed by an asymptomatic
period of 10–12 h. After this period, methanol metabo-
lites may cause nausea, vomiting, dizziness, headaches, re-
spiratory difficulty, abdominal pain, visual disturbances,
and metabolic acidosis (396). The main consequences of
methanol intoxication are metabolic acidosis, hyperosmo-
lality, formic acid, lactic acid, formaldehyde and ketones
accumulation in the organism, and retinal damage with
blindness (253). The high level of formic acid (formate) is
believed to be the main cause of the onset of the clinical
signs of methanol poisoning (23). Formate accumulates in
bodily fluids and tissues, leading to metabolic acidosis and
blood bicarbonate depletion (409). Formate inhibits mito-
chondrial cytochrome oxidase, resulting in tissue hypoxia;
this effect is enhanced when the pH is low. Notably, for-
mate injection per se leads to optic disk damage, indepen-
dently of acidosis; the ocular effects associated with meth-
anol poisoning appear to be due to hypoxia in areas of the
cerebral and distal optic nerves (276). One study directly
demonstrated this by intravenously injecting monkeys with
formate at concentrations similar to those observed in
methanol poisoning. The blood formate levels in these ani-
mals were constantly elevated for 2448 h, whereas the
blood pH maintained at physiological levels. All animals
developed ocular toxicity within 24 h, and its symptoms
were the same as those of methanol poisoning. Thus ocular
toxicity is caused by formate but not by methanol or acido-
sis (306). Further studies showed that acidosis correction
did not prevent ocular toxicity.
One of the first massive methanol poisoning outbreaks to be
documented was in the 1940s. During the Second World
War, more than 100 fatal cases were registered in the Ger-
man army. Lachenmeier (258) summarized further out-
breaks of methanol poisoning from surrogate alcohols that
have been reported in the scientific literature since the
1950s.
Modern toxicology handbooks recommend treating the
consequences of methanol intoxication when blood metha-
nol concentrations are more than 0.2 g/l in a nonacidotic
patient. However, it is likely that this threshold was chosen
arbitrarily (251), as this approach does not take into ac-
count the time passed since the moment the methanol-con-
taining liquid was ingested. For example, if the blood meth-
anol content is 0.2 g/l 12 h after ingestion, then the peak
level 1 h after ingestion could be estimated as more than 1
g/l. The same concentration of 0.2 g/l measured 24 or 48 h
after ingestion corresponds to peak levels of 2.15 or 4.2 g/l,
respectively (251). According to this approach, the amount
of ingested methanol might be significantly underestimated,
which could lead to improper treatment. Thus the time and
circumstances of the ingestion should be taken into account
whenever possible. The levels of blood formate should also
be assessed because the relative concentrations of methanol
and formate in the blood are functions of the amount of
methanol ingested. Different methods, including enzymatic
measurements and gas chromatography, for formic acid
analysis in both ante-mortem and postmortem blood sam-
ples were used to provide additional information to assist
with the interpretation of methanol fatalities (190, 495).
The best picture of toxic and lethal blood methanol concen-
trations was obtained from the analysis of postmortem
methanol and formate levels. Blood methanol and formate
contents estimated ante-mortem and postmortem were
shown to be very close. The case report study performed by
Jones (214) analyzed postmortem methanol and formate
concentrations in 73 cases; blood methanol levels varied
significantly, ranging from 0.7 to 7.9 g/l, whereas formate
concentrations showed a more narrow range between 0.53
and 1.40 g/l. In 97% of cases, the formate concentrations
ranged from 0.60 to 1.10 g/l. Similar ranges were detected
by Wallage (495). Another study of 15 cases of methanol
poisoning (no fatalities included) reported approximate
blood methanol levels ranging from 0.14 to 4.5 g/l and
formate levels ranging from 0.01 to 1.48 g/l. The authors
also noted that there was no correlation between methanol
and formate levels. Nonetheless, the levels of both com-
METABOLIC METHANOL
627Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
pounds are strongly recommended to be monitored in hos-
pitals for patients with methanol poisoning (190).
Methods of treatment for methanol poisoning are highly
dependent on how quickly treatment is initiated and the
stage of poisoning. Gastric lavage, vomiting induction, and
the use of activated charcoal can only be conducted within
1 h after methanol ingestion (253); after 1 h, the methanol
is fully absorbed by the gastrointestinal tract. For those
patients, therapy with inhibitors of subsequent methanol
metabolism can be applied. After methanol is rapidly and
completely absorbed by the gastrointestinal tract, it pene-
trates all bodily fluids. With the participation of liver ADH,
methanol then is metabolized to formaldehyde and formic
acid. The following methods can be applied as therapies for
methanol poisoning (399): 1) intravenous bicarbonate in-
jection to reduce metabolic acidosis; 2) the use of antidotes
such as ethanol and fomepizole (4-methylpyrazole) to ter-
minate the conversion of methanol to formic acid; and 3)
hemodialysis to remove methanol and formic acid from the
blood, used for uncomplicated poisoning with methanol
(316).
Therapy with the use of ethanol or fomepizole can be used
in cases in which the time of methanol intake is known and
the methanol levels in the blood rise above 10–20 mg/dl
or the osmolal gap is 10 mosmol/kgH
2
O. In cases in which
the time of methanol consumption is not known, any two of
the following criteria should be present: a bicarbonate con-
centration in the blood serum of 20 meq/l, an arterial
hydrogen rate (pH) less than 7.3, and an osmolal gap of
10 mosmol/kgH
2
O (22, 87, 227, 230). Treatment with
ethanol is the most accessible, due to cost effectiveness and
ease of use (peroral or intravenous), but to use ethanol as an
antidote, high serum levels of 100 mg/dl must be main-
tained. Thus it is not recommended for patients with liver
disease and stomach ulcers. Additionally, ethanol should be
used with caution in patients who use drugs that suppress
the central nervous system because ethanol may exacerbate
their condition (1, 23, 253).
Fomepizole, on the other hand, has a higher affinity for
ADH than ethanol (by close to 1,000 times) and can com-
pletely inhibit ADH1b at significantly lower serum concen-
trations (46, 316). Fomepizole concentrations of more than
0.8 mg/l (10 mM) in the serum provide permanent inhibi-
tion of ADH1b (23). For patients not undergoing hemodi-
alysis, the recommended doses of fomepizole are known
(392): an initial dose of 15 mg/kg, followed by 10 mg/kg
every 12 h, and then an increased dose of 15 mg/kg at 48 h,
followed by another 15 mg/kg dose after 12 h.
For patients undergoing hemodialysis, two options have
been proposed for fomepizole administration. In the first
case, the dosage remains the same as for patients not under-
going hemodialysis but with a reduced time interval be-
tween successive doses: the second dose is given 6 h after the
first dose, and subsequent doses are given every 4 h. In the
second case, continuous intravenous infusion of 1.0–1.5
mg·kg
1
·h
1
fomepizole is performed after the primary
dose. However, the use of fomepizole to treat methanol
poisoning is not possible for all patients at every clinic.
Most people with methanol poisoning are alcoholics who
cannot afford such an expensive drug; thus ethanol is more
commonly used in such cases (253, 392).
VII. GENES INVOLVED IN ENDOGENOUS
METHANOL CATABOLISM AND
HUMAN PATHOLOGY
A. Cluster of Genes Involved in Endogenous
Methanol Catabolism
Carbon dioxide and nitrogen oxide are signaling molecules
with small molecular weights that are widespread in the
environment. Methanol might be another example of a
small regulator of gene expression. Recently, a number of
genes were shown to undergo expression changes due to
increased methanol levels in the plasma of humans and
mice. Most of the genes affected by methanol fluctuation
are involved in ethanol metabolism (ADH1, ALDH2,
GSTO1, MGST3, GSTP1,CYP2E1), Alzheimer’s disease
pathogenesis (PSENEN, APOE, SNCA, MME), and hemo-
globin production (HBB2, HBA1) (246, 413). Although
ethanol was known to alter the gene expression of alcohol-
metabolizing enzymes in mammals, our data reveal the sig-
naling function of methanol or formaldehyde in mammals
for the first time. There are many examples, including form-
aldehyde, in which a product or substrate of a reaction can
affect the gene expression of an enzyme. Alcohol oxidase
from the methylotrophic yeast Pichia pastoris was shown to
undergo expression changes in the presence of methanol
(482a). Further investigation revealed a formaldehyde-sen-
sitive transcriptional factor that can alter the expression of
the methanol-oxidizing enzymes in yeasts (65, 374). At the
same time, increased methanol in the blood of mice led to
overexpression of the brain genes involved in formaldehyde
clearance (Aldh2,Gsto1,Mgst3,Gstp1) and downregulation
of Adh1, which participates in formaldehyde formation (246,
413). These changes in gene expression of brain tissue in
mice may control toxic formaldehyde concentrations.
However, Cyp2E1, which oxidizes ethanol and methanol
to aldehydes, was upregulated (246). Thus methanol or
formaldehyde can alter the expression of the genes involved
in methanol oxidation.
B. Alcoholism
Polymorphisms in genes that affect ethanol utilization are
of particular interest in heavy drinkers. Mutations in ADH
and ALDH genes that affect the enzymatic activity and
DOROKHOV ET AL.
628 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
protein structure could be associated with the risk of alco-
holism. Twin surveys revealed that the genetic influence on
the development of alcoholism was as important as that of
environmental factors (172, 173). Polymorphisms in en-
zymes such as ADH1b and ALDH2 that metabolize ethanol
are mostly associated with alcoholism. Genotypes with
ADH1 and ALDH2 alleles that lead to greater acetaldehyde
accumulation are protective against excessive ethanol
consumption. For example, the ADH1B*2 (Arg48His,
rs1229984) and ADH1B*3 (Arg370Cys, rs2066702) vari-
ants are able to oxidize ethanol to acetaldehyde much faster
(122) and are associated with a protective effect against
alcoholism in East Asian (122), African American (121),
and American Indian populations (372, 496).
Another way to increase acetaldehyde formation is to block
the activity of ALDH2, which is the primary enzyme that
oxidizes acetaldehyde to acetic acid. East Asian populations
with inactive ALDH2 accumulate acetaldehyde to a greater
extent while drinking, which confers a protective effect
against ethanol dependence (182). Conversely, carriers of
the SNPs in the ADH4 promoter were found to have a
higher risk of alcohol dependence (158, 294). Edenberg et
al. (123) demonstrated that one of the SNPs (75A) in-
creased the expression of corresponding gene compared
with the 75C variant (123). Individuals with cysteine at
the 75 position are three times more likely to develop
ethanol dependence (158).
C. Neurodegenerative Diseases
Thus far, the scheme of Alzheimer’s disease pathogenesis is
incomplete; however, there is strong evidence that oxidative
stress triggers and participates in Alzheimer’s disease devel-
opment. Mitochondrial dysfunction is strongly associated
with oxidative stress and neurodegenerative disorders, in-
cluding Alzheimer’s disease (279). Thus changes in mito-
chondrial proteins with antioxidative stress implications
may be involved in neurodegenerative diseases. Examples
include the cytochrome coxidizing complex, manganese
superoxide dismutase, and catalase and ALDH2, two en-
zymes that are important in alcohol metabolism. Whereas
catalase mainly oxidizes ethanol and methanol at high con-
centrations, ALDH2 is a central protein in the oxidation of
relational aldehydes. One of the major polymorphisms of
ALDH2 is the substitution of glutamate at the 487th posi-
tion by lysine. This variant, ALDH2*2, is widespread
among the Japanese population, in which up to 30% of
people are heterozygous (447). In individuals who carry at
least one ALDH2*2 allele, a lower K
m
results in higher
acetaldehyde concentrations, whereas ALDH2*2 homozy-
gotes show no ALDH2 activity (447). Because ALDH2 lo-
calizes to the mitochondrial matrix and is implicated in the
oxidation of aldehydes generated in oxidative stress,
ALDH2 was proposed to play a role in Alzheimer’s disease
pathogenesis. The low activity of ALDH2*2 causes the ac-
cumulation of highly toxic 4-hydroxy-2-nonenal (4-HNE),
which was shown to cause neuronal death (254) and was
observed in Alzheimer’s and Parkinson’s disease patients
(325, 404, 522). Analysis of the ALDH2*2 polymorphism
in the 472 Alzheimer’s disease patients and 472 nonde-
mented controls revealed a small but statistically significant
difference in the frequency of the allele (48.1 and 37.4%,
respectively) (221). Interesting, APOE-
4 homozygotes car-
rying at least one ALDH2*2 allele have a 31-fold higher
chance of developing Alzheimer’s disease than those with
neither allele (352). Because both proteins control the con-
centration of 4-HNE, the dysfunction of ALDH2*2 dra-
matically increases the levels of 4-HNE. While the apolipo-
protein E-4 (APOE-4) polymorphism is found in more
Alzheimer’s disease patients compared with healthy indi-
viduals than the ALDH2*2 allele, the coexistence of both
alleles dramatically increases the probability of developing
Alzheimer’s disease. Thus ALDH2*2 itself is not a strong
prognostic factor for Alzheimer’s disease, but dramatically
increases the risk of the Alzheimer’s disease development in
APOE-4homozygotes (352).
D. Cardiovascular Disease
Recent data revealed a crucial protective role of ALDH2 in
ischemic injury. ALDH2 is activated by protein kinase C-
via phosphorylation (71) and then detoxifies poisonous al-
dehydes such as 4-HNE and acetaldehyde (36, 88). Because
4-HNE can trigger necrotic cell death in the reperfused
myocardium (338) and blocks important metabolic pro-
teins such as glyceraldehyde 3-phosphate dehydrogenase
(475), Na
-K
-ATPase (417), and the 20S proteasome
(134), its deactivation by ALDH2 has a cytoprotective ef-
fect in ischemic patients. Additionally, ALDH2 participates
in the bioactivation of nitroglycerin, leading to increased
blood flow (321). Based on these data, ALDH2 is a prom-
ising target for the treatment of ischemic injury by activa-
tors such as ALDH2 activator 1 (Alda-1) and ALA (71,
171a). Alda-1 not only increases ALDH2 levels and restores
ALDH2*2 activities (53), but also prevents 4-HNE-medi-
ated blocking the function of ALDH2 in ischemic conditions
(71). Undoubtedly, the mechanisms regulating the transcrip-
tion, translation, and enzymatic activity of ALDH2 require
further investigation.
E. Tumoral Diseases
ADH1 and ALDH1 can oxidize a variety of the alcohol
metabolites, including aliphatic alcohols and retinol (vita-
min A). ADHs and ALDHs are implicated in developmen-
tal, apoptotic, metabolic, and signaling processes. The dys-
function of ADHs and ALDHs on different levels can lead
to or be associated with different disorders. Due to the
crucial role of ADH1 and ALDH1 in the retinoic acid path-
way, this pathway is of particular interest. The retinoic acid
METABOLIC METHANOL
629Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
pathway controls the processes of differentiation, prolifer-
ation, and apoptosis and was found to participate in the
development of various tumors (132, 280, 339).
1. Uterine fibroid
Among women, uterine fibroid (UF) is a widespread benign
tumor of the smooth muscle cells that rarely transforms into
malignant leiomyosarcoma. Due to uncontrolled fibroid
growth, UF can cause various pains, frequent urination, and
infertility in 1.0–2.4% of patients (109). Molecular factors
involved in UF development include retinoic acid, insulin
growth factor 2, transforming growth factor
pathways,
and extracellular matrix formation. In UF samples, ADH,
ALDH1, CRBP1, RXR
, and RXR
mRNA and protein
levels are significantly lower than in healthy controls (93,
529). ADH1 and ALDH1 oxidize retinol to retinal and
retinoic acid; thus decreased RA levels are observed in UF
samples. The reduced concentration of retinoic acid could
lead to typical tumor characteristics: decreased differentia-
tion and apoptosis together with increased proliferation
(529). Thus ADH and ALDH may play an important role in
UF development and could be targets in UF treatment.
2. Head and neck cancer
Acetaldehyde is a carcinogenic metabolite, and its concen-
tration is controlled by the enzymatic activities of ADHs,
ALDHs, catalase, and CYP2E1 (49). Because the main
source of acetaldehyde is ethanol consumption, this carcin-
ogen mostly affects the upper aerodigestive tract (408).
Variants of the ADH and ALDH genes that lead to changes
in the activities of their protein products are associated with
the development of head and neck, esophageal, and liver
cancers. Although variants that produce acetaldehyde in
greater concentration would be expected to increase the risk
of cancer, this is not always the case. Homozygotes for
the ADH1B*2 (Arg48His; rs1229984) and ADH1C*1
(Arg272Gln; rs1693482) allele variants were shown to pro-
duce 40- and 2.5-fold more acetaldehyde from ethanol, re-
spectively (39). However, meta-analyses revealed an inverse
correlation between the head and neck cancer risk and car-
rying the variants of ADH1C and ADH1B that metabolize
acetaldehyde faster (68). The authors provide three possible
explanations for this unexpected finding: 1) a decreased
opportunity for the oral microflora to produce acetalde-
hyde locally by the prolonged systemic circulation of etha-
nol, 2) the prevention of ethanol acting as a solvent for
other carcinogens, and 3) decreased amounts of ethanol
consumption due to the discomfort caused by the conse-
quent peak in systemic acetaldehyde. We propose that
formaldehyde might also affect the results: large concentra-
tions of formaldehyde can be found in human body after
alcoholic beverage consumption, and formaldehyde is
known to have a carcinogenic effect. Although little evi-
dence is available regarding formaldehyde causing NPC
(445), studies have demonstrated that chronic exposure to
formaldehyde causes squamous cell carcinoma in rats (323,
324, 444).
3. Colorectal cancer
ADH1 and ALDH may contribute to colorectal cancer
(CC). Biochemical tests reveal increased ADH1 activity in
CC samples compared with healthy controls (205). More-
over, ALDH activity in cancer cells is slightly lower than in
healthy tissues; the ratio between ADH1 and ALDH activ-
ities is 20.5:1.0 in cancer cells and 10:1 in healthy cells.
Acetaldehyde seems to affect the relationship between the
ethanol-oxidizing system and CC. In CC cells, carcinogenic
acetaldehyde accumulates roughly twofold faster than in
healthy cells, indicating that acetaldehyde may participate
in the progression of CC. Additionally, ADH1 activity in
the sera of CC patients is increased and rises with the stage
of cancer (206).
VIII. CONCLUDING COMMENTS
Over the past decade, significant progress has been made in
elucidating the functions and physiological roles of human
metabolic methanol. Moreover, recent advances in the field
of metabolic methanol research have provided insights into
the mechanisms by which low levels of formaldehyde are
maintained in human plasma. However, numerous aspects
of metabolic methanol synthesis and the regulation of meth-
anol-metabolizing genes are not understood.
Methanol is an integral and inevitable component of human
life. In the morning woods (193) or on a mowed meadow
(95a, 230, 505), inhaled air features a noticeable content of
plant methanol vapors. The blood of fasting individuals
contains small amounts of methanol and its oxidative prod-
uct, formaldehyde; the levels of these molecules increase
sharply even after the ingestion of vegetable pectin (112,
413). There are several metabolic sources of methanol (FIG-
URE 1). In addition to fruit and vegetables, methanol can be
consumed in soft drinks containing aspartame and alco-
holic beverages, which contain methanol as a product of the
fermentation of vegetative raw materials with PME. In ad-
dition to the processes of methylation-demethylation, an
important source of metabolic methanol is the intestinal
microflora. The detection of methanol even after inhibition
of liver ADH1b by 4-methylpyrazole (246, 403) or admin-
istration of ethanol-free methanol (413) clearly indicates
the existence of internal sources of methanol.
The mechanisms of phase I of the catabolism of methanol
share several features of ethanol catabolism. In humans, the
oxidation of methanol and ethanol requires several stages
of conversion. In the first stage, methanol is oxidized to
formaldehyde in three ways. The first is via oxidation by
CYP450s (57, 87, 495). The contribution of CYP450s to
DOROKHOV ET AL.
630 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
the metabolism of methanol and ethanol in the liver is low
(9%) and increases at higher doses of methanol and eth-
anol. The second form of methanol oxidation occurs via the
catalase-H
2
O
2
complex (64, 97). Its contribution to the
metabolism of methanol and ethanol in the liver is small
(1–2%) but increases at higher doses of methanol and
ethanol. The third means of methanol oxidation involves
ADH1b, which catalyzes up to 90% of methanol and eth-
anol oxidation in the liver (63, 168, 297). ADH1b activity
has not been detected in brain extracts (143). The enzymatic
reaction occurs in two stages. The first is the formation of
the methanol/ADH1b/NAD
ternary complex with the
participation of Zn. At this stage, methanol and ethanol
enter the complex equally effectively; thus ethanol func-
tions as an antidote for methanol poisoning (382). In the
second stage, hydride transfer occurs with the formation of
CHO NADH H
. Because ADH1b is involved in both
methanol and ethanol catabolism, ethanol functions as a
powerful competitive inhibitor at low concentrations, and
the enzyme has a strong preference for converting ethanol
to acetaldehyde over converting methanol to formaldehyde.
Methanol itself is not toxic to human cells; however, its
oxidative product, formaldehyde, is a toxin that is believed
to play a role in carcinogenesis and age-related neuronal
damage in the brain (472). Although formaldehyde is pri-
marily produced by the oxidation of methanol, there are
several sources of metabolic formaldehyde, including
SSAO-mediated oxidative deamination and the removal of
methyl groups from lysine residues in histones catalyzed by
histone demethylases.
The second phase of methanol catabolism, i.e., the conver-
sion of formaldehyde to formic acid, occurs via three mech-
anisms, including the participation of CYP2E1 (27, 183,
255, 289, 452), cytosolic AlDH1 and mitochondrial
AlDH2 (142, 297, 452, 453, 473, 485), and FDH (432).
There are two mechanisms by which metabolic formalde-
hyde can be maintained at low levels. The first is to decrease
ADH activity to prevent the oxidation of methanol to form-
aldehyde. The second method is the rapid and effective
oxidation of formaldehyde to the end products carbon di-
oxide and water. These two methods are implemented in
the brain and embryo, both of which are highly sensitive to
formaldehyde. ADH activity is decreased or absent in the
brain and the embryo (143), and increased FDH (432) and
ALDH2 (472) activity oxidize formaldehyde introduced via
the bloodstream or formed in situ.
In conclusion, human methanol and formaldehyde catabo-
lism is complex, multilayered, and highly effective. Meta-
bolic control can be achieved at the gene or protein level,
i.e., at the level of enzyme activity. Methanol is not poison-
ous itself but rather acts as a signaling molecule that regu-
lates the activity of a gene cluster involved in the mainte-
nance of a low level of formaldehyde (112, 246, 413). The
disruption of formaldehyde metabolism control may be a
causative factor in neurodegenerative diseases. Low levels
of formaldehyde have been associated with some human
pathologies. Increased formaldehyde content in the blood
and brain has been detected in neurological patients as well
as in the blood of the elderly, suggesting a disruption of the
genetic and biochemical mechanisms responsible for main-
taining low formaldehyde levels (463, 466). To determine
the beneficial or detrimental effects of methanol, we must
also consider the favorable role of fruits and raw vegetables
in human health. A vegetarian diet is the main source of
exogenous methanol in a healthy individual (112). The role
of methanol-generated pectin in atherosclerosis and cancer
prophylaxis is well-known (389). It has been suggested that
pectin can modulate detoxifying enzymes, stimulate the im-
mune system, modulate cholesterol synthesis, and act as an
antibacterial, antioxidant, or neuroprotective agent. There
is also increasing evidence to suggest that regular fruit and
vegetable consumption may play an important role in pre-
venting or delaying the onset of dementia, age-associated
cognitive decline, and Alzheimer’s disease (127).
Thus advances in modeling and an analysis of human meth-
anol metabolism will extend our knowledge of the role of
methanol in health and disease, permitting the customiza-
tion of existing and future therapeutic and prophylactic
modalities.
ACKNOWLEDGMENTS
Address for reprint requests and other correspondence:
Y. L. Dorokhov, A. N. Belozersky Institute of Physico-
Chemical Biology, Lomonosov Moscow State University,
Leninsky Gory 1, Laboratory Building A-40, Moscow
119991, Russia (e-mail: dorokhov@genebee.msu.su).
GRANTS
This work was supported by Russian Science Foundation
Grants 11-04-01152, 12-04-33016, and 14-04-00109; the
Russian Foundation for Basic Research; and a stipend from
the President of the Russian Federation for young scientists.
The funders had no role in data collection and analysis,
decision to publish, or preparation of the manuscript.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the authors.
REFERENCES
1. Abramson S, Singh AK. Treatment of the alcohol intoxications: ethylene glycol, meth-
anol and isopropanol. Curr Opin Nephrol Hypertens 9: 695–701, 2000.
METABOLIC METHANOL
631Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
2. Adas F, Berthou F, Picart D, Lozac’h P, Beaugé F, Amet Y. Involvement of cytochrome
P450 2E1 in the (omega-1)-hydroxylation of oleic acid in human and rat liver micro-
somes. J Lipid Res 39: 1210–1219, 1998.
3. Alnouti Y, Klaassen CD. Tissue distribution, ontogeny, and regulation of aldehyde
dehydrogenase (Aldh) enzymes mRNA by prototypical microsomal enzyme inducers
in mice. Toxicol Sci Off J Soc Toxicol 101: 51–64, 2008.
4. Amet Y, Berthou F, Goasduff T, Salaun JP, Le Breton L, Menez JF. Evidence that
cytochrome P450 2E1 is involved in the (omega-1)-hydroxylation of lauric acid in rat
liver microsomes. Biochem Biophys Res Commun 203: 1168–1174, 1994.
5. Anthon GE, Barrett DM. Characterization of the temperature activation of pectin
methylesterase in green beans and tomatoes. J Agric Food Chem 54: 204–211, 2006.
6. Anzenbacher P, Anzenbacherová E. Cytochromes P450 and metabolism of xenobi-
otics. Cell Mol Life Sci CMLS 58: 737–747, 2001.
7. Aragon CM, Amit Z. The effect of 3-amino-1,2,4-triazole on voluntary ethanol con-
sumption: evidence for brain catalase involvement in the mechanism of action. Neu-
ropharmacology 31: 709–712, 1992.
8. Aragon CM, Rogan F, Amit Z. Ethanol metabolism in rat brain homogenates by a
catalase-H
2
O
2
system. Biochem Pharmacol 44: 93–98, 1992.
9. Arcella D, Le Donne C, Piccinelli R, Leclercq C. Dietary estimated intake of intense
sweeteners by Italian teenagers. Present levels and projections derived from the
INRAN-RM-2001 food survey. Food Chem Toxicol Int J Publ Br Ind Biol Res Assoc 42:
677–685, 2004.
10. Arfman N, Watling EM, Clement W, Van Oosterwijk RJ, De Vries GE, Harder W,
Attwood MM, Dijkhuizen L. Methanol metabolism in thermotolerant methylotrophic
Bacillus strains involving a novel catabolic NAD-dependent methanol dehydrogenase
as a key enzyme. Arch Microbiol 152: 280–288, 1989.
11. Argenta LC, Fan X, Mattheis JP. Responses of “Fuji” apples to short and long duration
exposure to elevated CO
2
concentration. Postharvest Biol Technol 24: 13–24, 2002.
12. Argenta LC, Mattheis JP, Fan X, Finger FL. Production of volatile compounds by Fuji
apples following exposure to high CO
2
or low O
2
.J Agric Food Chem 52: 5957–5963,
2004.
13. Arici S, Karaman S, Dogru S, Cayli S, Arici A, Suren M, Karaman T, Kaya Z. Central
nervous system toxicity after acute oral formaldehyde exposure in rabbits: an exper-
imental study. Hum Exp Toxicol. In press, doi: 10.1177/0960327113514098.
14. Aubert J, Begriche K, Knockaert L, Robin MA, Fromenty B. Increased expression of
cytochrome P450 2E1 in nonalcoholic fatty liver disease: mechanisms and pathophys-
iological role. Clin Res Hepatol Gastroenterol 35: 630–637, 2011.
15. Avadhani NG, Sangar MC, Bansal S, Bajpai P. Bimodal targeting of cytochrome P450s
to endoplasmic reticulum and mitochondria: the concept of chimeric signals. FEBS J
278: 4218–4229, 2011.
16. Axelrod J, Daly J. Pituitary gland: enzymic formation of methanol from S-adenosylme-
thionine. Science 150: 892–893, 1965.
17. Badger TM, Huang J, Ronis M, Lumpkin CK. Induction of cytochrome P450 2E1 during
chronic ethanol exposure occurs via transcription of the CYP 2E1 gene when blood
alcohol concentrations are high. Biochem Biophys Res Commun 190: 780–785, 1993.
18. Bai J, Cederbaum AI. Overexpression of CYP2E1 in mitochondria sensitizes HepG2
cells to the toxicity caused by depletion of glutathione. J Biol Chem 281: 5128–5136,
2006.
19. Bai J, Cederbaum AI. Adenovirus-mediated expression of CYP2E1 produces liver
toxicity in mice. Toxicol Sci Off J Soc Toxicol 91: 365–371, 2006.
20. Baker GB, Sowa B, Todd KG. Amine oxidases and their inhibitors: what can they tell
us about neuroprotection and the development of drugs for neuropsychiatric disor-
ders? J Psychiatry Neurosci 32: 313–315, 2007.
21. Baker SC, Ferguson SJ, Ludwig B, Page MD, Richter OMH, van Spanning RJ. Molecular
genetics of the genus Paracoccus: metabolically versatile bacteria with bioenergetic
flexibility. Microbiol Mol Biol Rev 62: 1046–1078, 1998.
22. Baldwin IT, Halitschke R, Kessler A, Schittko U. Merging molecular and ecological
approaches in plant-insect interactions. Curr Opin Plant Biol 4: 351–358, 2001.
23. Barceloux DG, Bond GR, Krenzelok EP, Cooper H, Vale JA. American Academy of
Clinical Toxicology practice guidelines on the treatment of methanol poisoning. J
Toxicol Clin Toxicol 40: 415–446, 2002.
24. Batterman SA, Franzblau A. Time-resolved cutaneous absorption and permeation
rates of methanol in human volunteers. Int Arch Occup Environ Health 70: 341–351,
1997.
25. Batterman SA, Franzblau A, D’Arcy JB, Sargent NE, Gross KB, Schreck RM. Breath,
urine, and blood measurements as biological exposure indices of short-term inhala-
tion exposure to methanol. Int Arch Occup Environ Health 71: 325–335, 1998.
26. Bell LC, Guengerich FP. Oxidation kinetics of ethanol by human cytochrome P450
2E1. Rate-limiting product release accounts for effects of isotopic hydrogen substitu-
tion and cytochrome b5 on steady-state kinetics. J Biol Chem 272: 29643–29651,
1997.
27. Bell-Parikh LC, Guengerich FP. Kinetics of cytochrome P450 2E1-catalyzed oxidation
of ethanol to acetic acid via acetaldehyde. J Biol Chem 274: 23833–23840, 1999.
28. Bendtsen P, Jones AW, Helander A. Urinary excretion of methanol and 5-hy-
droxytryptophol as biochemical markers of recent drinking in the hangover state.
Alcohol Alcohol Oxf Oxfs 33: 431–438, 1998.
29. Bennett IL, Cary FH, Mitchell GL, Cooper MN. Acute methyl alcohol poisoning: a
review based on experiences in an outbreak of 323 cases. Medicine 32: 431–463,
1953.
30. Bergman M, Djaldetti M, Salman H, Bessler H. Effect of citrus pectin on malignant cell
proliferation. Biomed Pharmacother 64: 44–47, 2010.
31. Bethke G, Grundman RE, Sreekanta S, Truman W, Katagiri F, Glazebrook J. Arabidop-
sis pectin methylesterases contribute to immunity against Pseudomonas syringae.Plant
Physiol 164: 1093–1107, 2014.
32. Bianchi F, Careri M, Musci M, Mangia A. Fish and food safety: Determination of
formaldehyde in 12 fish species by SPME extraction and GC-MS analysis. Food Chem
100: 1049–1053, 2007.
33. Bicsak TA, Kann LR, Reiter A, Chase T Jr. Tomato alcohol dehydrogenase: purification
and substrate specificity. Arch Biochem Biophys 216: 605–615, 1982.
34. Bindler F, Voges E, Laugel P. The problem of methanol concentration admissible in
distilled fruit spirits. Food Addit Contam 5: 343–351, 1988.
35. Blakley RL. The Biochemistry of Folic Acid and Related Pteridines. London: North-
Holland, 1969.
36. Blasig IE, Grune T, Schonheit K, Rohde E, Jakstadt M, Haseloff RF, Siems WG. 4-Hy-
droxynonenal, a novel indicator of lipid peroxidation for reperfusion injury of the
myocardium. Am J Physiol Heart Circ Physiol 269: H14–H22, 1995.
37. Boleda MD, Saubi N, Farrés J, Parés X. Physiological substrates for rat alcohol dehy-
drogenase classes: aldehydes of lipid peroxidation, omega-hydroxyfatty acids, and
retinoids. Arch Biochem Biophys 307: 85–90, 1993.
38. Bonhomme-Florentin A. Degradation of hemicellulose and pectin by horse caecum
contents. Br J Nutr 60: 185–192, 1988.
39. Bosron WF, Li TK. Genetic polymorphism of human liver alcohol and aldehyde de-
hydrogenases, and their relationship to alcohol metabolism and alcoholism. Hepatol-
ogy 6: 502–510, 1986.
40. Bosron WF, Li TK, Dafeldecker WP, Vallee BL. Human liver pi-alcohol dehydroge-
nase: kinetic and molecular properties. Biochemistry 18: 1101–1105, 1979.
41. Bosron WF, Magnes LJ, Li TK. Kinetic and electrophoretic properties of native and
recombined isoenzymes of human liver alcohol dehydrogenase. Biochemistry 22:
1852–1857, 1983.
42. Bouchard M, Brunet RC, Droz PO, Carrier G. A biologically based dynamic model for
predicting the disposition of methanol and its metabolites in animals and humans.
Toxicol Sci Off J Soc Toxicol 64: 169–184, 2001.
43. Boyd A, Simon MI. Multiple electrophoretic forms of methyl-accepting chemotaxis
proteins generated by stimulus-elicited methylation in Escherichia coli.J Bacteriol 143:
809–815, 1980.
44. Boyle R. The Sceptical Chymist; or, Chymico-physical Doubts & Paradoxes, Touching the
Spagyrist’s Principles Commonly Call’d Hypostatical, as They Are Wont to Be Propos’d and
DOROKHOV ET AL.
632 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
Defended by the Generality of Alchymists. Whereunto Is Praemis’d Part of Another Dis-
course Relating to the Same Subject. London: F. Cadwell for F. Crooke, 1661.
45. Bradford BU, Seed CB, Handler JA, Forman DT, Thurman RG. Evidence that catalase
is a major pathway of ethanol oxidation in vivo: dose-response studies in deer mice
using methanol as a selective substrate. Arch Biochem Biophys 303: 172–176, 1993.
46. Brent J, McMartin K, Phillips S, Aaron C, Kulig K. Fomepizole for the treatment of
methanol poisoning. Intensive Care Med 344: 424–429, 2001.
47. Brien SE, Ronksley PE, Turner BJ, Mukamal KJ, Ghali WA. Effect of alcohol consump-
tion on biological markers associated with risk of coronary heart disease: systematic
review and meta-analysis of interventional studies. BMJ 342: d636, 2011.
48. Brilli F, Ruuskanen TM, Schnitzhofer R, Müller M, Breitenlechner M, Bittner V, Wohl-
fahrt G, Loreto F, Hansel A. Detection of plant volatiles after leaf wounding and
darkening by proton transfer reaction “time-of-flight” mass spectrometry (PTR-
TOF). PloS One 6: e20419, 2011.
49. Brooks PJ, Zakhari S. Acetaldehyde and the genome: Beyond nuclear DNA adducts
and carcinogenesis: Acetaldehyde and the Genome. Environ Mol Mutagen 55: 77–91,
2014.
50. Brooks RL, Shore JD. Effect of substrate structure on the rate of the catalytic step in
the liver alcohol dehydrogenase mechanism. Biochemistry 10: 3855–3858, 1971.
51. Brouns F, Theuwissen E, Adam A, Bell M, Berger A, Mensink RP. Cholesterol-lower-
ing properties of different pectin types in mildly hyper-cholesterolemic men and
women. Eur J Clin Nutr 66: 591–599, 2012.
52. Brown L, Rosner B, Willett WW, Sacks FM. Cholesterol-lowering effects of dietary
fiber: a meta-analysis. Am J Clin Nutr 69: 30–42, 1999.
53. Budas GR, Disatnik MH, Mochly-Rosen D. Aldehyde dehydrogenase 2 in cardiac
protection: a new therapeutic target? Trends Cardiovasc Med 19: 158–164, 2009.
54. Burd L, Blair J, Dropps K. Prenatal alcohol exposure, blood alcohol concentrations and
alcohol elimination rates for the mother, fetus and newborn. J Perinatol Off J Calif
Perinat Assoc 32: 652–659, 2012.
55. Burnell JC, Carr LG, Dwulet FE, Edenberg HJ, Li TK, Bosron WF. The human beta 3
alcohol dehydrogenase subunit differs from beta 1 by a Cys for Arg-369 substitution
which decreases NAD(H) binding. Biochem Biophys Res Commun 146: 1127–1133,
1987.
56. Byers JA. Attraction of bark beetles, Tomicus piniperda,Hylurgops palliatus, and Try-
podendron domesticum and other insects to short-chain alcohols and monoterpenes. J
Chem Ecol 18: 2385–2402, 1992.
57. Caro AA, Cederbaum AI. Oxidative stress, toxicology, and pharmacology of CYP2E1.
Annu Rev Pharmacol Toxicol 44: 27–42, 2004.
58. Casanova M, Heck HD, Everitt JI, Harrington WW, Popp JA. Formaldehyde concen-
trations in the blood of rhesus monkeys after inhalation exposure. Food Chem Toxicol
Int J Publ Br Ind Biol Res Assoc 26: 715–716, 1988.
59. Casanova-Schmitz M, David RM, Heck HD. Oxidation of formaldehyde and acetal-
dehyde by NAD
-dependent dehydrogenases in rat nasal mucosal homogenates.
Biochem Pharmacol 33: 1137–1142, 1984.
60. Cederbaum AI. CYP2E1–biochemical and toxicological aspects and role in alcohol-
induced liver injury. Mt Sinai J Med NY 73: 657–672, 2006.
61. Cederbaum AI. Role of CYP2E1 in ethanol-induced oxidant stress, fatty liver and
hepatotoxicity. Dig Dis Basel Switz 28: 802–811, 2010.
62. Cederbaum AI. CYP2E1 potentiates toxicity in obesity and after chronic ethanol
treatment. Drug Metabol Drug Interact 27: 125–144, 2012.
63. Cederbaum AI. Alcohol metabolism. Clin Liver Dis 16: 667–685, 2012.
64. Cederbaum AI, Qureshi A. Role of catalase and hydroxyl radicals in the oxidation of
methanol by rat liver microsomes. Biochem Pharmacol 31: 329–335, 1982.
65. Lin-Cereghino GP, Godfrey L, de la Cruz BJ, Johnson S, Khuongsathiene S, Tolstoru-
kov I, Yan M, Lin-Cereghino J, Veenhuis M, Subramani S, Cregg JM. Mxr1p, a key
regulator of the methanol utilization pathway and peroxisomal genes in Pichia pastoris.
Mol Cell Biol 26: 883–897, 2006.
66. Chalasani N, Gorski JC, Asghar MS, Asghar A, Foresman B, Hall SD, Crabb DW.
Hepatic cytochrome P450 2E1 activity in nondiabetic patients with nonalcoholic ste-
atohepatitis. Hepatology 37: 544–550, 2003.
67. Chang C, Meyerowitz EM. Molecular cloning and DNA sequence of the Arabidopsis
thaliana alcohol dehydrogenase gene. Proc Natl Acad Sci USA 83: 1408–1412, 1986.
68. Chang JS, Straif K, Guha N. The role of alcohol dehydrogenase genes in head and neck
cancers: a systematic review and meta-analysis of ADH1B and ADH1C.Mutagenesis
27: 275–286, 2012.
69. Charlton CG, Mack J. Substantia nigra degeneration and tyrosine hydroxylase deple-
tion caused by excess S-adenosylmethionine in the rat brain. Support for an excess
methylation hypothesis for parkinsonism. Mol Neurobiol 9: 149–161, 1994.
70. Chávez-Pacheco JL, Contreras-Zentella M, Membrillo-Hernández J, Arreguín-Espi-
noza R, Mendoza-Hernández G, Gómez-Manzo S, Escamilla JE. The quinohaemopro-
tein alcohol dehydrogenase from Gluconacetobacter xylinus: molecular and catalytic
properties. Arch Microbiol 192: 703–713, 2010.
71. Chen CH, Budas GR, Churchill EN, Disatnik MH, Hurley TD, Mochly-Rosen D.
Activation of aldehyde dehydrogenase-2 reduces ischemic damage to the heart. Sci-
ence 321: 1493–1495, 2008.
72. Chen K, Kazachkov M, Yu PH. Effect of aldehydes derived from oxidative deamination
and oxidative stress on beta-amyloid aggregation: pathological implications to Alzhei-
mer’s disease. J Neural Transm 114: 835–839, 2007.
73. Chen K, Maley J, Yu PH. Potential inplications of endogenous aldehydes in beta-
amyloid misfolding, oligomerization and fibrillogenesis. J Neurochem 99: 1413–1424,
2006.
74. Chen MH, Citovsky V. Systemic movement of a tobamovirus requires host cell pectin
methylesterase. Plant J Cell Mol Biol 35: 386–392, 2003.
75. Chen MH, Sheng J, Hind G, Handa AK, Citovsky V. Interaction between the tobacco
mosaic virus movement protein and host cell pectin methylesterases is required for
viral cell-to-cell movement. EMBO J 19: 913–920, 2000.
76. Chistoserdova L. C1 transfer enzymes and coenzymes linking methylotrophic bacte-
ria and methanogenic Archaea. Science 281: 99–102, 1998.
77. Chiva-Blanch G, Arranz S, Lamuela-Raventos RM, Estruch R. Effects of wine, alcohol
and polyphenols on cardiovascular disease risk factors: evidences from human studies.
Alcohol Alcohol Oxf Oxfs 48: 270–277, 2013.
78. Chuwers P, Osterloh J, Kelly T, D’Alessandro A, Quinlan P, Becker C. Neurobehav-
ioral effects of low-level methanol vapor exposure in healthy human volunteers.
Environ Res 71: 141–150, 1995.
79. Cloos PAC, Christensen J, Agger K, Helin K. Erasing the methyl mark: histone dem-
ethylases at the center of cellular differentiation and disease. Genes Dev 22: 1115–
1140, 2008.
80. Cohen E, Shalom Y, Rosenberger I. Postharvest ethanol buildup and off-flavor in
“Murcott” tangerine fruits. JAm Soc Hortic Sci 115: 775–778, 1990.
81. Cohen G, Sinet PM, Heikkila R. Ethanol oxidation by rat brain in vivo. Alcohol Clin Exp
Res 4: 366–370, 1980.
82. Colby J, Zatman LJ. Enzymological aspects of the pathways for trimethylamine oxida-
tion and C1 assimilation of obligate methylotrophs and restricted facultative methyl-
otrophs. Biochem J 148: 513–520, 1975.
83. Colonna-Cesari F, Perahia D, Karplus M, Eklund H, Brädén CI, Tapia O. Interdomain
motion in liver alcohol dehydrogenase. Structural and energetic analysis of the hinge
bending mode. J Biol Chem 261: 15273–15280, 1986.
84. Cook MR, Bergman FJ, Cohen HD, Gerkovich MM, Graham C, Harris RK, Siemann
LG. Effects of methanol vapor on human neurobehavioral measures. Res Rep Health
Eff Inst 1–45, 1991.
85. Cook RJ, Champion KM, Giometti CS. Methanol toxicity and formate oxidation in
NEUT2 mice. Arch Biochem Biophys 393: 192–198, 2001.
86. Coon MJ. Cytochrome P450: nature’s most versatile biological catalyst. Annu Rev
Pharmacol Toxicol 45: 1–25, 2005.
87. Coon MJ, Koop DR. Alcohol-inducible cytochrome P-450 (P-450ALC). Arch Toxicol
60: 16–21, 1987.
METABOLIC METHANOL
633Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
88. Cordis GA, Maulik N, Bagchi D, Engelman RM, Das DK. Estimation of the extent of
lipid peroxidation in the ischemic and reperfused heart by monitoring lipid metabolic
products with the aid of high-performance liquid chromatography. J Chromatogr 632,
1993.
89. Correa M, Salamone JD, Segovia KN, Pardo M, Longoni R, Spina L, Peana AT, Vinci S,
Acquas E. Piecing together the puzzle of acetaldehyde as a neuroactive agent. Neu-
rosci Biobehav Rev 36: 404–430, 2012.
90. Cosgrove DJ. Enzymes and other agents that enhance cell wall extensibility. Annu Rev
Plant Physiol Plant Mol Biol 50: 391–417, 1999.
91. Cox RB, Quayle JR. The autotrophic growth of Micrococcus denitrificans on methanol.
Biochem J 150: 569, 1975.
92. Crabb DW, Bosron WF, Li TK. Ethanol metabolism. Pharmacol Ther 34: 59 –73, 1987.
93. Csatlós É, Rigó J, Laky M, Brubel R, Joó GJ. The role of the alcohol dehydrogenase-1
(ADH1) gene in the pathomechanism of uterine leiomyoma. Eur J Obstet Gynecol
Reprod Biol 170: 492–496, 2013.
95. Davoli E, Cappellini L, Airoldi L, Fanelli R. Serum methanol concentrations in rats and
in men after a single dose of aspartame. Food Chem Toxicol Int J Publ Br Ind Biol Res
Assoc 24: 187–189, 1986.
95a.De Gouw JA, Howard CJ, Custer TG, Baker BM, Fall R. Proton-transfer chemical-
ionization mass spectrometry allows real-time analysis of volatile organic compounds
released from cutting and drying of Crops. Environ Sci Technol 34: 2640–2648, 2000.
95b.De Groot AC, Flyvholm MA, Lensen G, Menné T, Coenraads PJ. Formaldehyde-
releasers: relationship to formaldehyde contact allergy. Contact allergy to formalde-
hyde and inventory of formaldehyde-releasers. Contact Dermatitis 61: 63–85, 2009.
96. Deltour L, Foglio MH, Duester G. Metabolic deficiencies in alcohol dehydrogenase
Adh1,Adh3, and Adh4 null mutant mice. Overlapping roles of Adh1 and Adh4 in
ethanol clearance and metabolism of retinol to retinoic acid. J Biol Chem 274: 16796 –
16801, 1999.
97. Deng X, Deitrich RA. Putative role of brain acetaldehyde in ethanol addiction. Curr
Drug Abuse Rev 1: 3–8, 2008.
98. Dhareshwar SS, Stella VJ. Your prodrug releases formaldehyde: should you be con-
cerned? No! J Pharm Sci 97: 4184–4193, 2008.
99. Dickinson FM, Dalziel K. Substrate specificity and stereospecificity of alcohol dehy-
drogenases. Nature 214: 31–33, 1967.
100. Dickinson FM, Dalziel K. The specificities and configurations of ternary complexes of
yeast and liver alcohol dehydrogenases. Biochem J 104: 165–172, 1967.
101. Dierks A, Fischer K. Feeding responses and food preferences in the tropical, fruit-
feeding butterfly, Bicyclus anynana.J Insect Physiol 54: 1363–1370, 2008.
103. Diliberto EJ, Veiveros OH, Axelrod J. Subcellualr distribution of protein carboxym-
ethylase and its endogenous substrates in the adrenal medulla: possible role in exci-
tation-secretion coupling. Proc Natl Acad Sci USA 73: 4050–4054, 1976.
104. Diliberto EJ, Axelrod J. Characterization and substrate specificity of a protein car-
boxymethylase in the pituitary gland. Proc Natl Acad Sci USA 71: 1701–1704, 1974.
105. Diliberto EJ, Axelrod J. Regional and subcellular distribution of protein carboxymethy-
lase in brain and other tissues. J Neurochem 26: 1159–1165, 1976.
106. Dixit S, Upadhyay SK, Singh H, Sidhu OP, Verma PC. Enhanced methanol production
in plants provides broad spectrum insect resistance. PloS One 8: e79664, 2013.
107. Dolferus R, Jacobs M, Peacock WJ, Dennis ES. Differential interactions of promoter
elements in stress responses of the Arabidopsis Adh gene. Plant Physiol 105: 1075–
1087, 1994.
108. Dominy NJ. Fruits, fingers, and fermentation: the sensory cues available to foraging
primates. Integr Comp Biol 44: 295–303, 2004.
109. Donnez J, Jadoul P. What are the implications of myomas on fertility? A need for a
debate? Hum Reprod 17: 1424–1430, 2002.
110. Dorman DC, Moss OR, Farris GM, Janszen D, Bond JA, Medinsky MA. Pharmacoki-
netics of inhaled [
14
C]methanol and methanol-derived [
14
C]formate in normal and
folate-deficient cynomolgus monkeys. Toxicol Appl Pharmacol 128: 229–238, 1994.
111. Dorokhov YL, Komarova TV, Petrunia IV, Frolova OY, Pozdyshev DV, Gleba YY.
Airborne signals from a wounded leaf facilitate viral spreading and induce antibacterial
resistance in neighboring plants. PLoS Pathog 8: e1002640, 2012.
112. Dorokhov YL, Komarova TV, Petrunia IV, Kosorukov VS, Zinovkin RA, Shindyapina
AV, Frolova OY, Gleba YY. Methanol may function as cross-kingdom signal. PLoS One
7: e36122, 2012.
113. Dorokhov YL, Mäkinen K, Frolova OY, Merits A, Saarinen J, Kalkkinen N, Atabekov
JG, Saarma M. A novel function for a ubiquitous plant enzyme pectin methylesterase:
the host-cell receptor for the tobacco mosaic virus movement protein. FEBS Lett 461:
223–228, 1999.
114. Dorokhov YL, Skurat EV, Frolova OY, Gasanova TV, Ivanov PA, Ravin NV, Skryabin
KG, Mäkinen KM, Klimyuk VI, Gleba YY, Atabekov JG. Role of the leader sequence in
tobacco pectin methylesterase secretion. FEBS Lett 580: 3329–3334, 2006.
115. Downie A, Miyazaki S, Bohnert H, John P, Coleman J, Parry M, Haslam R. Expression
profiling of the response of Arabidopsis thaliana to methanol stimulation. Phytochem-
istry 65: 2305–2316, 2004.
116. Dudek M, Knutelska J, Bednarski M, Nowi´
nski L, Zygmunt M, Bilska-Wilkosz A, Iciek
M, Otto M, Z
˙ytka I, Sapa J, Włodek L, Filipek B. Alpha lipoic acid protects the heart
against myocardial post ischemia-reperfusion arrhythmias via KATP channel activa-
tion in isolated rat hearts. Pharmacol Rep 66: 499–504, 2014.
117. Dudley R. Fermenting fruit and the historical ecology of ethanol ingestion: is alcohol-
ism in modern humans an evolutionary hangover? Addiction 97: 381–388, 2002.
118. Dudley R. Ethanol, fruit ripening, and the historical origins of human alcoholism in
primate frugivory. Integr Comp Biol 44: 315–323, 2004.
119. Dunkel P, Gelain A, Barlocco D, Haider N, Gyires K, Sperlágh B, Magyar K, Maccioni
E, Fadda A, Mátyus P. Semicarbazide-sensitive amine oxidase/vascular adhesion pro-
tein 1: recent developments concerning substrates and inhibitors of a promising
therapeutic target. Curr Med Chem 15: 1827–1839, 2008.
120. Edenberg HJ. Regulation of the mammalian alcohol dehydrogenase genes. Prog Nu-
cleic Acid Res Mol Biol 64: 295–341, 2000.
121. Edenberg HJ. Association of alcohol dehydrogenase genes with alcohol dependence:
a comprehensive analysis. Hum Mol Genet 15: 1539–1549, 2006.
122. Edenberg HJ. The genetics of alcohol metabolism: role of alcohol dehydrogenase and
aldehyde dehydrogenase variants. Alcohol Res Health 30: 5–13, 2007.
123. Edenberg HJ, Jerome RE, Li M. Polymorphism of the human alcohol dehydrogenase 4
(ADH4) promoter affects gene expression. Pharmacogenetics 9: 25–30, 1999.
124. Elleuche S, Antranikian G. Bacterial group III alcohol dehydrogenases-function, evo-
lution and biotechnological applications [Online]. OA Alcohol 1, 2013, http://www.
oapublishinglondon.com/images/article/pdf/1393705703.pdf (29 Jul. 2014).
125. El-Rashidy R, Niazi S. A new metabolite of butylated hydroxyanisole in man. Biopharm
Drug Dispos 4: 389–396, 1983.
126. Eriksen SP, Kulkarni AB. Methanol in normal human breath. Science 141: 639–640,
1963.
127. Essa MM, Vijayan RK, Castellano-Gonzalez G, Memon MA, Braidy N, Guillemin GJ.
Neuroprotective effect of natural products against Alzheimer’s disease. Neurochem
Res 37: 1829–1842, 2012.
128. Eto K, Asada T, Arima K, Makifuchi T, Kimura H. Brain hydrogen sulfide is severely
decreased in Alzheimer’s disease. Biochem Biophys Res Commun 293: 1485–1488,
2002.
129. Ettwig KF, Butler MK, Le Paslier D, Pelletier E, Mangenot S, Kuypers MMM, Schreiber
F, Dutilh BE, Zedelius J, de Beer D, Gloerich J, Wessels HJCT, van Alen T, Luesken F,
Wu ML, van de Pas-Schoonen KT, Op den Camp HJM, Janssen-Megens EM, Francoijs
K-J, Stunnenberg H, Weissenbach J, Jetten MSM, Strous M. Nitrite-driven anaerobic
methane oxidation by oxygenic bacteria. Nature 464: 543–548, 2010.
130. European Parliament Council. Directive 94/35/EC of 30 June 1994 on sweeteners for
use in foodstuffs. Off J 237 10.9.199.: 3–12, 1994.
130a.European Parliament Council. Regulation (EC) No 110/2008 of the European Parlia-
ment and of the Council. Off J Eur Union 51: 16–50, 2008.
DOROKHOV ET AL.
634 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
131. Fall R. Abundant oxygenates in the atmosphere: a biochemical perspective. Chem Rev
103: 4941–4952, 2003.
132. Farias EF, Ong DE, Ghyselinck NB, Nakajo S, Kuppumbatti YS, Mira y Lopez R.
Cellular retinol-binding protein I, a regulator of breast epithelial retinoic acid receptor
activity, cell differentiation, and tumorigenicity. J Natl Cancer Inst 97: 21–29, 2005.
133. Farneti B, Cristescu SM, Costa G, Harren FJM, Woltering EJ. Rapid tomato volatile
profiling by using proton-transfer reaction mass spectrometry (PTR-MS). J Food Sci 77:
C551–559, 2012.
134. Farout L, Mary J, Vinh J, Szweda LI, Friguet B. Inactivation of the proteasome by
4-hydroxy-2-nonenal is site specific and dependant on 20S proteasome subtypes. Arch
Biochem Biophys 453: 135–142, 2006.
135. Farrar C, Clarke S. Altered levels of S-adenosylmethionine and S-adenosylhomocys-
teine in the brains of L-isoaspartyl (D-aspartyl) O-methyltransferase-deficient mice. J
Biol Chem 277: 27856–27863, 2002.
136. FDA. Food additives permitted for direct addition to food for human consumption:
aspartame. Fed Reg (Food and Drug Administration) 61: 33654–33656, 1996.
137. Ferguson CS, Miksys S, Palmour RM, Tyndale RF. Ethanol self-administration and
nicotine treatment induce brain levels of CYP2B6 and CYP2E1 in African green mon-
keys. Neuropharmacology 72: 74–81, 2013.
138. Ferguson CS, Tyndale RF. Cytochrome P450 enzymes in the brain: emerging evi-
dence of biological significance. Trends Pharmacol Sci 32: 708–714, 2011.
139. Fernández MR, Biosca JA, Parés X. S-nitrosoglutathione reductase activity of human
and yeast glutathione-dependent formaldehyde dehydrogenase and its nuclear and
cytoplasmic localisation. Cell Mol Life Sci 60: 1013–1018, 2003.
140. Ferrari S, Savatin DV, Sicilia F, Gramegna G, Cervone F, Lorenzo GD. Oligogalac-
turonides: plant damage-associated molecular patterns and regulators of growth and
development. Front Plant Sci 4: 49, 2013.
141. Freeling M, Bennett DC. Maize Adh1. Annu Rev Genet 19: 297–323, 1985.
142. Friedenson B. A common environmental carcinogen unduly affects carriers of cancer
mutations: carriers of genetic mutations in a specific protective response are more
susceptible to an environmental carcinogen. Med Hypotheses 77: 791–797, 2011.
143. Galter D, Carmine A, Buervenich S, Duester G, Olson L. Distribution of class I, III and
IV alcohol dehydrogenase mRNAs in the adult rat, mouse and human brain. Eur J
Biochem 270: 1316–1326, 2003.
144. Gaziano JM, Gaziano TA, Glynn RJ, Sesso HD, Ajani UA, Stampfer MJ, Manson JE,
Hennekens CH, Buring JE. Light-to-moderate alcohol consumption and mortality in
the Physicians’ Health Study enrollment cohort. J Am Coll Cardiol 35: 96–105, 2000.
145. Gill K, Menez JF, Lucas D, Deitrich RA. Enzymatic production of acetaldehyde from
ethanol in rat brain tissue. Alcohol Clin Exp Res 16: 910–915, 1992.
146. Glinsky VV, Raz A. Modified citrus pectin anti-metastatic properties: one bullet, mul-
tiple targets. Carbohydr Res 344: 1788–1791, 2009.
147. Glymour MM. Alcohol and cardiovascular disease. BMJ 349: g4334, 2014.
148. Gnekow B, Ough CS. Methanol in wines and musts source and amounts. Am J Enol
Vitic 27: 1–6, 1976.
149. Goeppert A, Czaun M, Jones JP, Surya Prakash GK, Olah GA. Recycling of carbon
dioxide to methanol and derived products: closing the loop. Chem Soc Rev. In press,
doi: 10.1039/c4cs00122b.
150. Gonzalez FJ. The 2006 Bernard B Brodie Award Lecture. Cyp2e1 Drug Metab Dispos
Biol Fate Chem 35: 1–8, 2007.
151. Gout E, Aubert S, Bligny R, Rébeillé F, Nonomura AR, Benson AA, Douce R. Metab-
olism of methanol in plant cells. Carbon-13 nuclear magnetic resonance studies. Plant
Physiol 123: 287–296, 2000.
153. Greizerstein HB. Congener contents of alcoholic beverages. J Stud Alcohol 42: 1030 –
1037, 1981.
155. Gubisne-Haberle D, Hill W, Kazachkov M, Richardson JS, Yu PH. Protein cross-
linkage induced by formaldehyde derived from semicarbazide-sensitive amine oxi-
dase-mediated deamination of methylamine. J Pharmacol Exp Ther 310: 1125–1132,
2004.
156. Guenther A, Hewitt CN, Erickson D, Fall R, Geron C, Graedel T, Harley P, Klinger L,
Lerdau M, Mckay WA, Pierce T, Scholes B, Steinbrecher R, Tallamraju R, Taylor J,
Zimmerman P. A global model of natural volatile organic compound emissions. J
Geophys Res Atmospheres 100: 8873–8892, 1995.
157. Guerin MR, Higgins CE, Griest WH. The analysis of the particulate and vapour phases
of tobacco smoke. IARC Sci Publ 115–139, 1987.
158. Guindalini C, Scivoletto S, Ferreira RGM, Breen G, Zilberman M, Peluso MA, Zatz M.
Association of genetic variants in alcohol dehydrogenase 4 with alcohol dependence in
Brazilian patients. Am J Psychiatry 162: 1005–1007, 2005.
159. Gulati A, Nath C, Shanker K, Srimal RC, Dhawan KN, Bhargava KP. Effect of alcohols
on the permeability of blood-brain barrier. Pharmacol Res Commun 17: 85–93, 1985.
160. Gunness P, Gidley MJ. Mechanisms underlying the cholesterol-lowering properties of
soluble dietary fibre polysaccharides. Food Funct 1: 149–155, 2010.
161. Guo JM, Liu AJ, Zang P, Dong WZ, Ying L, Wang W, Xu P, Song XR, Cai J, Zhang SQ,
Duan JL, Mehta JL, Su DF. ALDH2 protects against stroke by clearing 4-HNE. Cell Res
23: 915–930, 2013.
162. Guo L, Zeng XY, Wang DY, Li GQ. Methanol metabolism in the Asian corn borer,
Ostrinia furnacalis (Guenée) (Lepidoptera: Pyralidae). J Insect Physiol 56: 260–265,
2010.
163. Hanson RS, Hanson TE. Methanotrophic bacteria. Microbiol Rev 60: 439–471, 1996.
164. Hansson T, Tindberg N, Ingelman-Sundberg M, Köhler C. Regional distribution of
ethanol-inducible cytochrome P450 IIE1 in the rat central nervous system. Neurosci-
ence 34: 451–463, 1990.
165. Harholt J, Suttangkakul A, Vibe Scheller H. Biosynthesis of pectin. Plant Physiol 153:
384–395, 2010.
166. Harrington AA, Kallio RE. Oxidation of methanol and formaldehyde by Pseudomonas
methanica.Can J Microbiol 6, 1960.
167. Harris C, Dixon M, Hansen JM. Glutathione depletion modulates methanol, formal-
dehyde and formate toxicity in cultured rat conceptuses. Cell Biol Toxicol 20: 133–145,
2004.
168. Harris C, Wang SW, Lauchu JJ, Hansen JM. Methanol metabolism and embryotoxicity
in rat and mouse conceptuses: comparisons of alcohol dehydrogenase (ADH1), form-
aldehyde dehydrogenase (ADH3), and catalase. Reprod Toxicol 17: 349–357, 2003.
169. Haseba T, Ohno Y. A new view of alcohol metabolism and alcoholism–role of the
high-K
m
Class III alcohol dehydrogenase (ADH3). Int J Environ Res Public Health 7:
1076–1092, 2010.
170. Haseba T, Sugimoto J, Sato S, Abe Y, Ohno Y. Phytophenols in whisky lower blood
acetaldehyde level by depressing alcohol metabolism through inhibition of alcohol
dehydrogenase 1 (class I) in mice. Metabolism 57: 1753–1759, 2008.
171. Hayashi T, Reece CA, Shibamoto T. Gas chromatographic determination of formal-
dehyde in coffee via thiazolidine derivative. J Assoc Anal Chem 69: 101–105, 1986.
171a.He L, Liu B, Dai Z, Zhang HF, Zhang YS, Luo XJ, Ma QL, Peng J. Alpha lipoic acid
protects heart against myocardial ischemia-reperfusion injury through a mechanism
involving aldehyde dehydrogenase 2 activation. Eur J Pharmacol 678: 32–38, 2012.
171b.He R, Lu J, Miao J. Formaldehyde stress. Sci China Life Sci 53: 1399–1404, 2010.
172. Heath AC, Bucholz KK, Madden PAF, Dinwiddie SH, Slutske WS, Bierut LJ, Statham
DJ, Dunne MP, Whitfield JB, Martin NG. Genetic and environmental contributions to
alcohol dependence risk in a national twin sample: consistency of findings in women
and men. Psychol Med 27: 1381–1396, 1997.
173. Heath AC, Martin NG. Genetic influences on alcohol consumption patterns and
problem drinking: results from the Australian NH&MRC twin panel follow-up survey.
Ann NY Acad Sci 708: 72–85, 1994.
174. Heck HD, Casanova M. The implausibility of leukemia induction by formaldehyde: a
critical review of the biological evidence on distant-site toxicity. Regul Toxicol Pharma-
col 40: 92–106, 2004.
175. Heck HD, Casanova M, Starr TB. Formaldehyde toxicity–new understanding. Crit Rev
Toxicol 20: 397–426, 1990.
METABOLIC METHANOL
635Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
176. Heck HD, Casanova-Schmitz M, Dodd PB, Schachter EN, Witek TJ, Tosun T. Form-
aldehyde (CH
2
O) concentrations in the blood of humans and Fischer-344 rats ex-
posed to CH
2
O under controlled conditions. Am Ind Hyg Assoc J 46: 1–3, 1985.
177. Heikes BG, Chang W, Pilson MEQ, Swift E, Singh HB, Guenther A, Jacob DJ, Field BD,
Fall R, Riemer D, Brand L. Atmospheric methanol budget and ocean implication. Glob
Biogeochem Cycles 16: 1133, 2002.
178. Heit C, Dong H, Chen Y, Thompson DC, Deitrich RA, Vasiliou VK. The role of
CYP2E1 in alcohol metabolism and sensitivity in the central nervous system. Subcell
Biochem 67: 235–247, 2013.
181. Heyn H, Li N, Ferreira HJ, Moran S, Pisano DG, Gomez A, Diez J, Sanchez-Mut JV,
Setien F, Carmona FJ, Puca AA, Sayols S, Pujana MA, Serra-Musach J, Iglesias-Platas I,
Formiga F, Fernandez AF, Fraga MF, Heath SC, Valencia A, Gut IG, Wang J, Esteller M.
Distinct DNA methylomes of newborns and centenarians. Proc Natl Acad Sci USA 109:
10522–10527, 2012.
182. Higuchi S, Matsushita S, Masaki T, Yokoyama A, Kimura M, Suzuki G, Mochizuki H.
Influence of genetic variations of ethanol-metabolizing enzymes on phenotypes of
alcohol-related disorders. Ann NY Acad Sci 1025: 472–480, 2004.
183. Hlavica P. Models and mechanisms of O-O bond activation by cytochrome P450. A
critical assessment of the potential role of multiple active intermediates in oxidative
catalysis. Eur J Biochem 271: 4335–4360, 2004.
184. Holmes MV, Dale CE, Zuccolo L, Silverwood RJ, Guo Y, Ye Z, Prieto-Merino D,
Dehghan A, Trompet S, Wong A, Cavadino A, Drogan D, Padmanabhan S, Li S,
Yesupriya A, Leusink M, Sundstrom J, Hubacek JA, Pikhart H, Swerdlow DI, Panay-
iotou AG, Borinskaya SA, Finan C, Shah S, Kuchenbaecker KB, Shah T, Engmann J,
Folkersen L, Eriksson P, Ricceri F, Melander O, Sacerdote C, Gamble DM, Rayaprolu
S, Ross OA, McLachlan S, Vikhireva O, Sluijs I, Scott RA, Adamkova V, Flicker L,
Bockxmeer van FM, Power C, Marques-Vidal P, Meade T, Marmot MG, Ferro JM,
Paulos-Pinheiro S, Humphries SE, Talmud PJ, Mateo Leach I, Verweij N, Linneberg A,
Skaaby T, Doevendans PA, Cramer MJ, van der Harst P, Klungel OH, Dowling NF,
Dominiczak AF, Kumari M, Nicolaides AN, Weikert C, Boeing H, Ebrahim S, Gaunt
TR, Price JF, Lannfelt L, Peasey A, Kubinova R, Pajak A, Malyutina S, Voevoda MI,
Tamosiunas A, Maitland-van der Zee AH, Norman PE, Hankey GJ, Bergmann MM,
Hofman A, Franco OH, Cooper J, Palmen J, Spiering W, de Jong PA, Kuh D, Hardy R,
Uitterlinden AG, Ikram MA, Ford I, Hyppönen E, Almeida OP, Wareham NJ, Khaw
KT, Hamsten A, Husemoen LLN, Tjønneland A, Tolstrup JS, Rimm E, Beulens JWJ,
Verschuren WMM, Onland-Moret NC, Hofker MH, Wannamethee SG, Whincup PH,
Morris R, Vicente AM, Watkins H, Farrall M, Jukema JW, Meschia J, Cupples LA, Sharp
SJ, Fornage M, Kooperberg C, LaCroix AZ, Dai JY, Lanktree MB, Siscovick DS,
Jorgenson E, Spring B, Coresh J, Li YR, Buxbaum SG, Schreiner PJ, Ellison RC, Tsai MY,
Patel SR, Redline S, Johnson AD, Hoogeveen RC, Hakonarson H, Rotter JI, Boer-
winkle E, de Bakker PIW, Kivimaki M, Asselbergs FW, Sattar N, Lawlor DA, Whittaker
J, Davey Smith G, Mukamal K, Psaty BM, Wilson JG, Lange LA, Hamidovic A, Hingo-
rani AD, Nordestgaard BG, Bobak M, Leon DA, Langenberg C, Palmer TM, Reiner
AP, Keating BJ, Dudbridge F, Casas JP, InterAct Consortium. Association between
alcohol and cardiovascular disease: Mendelian randomisation analysis based on indi-
vidual participant data. BMJ 349: g4164, 2014.
185. Höög JO, Hedén LO, Larsson K, Jörnvall H, von Bahr-Lindström H. The gamma 1 and
gamma 2 subunits of human liver alcohol dehydrogenase cDNA structures, two amino
acid replacements, and compatibility with changes in the enzymatic properties. Eur J
Biochem 159: 215–218, 1986.
186. Höög JO, Ostberg LJ. Mammalian alcohol dehydrogenases–a comparative investiga-
tion at gene and protein levels. Chem Biol Interact 191: 2–7, 2011.
187. Hoondal GS, Tiwari RP, Tewari R, Dahiya N, Beg QK. Microbial alkaline pectinases
and their industrial applications: a review. Appl Microbiol Biotechnol 59: 409–418,
2002.
188. Horton VL, Higuchi MA, Rickert DE. Physiologically based pharmacokinetic model for
methanol in rats, monkeys, and humans. Toxicol Appl Pharmacol 117: 26–36, 1992.
189. Hou H, Yu H. Structural insights into histone lysine demethylation. Curr Opin Struct
Biol 20: 739–748, 2010.
190. Hovda KE, Urdal P, Jacobsen D. Increased serum formate in the diagnosis of methanol
poisoning. J Anal Toxicol 29: 586–588, 2005.
191. Hua Q, He RQ. Effect of phosphorylation and aggregation on tau binding to DNA.
Protein Pept Lett 9: 349–357, 2002.
192. Hurley TD, Edenberg HJ. Genes encoding enzymes involved in ethanol metabolism.
Alcohol Res Curr Rev 34: 339–344, 2012.
193. Huve K, Christ M, Kleist E, Uerlings R, Niinemets U, Walter A, Wildt J. Simultaneous
growth and emission measurements demonstrate an interactive control of methanol
release by leaf expansion and stomata. J Exp Bot 58: 1783–1793, 2007.
194. Hyun TK, Eom SH, Yu CY, Roitsch T. Hovenia dulcis–an Asian traditional herb. Planta
Med 76: 943–949, 2010.
195. Iborra FJ, Renau-Piqueras J, Portoles M, Boleda MD, Guerri C, Pares X. Immunocy-
tochemical and biochemical demonstration of formaldhyde dehydrogenase (class III
alcohol dehydrogenase) in the nucleus. J Histochem Cytochem 40: 1865–1878, 1992.
196. Igarashi K, Katoh Y. Metabolic aspects of epigenome: coupling of S-adenosylmethio-
nine synthesis and gene regulation on chromatin by SAMIT module. Subcell Biochem
61: 105–118, 2013.
197. Isse T, Matsuno K, Oyama T, Kitagawa K, Kawamoto T. Aldehyde dehydrogenase 2
gene targeting mouse lacking enzyme activity shows high acetaldehyde level in blood,
brain, and liver after ethanol gavages. Alcohol Clin Exp Res 29: 1959–1964, 2005.
198. Jackson CL, Dreaden TM, Theobald LK, Tran NM, Beal TL, Eid M, Gao MY, Shirley
RB, Stoffel MT, Kumar MV, Mohnen D. Pectin induces apoptosis in human prostate
cancer cells: correlation of apoptotic function with pectin structure. Glycobiology 17:
805–819, 2007.
199. Jacob DJ, Field BD, Li Q, Blake DR, de Gouw J, Warneke C, Hansel A, Wisthaler A,
Singh HB, Guenther A. Global budget of methanol: constraints from atmospheric
observations. J Geophys Res Atmospheres 110: D08303, 2005.
200. Jacobsen D, Sebastian CS, Dies DF, Breau RL, Spann EG, Barron SK, McMartin KE.
Kinetic interactions between 4-methylpyrazole and ethanol in healthy humans. Alcohol
Clin Exp Res 20: 804–809, 1996.
201. Jalkanen S, Salmi M. Cell surface monoamine oxidases: enzymes in search of a func-
tion. EMBO J20: 3893–3901, 2001.
202. Jalkanen S, Salmi M. VAP-1 and CD73, endothelial cell surface enzymes in leukocyte
extravasation. Arterioscler Thromb Vasc Biol 28: 18–26, 2008.
203. Jamal M, Ameno K, Uekita I, Kumihashi M, Wang W, Ijiri I. Catalase mediates acetal-
dehyde formation in the striatum of free-moving rats. Neurotoxicology 28: 1245–1248,
2007.
204. Jelski W, Szmitkowski M. Alcohol dehydrogenase (ADH) and aldehyde dehydroge-
nase (ALDH) in the cancer diseases. Clin Chim Acta 395: 1–5, 2008.
205. Jelski W, Zalewski B, Chrostek L, Szmitkowski M. The activity of class I, II, III, and IV
alcohol dehydrogenase isoenzymes and aldehyde dehydrogenase in colorectal cancer.
Dig Dis Sci 49: 977–981, 2004.
206. Jelski W, Zalewski B, Chrostek L, Szmitkowski M. Alcohol dehydrogenase (ADH)
isoenzymes and aldehyde dehydrogenase (ALDH) activity in the sera of patients with
colorectal cancer. Clin Exp Med 7: 154–157, 2007.
207. Jensen EN, Buch-Andersen T, Ravn-Haren G, Dragsted LO, Jensen EN, Buch-Ander-
sen T, Ravn-Haren G, Dragsted LO. Mini-review: the effects of apples on plasma
cholesterol levels and cardiovascular risk–a review of the evidence. J Hortic Sci Bio-
technol 34–41, 2009.
208. Jensen NS, Canale-Parola E. Nutritionally limited pectinolytic bacteria from the hu-
man intestine. Appl Environ Microbiol 50: 172–173, 1985.
209. Jensen NS, Canale-Parola E. Bacteroides pectinophilus sp. nov and Bacteroides galac-
turonicus sp nov: two pectinolytic bacteria from the human intestinal tract. Appl Envi-
ron Microbiol 52: 880–887, 1986.
210. Jiang ZJ, Richardson JS, Yu PH. The contribution of cerebral vascular semicarbazide-
sensitive amine oxidase to cerebral amyloid angiopathy in Alzheimer’s disease. Neu-
ropathol Appl Neurobiol 34: 194–204, 2008.
211. Jokelainen K, Matysiak-Budnik T, Mäkisalo H, Höckerstedt K, Salaspuro M. High
intracolonic acetaldehyde values produced by a bacteriocolonic pathway for ethanol
oxidation in piglets. Gut 39: 100–104, 1996.
212. Jones AW. Abnormally high concentrations of methanol in breath: a useful biochem-
ical marker of recent heavy drinking. Clin Chem 32: 1241–1242, 1986.
DOROKHOV ET AL.
636 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
213. Jones AW. Elimination half-life of methanol during hangover. Pharmacol Toxicol 60:
217–220, 1987.
214. Jones GR, Singer PP, Rittenbach K. The relationship of methanol and formate concen-
trations in fatalities where methanol is detected. J Forensic Sci 52: 1376–1382, 2007.
215. Jörnvall H, Hempel J, Vallee BL, Bosron WF, Li TK. Human liver alcohol dehydroge-
nase: amino acid substitution in the beta 2 beta 2 Oriental isozyme explains functional
properties, establishes an active site structure, and parallels mutational exchanges in
the yeast enzyme. Proc Natl Acad Sci USA 81: 3024–3028, 1984.
216. Joshi M, Tyndale RF. Regional and cellular distribution of CYP2E1 in monkey brain and
its induction by chronic nicotine. Neuropharmacology 50: 568–575, 2006.
217. Julia` P, Farrés J, Parés X. Characterization of three isoenzymes of rat alcohol dehy-
drogenase. Tissue distribution and physical and enzymatic properties. Eur J Biochem
162: 179–189, 1987.
218. Just W, Zeller J, Riegert C, Speit G. Genetic polymorphisms in the formaldehyde
dehydrogenase gene and their biological significance. Toxicol Lett 207: 121–127, 2011.
219. Kalász H, Magyar K, Szo˝ke É, Adeghate E, Adem A, Hasan MY, Nurulain SM, Tekes K.
Metabolism of selegiline [()-deprenyl]. Curr Med Chem 21: 1522–1530, 2014.
220. Kalhan SC. One-carbon metabolism, fetal growth and long-term consequences.
Nestlé Nutr Inst Workshop Ser 74: 127–138, 2013.
221. Kamino K, Nagasaka K, Imagawa M, Yamamoto H, Yoneda H, Ueki A, Kitamura S,
Namekata K, Miki T, Ohta S. Deficiency in mitochondrial aldehyde dehydrogenase
increases the risk for late-onset Alzheimer’s disease in the Japanese population.
Biochem Biophys Res Commun 273: 192–196, 2000.
222. Kaminski J, Atwal AS, Mahadevan S. Determination of formaldehyde in fresh and retail
milk by liquid column chromatography. J AOAC Int 76: 1010–1013, 1993.
223. Kamlay MT, Shore JD. Transient kinetic studies of substrate inhibition in the horse
liver alcohol dehydrogenase reaction. Arch Biochem Biophys 222: 59–66, 1983.
224. Kang G, Bae KY, Kim SW, Kim J, Shin HY, Kim JM, Shin IS, Yoon JS, Kim JK. Effect of
the allelic variant of alcohol dehydrogenase ADH1B*2 on ethanol metabolism. Alcohol
Clin Exp Res 38: 1502–1509, 2014.
225. Kano M, Ishikawa F, Matsubara S, Kikuchi-Hayakawa H, Shimakawa Y. Soymilk prod-
ucts affect ethanol absorption and metabolism in rats during acute and chronic ethanol
intake. J Nutr 132: 238–244, 2002.
226. Kaphalia L, Boroumand N, Hyunsu J, Kaphalia BS, Calhoun WJ. Ethanol metabolism,
oxidative stress, and endoplasmic reticulum stress responses in the lungs of hepatic
alcohol dehydrogenase deficient deer mice after chronic ethanol feeding. Toxicol Appl
Pharmacol 277: 109–117, 2014.
227. Kaphalia L, Calhoun WJ. Alcoholic lung injury: metabolic, biochemical and immuno-
logical aspects. Toxicol Lett 222: 171–179, 2013.
228. Karahanian E, Quintanilla ME, Tampier L, Rivera-Meza M, Bustamante D, Gonzalez-
Lira V, Morales P, Herrera-Marschitz M, Israel Y. Ethanol as a prodrug: brain metab-
olism of ethanol mediates its reinforcing effects. Alcohol Clin Exp Res 35: 606–612,
2011.
229. Karinje KU, Ogata M. Methanol metabolism in acatalasemic mice. Physiol Chem Phys
Med NMR 22: 193–198, 1990.
230. Karl T, Guenther A, Lindinger C, Jordan A, Fall R, Lindinger W. Eddy covariance
measurements of oxygenated volatile organic compound fluxes from crop harvesting
using a redesigned proton-transfer-reaction mass spectrometer. J Geophys Res Atmo-
spheres 106: 24157–24167, 2001.
231. Kavet R, Nauss KM. The toxicity of inhaled methanol vapors. Crit Rev Toxicol 21:
21–50, 1990.
232. Kay RM. Dietary fiber. J Lipid Res 23: 221–242, 1982.
233. Kazachkov M, Chen K, Babiy S, Yu PH. Evidence for in vivo scavenging by aminogua-
nidine of formaldehyde produced via semicarbazide-sensitive amine oxidase-medi-
ated deamination. J Pharmacol Exp Ther 322: 1201–1207, 2007.
234. Keilin D, Hartree EF. Properties of catalysis of coupled oxidation of alcohols. Biochem
J39: 293–301, 1945.
235. Keung WM, Vallee BL. Daidzin: a potent, selective inhibitor of human mitochondrial
aldehyde dehydrogenase. Proc Natl Acad Sci USA 90: 1247–1251, 1993.
236. Khlifi R, Messaoud O, Rebai A, Hamza-Chaffai A. Polymorphisms in the human cyto-
chrome P450 and arylamine N-acetyltransferase: susceptibility to head and neck can-
cers. BioMed Res Int 2013: 582768, 2013.
237. Kimmerer TW, MacDonald RC. Acetaldehyde and ethanol biosynthesis in leaves of
plants. Plant Physiol 84: 1204–1209, 1987.
238. Kim SH, Cho SK, Min TS, Kim Y, Yang SO, Kim HS, Hyun SH, Kim H, Kim YS, Choi
HK. Ameliorating effects of mango (Mangifera indica L.) fruit on plasma ethanol level in
a mouse model assessed with H-NMR based metabolic profiling. J Clin Biochem Nutr
48: 214–221, 2011.
239. Kim SY, Millet DB, Hu L, Mohr MJ, Griffis TJ, Wen D, Lin JC, Miller SM, Longo M.
Constraints on carbon monoxide emissions based on tall tower measurements in the
US Upper Midwest. Environ Sci Technol 47: 8316–8324, 2013.
241. Kloner RA, Rezkalla SH. To drink or not to drink? That is the question. Circulation 116:
1306–1317, 2007.
242. Knockaert L, Fromenty B, Robin MA. Mechanisms of mitochondrial targeting of cy-
tochrome P450 2E1: physiopathological role in liver injury and obesity. FEBS J 278:
4252–4260, 2011.
243. Koivusalo M, Baumann M, Uotila L. Evidence for the identity of glutathione-depen-
dent formaldehyde dehydrogenase and class III alcohol dehydrogenase. FEBS Lett 257:
105–109, 1989.
244. Kolb S. Aerobic methanol-oxidizing bacteria in soil. FEMS Microbiol Lett 300: 1–10,
2009.
245. Komarova TV, Citovsky V, Dorokhov YL. Pectin methylesterase enhances tomato
bushy stunt virus P19 RNA silencing suppressor effects. In: RNAi Technology. Boca
Raton, FL: CRC, 2014, p. 125–134, http://www.crcnetbase.com/doi/abs/10.1201/
b10882-8.
246. Komarova TV, Petrunia IV, Shindyapina AV, Silachev DN, Sheshukova EV, Kiryanov
GI, Dorokhov YL. Endogenous methanol regulates mammalian gene activity. PLoS One
9: e90239, 2014.
247. Komarova TV, Pozdyshev DV, Petrunia IV, Sheshukova EV, Dorokhov YL. Pectin
methylesterase-generated methanol may be involved in tobacco leaf growth. Bio-
chemistry 79: 102–110, 2014.
248. Komarova TV, Sheshukova EV, Dorokhov YL. Cell wall methanol as a signal in plant
immunity. Front Plant Sci 5: 101, 2014.
249. Konstandi M, Johnson EO, Lang MA. Consequences of psychophysiological stress on
cytochrome P450-catalyzed drug metabolism. Neurosci Biobehav Rev 45C: 149–167,
2014.
250. Körner E, von Dahl CC, Bonaventure G, Baldwin IT. Pectin methylesterase NaPME1
contributes to the emission of methanol during insect herbivory and to the elicitation
of defence responses in Nicotiana attenuata.J Exp Bot 60: 2631–2640, 2009.
251. Kostic MA, Dart RC. Rethinking the toxic methanol level. Clin Toxicol 41: 793–800,
2003.
252. Krau SD. Cytochrome P450 part 3: drug interactions: essential concepts and consid-
erations. Nurs Clin North Am 48: 697–706, 2013.
253. Kraut JA, Kurtz I. Toxic alcohol ingestions: clinical features, diagnosis, and manage-
ment. Clin J Am Soc Nephrol 3: 208–225, 2008.
254. Kruman II, Mattson MP. Pivotal role of mitochondrial calcium uptake in neural cell
apoptosis and necrosis. J Neurochem 72: 529–540, 1999.
255. Kumar D, de Visser SP, Sharma PK, Derat E, Shaik S. The intrinsic axial ligand effect on
propene oxidation by horseradish peroxidase versus cytochrome P450 enzymes. J
Biol Inorg Chem 10: 181–189, 2005.
256. Ku RH, Billings RE. Relationships between formaldehyde metabolism and toxicity and
glutathione concentrations in isolated rat hepatocytes. Chem Biol Interact 51: 25–36,
1984.
257. Lachenmeier DW, Kanteres F, Rehm J. Carcinogenicity of acetaldehyde in alcoholic
beverages: risk assessment outside ethanol metabolism. Addict 104: 533–550, 2009.
METABOLIC METHANOL
637Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
258. Lachenmeier DW, Rehm J, Gmel G. Surrogate alcohol: what do we know and where
do we go? Alcohol Clin Exp Res 31: 1613–1624, 2007.
259. Laethem RM, Balazy M, Falck JR, Laethem CL, Koop DR. Formation of 19(S)-, 19(R)-,
and 18(R)-hydroxyeicosatetraenoic acids by alcohol-inducible cytochrome P450 2E1.
J Biol Chem 268: 12912–12918, 1993.
260. Lagaert S, Beliën T, Volckaert G. Plant cell walls: protecting the barrier from degra-
dation by microbial enzymes. Semin Cell Dev Biol 20: 1064–1073, 2009.
261. Lampe JW. Health effects of vegetables and fruit: assessing mechanisms of action in
human experimental studies. Am J Clin Nutr 70: 475s–490s, 1999.
262. Laniewska-Dunaj M, Jelski W, Orywal K, Kochanowicz J, Rutkowski R, Szmitkowski
M. The activity of class I, II, III and IV of alcohol dehydrogenase (ADH) isoenzymes and
aldehyde dehydrogenase (ALDH) in brain cancer. Neurochem Res 38: 1517–1521,
2013.
263. Leaf G, Zatman LJ. A study of the conditions under which methanol may exert a toxic
hazard in industry. Br J Ind Med 9: 19–31, 1952.
264. Ledesma JC, Aragon CMG. Acquisition and reconditioning of ethanol-induced condi-
tioned place preference in mice is blocked by the H2O2 scavenger alpha lipoic acid.
Psychopharmacology 226: 673–685, 2013.
265. Lee ES, Chen H, Hardman C, Simm A, Charlton C. Excessive S-adenosyl-L-methio-
nine-dependent methylation increases levels of methanol, formaldehyde and formic
acid in rat brain striatal homogenates: possible role in S-adenosyl-L-methionine-in-
duced Parkinson’s disease-like disorders. Life Sci 83: 821–827, 2008.
266. Lee MY, Mukherjee N, Pakstis AJ, Khaliq S, Mohyuddin A, Mehdi SQ, Speed WC, Kidd
JR, Kidd KK. Global patterns of variation in allele and haplotype frequencies and
linkage disequilibrium across the CYP2E1 gene. Pharmacogenomics J 8: 349 –356, 2008.
267. Lee SL, Chau GY, Yao CT, Wu CW, Yin SJ. Functional assessment of human alcohol
dehydrogenase family in ethanol metabolism: significance of first-pass metabolism.
Alcohol Clin Exp Res 30: 1132–1142, 2006.
268. Lee SL, Wang MF, Lee AI, Yin SJ. The metabolic role of human ADH3 functioning as
ethanol dehydrogenase. FEBS Lett 544: 143–147, 2003.
269. Lester D. Endogenous ethanol: a review. Q J Stud Alcohol 22: 554–574, 1961.
270. Levey DJ. The evolutionary ecology of ethanol production and alcoholism. Integr
Comp Biol 44: 284–289, 2004.
271. Lieber CS. Microsomal ethanol-oxidizing system (MEOS): the first 30 years (1968
1998)–a review. Alcohol Clin Exp Res 23: 991–1007, 1999.
272. Lieber CS. The discovery of the microsomal ethanol oxidizing system and its physio-
logic and pathologic role. Drug Metab Rev 36: 511–529, 2004.
273. Lieber CS, Cao Q, DeCarli LM, Leo MA, Mak KM, Ponomarenko A, Ren C, Wang X.
Role of medium-chain triglycerides in the alcohol-mediated cytochrome P450 2E1
induction of mitochondria. Alcohol Clin Exp Res 31: 1660–1668, 2007.
274. Lieber CS, DeCarli LM. Ethanol oxidation by hepatic microsomes: adaptive increase
after ethanol feeding. Science 162: 917–918, 1968.
275. Lieber CS, DeCarli LM. The role of the hepatic microsomal ethanol oxidizing system
(MEOS) for ethanol metabolism in vivo. J Pharmacol Exp Ther 181: 279–287, 1972.
276. Liesivuori J, Savolainen H. Methanol and formic acid toxicity: biochemical mecha-
nisms. Pharmacol Toxicol 69: 157–163, 1991.
277. Li MH, Tang JP, Zhang P, Li X, Wang CY, Wei HJ, Yang XF, Zou W, Tang XQ.
Disturbance of endogenous hydrogen sulfide generation and endoplasmic reticulum
stress in hippocampus are involved in homocysteine-induced defect in learning and
memory of rats. Behav Brain Res 262: 35–41, 2014.
278. Lindinger W, Taucher J, Jordan A, Hansel A, Vogel W. Endogenous production of
methanol after the consumption of fruit. Alcohol Clin Exp Res 21: 939–943, 1997.
279. Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative
diseases. Nature 443: 787–795, 2006.
280. Lin RJ, Sternsdorf T, Tini M, Evans RM. Transcriptional regulation in acute promyelo-
cytic leukemia. Oncogene 20: 7204–7215, 2001.
281. Lin S, Hanson RE, Cronan JE. Biotin synthesis begins by hijacking the fatty acid syn-
thetic pathway. Nat Chem Biol 6: 682–688, 2010.
282. Lionetti V, Cervone F, Bellincampi D. Methyl esterification of pectin plays a role
during plant-pathogen interactions and affects plant resistance to diseases. J Plant
Physiol 169: 1623–1630, 2012.
283. Lionetti V, Raiola A, Camardella L, Giovane A, Obel N, Pauly M, Favaron F, Cervone
F, Bellincampi D. Overexpression of pectin methylesterase inhibitors in Arabidopsis
restricts fungal infection by Botrytis cinerea.Plant Physiol 143: 1871–1880, 2007.
284. Lionetti V, Raiola A, Cervone F, Bellincampi D. Transgenic expression of pectin
methylesterase inhibitors limits tobamovirus spread in tobacco and Arabidopsis.Mol
Plant Pathol 15: 265–274, 2014.
285. Li RJ, Ji WQ, Pang JJ, Wang JL, Chen YG, Zhang Y. Alpha-lipoic acid ameliorates
oxidative stress by increasing aldehyde dehydrogenase-2 activity in patients with
acute coronary syndrome. Tohoku J Exp Med 229: 45–51, 2013.
286. Li TK. Enzymology of human alcohol metabolism. Adv Enzymol Relat Areas Mol Biol 45:
427–483, 1977.
287. Liu J, Liu FY, Tong ZQ, Li ZH, Chen W, Luo WH, Li H, Luo HJ, Tang Y, Tang JM, Cai
J, Liao FF, Wan Y. Lysine-specific demethylase 1 in breast cancer cells contributes to
the production of endogenous formaldehyde in the metastatic bone cancer pain
model of rats. PloS One 8: e58957, 2013.
288. Liu XH, Pan H, Mazur P. Permeation and toxicity of ethylene glycol and methanol in
larvae of Anopheles gambiae.J Exp Biol 206: 2221–2228, 2003.
289. Liu X, Wang Y, Han K. Systematic study on the mechanism of aldehyde oxidation to
carboxylic acid by cytochrome P450. J Biol Inorg Chem 12: 1073–1081, 2007.
290. Louvet R, Cavel E, Gutierrez L, Guénin S, Roger D, Gillet F, Guerineau F, Pelloux J.
Comprehensive expression profiling of the pectin methylesterase gene family during
silique development in Arabidopsis thaliana.Planta 224: 782–791, 2006.
291. Lowe ED, Gao GY, Johnson LN, Keung WM. Structure of daidzin, a naturally occur-
ring anti-alcohol-addiction agent, in complex with human mitochondrial aldehyde
dehydrogenase. J Med Chem 51: 4482–4487, 2008.
292. Lu J, Miao J, Su T, Liu Y, He R. Formaldehyde induces hyperphosphorylation and
polymerization of Tau protein both in vitro and in vivo. Biochim Biophys Acta 1830:
4102–4116, 2013.
293. Lund ED, Kirkland CL, Shaw PE. Methanol, ethanol, and acetaldehyde contents of
citrus products. J Agric Food Chem 29: 361–366, 1981.
294. Luo X, Kranzler HR, Zuo L, Yang B, Lappalainen J, Gelernter J. ADH4 gene variation is
associated with alcohol and drug dependence: results from family controlled and
population-structured association studies. Pharmacogenet Genomics 15: 755–768,
2005.
295. Lu Y, Cederbaum AI. CYP2E1 and oxidative liver injury by alcohol. Free Radic Biol Med
44: 723–738, 2008.
296. Lu Y, He HJ, Zhou J, Miao JY, Lu J, He YG, Pan R, Wei Y, Liu Y, He RQ. Hyperphos-
phorylation results in tau dysfunction in DNA folding and protection. J Alzheimers Dis
37: 551–563, 2013.
297. MacAllister SL, Choi J, Dedina L, O’Brien PJ. Metabolic mechanisms of methanol/
formaldehyde in isolated rat hepatocytes: carbonyl-metabolizing enzymes versus ox-
idative stress. Chem Biol Interact 191: 308–314, 2011.
298. Magnuson BA, Burdock GA, Doull J, Kroes RM, Marsh GM, Pariza MW, Spencer PS,
Waddell WJ, Walker R, Williams GM. Aspartame: a safety evaluation based on current
use levels, regulations, and toxicological and epidemiological studies. Crit Rev Toxicol
37: 629–727, 2007.
299. Majchrowicz E, Mendelson JH. Blood methanol concentrations during experimentally
induced ethanol intoxication in alcoholics. J Pharmacol Exp Ther 179: 293–300, 1971.
300. Maleknia SD, Vail TM, Cody RB, Sparkman DO, Bell TL, Adams MA. Temperature-
dependent release of volatile organic compounds of eucalypts by direct analysis in real
time (DART) mass spectrometry. Rapid Commun Mass Spectrom 23: 2241–2246,
2009.
301. Mani JC, Pietruszko R, Theorell H. Methanol activity o alcohol dehydrogenases from
human liver, horse liver, and yeast. Arch Biochem Biophys 140: 52–59, 1970.
DOROKHOV ET AL.
638 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
302. Mansfield CT, Hodge BT, Hege RB, Hamlin WC. Analysis of formaldehyde in tobacco
smoke by high performance liquid chromatography. J Chromatogr Sci 15: 301–302,
1977.
303. Markovic O, Cederlund E, Griffiths WJ, Lipka T, Jörnvall H. Characterization of carrot
pectin methylesterase. Cell Mol Life Sci 59: 513–518, 2002.
304. Markovic O, Janecek S. Pectin methylesterases: sequence-structural features and
phylogenetic relationships. Carbohydr Res 339: 2281–2295, 2004.
305. Marmot M, Brunner E. Alcohol and cardiovascular disease: the status of the U-shaped
curve. BMJ 303: 565–568, 1991.
306. Martin-Amat G, McMartin KE, Hayreh SS, Hayreh MS, Tephly TR. Methanol poison-
ing: ocular toxicity produced by formate. Toxicol Appl Pharmacol 45: 201–208, 1978.
307. Martin S, Lenz EM, Temesi D, Wild M, Clench MR. Reaction of homopiperazine with
endogenous formaldehyde: a carbon hydrogen addition metabolite/product identified
in rat urine and blood. Drug Metab Dispos Biol Fate Chem 40: 1478–1486, 2012.
308. Matsuo Y, Yokoyama R, Yokoyama S. The genes for human alcohol dehydrogenases
beta 1 and beta 2 differ by only one nucleotide. Eur J Biochem 183: 317–320, 1989.
309. Mazeh S, Korine C, Pinshow B, Dudley R. The influence of ethanol on feeding in the
frugivorous yellow-vented bulbul (Pycnonotus xanthopygos). Behav Processes 77: 369 –
375, 2008.
310. McCarty MF. Nutraceutical strategies for ameliorating the toxic effects of alcohol.
Med Hypotheses 80: 456–462, 2013.
311. McCord CP. Toxicity of methyl alcohol (methanol) following skin absorption and
inhalation. Ind Eng Chem 23: 931–936, 1931.
312. McGovern PE, Zhang J, Tang J, Zhang Z, Hall GR, Moreau RA, Nuñez A, Butrym ED,
Richards MP, Wang CS, Cheng G, Zhao Z, Wang C. Fermented beverages of pre- and
proto-historic China. Proc Natl Acad Sci USA 101: 17593–17598, 2004.
313. McGregor NR. Pueraria lobata (Kudzu root) hangover remedies and acetaldehyde-
associated neoplasm risk. Alcohol 41: 469–478, 2007.
314. McGuire SUS. Department of Agriculture and US Department of Health and Human
Services, Dietary Guidelines for Americans, 2010 7th Edition, Washington, DC: US
Government Printing Office, January 2011. Adv Nutr 2: 293–294, 2011.
315. Medeiros PM, Simoneit BRT. Source profiles of organic compounds emitted upon
combustion of green vegetation from temperate climate forests. Environ Sci Technol
42: 8310–8316, 2008.
316. Mégarbane B, Borron SW, Baud FJ. Current recommendations for treatment of
severe toxic alcohol poisonings. Intensive Care Med 31: 189–195, 2005.
316a.Methanol 196. (EHC, 1997) (Online), 2014. http://www.inchem.org/documents/
ehc/ehc/ehc196.htm.
317. Michaud V, Frappier M, Dumas MC, Turgeon J. Metabolic activity and mRNA levels of
human cardiac CYP450s involved in drug metabolism. PloS One 5: e15666, 2010.
318. Micheli F. Pectin methylesterases: cell wall enzymes with important roles in plant
physiology. Trends Plant Sci 6: 414–419, 2001.
319. Mokhtar SH, Bakhuraysah MM, Cram DS, Petratos S. The beta-amyloid protein of
Alzheimer’s disease: communication breakdown by modifying the neuronal cytoskel-
eton. Int J Alzheimers Dis 2013: 910502, 2013.
320. Molotkov A, Fan X, Deltour L, Foglio MH, Martras S, Farrés J, Parés X, Duester G.
Stimulation of retinoic acid production and growth by ubiquitously expressed alcohol
dehydrogenase Adh3. Proc Natl Acad Sci USA 99: 5337–5342, 2002.
321. Moncada S, Palmer RM, Higgs EA. Nitric oxide: physiology, pathophysiology, phar-
macology. Pharmacol Rev 43: 109–142, 1991.
322. Monte WC. Methanol: a chemical Trojan horse as the root of the inscrutable U. Med
Hypotheses 74: 493–496, 2010.
323. Monticello TM, Morgan KT, Hurtt ME. Unit length as the denominator for quantita-
tion of cell proliferation in nasal epithelia. Toxicol Pathol 18: 24–31, 1990.
324. Monticello TM, Swenberg JA, Gross EA, Leininger JR, Kimbell JS, Seilkop S, Starr TB,
Gibson JE, Morgan KT. Correlation of regional and nonlinear formaldehyde-induced
nasal cancer with proliferating populations of cells. Cancer Res 56: 1012–1022, 1996.
325. Montine TJ, Neely MD, Quinn JF, Beal MF, Markensbery WR, Roberts IILJ, Morrow
JD. Serial review: causes and consequences of oxidative stress in Alzheimer’s disease.
Free Rad Biol Med 33: 620–626, 2002.
326. Moody DM. The blood-brain barrier and blood-cerebral spinal fluid barrier. Semin
Cardiothorac Vasc Anesth 10: 128–131, 2006.
327. Moore PK, Bhatia M, Moochhala S. Hydrogen sulfide: from the smell of the past to the
mediator of the future? Trends Pharmacol Sci 24: 609–611, 2003.
328. Mori O, Haseba T, Kameyama K, Shimizu H, Kudoh M, Ohaki O, Arai Y, Yamazaki M,
Asano G. Histological distribution of class III alcohol dehydrogenase in human brain.
Brain Res 852: 186–190, 2000.
329. Morton BR, Gaut BS, Clegg MT. Evolution of alcohol dehydrogenase genes in the palm
and grass families. Proc Natl Acad Sci USA 93: 11735–11739, 1996.
330. Morton MS, Arisaka O, Miyake N, Morgan LD, Evans BAJ. Phytoestrogen concentra-
tions in serum from Japanese men and women over forty years of age. J Nutr 132:
3168–3171, 2002.
331. Moshonas MG, Shaw PE. Compounds new to essential orange oil from fruit treated
with abscission chemicals. J Agric Food Chem 26: 1288–1293, 1978.
332. Moulis JM, Holmquist B, Vallee BL. Hydrophobic anion activation of human liver chi
chi alcohol dehydrogenase. Biochemistry 30: 5743–5749, 1991.
333. Movva R, Figueredo VM. Alcohol and the heart: to abstain or not to abstain? Int J
Cardiol 164: 267–276, 2013.
334. Muggironi G, Fois GR, Diana M. Ethanol-derived acetaldehyde: pleasure and pain of
alcohol mechanism of action. Front Behav Neurosci 7: 87, 2013.
335. Mukamal KJ, Chen CM, Rao SR, Breslow RA. Alcohol consumption and cardiovascular
mortality among US adults, 1987 to 2002. J Am Coll Cardiol 55: 1328–1335, 2010.
336. Mukerjee N, Pietruszko R. Human mitochondrial aldehyde dehydrogenase substrate
specificity: comparison of esterase with dehydrogenase reaction. Arch Biochem Bio-
phys 299: 23–29, 1992.
337. Murray RP, Connett JE, Tyas SL, Bond R, Ekuma O, Silversides CK, Barnes GE.
Alcohol volume, drinking pattern, and cardiovascular disease morbidity and mortality:
is there a U-shaped function? Am J Epidemiol 155: 242–248, 2002.
338. Nakagawa T, Shimizu S, Watanabe T, Yamaguchi O, Otsu K, Yamagata H, Inohara H,
Kubo T, Tsujimoto Y. Cyclophilin D-dependent mitochondrial permeability transition
regulates some necrotic but not apoptotic cell death. Nature 434: 652–658, 2005.
339. Napoli JL. Biochemical pathways of retinoid transport, metabolism, and signal trans-
duction. Clin Immunol Immunopathol 80: S52–S62, 1996.
340. Nebert DW, Adesnik M, Coon MJ, Estabrook RW, Gonzalez FJ, Guengerich FP,
Gunsalus IC, Johnson EF, Kemper B, Levin W. The P450 gene superfamily: recom-
mended nomenclature. DNA Mary Ann Liebert Inc 6: 1–11, 1987.
341. Nebert DW, Wikvall K, Miller WL. Human cytochromes P450 in health and disease.
Philos Trans R Soc Lond B Biol Sci 368: 20120431, 2013.
342. Nelson DR, Koymans L, Kamataki T, Stegeman JJ, Feyereisen R, Waxman DJ, Water-
man MR, Gotoh O, Coon MJ, Estabrook RW, Gunsalus IC, Nebert DW. P450 super-
family: update on new sequences, gene mapping, accession numbers and nomencla-
ture. Pharmacogenetics 6: 1–42, 1996.
343. Nemecek-Marshall M, MacDonald RC, Franzen JJ, Wojciechowski CL, Fall R. Meth-
anol emission from leaves (enzymatic detection of gas-phase methanol and relation of
methanol fluxes to stomatal conductance and leaf development). Plant Physiol 108:
1359–1368, 1995.
344. Nie CL, Wei Y, Chen X, Liu YY, Dui W, Liu Y, Davies MC, Tendler SJB, He RG.
Formaldehyde at low concentration induces protein tau into globular amyloid-like
aggregates in vitro and in vivo. PloS One 2: e629, 2007.
345. Niederhut MS, Gibbons BJ, Perez-Miller S, Hurley TD. Three-dimensional struc-
tures of the three human class I alcohol dehydrogenases. Protein Sci 10: 697–706,
2001.
346. Nonomura AM, Benson AA. The path of carbon in photosynthesis: improved crop
yields with methanol. Proc Natl Acad Sci USA 89: 9794–9798, 1992.
METABOLIC METHANOL
639Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
347. Norberg A, Jones AW, Hahn RG, Gabrielsson JL. Role of variability in explaining
ethanol pharmacokinetics: research and forensic applications. Clin Pharmacokinet 42:
1–31, 2003.
348. Nosova T, Jousimies-Somer H, Kaihovaara P, Jokelainen K, Heine R, Salaspuro M.
Characteristics of alcohol dehydrogenases of certain aerobic bacteria representing
human colonic flora. Alcohol Clin Exp Res 21: 489–494, 1997.
349. Nosti-Palacios R, Gómez-Garduño J, Molina-Ortiz D, Calzada-León R, Dorado-
González VM, Vences-Mejía A. Aspartame administration and insulin treatment al-
tered brain levels of CYP2E1 and CYP3A2 in streptozotocin-induced diabetic rats. Int
J Toxicol 33: 325–331, 2014.
350. Novak RF, Woodcroft KJ. The alcohol-inducible form of cytochrome P450 (CYP 2E1):
role in toxicology and regulation of expression. Arch Pharm Res 23: 267–282, 2000.
351. Nursten HE, Williams AA. Fruit aromas: a survey of components identified. Chem Ind
12: 486–497, 1967.
352. Ohta S, Ohsawa I. Dysfunction of mitochondria and oxidative stress in the pathogen-
esis of Alzheimer’s disease: on defects in the cytochrome coxidase complex and
aldehyde detoxification. J Alzheimers Dis 9: 155–166, 2006.
353. Oikawa PY, Giebel BM, Sternberg da SLO L, Li L, Timko MP, Swart PK, Riemer DD,
Mak JE, Lerdau MT. Leaf and root pectin methylesterase activity and
13
C/
12
C stable
isotopic ratio measurements of methanol emissions give insight into methanol pro-
duction in Lycopersicon esculentum. New Phytol 191: 1031–1040, 2011.
354. OIV. International Code of Oenological Practices. Paris: International Organisation of
Vine and Wine, 2014.
355. O’Keefe JH, Bhatti SK, Bajwa A, DiNicolantonio JJ, Lavie CJ. Alcohol and cardiovas-
cular health: the dose makes the poisonѧor the remedy. Mayo Clin Proc 89: 382–393,
2014.
356. Olah GA, Prakash GKS, Goeppert A. Anthropogenic chemical carbon cycle for a
sustainable future. J Am Chem Soc 133: 12881–12898, 2011.
357. Olszowy Z, Plewka A, Czech E, Nowicka J, Plewka D, Nowaczyk G, Kami´
nski M.
Effect of L-carnitine supplementation on xenobiotic-metabolizing hepatic enzymes
exposed to methanol. Exp Toxicol Pathol 57: 427–435, 2006.
358. Op den Camp H.JM, Islam T, Stott MB, Harhangi HR, Hynes A, Schouten S, Jetten
MSM, Birkeland N-K, Pol A, Dunfield PF. Environmental, genomic, and taxonomic
perspectives on methanotrophic Verrucomicrobia. Environ Microbiol Rep 1: 293–306,
2009.
359. Osier MV, Pakstis AJ, Goldman D, Edenberg HJ, Kidd JR, Kidd KK. A proline-threonine
substitution in codon 351 of ADH1C is common in Native Americans. Alcohol Clin Exp
Res 26: 1759–1763, 2002.
360. Osier MV, Pakstis AJ, Soodyall H, Comas D, Goldman D, Odunsi A, Okonofua F,
Parnas J, Schulz LO, Bertranpetit J, Bonne-Tamir B, Lu RB, Kidd JR, Kidd KK. A global
perspective on genetic variation at the ADH genes reveals unusual patterns of linkage
disequilibrium and diversity. Am J Hum Genet 71: 84–99, 2002.
361. Ostberg LJ, Strömberg P, Hedberg JJ, Persson B, Höög JO. Analysis of mammalian
alcohol dehydrogenase 5 (ADH5): characterisation of rat ADH5 with comparisons to
the corresponding human variant. Chem Biol Interact 202: 97–103, 2013.
362. Ostrovsky Y. Endogenous ethanol–its metabolic, behavioral and biomedical signifi-
cance. Alcohol 3: 239–247, 1986.
363. O’Sullivan J, Unzeta M, Healy J, O’Sullivan MI, Davey G, Tipton KF. Semicarbazide-
sensitive amine oxidases: enzymes with quite a lot to do. Neurotoxicology 25: 303–315,
2004.
364. Paine A, Davan AD. Defining a tolerable concentration of methanol in alcoholic drinks.
Hum Exp Toxicol 20: 563–568, 2001.
365. Pallas M, Camins A, Smith MA, Perry G, Lee H, Casadesus G. From aging to Alzhei-
mer’s disease: unveiling “the switch” with the senescence-accelerated mouse model
(SAMP8). J Alzheimers Dis 15: 615–624, 2008.
366. Patra J, Taylor B, Irving H, Roerecke M, Baliunas D, Mohapatra S, Rehm J. Alcohol
consumption and the risk of morbidity and mortality for different stroke types–a
systematic review and meta-analysis. BMC Public Health 10: 258, 2010.
367. Payasi A, Mishra NN, Chaves ALS, Singh R. Biochemistry of fruit softening: an over-
view. Physiol Mol Biol Plants 15: 103–113, 2009.
368. Peana AT, Muggironi G, Fois G, Diana M. Alpha-lipoic acid reduces ethanol self-
administration in rats. Alcohol Clin Exp Res 37: 1816–1822, 2013.
369. Peaucelle A, Braybrook S, Höfte H. Cell wall mechanics and growth control in plants:
the role of pectins revisited. Front Plant Sci 3: 121, 2012.
370. Pellizzari E, Hartwell T, Harris BH III, Waddell R, Whitaker D, Erickson M. Purgeable
organic compounds in mother’s milk. Bull Environ Contam Toxicol 28: 322–328, 1982.
371. Pelloux J, Rustérucci C, Mellerowicz EJ. New insights into pectin methylesterase
structure and function. Trends Plant Sci 12: 267–277, 2007.
372. Peng Q, Gizer IR, Libiger O, Bizon C, Wilhelmsen KC, Schork NJ, Ehlers CL. Associ-
ation and ancestry analysis of sequence variants in ADH and ALDH using alcohol-
related phenotypes in a Native American community sample. Am J Med Genet Part B
Neuropsychiatr Genet, doi:10.1002/ajmg.b.32272.
373. Penning R, van Nuland M, Fliervoet a L. L, Olivier B, Verster JC. The pathology of
alcohol hangover. Curr Drug Abuse Rev 3: 68–75, 2010.
374. Pereira GG, Hollenberg CP. Conserved regulation of the Hansenula polymorpha MOX
promoter in Saccharomyces cerevisiae reveals insights in the transcriptional activation
by Adrlp. Eur J Biochem 238: 181–191, 1996.
375. Person RE, Chen H, Fantel AG, Juchau MR. Enzymic catalysis of the accumulation of
acetaldehyde from ethanol in human prenatal cephalic tissues: evaluation of the rela-
tive contributions of CYP2E1, alcohol dehydrogenase, and catalase/peroxidases. Al-
cohol Clin Exp Res 24: 1433–1442, 2000.
376. Pettersson G. Liver alcohol dehydrogenase. CRC Crit Rev Biochem 21: 349 –389, 1987.
377. Piasecki TM, Robertson BM, Epler AJ. Hangover and risk for alcohol use disorders:
existing evidence and potential mechanisms. Curr Drug Abuse Rev 3: 92–102, 2010.
378. Pietruszko R. Human liver alcohol dehydrogenase–inhibition of methanol activity by
pyrazole, 4-methylpyrazole, 4-hydroxymethylpyrazole and 4-carboxypyrazole.
Biochem Pharmacol 24: 1603–1607, 1975.
379. Pocker Y. Bioinorganic and bioorganic studies of liver alcohol dehydrogenase. Chem
Biol Interact 130–132: 383–393, 2001.
380. Pocker Y, Li H. Kinetics and mechanism of methanol and formaldehyde interconver-
sion and formaldehyde oxidation catalyzed by liver alcohol dehydrogenase. Adv Exp
Med Biol 284: 315–325, 1991.
381. Pocker Y, Li H, Page JD. Liver alcohol dehydrogenase: metabolic and energetic as-
pects. Alcohol Alcohol Oxf Suppl 1: 181–185, 1987.
382. Pocker Y, Page JD, Li H, Bhat CC. Ternary complexes of liver alcohol dehydrogenase.
Chem Biol Interact 130–132: 371–381, 2001.
383. Pogorelko G, Lionetti V, Bellincampi D, Zabotina O. Cell wall integrity: targeted
post-synthetic modifications to reveal its role in plant growth and defense against
pathogens. Plant Signal Behav 8: 2013.
384. Prasanna V, Prabha TN, Tharanathan RN. Fruit ripening phenomena–an overview.
Crit Rev Food Sci Nutr 47: 1–19, 2007.
385. Pronko PS, Velichko MG, Moroz AR, Rubanovich NN. Low-molecular-weight metab-
olites relevant to ethanol metabolism: correlation with alcohol withdrawal severity
and utility for identification of alcoholics. Alcohol 32: 761–768, 1997.
386. Qiang M, Xiao R, Su T, Wu BB, Tong ZQ, Liu Y, He RQ. A novel mechanism for
endogenous formaldehyde elevation in SAMP8 mouse (Online). J Alzheimers Dis,
http://iospress.metapress.com/index/78428752K465446U.pdf.
387. Quayle JR, Ferenci T. Evolutionary aspects of autotrophy. Microbiol Rev 42: 251, 1978.
388. Rajamani R, Muthuvel A, Manikandan S, Srikumar R, Sheeladevi R. Efficacy of DL-alpha-
lipoic acid on methanol induced free radical changes, protein oxidative damages and
hsp70 expression in folate deficient rat nervous tissue. Chem Biol Interact 167: 161–
167, 2007.
389. Ravn-Haren G, Dragsted LO, Buch-Andersen T, Jensen EN, Jensen RI, Németh-
Balogh M, Paulovicsová B, Bergström A, Wilcks A, Licht TR, Markowski J, Bügel S.
Intake of whole apples or clear apple juice has contrasting effects on plasma lipids in
healthy volunteers. Eur J Nutr 52: 1875–1889, 2013.
390. Read KA, Carpenter LJ, Arnold SR, Beale R, Nightingale PD, Hopkins JR, Lewis AC,
Lee JD, Mendes L, Pickering SJ. Multiannual observations of acetone, methanol, and
DOROKHOV ET AL.
640 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
acetaldehyde in remote tropical atlantic air: implications for atmospheric OVOC
budgets and oxidative capacity. Environ Sci Technol 46: 11028–11039, 2012.
391. Restani P, Galli CL. Oral toxicity of formaldehyde and its derivatives. Crit Rev Toxicol
21: 315–328, 1991.
392. Rietjens SJ, de Lange DW, Meulenbelt J. Ethylene glycol or methanol intoxication:
which antidote should be used, fomepizole or ethanol? Neth J Med 72: 73–79, 2014.
393. Riveros-Rosas H, Julian-Sanchez A, Pinã E. Enzymology of ethanol and acetaldehyde
metabolism in mammals. Arch Med Res 28: 453–471, 1997.
394. Rizak JD, Ma Y, Hu X. Is formaldehyde the missing link in AD pathology? The differ-
ential aggregation of amyloid-beta with APOE isoforms in vitro. Curr Alzheimer Res 11:
461–468, 2014.
395. Ro Y, Kim E, Kim Y. Enzyme activities related to the methanol oxidation of Mycobac-
terium sp. strain JC1 DSM 3803. J Microbiol 38: 209–217, 2000.
396. Roe O. The metabolism and toxicity of methanol. Pharmacol Rev 7: 399–412, 1955.
397. Ronis MJ, Huang J, Crouch J, Mercado C, Irby D, Valentine CR, Lumpkin CK, Ingel-
man-Sundberg M, Badger TM. Cytochrome P450 CYP 2E1 induction during chronic
alcohol exposure occurs by a two-step mechanism associated with blood alcohol
concentrations in rats. J Pharmacol Exp Ther 264: 944–950, 1993.
398. Salaspuro MP, Lieber CS. Non-uniformity of blood ethanol elimination: its exaggera-
tion after chronic consumption. Ann Clin Res 10: 294–297, 1978.
399. Sanaei-Zadeh H, Kazemi Esfeh S, Zamani N, Jamshidi F, Shadnia S. Hyperglycemia is
a strong prognostic factor of lethality in methanol poisoning. J Med Toxicol 7: 189 –194,
2011.
400. Sánchez-Catalán MJ, Hipólito L, Guerri C, Granero L, Polache A. Distribution and
differential induction of CYP2E1 by ethanol and acetone in the mesocorticolimbic
system of rat. Alcohol 43: 401–407, 2008.
401. Sánchez F, Korine C, Steeghs M, Laarhoven LJ, Cristescu SM, Harren FJM, Dudley R,
Pinshow B. Ethanol and methanol as possible odor cues for Egyptian fruit bats (Rou-
settus aegyptiacus). J Chem Ecol 32: 1289–1300, 2006.
402. San José B, van de Mheen H, van Oers JA, Mackenbach JP, Garretsen HF. The
U-shaped curve: various health measures and alcohol drinking patterns. J Stud Alcohol
60: 725–731, 1999.
403. Sarkola T, Eriksson CJ. Effect of 4-methylpyrazole on endogenous plasma ethanol and
methanol levels in humans. Alcohol Clin Exp Res 25: 513–516, 2001.
404. Sayre LM, Zelasko DA, Harris PL, Perry G, Salomon RG, Smith MA. 4-Hydroxynon-
enal-derived advanced lipid peroxidation end products are increased in Alzheimer’s
disease. J Neurochem 68: 2092–2097, 1997.
405. Schehl B, Lachenmeier D, Senn T, Heinisch JJ. Effect of the stone content on the
quality of plum and cherry spirits produced from mash fermentations with commer-
cial and laboratory yeast strains. J Agric Food Chem 53: 8230–8238, 2005.
406. Schwab W, Davidovich-Rikanati R, Lewinsohn E. Biosynthesis of plant-derived flavor
compounds. Plant J Cell Mol Biol 54: 712–732, 2008.
407. Seco R, Peñuelas J, Filella I. Short-chain oxygenated VOCs: emission and uptake by
plants and atmospheric sources, sinks, and concentrations. Atmos Environ 41: 2477–
2499, 2007.
408. Secretan B, Straif K, Baan R, Grosse Y, El Ghissassi F, Bouvard V, Benbrahim-Tallaa L,
Guha N, Freeman C, Galichet L, Cogliano V. A review of human carcinogens–Part E:
tobacco, areca nut, alcohol, coal smoke, and salted fish. Lancet Oncol 10: 1033–1034,
2009.
409. Sejersted OM, Jacobsen D, Ovrebø S, Jansen H. Formate concentrations in plasma
from patients poisoned with methanol. Acta Med Scand 213: 105–110, 1983.
410. Shcherbakova LN, Tel’pukhov VI, Trenin SO, Bashilov IA, Lapkina TI. Permeability of
the blood-brain barrier to intra-arterial formaldehyde. Biulleten Eksp Biol Meditsiny
102: 573–575, 1986.
411. Sheehan MC, Bailey CJ, Dowds BC, McConnell DJ. A new alcohol dehydrogenase,
reactive towards methanol, from Bacillus stearothermophilus.Biochem J 252: 661– 666,
1988.
412. Shinagawa E, Toyama H, Matsushita K, Tuitemwong P, Theeragool G, Adachi O. A
novel type of formaldehyde-oxidizing enzyme from the membrane of Acetobacter sp.
SKU 14. Biosci Biotechnol Biochem 70: 850–857, 2006.
413. Shindyapina AV, Petrunia IV, Komarova TV, Sheshukova EV, Kosorukov VS, Kiryanov
GI, Dorokhov YL. Dietary methanol regulates human gene activity. PloS One 9:
e102837, 2014.
414. Shinohara K, Ohashi Y, Kawasumi K, Terada A, Fujisawa T. Effect of apple intake on
fecal microbiota and metabolites in humans. Anaerobe 16: 510–515, 2010.
415. Shirreffs SM, Merson SJ, Fraser SM, Archer DT. The effects of fluid restriction on
hydration status and subjective feelings in man. Br J Nutr 91: 951–958, 2004.
416. Shore JD, Gilleland MJ. Binding and kinetic studies of liver alcohol dehydrogenase-
coenzyme-pyrazole complexes. J Biol Chem 245: 3422–3425, 1970.
417. Siems WG, Hapner SJ, van Kuijk FJ. 4-Hydroxynonenal inhibits Na
-K
-ATPase. Free
Radic Biol Med 20: 215–223, 1996.
418. Siragusa RJ, Cerda JJ, Baig MM, Burgin CW, Robbins FL. Methanol production from
the degradation of pectin by human colonic bacteria. Am J Clin Nutr 47: 848–851,
1988.
419. Skrzydlewska E. Toxicological and metabolic consequences of methanol poisoning.
Toxicol Mech Methods 13: 277–293, 2003.
420. Sládek NE. Human aldehyde dehydrogenases: potential pathological, pharmacologi-
cal, and toxicological impact. J Biochem Mol Toxicol 17: 7–23, 2003.
421. Small DH, Gasperini R, Vincent AJ, Hung AC, Foa L. The role of Abeta-induced
calcium dysregulation in the pathogenesis of Alzheimer’s disease. J Alzheimers Dis 16:
225–233, 2009.
422. Snyder SH, Axelrod J, Smith OD, Pucci GL. Formation of methanol by an enzyme in an
ectopic pinealoma. Nature 215: 773–774, 1967.
423. Soffritti M, Padovani M, Tibaldi E, Falcioni L, Manservisi F, Belpoggi F. The carcino-
genic effects of aspartame: the urgent need for regulatory re-evaluation. Am J Ind Med
57: 383–397, 2014.
424. Song MS, Baker GB, Dursun SM, Todd KG. The antidepressant phenelzine protects
neurons and astrocytes against formaldehyde-induced toxicity. J Neurochem 114:
1405–1413, 2010.
425. Songur A, Ozen OA, Sarsilmaz M. The toxic effects of formaldehyde on the nervous
system. Rev Environ Contam Toxicol 203: 105–118, 2010.
426. Songur A, Sarsilmaz M, Ozen O, Sahin S, Koken R, Zararsiz I, Ilhan N. The effects
of inhaled formaldehyde on oxidant and antioxidant systems of rat cerebellum
during the postnatal development process. Toxicol Mech Methods 18: 569–574,
2008.
427. Sophos NA, Vasiliou V. Aldehyde dehydrogenase gene superfamily: the 2002 update.
Chem Biol Interact 143–144: 5–22, 2003.
428. Sorrentino P, Iuliano A, Polverino A, Jacini F, Sorrentino G. The dark sides of amyloid
in Alzheimer’s disease pathogenesis. FEBS Lett 588: 641–652, 2014.
429. Sperling R, Johnson K. Biomarkers of Alzheimer disease: current and future applica-
tions to diagnostic criteria. Contin Minneap 19: 325–338, 2013.
430. Staab CA, Alander J, Brandt M, Lengqvist J, Morgenstern R, Grafström RC, Höög JO.
Reduction of S-nitrosoglutathione by alcohol dehydrogenase 3 is facilitated by sub-
strate alcohols via direct cofactor recycling and leads to GSH-controlled formation of
glutathione transferase inhibitors. Biochem J 413: 493–504, 2008.
431. Staab CA, Alander J, Morgenstern R, Grafström RC, Höög JO. The Janus face of
alcohol dehydrogenase 3. Chem Biol Interact 178: 29–35, 2009.
432. Staab CA, Hellgren M, Höög JO. Medium- and short-chain dehydrogenase/reductase
gene and protein families: dual functions of alcohol dehydrogenase 3: implications
with focus on formaldehyde dehydrogenase and S-nitrosoglutathione reductase ac-
tivities. Cell Mol Life Sci 65: 3950–3960, 2008.
433. Stegink LD, Brummel MC, McMartin K, Martin-Amat G, Filer LJ, Baker GL, Tephly
TR. Blood methanol concentrations in normal adult subjects administered abuse
doses of aspartame. J Toxicol Environ Health 7: 281–290, 1981.
METABOLIC METHANOL
641Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
434. Stegink LD, Filer LJ, Bell EF, Ziegler EE, Tephly TR. Effect of repeated ingestion of
aspartame-sweetened beverage on plasma amino acid, blood methanol, and blood
formate concentrations in normal adults. Metabolism 38: 357–363, 1989.
435. Stewart MJ, Malek K, Crabb DW. Distribution of messenger RNAs for aldehyde
dehydrogenase 1, aldehyde dehydrogenase 2, and aldehyde dehydrogenase 5 in hu-
man tissues. J Investig Med 44: 42–46, 1996.
436. Strolin Benedetti M, Tipton KF, Whomsley R. Amine oxidases and monooxygenases in
the in vivo metabolism of xenobiotic amines in humans: has the involvement of amine
oxidases been neglected? Fundam Clin Pharmacol 21: 467–480, 2007.
437. Strommer J. The plant ADH gene family. Plant J Cell Mol Biol 66: 128–142, 2011.
438. Sun A, Ren J. ALDH2, a novel protector against stroke? Cell Res 23: 874–875, 2013.
439. Sun P, Ding H, Liang M, Li X, Mo W, Wang X, Liu Y, He R, Hua Q. Neuroprotective
effects of geniposide in SH-SY5Y cells and primary hippocampal neurons exposed to
A
42. BioMed Res Int 2014: 1–11, 2014.
440. Suomalainen H. Aroma of Beer, Wine, and Distilled Alcoholic Beverages. New York:
Springer Science & Business Media, 1983.
441. Su T, Wei Y, He RQ. Assay of brain endogenous formaldehyde with 2,4-dinitrophe-
nylhydrazine through UV-HPLC. Prog Biochem Biophys 38: 1171–1177, 2012.
442. Sweeting JN, Siu M, McCallum GP, Miller L, Wells PG. Species differences in methanol
and formic acid pharmacokinetics in mice, rabbits and primates. Toxicol Appl Pharma-
col 247: 2010.
443. Sweeting JN, Siu M, Wiley MJ, Wells PG. Species- and strain-dependent teratogenicity
of methanol in rabbits and mice. Reprod Toxicol 31: 50–58, 2011.
444. Swenberg JA, Kerns WD, Mitchell RI, Gralla EJ, Pavkov KL. Induction of squamous cell
carcinomas of the rat nasal cavity by inhalation exposure to formaldehyde vapor.
Cancer Res 40: 3398–3402, 1980.
445. Swenberg JA, Moeller BC, Lu K, Rager JE, Fry RC, Starr TB. Formaldehyde carcino-
genicity research: 30 years and counting for mode of action, epidemiology, and cancer
risk assessment. Toxicol Pathol 41: 181–189, 2013.
446. Swift R, Davidson D. Alcohol hangover: mechanisms and mediators. Alcohol Health
Res World 22: 54–60, 1998.
447. Takeshita T, Morimoto K, Mao X, Hashimoto T, Furuyama J. Characterization of the
three genotypes of low K
m
aldehyde dehydrogenase in a Japanese population. Hum
Genet 94: 217–223, 1994.
448. Tang XQ, Fang HR, Zhou CF, Zhuang YY, Zhang P, Gu HF, Hu B. A novel mechanism
of formaldehyde neurotoxicity: inhibition of hydrogen sulfide generation by promot-
ing overproduction of nitric oxide. PloS One 8: e54829, 2013.
449. Tang XQ, Shen XT, Huang YE, Chen RQ, Ren YK, Fang HR, Zhuang YY, Wang CY.
Inhibition of endogenous hydrogen sulfide generation is associated with homocys-
teine-induced neurotoxicity: role of ERK1/2 activation. J Mol Neurosci 45: 60–67,
2011.
450. Tang XQ, Zhuang YY, Zhang P, Fang HR, Zhou CF, Gu HF, Zhang H, Wang CY.
Formaldehyde impairs learning and memory involving the disturbance of hydrogen
sulfide generation in the hippocampus of rats. J Mol Neurosci 49: 140–149, 2013.
451. Taucher J, Lagg A, Hansel A, Vogel W, Lindinger W. Methanol in human breath.
Alcohol Clin Exp Res 19: 1147–1150, 1995.
452. Teng S, Beard K, Pourahmad J, Moridani M, Easson E, Poon R, O’Brien PJ. The
formaldehyde metabolic detoxification enzyme systems and molecular cytotoxic
mechanism in isolated rat hepatocytes. Chem Biol Interact 130–132: 285–296, 2001.
453. Tephly TR. The toxicity of methanol. Life Sci 48: 1031–1041, 1991.
454. Thelen K, Dressman JB. Cytochrome P450-mediated metabolism in the human gut
wall. J Pharm Pharmacol 61: 541–558, 2009.
455. Theorell H, Yonetani T. Liver alcohol dehydrogenase-DPN-pyrazole complex: a
model of a ternary intermediate in the enzyme reaction. Biochem Z 338: 537–553,
1963.
456. Theuwissen E, Mensink RP. Water-soluble dietary fibers and cardiovascular disease.
Physiol Behav 94: 285–292, 2008.
457. Thompson CM, Ceder R, Grafström RC. Formaldehyde dehydrogenase: beyond
phase I metabolism. Toxicol Lett 193: 1–3, 2010.
458. Thorogood M. Vegetarianism, coronary disease risk factors and coronary heart dis-
ease. Curr Opin Lipidol 5: 17–21, 1994.
459. Tillonen J, Kaihovaara P, Jousimies-Somer H, Heine R, Salaspuro M. Role of catalase in
in vitro acetaldehyde formation by human colonic contents. Alcohol Clin Exp Res 22:
1113–1119, 1998.
460. Ting HR, Qiang Min LI. Chronic dehydration and regularly drinking water to mitigate
age-related cognitive impairment. Acta Neuropharmacol 2: 43, 2012.
461. Toews ML, Adler J. Methanol formation in vivo from methylated chemotaxis proteins
in Escherichia coli.J Biol Chem 254: 1761–1764, 1979.
462. Tolstrup JS, Stephens R, Grønbaek M. Does the severity of hangovers decline with
age? Survey of the incidence of hangover in different age groups. Alcohol Clin Exp Res
38: 466–470, 2014.
463. Tong Z, Han C, Luo W, Li H, Luo H, Qiang M, Su T, Wu B, Liu Y, Yang X, Wan Y, Cui
D, He R. Aging-associated excess formaldehyde leads to spatial memory deficits. Sci
Rep 3: 1807, 2013.
464. Tong Z, Han C, Luo W, Wang X, Li H, Luo H, Zhou J, Qi J, He R. Accumulated
hippocampal formaldehyde induces age-dependent memory decline. Age 35: 583–
596, 2013.
465. Tong Z, Han C, Qiang M, Wang W, Lv J, Zhang S, Luo W, Li H, Luo H, Zhou J, Wu B,
Su T, Yang X, Wang X, Liu Y, He R. Age-related formaldehyde interferes with DNA
methyltransferase function, causing memory loss in Alzheimer’s disease. Neurobiol
Aging 36: 100–110, 2014.
466. Tong Z, Zhang J, Luo W, Wang W, Li F, Li H, Luo H, Lu J, Zhou J, Wan Y, He R. Urine
formaldehyde level is inversely correlated to mini mental state examination scores in
senile dementia. Neurobiol Aging 32: 31–41, 2011.
467. Trocho C, Pardo R, Rafecas I, Virgili J, Remesar X, Fernández-López JA, Alemany M.
Formaldehyde derived from dietary aspartame binds to tissue components in vivo.
Life Sci 63: 337–349, 1998.
468. Truswell AS. Meta-analysis of the cholesterol-lowering effects of dietary fiber. Am J
Clin Nutr 70: 942–943, 1999.
469. Tsuchiya K, Hayashi Y, Onodera M, Hasegawa T. Toxicity of formaldehyde in exper-
imental animals–concentrations of the chemical in the elution from dishes of formal-
dehyde resin in some vegetables. Keio J Med 24: 19–37, 1975.
470. Tsukada Y, Fang J, Erdjument-Bromage H, Warren ME, Borchers CH, Tempst P,
Zhang Y. Histone demethylation by a family of JmjC domain-containing proteins.
Nature 439: 811–816, 2006.
471. Tulpule K, Dringen R. Formate generated by cellular oxidation of formaldehyde ac-
celerates the glycolytic flux in cultured astrocytes. Glia 60: 582–593, 2012.
472. Tulpule K, Dringen R. Formaldehyde in brain: an overlooked player in neurodegen-
eration? J Neurochem 127: 7–21, 2013.
473. Tulpule K, Hohnholt MC, Dringen R. Formaldehyde metabolism and formaldehyde-
induced stimulation of lactate production and glutathione export in cultured neurons.
J. Neurochem, doi:10.1111/jnc.12170.
474. Turner C, Španeˇl P, Smith D. A longitudinal study of methanol in the exhaled breath
of 30 healthy volunteers using selected ion flow tube mass spectrometry, SIFT-MS.
Physiol Meas 27: 637–648, 2006.
475. Uchida K, Stadtman E. Covalent attachment of 4-hydroxynonenal to glyceraldehyde-
3-phosphate dehydrogenase. A possible involvement of intra-and intermolecular
cross-linking reaction. J Biol Chem 268: 6388–6393, 1993.
476. Uebelacker M, Lachenmeier DW. Quantitative determination of acetaldehyde in
foods using automated digestion with simulated gastric fluid followed by headspace
gas chromatography. J Autom Methods Manag Chem 2011: 1–13, 2011.
477. Umar S, Morris AP, Kourouma F, Sellin JH. Dietary pectin and calcium inhibit colonic
proliferation in vivo by differing mechanisms. Cell Prolif 36: 361–375, 2003.
477a.United States Code of Federal Regulations. Title 27: Alcohol, Tobacco Products, and
Firearms, chapter 1, subpart C. Standards of Identity for distilled spirits, 2003.
DOROKHOV ET AL.
642 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
477b.United States Code of Federal Regulations. Title 21: Food and Drugs, section 172.804
(Online), 2005. http://www.accessdata.fda.gov/scripts/cdrh/cfdocs/cfCFR/
CFRSearch.cfm?fr172.804.
478. Unzeta M, Solé M, Boada M, Hernández M. Semicarbazide-sensitive amine oxidase
(SSAO) and its possible contribution to vascular damage in Alzheimer’s disease. J
Neural Transm 114: 857–862, 2007.
479. Uotila L, Koivusalo M. Formaldehyde dehydrogenase from human liver. Purification,
properties, and evidence for the formation of glutathione thiol esters by the enzyme.
J Biol Chem 249: 7653–7663, 1974.
480. Uotila L, Koivusalo M. Purification and properties of S-formylglutathione hydrolase
from human liver. J Biol Chem 249: 7664–7672, 1974.
481. Upadhyay RK. Drug delivery systems, CNS protection, and the blood brain barrier.
BioMed Res Int 2014: 1–37, 2014.
482. Valencia-Olvera AC, Morán J, Camacho-Carranza R, Prospéro-García O, Espinosa-
Aguirre JJ. CYP2E1 induction leads to oxidative stress and cytotoxicity in glutathione-
depleted cerebellar granule neurons. Toxicol Vitro 28: 1206–1214, 2014.
482a.Van der Klei IJ, Harder W, Veenhuis M. Biosynthesis and assembly of alcohol oxidase,
a peroxisomal matrix protein in methylotrophic yeasts: a review. Yeast 7, 1991.
482b.Van Dijken JP, Otto R, Harder W. Oxidation of methanol, formaldehyde and formate
by catalase purified from methanol-grown Hansenula polymorpha.Arch Microbiol 106:
221–226, 1975.
483. Varma MVS, Feng B, Obach RS, Troutman MD, Chupka J, Miller HR, El-Kattan A.
Physicochemical determinants of human renal clearance. J Med Chem 52: 4844 –4852,
2009.
484. Vasiliou V, Nebert DW. Analysis and update of the human aldehyde dehydrogenase
(ALDH) gene family. Hum Genomics 2: 138–143, 2005.
485. Vasiliou V, Pappa A, Estey T. Role of human aldehyde dehydrogenases in endobiotic
and xenobiotic metabolism. Drug Metab Rev 36: 279–299, 2004.
486. Vasiliou V, Ziegler TL, Bludeau P, Petersen DR, Gonzalez FJ, Deitrich RA. CYP2E1
and catalase influence ethanol sensitivity in the central nervous system. Pharmacogenet
Genomics 16: 51–58, 2006.
487. Verster JC, Penning R. Treatment and prevention of alcohol hangover. Curr Drug
Abuse Rev 3: 103–109, 2010.
488. Volpi C, Janni M, Lionetti V, Bellincampi D, Favaron F, D’Ovidio R. The ectopic
expression of a pectin methyl esterase inhibitor increases pectin methyl esterification
and limits fungal diseases in wheat. Mol Plant Microbe Interact 24: 1012–1019, 2011.
488a.Von Dahl CC, Hävecker M, Schlögl R, Baldwin IT. Caterpillar-elicited methanol emis-
sion: a new signal in plant-herbivore interactions? Plant J Cell Mol Biol 46: 948–960,
2006.
489. Vonwartburg JP, Bethune JL, Vallee BL. Human liver-alcohol dehydrogenase. Kinetic
and physicochemical properties. Biochemistry 3: 1775–1782, 1964.
490. Vree TB, Verwey-van Wissen CP. Pharmacokinetics and metabolism of codeine in
humans. Biopharm Drug Dispos 13: 445–460, 1992.
491. Wagner F. Methanol: a fermentation substrate. Cell Mol Life Sci 33: 110–113, 1977.
492. Wagner FW, Burger AR, Vallee BL. Kinetic properties of human liver alcohol dehy-
drogenase: oxidation of alcohols by class I isoenzymes. Biochemistry 22: 1857–1863,
1983.
493. Wales MR, Fewson CA. NADP-dependent alcohol dehydrogenases in bacteria and
yeast: purification and partial characterization of the enzymes from Acinetobacter
calcoaceticus and Saccharomyces cerevisiae.Microbiology 140: 173–183, 1994.
494. Walker RK, Cousins VM, Umoh NA, Jeffress MA, Taghipour D, Al-Rubaiee M, Haddad
GE. The good, the bad, and the ugly with alcohol use and abuse on the heart. Alcohol
Clin Exp Res 37: 1253–1260, 2013.
495. Wallage HR, Watterson JH. Formic acid and methanol concentrations in death inves-
tigations. J Anal Toxicol 32: 241–247, 2008.
496. Wall TL, Carr LG, Ehlers CL. Protective association of genetic variation in alcohol
dehydrogenase with alcohol dependence in Native American Mission Indians. Am J
Psychiatry 160: 41–46, 2003.
497. Walport LJ, Hopkinson RJ, Schofield CJ. Mechanisms of human histone and nucleic
acid demethylases. Curr Opin Chem Biol 16: 525–534, 2012.
498. Wang B, Wang J, Zhou S, Tan S, He X, Yang Z, Xie YC, Li S, Zheng C, Ma X. The
association of mitochondrial aldehyde dehydrogenase gene (ALDH2) polymorphism
with susceptibility to late-onset Alzheimer’s disease in Chinese. J Neurol Sci 268:
172–175, 2008.
499. Wang J, Du H, Jiang L, Ma X, de Graaf RA, Behar KL, Mason GF. Oxidation of ethanol
in the rat brain and effects associated with chronic ethanol exposure. Proc Natl Acad
Sci USA 110: 14444–14449, 2013.
500. Wang R. Two’s company, three’s a crowd: can H
2
S be the third endogenous gaseous
transmitter? FASEB J 16: 1792–1798, 2002.
501. Wang R. Hydrogen sulfide: the third gasotransmitter in biology and medicine. Antioxid
Redox Signal 12: 1061–1064, 2010.
502. Wang RS, Nakajima T, Kawamoto T, Honma T. Effects of aldehyde dehydrogenase-2
genetic polymorphisms on metabolism of structurally different aldehydes in human
liver. Drug Metab Dispos Biol Fate Chem 30: 69–73, 2002.
503. Wang SP, He GL, Chen RR, Li F, Li GQ. The involvement of cytochrome P450
monooxygenases in methanol elimination in Drosophila melanogaster larvae. Arch In-
sect Biochem Physiol 79: 264–275, 2012.
504. Wang SP, Hu XX, Meng QW, Muhammad SA, Chen RR, Li F, Li GQ. The involvement
of several enzymes in methanol detoxification in Drosophila melanogaster adults. Comp
Biochem Physiol B Biochem Mol Biol 166: 7–14, 2013.
505. Warneke C, Luxembourg SL, de Gouw JA, Rinne HJI, Guenther AB, Fall R. Disjunct
eddy covariance measurements of oxygenated volatile organic compounds fluxes
from an alfalfa field before and after cutting. J Geophys Res Atmospheres 107: ACH
6-1–ACH 6–10, 2002.
506. Weltman MD, Farrell GC, Hall P, Ingelman-Sundberg M, Liddle C. Hepatic cyto-
chrome P450 2E1 is increased in patients with nonalcoholic steatohepatitis. Hepatol-
ogy 27: 128–133, 1998.
507. Westerlund M, Galter D, Carmine A, Olson L. Tissue- and species-specific expression
patterns of class I, III, and IV Adh and Aldh 1 mRNAs in rodent embryos. Cell Tissue Res
322: 227–236, 2005.
508. White RE, Coon MJ. Oxygen activation by cytochrome P-450. Annu Rev Biochem 49:
315–356, 1980.
509. Wiens F, Zitzmann A, Lachance MA, Yegles M, Pragst F, Wurst FM, von Holst D, Guan
SL, Spanagel R. Chronic intake of fermented floral nectar by wild treeshrews. Proc
Natl Acad Sci USA 105: 10426–10431, 2008.
510. Wiese JG, Shlipak MG, Browner WS. The alcohol hangover. Ann Intern Med 132:
897–902, 2000.
511. Willats WG, McCartney L, Mackie W, Knox JP. Pectin: cell biology and prospects for
functional analysis. Plant Mol Biol 47: 9–27, 2001.
512. Williams LB, Canfield B, Voglesonger KM, Holloway JR. Organic molecules formed in
a “primordial womb.” Geology 33: 913–916, 2005.
513. Wolf S, Mouille G, Pelloux J. Homogalacturonan methyl-esterification and plant de-
velopment. Mol Plant 2: 851–860, 2009.
514. Wolin MJ. Fermentation in the rumen and human large intestine. Science 213: 1463–
1468, 1981.
515. Woolf D, Lehmann J, Fisher EM, Angenent LT. Biofuels from pyrolysis in perspective:
trade-offs between energy yields and soil-carbon additions. Environ Sci Technol 48:
6492–6499, 2014.
516. Wu D, Cederbaum AI. Alcohol, oxidative stress, and free radical damage. Alcohol Res
Health 27: 277–284, 2003.
517. Yang M, Lu J, Miao J, Rizak J, Yang J, Zhai R, Zhou J, Qu J, Wang J, Yang S, Ma Y, Hu X,
He R. Alzheimer’s disease and methanol toxicity (part 1): chronic methanol feeding
led to memory impairments and tau hyperphosphorylation in mice. J Alzheimers Dis
41: 1117–1129, 2014.
518. Yang M, Miao J, Rizak J, Zhai R, Wang Z, Anwar TH, Li T, Zheng N, Wu S, Zheng Y.
Alzheimer’s disease and methanol toxicity (part 2): lessons from Four rhesus ma-
METABOLIC METHANOL
643Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
caques (Macaca mulatta) chronically fed methanol (Online). J. Alzheimers Dis,
http://iospress.metapress.com/index/PU35QT82R4L05226.pdf.
519. Yang M, Nightingale PD, Beale R, Liss PS, Blomquist B, Fairall C. Atmospheric deposition of
methanol over the Atlantic Ocean. Proc Natl Acad Sci USA 110: 20034 –20039, 2013.
520. Yang ZN, Bosron WF, Hurley TD. Structure of human chi chi alcohol dehydrogenase: a
glutathione-dependent formaldehyde dehydrogenase. J Mol Biol 265: 330 –343, 1997.
521. Yasueda H, Kawahara Y, Sugimoto S. Bacillus subtilis yckG and yckF encode two key
enzymes of the ribulose monophosphate pathway used by Methylotrophs, and yckH
is required for their expression. J Bacteriol 181: 7154–7160, 1999.
522. Yoritaka A, Hattori N, Uchida K, Tanaka M, Stadtman ER, Mizuno Y. Immunohisto-
chemical detection of 4-hydroxynonenal protein adducts in Parkinson disease. Proc
Natl Acad Sci 93: 2696–2701, 1996.
523. Yound JM, Wang JH. The nature of binding of competitive inhibitors to alcohol dehy-
drogenases. J Biol Chem 246: 2815–2821, 1971.
524. Yraola F, Zorzano A, Albericio F, Royo M. Structure-activity relationships of SSAO/
VAP-1 arylalkylamine-based substrates. Chem Med Chem 4: 495–503, 2009.
525. Yu PH. Involvement of cerebrovascular semicarbazide-sensitive amine oxidase in the patho-
genesis of Alzheimer’s disease and vascular dementia. Med Hypotheses 57: 175–179, 2001.
526. Yu PH, Deng Y. Potential cytotoxic effect of chronic administration of creatine, a nutrition
supplement to augment athletic performance. Med Hypotheses 54: 726 –728, 2000.
527. Yu PH, Wright S, Fan EH, Lun ZR, Gubisne-Harberle D. Physiological and pathological
implications of semicarbazide-sensitive amine oxidase. Biochim Biophys Acta 1647:
193–199, 2003.
528. Yurimoto H, Hirai R, Yasueda H, Mitsui R, Sakai Y, Kato N. The ribulose monophos-
phate pathway operon encoding formaldehyde fixation in a thermotolerant methyl-
otroph, Bacillus brevis S1. FEMS Microbiol Lett 214: 189–193, 2002.
529. Zaitseva M, Vollenhoven BJ, Rogers PAW. Retinoic acid pathway genes show signifi-
cantly altered expression in uterine fibroids when compared with normal myome-
trium. Mol Hum Reprod 13: 577–585, 2007.
530. Zanger UM, Schwab M. Cytochrome P450 enzymes in drug metabolism: regulation of
gene expression, enzyme activities, and impact of genetic variation. Pharmacol Ther
138: 103–141, 2013.
531. Zeng J, Konopka G, Hunt BG, Preuss TM, Geschwind D, Yi SV. Divergent whole-
genome methylation maps of human and chimpanzee brains reveal epigenetic basis of
human regulatory evolution. Am J Hum Genet 91: 455–465, 2012.
532. Zhang RH, Gao JY, Guo HT, Scott GI, Eason AR, Wang XM, Ren J. Inhibition of
CYP2E1 attenuates chronic alcohol intake-induced myocardial contractile dysfunc-
tion and apoptosis. Biochim Biophys Acta 1832: 128–141, 2013.
533. Zhang W, Lu D, Dong W, Zhang L, Zhang X, Quan X, Ma C, Lian H, Zhang L.
Expression of CYP2E1 increases oxidative stress and induces apoptosis of cardiomy-
ocytes in transgenic mice. FEBS J 278: 1484–1492, 2011.
534. Zhang Y, Liu X, McHale C, Li R, Zhang L, Wu Y, Ye X, Yang X, Ding S. Bone marrow
injury induced via oxidative stress in mice by inhalation exposure to formaldehyde.
PloS One 8: e74974, 2013.
535. Zimatkin SM. Histochemical study of aldehyde dehydrogenase in the rat CNS. J
Neurochem 56: 1–11, 1991.
536. Zimatkin SM, Buben AL. Ethanol oxidation in the living brain. Alcohol 42: 529–532,
2007.
537. Zimatkin SM, Pronko SP, Vasiliou V, Gonzalez FJ, Deitrich RA. Enzymatic mechanisms
of ethanol oxidation in the brain. Alcohol Clin Exp Res 30: 1500–1505, 2006.
538. Zuber R, Anzenbacherová E, Anzenbacher P. Cytochromes P450 and experimental
models of drug metabolism. JCell Mol Med 6: 189–198, 2002.
DOROKHOV ET AL.
644 Physiol Rev VOL 95 APRIL 2015 www.prv.org
on April 2, 2015Downloaded from
doi:10.1152/physrev.00034.2014 95:603-644, 2015.Physiol Rev
Tatiana V. Komarova
Yuri L. Dorokhov, Anastasia V. Shindyapina, Ekaterina V. Sheshukova and
Physiological Roles
Metabolic Methanol: Molecular Pathways and
You might find this additional info useful...
518 articles, 112 of which can be accessed free at:This article cites /content/95/2/603.full.html#ref-list-1
including high resolution figures, can be found at:Updated information and services /content/95/2/603.full.html
can be found at:Physiological Reviewsabout Additional material and information
http://www.the-aps.org/publications/prv
This information is current as of April 2, 2015.
website at http://www.the-aps.org/.
MD 20814-3991. Copyright © 2015 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit our
published quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda
provides state of the art coverage of timely issues in the physiological and biomedical sciences. It isPhysiological Reviews
on April 2, 2015Downloaded from
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Cell proliferation data, generally based on a labeling index (LI), provide a valuable endpoint for assessment of toxic and potentially carcinogenic responses in laboratory animals. Measurement of the LI is time consuming because of the large number of cells that need to be counted to determine the denominator. In respiratory mucosa, the total cell count of the surface epithelia may be altered in response to treatment, either through cell loss or increases in cell number (e.g., hyperplasia). As an alternative to the more conventional LI, the present studies were carried out to assess the value of expressing cell proliferation in nasal epithelia as a unit length labeling index (ULLI), defined as labeled cells per mm of basement membrane. Rats were exposed by inhalation to formaldehyde or methyl bromide, and changes in cell proliferation were determined in the respiratory and olfactory epithelia, respectively, using both total cell count and basement membrane length as denominators. Total cell counts were clearly influenced by treatment, while basement membrane length was not. Both methods revealed similar treatment-induced effects on cell proliferation, and in fact were highly correlated (R ≥ 0.92, p < 0.001). It was concluded that the ULLI method provides an effective alternative to total cell counts and the LI method. This approach is not influenced by alterations in the total cell population, and has the benefit of being less labor intensive than LI determinations.
Article
Paracoccus denitrificans and its near relative Paracoccus versutus (formerly known as Thiobacillus versutus) have been attracting increasing attention because the aerobic respiratory system of P. denitrificans has long been regarded as a model for that of the mitochondrion, with which there are many components (e.g., cytochrome aa(3) oxidase) in common. Members of the genus exhibit a great range of metabolic flexibility, particularly with respect to processes involving respiration. Prominent examples of flexibility are the use in denitrification of nitrate, nitrite, nitrous oxide, and nitric oxide as alternative electron acceptors to oxygen and the ability to use C-1 compounds (e.g., methanol and methylamine) as electron donors to the respiratory chains. The proteins required for these respiratory processes are not constitutive, and the underlying complex regulatory systems that regulate their expression are beginning to be unraveled. There has been uncertainty about whether. transcription in a member of the alpha-3 Proteobacteria such as P. denitrificans involves a conventional sigma(70)-type RNA polymerase, especially since canonical -35 and -10 DNA binding sites have not been readily identified. In this review, we argue that many genes, in particular those encoding constitutive proteins, may be under the control of a sigma(70) RNA polymerase very closely related to that of Rhodobacter capsulatus. While the main focus is on the structure and regulation of genes coding for products involved in respiratory processes in Paracoccus, the current state of knowledge of the components of such respiratory pathways, and their biogenesis, is also reviewed.
Article
Mycobacterium sp. strain JC1 DSM 3803 grown in methanol showed no methanol dehydrogenase or oxidase activities found in most methylotrophic bacteria and yeasts, respectively. Even though the methanol-grown cells exhibited a little methanol-dependent oxidation by cytochrome c-dependent methanol dehydrogenase and alcohol dehydrogenase, they were not the key enzymes responsible for the methanol oxidation of the cells, in that the cells contained no c-type cytochrome and the methanol oxidizing activity from the partially purified alcohol dehydrogenase was too low, respectively. In substrate switching experiments, we found that only a catalase-peroxidase among the three types of catalase found in glucose-grown cells was highly expressed in the methanol-grown cells and that its activity was relatively high during the exponential growth phase in Mycobacterium sp. JC1. Therefore, we propose that catalase-peroxidase is an essential enzyme responsible for the methanol metabolism directly or indirectly in Mycobacterium sp. JC1.