Sei sulla pagina 1di 15

JOURNAL OF NEUROCHEMISTRY | 2013 | 127 | 721 doi: 10.1111/jnc.

12356

*Indian Institute of Science Education and Research, Pashan, Pune, India


Centre for Biomolecular Interactions Bremen, University of Bremen, Bremen, Germany
Centre for Environmental Research and Sustainable Technology, Bremen, Germany

Abstract Although moderate concentrations of formaldehyde are not


Formaldehyde is an environmental pollutant that is also acutely toxic for brain cells, exposure to formaldehyde
generated in substantial amounts in the human body during severely affects their metabolism as demonstrated by the
normal metabolism. This aldehyde is a well-established formaldehyde-induced acceleration of glycolytic ux and by
neurotoxin that affects memory, learning, and behavior. In the rapid multidrug resistance protein 1-mediated export of
addition, in several pathological conditions, including Alzhei- glutathione from both astrocytes and neurons. These formal-
mers disease, an increase in the expression of formaldehyde- dehyde-induced alterations in the metabolism of brain cells
generating enzymes and elevated levels of formaldehyde in may contribute to the impaired cognitive performance
brain have been reported. This article gives an overview on observed after formaldehyde exposure and to the neurode-
the current knowledge on the generation and metabolism of generation in diseases that are associated with increased
formaldehyde in brain cells as well as on formaldehyde- formaldehyde levels in brain.
induced alterations in metabolic processes. Brain cells have Keywords: formaldehyde, glutathione, glycolysis, metabolism,
the potential to generate and to dispose formaldehyde. In neurodegeneration, neurotoxicity.
culture, both astrocytes and neurons efciently oxidize form- J. Neurochem. (2013) 127, 721.
aldehyde to formate which can be exported or further oxidized.

Formaldehyde chemistry contains 35% formaldehyde, while for xation of tissues,


tissue sections, or cultured cells, a 4% formaldehyde
Formaldehyde (HCHO) is the simplest aldehyde that is also solution is frequently used (Kiernan 2000). Such a 4%
known as methanal. This compound was rst described in formaldehyde solution contains the aldehyde in a concen-
1855 by Alexander Butlerov, while its chemical synthesis by tration of above 1 M. Thus, the concentrations of formal-
methanol dehydration was rst achieved in 1867 by August dehyde that are used for technical processes are several
Wilhelm von Hofmann (Salthammer et al. 2010). In the orders of magnitude higher than the concentrations of
following decades, the properties of formaldehyde were
extensively studied and this compound was one of the Received May 30, 2013; revised manuscript received June 12, 2013;
earliest to obtain a CAS registry number (50-00-0). Form- accepted June 21, 2013.
aldehyde is highly reactive. It can undergo hydration and Address correspondence and reprint requests to Dr. Ralf Dringen,
forms hemiacetals with alcohols or thiohemiacetals with Centre for Biomolecular Interactions Bremen, University of Bremen,
PO. Box 330440, D-28334 Bremen, Germany.
thiols. Formaldehyde also reacts with amines to form Schiff E-mail: ralf.dringen@uni-bremen.de
bases and cross-links proteins by forming methylene bridges Abbreviations used: AD, Alzheimers disease; ADH, alcohol dehy-
between amino groups (Metz et al. 2004, 2006). This high drogenase; ALDH, aldehyde dehydrogenase; GSH, glutathione; GSSG,
reactivity of formaldehyde is the reason for its extensive use glutathione disulde; JHDM, JmjC domain-containing histone demeth-
in industries (Tang et al. 2009). ylases; LSD, lysine-specic demethylase; MCT, monocarboxylate
transporter; Mrp, multidrug resistance protein; MS, multiple sclerosis;
Due to its protein cross-linking ability, formaldehyde is MTHFD, methylene tetrahydrofolate dehydrogenase; SSAO, semicar-
frequently used for tissue preservation and xation (Nazar- bazide-sensitive amine oxidases; THF, tetrahydrofolate; VAP, vascular
ian et al. 2009). Formalin solution that is used in pathology adhesion protein.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21 7


8 K. Tulpule and R. Dringen

formaldehyde (0.10.4 mM) that are found in body uids the majority of methanol oxidation in rats has been reported
and tissues under normal and pathological conditions (Heck to be mediated by catalase (Tephly 1991; Skrzydlewska
and Casanova 2004; Tong et al. 2013a). 2003).
Another endogenous source of formaldehyde are semicar-
bazide-sensitive amine oxidases (SSAO) which represent a
Endogenous and exogenous sources of
group of copper-containing amine oxidases that are inhibited
formaldehyde
by semicarbazide and most of them contain topa-quinone at
Formaldehyde exposure is caused by the generation of this their catalytic centre (Jalkanen and Salmi 2001; Yu et al.
aldehyde within the body and can also be a consequence of 2003). Oxidative deamination of methylamine by SSAO
contact with elevated levels of environmental formaldehyde generates formaldehyde together with ammonia and hydro-
(Fig. 1). Some of the endogenous enzymatic reactions that gen peroxide (Yu et al. 2003; OSullivan et al. 2004). In
generate formaldehyde as well as exogenous sources of mammals, SSAO are either membrane-associated or circulate
formaldehyde are described below. in a soluble form in the vascular system (Jalkanen and Salmi
Formaldehyde is the oxidation product of methanol. This 2001). Among the SSAO, the vascular adhesion protein
alcohol can be generated within the body by hydrolysis of (VAP) 1 is one of the most extensively studied members of
protein carboxymethyl esters either non-enzymatically or this group of enzymes (Smith and Vainio 2007; Jalkanen and
catalyzed by methylesterases (Lee et al. 2008). In addition, Salmi 2008).
accidental or intentional intake of methanol will further Formaldehyde is also generated as by-product of reactions
expose the body to this alcohol. In cells, methanol is oxidized catalyzed by lysine-specic demethylase (LSD) 1 and JmjC
to formaldehyde by alcohol dehydrogenase (ADH) 1, by domain-containing histone demethylases (JHDM) (Cloos
catalase or by a non-enzymatic reaction of methanol with et al. 2008; Hou and Yu 2010). These enzymes remove
hydroxyl radicals (Harris et al. 2003; MacAllister et al. methyl groups from lysine residues in histones, thereby
2011). In humans and primates, ADH1 appears to be altering the chromatin structure (Cheng and Zhang 2007;
predominately responsible for methanol oxidation, while Cloos et al. 2008; Hou and Yu 2010; Izzo and Schneider
2010). LSD1 is a avin-containing enzyme that selectively
demethylates the mono- or dimethylated lysine residue in
position 4 of histone H3 (Forneris et al. 2009; Hou and Yu
2010). On the other hand, JHDM can remove methyl groups
from mono-, di-, or trimethylated lysine residues and require
Fe2+ and a-ketoglutarate as cofactors (Cloos et al. 2008; Hou
and Yu 2010).
In addition to endogenous sources, the body can also
encounter environmental formaldehyde, since a number of
commonly used products contain either formaldehyde or
formaldehyde-releasing substances (Sasseville 2004;
de Groot et al. 2009). Some examples of such products are
construction materials, agricultural fertilizers, fumigants,
paints, cosmetics, antiperspirants, polish, cleaning agents,
and toiletries (Sasseville 2004; de Groot et al. 2009, 2010).
In addition, formaldehyde can be produced and released from
burning of wood, coal, tobacco, natural gas, and kerosene
(de Groot et al. 2009; Laitinen et al. 2010). Moreover, foods
like coffee, codsh, meat, poultry, and maple syrup naturally
contain formaldehyde (Dhareshwar and Stella 2008; de
Groot et al. 2009). Thus, this ubiquitously present compound
can enter the human body by inhalation, ingestion, or entry
through the skin.
One pertinent question is whether exogenous formalde-
Fig. 1 Endogenous and exogenous sources of formaldehyde (HCHO)
and pathways involved in cellular formaldehyde disposal. For details
hyde can pose a big threat to the central nervous system by
see text. ADH, alcohol dehydrogenase; ALDH, aldehyde dehydroge- entering the blood and ultimately reaching the brain after
nase; cy, cytosolic; JHDM, JmjC domain-containing histone demeth- crossing the bloodbrain barrier. In healthy individuals, the
ylases; LSD, lysine-specic demethylase; mt, mitochondrial; MTHFD, formaldehyde concentration in the blood is around 0.1 mM
methylene tetrahydrofolate dehydrogenase; SSAO, semicarbazide- (Heck and Casanova 2004) and that in the brain is
sensitive amine oxidases; VAP, vascular adhesion protein. 0.20.4 mM (Tong et al. 2013a). Inhalation of moderate

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 9

doses of formaldehyde does not severely increase the ADH3 is likely to be especially important for the oxidation
formaldehyde level in blood (Heck et al. 1985; Franks of low concentrations of formaldehyde.
2005). This is expected as the formaldehyde-oxidizing The formate generated by formaldehyde oxidation can
enzymes ADH3 and aldehyde dehydrogenase (ALDH) 2 undergo further oxidization to carbon dioxide in a metabolic
(Fig. 1) are ubiquitously expressed in all tissues (Nishimura pathway involving tetrahydrofolate (THF), wherein formate
and Naito 2006; Alnouti and Klaassen 2008) and will is rst converted to 10-formyl THF (Fig. 1) in an ATP-
quickly clear a low excess of environmentally derived dependent reaction (Skrzydlewska 2003; Krupenko 2009;
formaldehyde. However, exposure to high concentrations of Krupenko et al. 2010). This reaction is catalyzed either by
exogenous formaldehyde that exceeds the peripheral form- the cytosolic methylene tetrahydrofolate dehydrogenase
aldehyde oxidation capacity will elevate the normal tolerable (MTHFD) 1 or by its mitochondrial isoform MTHFD1L
concentration of formaldehyde in the blood and could lead (Tibbetts and Appling 2010). 10-formyl THF is subsequently
to neural damage. Indeed, exposure to exogenous formal- oxidized by the cytosolic 10-formyl THF dehydrogenase,
dehyde has been reported to cause neurotoxicity in humans also known as ALDH1L1, or its mitochondrial isoform
and animals and the extent of damage depends on the dose ALDH1L2 to carbon dioxide. Both enzymes use NADP+ as a
of formaldehyde and the duration of the exposure (Kilburn co-factor and regenerate THF (Skrzydlewska 2003; Kru-
et al. 1985a, b; Songur et al. 2008, 2010). Especially, penko 2009; Krupenko et al. 2010). Although formate
individuals who carry functional polymorphisms in the oxidation takes place predominantly by the THF-dependent
genes encoding for formaldehyde-metabolizing enzymes, pathway, catalase-mediated oxidation of formate has also
ADH3 or ALDH2, which are discussed to be associated with been reported (Cook et al. 2001; Skrzydlewska 2003).
reduced formaldehyde-oxidizing capacity (Hedberg et al. Formaldehyde metabolism (Fig. 1) is best studied for the
2001; Wang et al. 2002), may be more vulnerable to liver (Skrzydlewska 2003; Tibbetts and Appling 2010), but it
neural damage by endogenously generated or environmental is very likely that other organs including the brain will also
formaldehyde. use the enzymatic pathways that are well known for
formaldehyde metabolism in liver. In brain, at least all the
enzymes required for complete formaldehyde oxidation are
Metabolism of formaldehyde
expressed (Table 1).
Despite of the multiple endogenous and exogenous sources Differences in the rate of formaldehyde metabolism have
of formaldehyde, a low physiological level of formaldehyde been described between species for the formaldehyde
in body uids and tissue is maintained by the continuous metabolism. For example, formate is metabolized at a slower
action of cellular formaldehyde-metabolizing enzymes rate in the liver of monkeys and humans compared to rats,
(Fig. 1). ADH1 is considered to play a negligible role in partly because rats have a higher hepatic THF content
formaldehyde reduction to methanol because of its very high (Tephly 1991; Skrzydlewska 2003). Also, species-specic
KM-value for formaldehyde (about 30 mM) (Skrzydlewska differences in the kinetic parameters of the enzymes involved
2003). The formaldehyde oxidation product formate is in formaldehyde metabolism may contribute to the different
generated by two independent pathways that are mediated rates of formaldehyde oxidation observed and subsequently
by either the mitochondrial ALDH2 or the cytosolic ADH3 may determine the consequences of an exposure to formal-
(Teng et al. 2001; Friedenson 2011; MacAllister et al. dehyde and/or it metabolites.
2011). ADH3, also known as glutathione (GSH)-dependent
formaldehyde dehydrogenase, oxidizes formaldehyde to
Generation and oxidation of formaldehyde
formate in a two-step process (Harris et al. 2003; Staab
in brain cells
et al. 2009; Thompson et al. 2010; MacAllister et al. 2011).
In the rst step, GSH reacts with formaldehyde in an Several reports have demonstrated that the enzymes required
enzyme-independent manner to form S-hydroxymethyl GSH to produce or metabolize formaldehyde are expressed in the
that is subsequently used as ADH3 substrate to generate S- brain on the mRNA or protein level (Table 1). Of these
formyl GSH (Harris et al. 2003; Staab et al. 2009; Thomp- enzymes, only the expression of ADH1 in the brain has been
son et al. 2010; MacAllister et al. 2011). The conjugate S- controversially discussed, since this dehydrogenase was not
formyl GSH is hydrolyzed by a thiolase to generate formate detected in brain by some investigators (Julia et al. 1987;
and GSH (Teng et al. 2001; Harris et al. 2003; MacAllister Galter et al. 2003). Despite the presence of ADH1 mRNA in
et al. 2011). Unlike ADH3, the reaction catalyzed by cultured neural cells, methanol generation was not found for
ALDH2 is a single-step GSH-independent process (Teng formaldehyde-exposed cultured brain cells (Tulpule and
et al. 2001; MacAllister et al. 2011). Since ADH3 has a Dringen 2012; Tulpule et al. 2013), suggesting that oxida-
very low KM-value for S-hydroxymethyl GSH (less than tion to formate is the preferred pathway of formaldehyde
10 lM) compared to that of ALDH2 for formaldehyde (0.2 metabolism in brain cells. Cultured astrocytes and neurons
0.5 mM) (Casanova-Schmitz et al. 1984; Heck et al. 1990), contain the mRNAs for SSAO and LSD1 as well as for the

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


10 K. Tulpule and R. Dringen

Table 1 Formaldehyde-producing and formaldehyde-metabolizing enzymes in the brain

Species

Enzymes Rat Mouse Human

Formaldehyde generation
ADH1 Martinez et al. (2001)
Catalase Zimatkin and Lindros (1996); Schad et al. (2003) Meinerz et al. (2013) van Horssen et al. (2008)
SSAO/VAP1 Obata and Yamanaka (2000) Ferrer et al. (2002); Unzeta et al. (2007);
Valente et al. (2012)
LSD1 Zibetti et al. (2010) Zhang et al. (2010) Zibetti et al. (2010)
JHDM Wolf et al. (2007) Fukuda et al. (2011) Wolf et al. (2007)
Formaldehyde oxidation
ADH3 Julia et al. (1987); Iborra et al. (1992); Galter et al. (2003) Galter et al. (2003)
Galter et al. (2003)
ALDH2 Guo et al. (2013) Alnouti and Klaassen (2008) Stewart et al. (1996)
Formate oxidation
MTHFD1 Thigpen et al. (1990) MacFarlane et al. (2009) Fountoulakis et al. (2003)
MTHFD1L Prasannan et al. (2003)
ALDH1L1 Neymeyer et al. (1997); Anthony and Heintz (2007) Cahoy et al. (2008) Oldham et al. (2008)
ALDH1L2 Krupenko et al. (2010)

enzymes involved in formaldehyde metabolism (Tulpule and these cells to further oxidize formate to carbon dioxide
Dringen 2012; Tulpule et al. 2013). These studies indicate (Fig. 1). Although the putative formate exporters, GABA-
that formaldehyde may be produced locally in the brain and gated channels (Mason et al. 1990) and monocarboxylate
that among the different types of brain cells at least astrocytes transporter (MCT) 1 (Moschen et al. 2012) are expressed in
and neurons have the potential to generate and oxidize both astrocytes and neurons (Debernardi et al. 2003; Olsen
formaldehyde. and Sieghart 2009; Lee et al. 2011; Velez-Fort et al. 2011),
Acute formaldehyde exposure in concentrations of up to the expression level of MCT1 in neurons has been reported
1 mM for up to 3 h does not cause severe toxicity in cultured to be very low (Debernardi et al. 2003). However, if poor
astrocytes or neurons (Song et al. 2010; Tulpule and Dringen export of formate would be the only reason behind the lower
2011, 2012; Tulpule et al. 2013). A rapid metabolism of extracellular accumulation of this metabolite in cultured
cellular formaldehyde may contribute to the resistance of neurons, these cells should accumulate large amounts of
cultured brain cells to formaldehyde toxicity, since formal- formaldehyde-derived formate, which is not the case (Tulp-
dehyde has been reported to be more cytotoxic than its ule et al. 2013). Thus, the lower extracellular accumulation
metabolites, methanol and formate (Oyama et al. 2002; Lee of formaldehyde-derived formate in cultured neurons com-
et al. 2008). Both, cultured astrocytes and neurons clear pared to cultured astrocytes is likely to be predominantly
exogenously applied formaldehyde with a similar rate of caused by oxidation of formaldehyde-derived cellular
around 0.2 lmol/(h 9 mg) (Tulpule and Dringen 2012; formate to carbon dioxide. The enzymes involved in the
Tulpule et al. 2013) which is about 20% of the formaldehyde oxidation of 10-formyl THF require NADP+ as electron
oxidation rate reported for liver cells (Dicker and Cederbaum acceptor (Krupenko 2009; Krupenko et al. 2010), and the
1984). The KM-value for formaldehyde clearance by cultured availability of NADP+ in cytosol and mitochondria depends
astrocytes is around 0.19 mM, suggesting that both the on the pathways involved in NADPH consumption and
cytosolic ADH3 and mitochondrial ALDH2 could contribute NADPH regeneration. As such pathways differ between
to formaldehyde oxidation (Tulpule and Dringen 2012). astrocytes and neurons (Dringen et al. 2007), the NADP+
Although cultured astrocytes and neurons have compara- availability could also contribute to the differences observed
ble rates of formaldehyde clearance, the metabolic fate of the in formate release from astrocytes and neurons that were
disposed formaldehyde differs between these two types of exposed to formaldehyde (Tulpule and Dringen 2012;
neural cells. Although astrocytes convert the majority Tulpule et al. 2013).
(> 90%) of formaldehyde to formate that is subsequently
exported from the cells (Tulpule and Dringen 2012), only
Alterations of the metabolism of brain
about 25% of the formaldehyde cleared by cultured neurons
cells upon exposure to formaldehyde
is detected as extracellular formate (Tulpule et al. 2013). The
underlying reason for this difference might be a poor export A large number of adverse consequences have been reported
of formate from cultured neurons and/or a higher capacity of for an exposure of brain cells to formaldehyde in vivo and

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 11

Table 2 Consequences of a formaldehyde exposure of rodent brain cells in vivo and in vitro

References

In vivo
Decrease in the number of neuron Gurel et al. (2005); Aslan et al. (2006); Sarsilmaz et al. (2007)
Decreased level of GSH Lu et al. (2008)
Lowered levels of superoxide dismutase and catalase Gurel et al. (2005); Zararsiz et al. (2006, 2007, 2011); Lu et al. (2008);
Songur et al. (2008)
Increase in levels of nitric oxide, malondialdehyde Gurel et al. (2005); Zararsiz et al. (2006, 2007, 2011); Lu et al. (2008);
and protein carbonyls Songur et al. (2008)
Increase in apoptotic events Zararsiz et al. (2006, 2007)
Decit in memory and learning Pitten et al. (2000); Usanmaz et al. (2002); Malek et al. (2003);
Sorg et al. (2004); Lu et al. (2008); Turkoglu et al. (2008);
Tong et al. (2011, 2013a, b)
In vitro
Elevated glycolysis in neurons and astrocytes Tulpule and Dringen (2012); Tulpule et al. (2013)
Mrp1-stimulated GSH export from neurons and astrocytes Tulpule and Dringen (2011); Tulpule et al. (2013)
Decreased glutamate uptake in cultured astrocytes Song et al. (2010)
Lower expression of neuronal NMDA receptor subunits Tong et al. (2013a)

The articles by Lu et al. (2008), Usanmaz et al. (2002), and Tong et al. (2011) describe data that have been obtained on mice, whereas all other
studies were performed on rats or rat brain cells.

in vitro (Table 2). Recently, it was demonstrated that c oxidase (Nicholls 1975; Wallace et al. 1997). This view is
formaldehyde in the concentration range between 0.1 mM supported by the observation that incubation of astrocytes
and 1 mM strongly affects basal metabolic properties of with formaldehyde for 90 min is required for the accelerated
cultured astrocytes and neurons, that is, formaldehyde lactate release to persist even after removal of formaldehyde
stimulates glycolytic ux and the export of the antioxidative (Tulpule and Dringen 2012). This long delay most likely
tripeptide GSH from brain cells. reects the slow mitochondrial accumulation of formalde-
hyde-derived formate to concentrations that are sufcient to
Formaldehyde-stimulated glycolysis inactivate respiration, as most of the formate is efciently
Astrocytes are more glycolytic than neurons (Bola~ nos et al. exported from astrocytes. Moreover, the persistent lactate
2010), a feature which has been attributed to expression of release of astrocytes exposed to formaldehyde was not
the glycolysis-promoting enzyme PFKFB3 in astrocytes further enhanced by application of azide, an inhibitor of
(Herrero-Mendez et al. 2009), an inhibited pyruvate dehy- mitochondrial cytochrome c oxidase (Tulpule and Dringen
drogenase complex (Halim et al. 2010) and a low rate of 2012). Thus, formaldehyde-derived formate is likely to
NADH shuttling into mitochondria in astrocytes (Berkich stimulate glycolytic ux as a consequence of an inhibited
et al. 2007; Neves et al. 2012). Despite the differences in respiration, as also other inhibitors of respiratory chain
basal rates of glucose consumption and lactate release in complexes stimulate glycolytic lactate production in cultured
cultured astrocytes and neurons, application of formaldehyde astrocytes and neurons (Pauwels et al. 1985; Scheiber and
signicantly increases these rates in both types of brain cells Dringen 2011).
(Tulpule and Dringen 2012; Tulpule et al. 2013). However,
the extent of stimulation of glycolytic ux in formaldehyde- Formaldehyde-accelerated glutathione export
exposed cells compared to the basal condition differs GSH is an important antioxidant (Lushchak 2012; Schmidt
between the culture types investigated. For example, at a and Dringen 2012; Lu 2013) that is also involved in the
formaldehyde concentration of 0.5 mM, the lactate release formaldehyde oxidation catalyzed by ADH3 (Fig. 1). Under
and glucose consumption rates were doubled in cultured basal conditions, cultured astrocytes and neurons as well as
neurons (Tulpule et al. 2013), while this concentration of cells of the oligodendroglial cell line OLN-93 export GSH,
formaldehyde did not affect glycolysis in cultured astrocytes although with variable rates (Tulpule and Dringen 2011;
(Tulpule and Dringen 2012). Astrocytes had to be exposed to Tulpule et al. 2012, 2013). Formaldehyde treatment stimu-
1 mM formaldehyde to elevate glycolysis by 50% (Tulpule lated GSH export from all three types of cultured neural cells
and Dringen 2012). without severely altering the ratio of GSH to glutathione
The accelerated glycolysis in formaldehyde-exposed neu- disulde (GSSG) (Tulpule and Dringen 2011; Tulpule et al.
ral cells is likely to be caused by the formaldehyde-derived 2012, 2013). This accelerated GSH export from formalde-
formate which is known to inhibit mitochondrial cytochrome hyde-treated neural cells is mediated by multidrug resistance

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


12 K. Tulpule and R. Dringen

protein (Mrp) 1 (Tulpule and Dringen 2011; Tulpule et al. has been reported for cultured astrocytes treated with
2012, 2013). Mrp1 is a member of ATP-binding cassette bilirubin (Gennuso et al. 2004).
transporters and transports, besides GSH, a wide array of Mrp1 efciently exports GSH conjugates (Keppler 2011;
substrates including GSSG and GSH conjugates (Keppler Yin and Zhang 2011). As the formaldehyde metabolism in
2011; Yin and Zhang 2011). The potential of formaldehyde neural cells involves the generation of the GSH conjugates
to accelerate GSH export differs between different brain cell S-hydroxymethyl GSH and S-formyl GSH (Fig. 1), these
culture types. For example, exposure to 0.5 mM formalde- conjugates could also serve as substrates of Mrp1 (Fig. 2b).
hyde increased the respective GSH export rates of cultured Since both conjugates are known to be labile (Ahmed and
astrocytes, neurons and OLN-93 cells by 10-, 5- and 20-fold, Ahmed 1978; Uotila 1981), they are likely to disintegrate
respectively (Tulpule and Dringen 2011; Tulpule et al. 2012, into GSH and formaldehyde or formate immediately after
2013). However, half-maximal cellular GSH depletions were being exported.
observed at similar incubation parameters for all types of Direct experimental evidence that discriminates between
neural cells after incubation for 1 h with 0.3 mM formalde- the potential two mechanisms (Fig. 2) that may be involved
hyde (Tulpule and Dringen 2011; Tulpule et al. 2012, 2013). in the formaldehyde-induced accelerated GSH export via
Formaldehyde exposure does not impair the capacity of Mrp1 is missing so far. However, determination of the
neural cells to synthesize GSH. At least formaldehyde-treated kinetic parameters for the GSH export from astrocytes
neurons restored their cellular GSH levels after application of revealed that the KM-values of the basal as well as the
amino acid precursors for GSH synthesis (Tulpule et al. formaldehyde-accelerated GSH export from astrocytes are
2013). identical (about 100 nmol/mg or 25 mM), but that the
The molecular mechanism involved in the formaldehyde- Vmax-value for the stimulated GSH export is eightfold higher
accelerated Mrp1-mediated GSH export from neural cells is than that for the basal GSH export (Tulpule et al. 2012).
not resolved so far. Since the stimulation of GSH export is These data suggest that at least for formaldehyde-treated
observed within minutes after formaldehyde application astrocytes GSH rather than a GSH conjugate is exported via
(Tulpule and Dringen 2011; Tulpule et al. 2012, 2013), Mrp1, since the KM-values of Mrp1 for its substrate GSH are
de novo synthesis of Mrp1 is unlikely to explain the normally higher than 5 mM, while that for GSH conjugates
stimulated GSH efux. Furthermore, the nding that removal are below 1 mM (Burg et al. 2002; Cole and Deeley 2006;
of formaldehyde instantly decelerates the stimulated GSH Deeley and Cole 2006).
export (Tulpule and Dringen 2011; Tulpule et al. 2012, Application of formaldehyde does not deprive the cells
2013) indicates that the mechanism responsible for formal- completely of their GSH and about 5% residual GSH still
dehyde-accelerated GSH export is quickly reversible. remains within neural cells (Tulpule and Dringen 2011;
Assuming that cellular GSH is the transported Mrp1 Tulpule et al. 2012, 2013). In cultured astrocytes, this low
substrate (Fig. 2a), formaldehyde could stimulate GSH cellular GSH content represents a residual GSH concentra-
export by a reversible, covalent activation of this transporter. tion of about 0.4 mM (Dringen and Hamprecht 1998) which
Alternatively, a formaldehyde-induced recruitment of intra- will be sufcient to drive ADH3-catalyzed GSH-dependent
cellular Mrp1 molecules into the cell membrane could formaldehyde oxidation, since the KM-value of ADH3 for
explain the accelerated GSH export. Such a reversible S-hydroxymethyl GSH is less than 10 lM (Casanova-
translocation of Mrp1 from the Golgi to the cell surface Schmitz et al. 1984; Heck et al. 1990) and this reaction

(a) (b)

Fig. 2 Potential mechanisms involved in


formaldehyde-stimulated glutathione (GSH)
export from brain cells. (a) Formaldehyde
directly stimulates Mrp1-mediated GSH export.
(b) The GSH conjugates S-hydroxymethyl
GSH and/or S-formyl GSH which are
intermediates of cellular formaldehyde
metabolism are exported by Mrp1. The
labile conjugates immediately disintegrate
after export to generate GSH.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 13

involves recycling of GSH (Fig. 1). Thus, the stimulated Increased expression of formaldehyde-generating enzymes
GSH export is unlikely to compromise GSH-dependent (Table 3) as well as elevated formaldehyde levels have also
formaldehyde oxidation. been reported in brains of patients suffering from neurode-
generative diseases like Alzheimers disease (AD) or multi-
ple sclerosis (MS) (Khokhlov et al. 1989 cited in Miao and
Evidence for the role of formaldehyde in pathology
He 2012; Tong et al. 2011, 2013a). Some hypotheses have
In healthy individuals, the formaldehyde concentration in the been postulated that link the increase in formaldehyde level
blood has been reported to be around 0.1 mM (Heck and to neuropathology. For example, some human subjects who
Casanova 2004) while that in the brain is about 0.2 mM suffered from methanol poisoning developed symptoms of
(hippocampus) and 0.4 mM (cortex) (Tong et al. 2013a). MS which has been discussed to be an effect of methanol
These levels of formaldehyde represent the normal phy- oxidation to formaldehyde and the subsequent modication
siological balance between formaldehyde-generating and of proteins resulting in an immune reaction (Schwyzer and
formaldehyde-disposing processes. However, an increased Henzi 1983; Henzi 1984). Along that line it was discussed
activity of formaldehyde-generating enzymes or an acute that formaldehyde methylates proteins like tau (in AD) or
exposure to high amounts of exogenous formaldehyde myelin basic protein (in MS) which in turn elicits an immune
without a concurrent elevation in the capacity to clear response by the body that is characteristic for these diseases
formaldehyde will raise formaldehyde level in the body and (Monte 2010; Lu et al. 2013). Also, inhibition of SSAO in a
will lead to formaldehyde stress (He et al. 2010). Indeed, an murine model of MS has been shown to reduce the incidence
increased expression/activity of the formaldehyde-generating and severity of this disease (Wang et al. 2006) which could,
enzymes VAP1/SSAO, LSD1 and JHDM has been reported at least partly, be the consequence of a lowered formaldehyde
for various diseases (Table 3). While a broad spectrum of generation. Moreover, formaldehyde exposure has been
pathological conditions are associated with elevated levels of implicated to be a risk factor for the development of
VAP1/SSAO, an increase in the expression of the histone amyotrophic lateral sclerosis (Weisskopf et al. 2009), a
demethylases has especially been observed in different types disease that is characterized by degeneration of motor
of cancer (Table 3). The elevated expression of formalde- neurons (Kiernan et al. 2011).
hyde-generating enzymes is accompanied by increased
formaldehyde levels in diabetic rats (Tong et al. 2013a), in
Formaldehyde-induced alterations in neural
cancer tissue (Tong et al. 2010) and in some human cancer
metabolism as potential contributors to
cell lines (Kato et al. 2001; Tong et al. 2010).
neurodegeneration
Figure 3 summarizes the current knowledge on formalde-
Table 3 Elevation in expression or activity of formaldehyde-generat-
hyde metabolism and on formaldehyde-induced alterations in
ing enzymes in human diseases
the glucose and GSH metabolism of neural cells. The
potential of cultured brain cells to efciently metabolize
Enzyme Disease References
formaldehyde suggests that also the cells in brain deal quite
SSAO/VAP1 Alzheimers disease Ferrer et al. (2002); del Mar well with the moderate amounts of formaldehyde that are
Hernandez et al. (2005); generated under physiological conditions. Similar to liver
Unzeta et al. (2007) cells, brain cells are likely to use both cytosolic and
Multiple sclerosis Airas et al. (2006) mitochondrial pathways for formaldehyde oxidation to
Heart disease Boomsma et al. (2000, 2005)
formate and further to carbon dioxide (Figs 1 and 3).
Diabetes mellitus Meszaros et al. (1999);
Cultured brain cells efciently produce and export glyco-
and diabetic Gronvall-Nordquist
lytically generated lactate and also release GSH into the
complications et al. (2001); Karadi et al.
(2002); Boomsma et al.
medium, although the basal rates of glycolysis and GSH
(2005); Obata (2006) export differ between different types of neural cells (Tulpule
Chronic liver disease Kurkijarvi et al. (2000) and Dringen 2011, 2012; Tulpule et al. 2012, 2013). These
LSD1/JHDM Sarcoma Schildhaus et al. (2011); pathways are not affected by low concentrations of formal-
Bennani-Baiti et al. (2012) dehyde, but as soon as formaldehyde levels are increased in
Peripheral nerve Schildhaus et al. (2011) pathological conditions, an accelerated generation of formate
sheath tumor is likely to stimulate glycolytic ux by inhibition of the
Neuroblastoma Schulte et al. (2009) mitochondrial respiration (Fig. 3). In addition, an excess of
Bladder cancer Hayami et al. (2010, 2011)
formaldehyde deprives brain cells of GSH by stimulating
Breast cancer Lim et al. (2010)
Mrp1-mediated GSH export (Fig. 3). Although, caution should
Prostate cancer Kahl et al. (2006); Xiang
be exercised while extrapolating in vitro data to the situation
et al. (2007)
in the brain, a speculation on potential consequences of

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


14 K. Tulpule and R. Dringen

Fig. 3 Metabolic consequences of a formaldehyde exposure in generated formate is exported, while a fraction is further oxidized to
cultured brain cells. Exogenous formaldehyde is entering brain cells carbon dioxide. Remaining cellular formate is likely to inhibit mito-
most likely by diffusion through the cell membrane and is oxidized chondrial cytochrome c oxidase which leads to accelerated glycolytic
within the cell to formate either in a glutathione (GSH)-dependent ux. Formaldehyde also induces a rapid Mrp1-mediated GSH export
reaction that is mediated by cytosolic alcohol dehydrogenase (ADH) 3 from brain cells. Small black squares indicate transporters that are
or by the mitochondrial aldehyde dehydrogenase (ALDH) 2. Part of the required for membrane transport of the indicated metabolites.

Fig. 4 Potential consequences of an


excess of formaldehyde in brain. Presence
of excess of formaldehyde or formaldehyde-
derived metabolites will acutely modulate
metabolic pathways of brain cells (light gray
squares) which are likely to cause delayed,
indirect consequences (dark gray squares)
that nally lead to the adverse effects
reported for formaldehyde exposure.

elevated formaldehyde levels in brain on the cellular metab- released from brain cells or inactivates mitochondrial cyto-
olism is tempting, especially since the formaldehyde concen- chrome c oxidase. An inhibition of the mitochondrial
trations that have been shown to alter metabolic properties of respiratory chain will stimulate glycolytic ux in the brain
cultured brain cells (0.11 mM) are in the concentration cells to, at least transiently, meet their energy demand.
range reported for the normal brain (0.20.4 mM). Thus, mild However, prolonged exposure to formaldehyde is likely to
elevations in brain formaldehyde concentrations could already result in energy crisis that in turn will disrupt the functions of
strongly affect energy and GSH metabolism of this organ. brain cells. This may also be the underlying mechanism of
The potential pathological implications of metabolic the neurotoxicity of formate in hippocampal brain slices
changes exerted by excess of formaldehyde in the brain are (Kapur et al. 2007). Besides this impairment of energy
shown in Fig. 4. Astrocytes and neurons in brain are likely to metabolism, formaldehyde-induced accumulation of both
efciently metabolize an excess of formaldehyde, as also formate and lactate in the brain would cause cerebral acidosis
reported for brain homogenates (Iborra et al. 1992). Subse- (Skrzydlewska 2003; Rose 2010) which would subsequently
quently, the formate generated from formaldehyde is either induce astrocytic swelling, impairment of neuronal signal

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 15

transmission and neurological decits (Staub et al. 1993; Li apoptotic Bax protein increases (Tang et al. 2012). Also, the
et al. 2011; Zhao et al. 2011). expression of the rate-limiting enzyme in dopamine synthesis,
Exposure to high levels of formaldehyde will cause GSH tyrosine hydroxylase, is lowered in PC12 cells after exposure
depletion in brain cells together with GSH accumulation in to formaldehyde (Lee et al. 2008). Further studies are now
the extracellular space. As GSH is involved in important required to investigate the signaling pathways that link the
cellular functions in the brain like protection against reactive acute formaldehyde-induced metabolic alterations to the
oxygen species and detoxication of xenobiotics (Lushchak known brain pathology of an excess of formaldehyde
2012; Schmidt and Dringen 2012; Lu 2013), GSH depletion (Table 2)
may contribute to the severe oxidative stress reported for Conditions such as aging and diseases like MS and AD
brain after prolonged exposure to formaldehyde (Zararsiz which are associated with increased levels of formaldehyde
et al. 2006, 2007, 2011; Songur et al. 2008). A loss in in brain (Khokhlov et al. 1989 cited in Miao and He 2012;
cellular GSH would under normal conditions be compen- Tong et al. 2011, 2013a, b) show impaired mitochondrial
sated by increased GSH synthesis. However, lactacidosis function (Sullivan and Brown 2005; Mahad et al. 2008;
caused by the formaldehyde-induced production of lactate Boumezbeur et al. 2010; Leuner et al. 2012) together with
(Skrzydlewska 2003; Rose 2010) impairs GSH synthesis an increase in brain lactate content (Parnetti et al. 2000; Ross
(Lewerenz et al. 2010) and cellular GSH levels are likely to et al. 2010; Paling et al. 2011). Moreover, ageing, MS and
remain low. Thus, chronic exposure to formaldehyde may AD have been connected with oxidative stress in the brain
render brain cells incapable of fully restoring their cellular (Haider et al. 2011; van Horssen et al. 2011; Belkacemi
GSH levels. and Ramassamy 2012; Sohal and Orr 2012; Steele and
The formaldehyde-induced accumulation of extracellular Robinson 2012). These reports strengthen the view that
GSH in brain can also be detrimental, since GSH has been formaldehyde may, at least to some extent, have a role in the
suggested to act as a neurotransmitter and neuromodulator at initiation and/or progression of pathological symptoms of
glutamate receptors (Janaky et al. 2007) which play impor- neurodegenerative conditions (Yu 2001; Monte 2010). An
tant roles in memory and learning (Davis et al. 2013; adequate supply of lactate to neurons has been shown to
Mukherjee and Manahan-Vaughan 2013). Also, accelerated foster memory formation (Suzuki et al. 2011), while GSH
extracellular GSH hydrolysis by the astrocytic ectoenzyme depletion in the brain has been demonstrated to result in
c-GT (Dringen et al. 1997) caused by the increased extra- behavioral changes (Steullet et al. 2010). Thus, the formal-
cellular GSH concentration would generate the neurotrans- dehyde-induced alterations in glucose and GSH metabolism
mitter glutamate (Fernandez-Fernandez et al. 2012; Schmidt may contribute to the decits in behavior, cognition and
and Dringen 2012). Thus, excessive accumulation of extra- learning observed in formaldehyde-exposed animals (Pitten
cellular GSH as well as GSH-derived glutamate may cause et al. 2000; Malek et al. 2003; Lu et al. 2008; Tong et al.
excitotoxicity which has at least been demonstrated in vitro 2011, 2013a, b)
(Regan and Guo 1999a, b).
To address the molecular mechanisms that are involved in
Conclusions and future perspectives
the development of adverse neural effects of an elevated
concentration of formaldehyde, it has to be discriminated In conclusion, elevation of brain formaldehyde levels is
between direct and indirect consequences of formaldehyde likely to alter brain cell metabolism which may affect the
exposure. Acute exposure of neural cells to formaldehyde function of this vital organ. Although some studies have
and/or the rapid generation of formaldehyde-derived metab- correlated that neurodegenerative conditions are associated
olites will directly affect basal metabolic parameters (Fig. 4, with increased levels of formaldehyde in the brain and others
light gray squares), which may subsequently lead to indirect, have connected such diseases with impaired energy metab-
delayed consequences (Fig. 4, dark gray squares). Little is olism and oxidative stress, a direct causal link between
known so far on the mechanisms that link acute direct formaldehyde, impaired metabolism and oxidative stress
consequences of a formaldehyde exposure, such as acceler- remains to be demonstrated. Interestingly, resveratrol which
ated glycolysis or GSH export, to the known adverse effects is known to be neuroprotective for AD (Richard et al. 2011;
of formaldehyde on neural cells (Table 2). Activation of Li et al. 2012) is a formaldehyde scavenger (Tyihak and
signaling cascades as well as alterations in protein expression Kiraly-Veghely 2008), suggesting that the benecial effects
are likely to be involved in the development of the delayed of resveratrol could also include removal of excess formal-
indirect effects of an exposure to excess of formaldehyde. dehyde. Further studies that will combine the quantication
For example, formaldehyde-exposed neuronal PC12 cells of formaldehyde levels in post-mortem brains with metab-
show endoplasmic reticulum stress, decreased levels of the olite proles and analysis of oxidative stress markers are now
antioxidant proteins thioredoxin and paraoxonase 1 (Tang required to provide further experimental evidence for a direct
et al. 2011; Luo et al. 2012) and a decreased expression of contribution of formaldehyde in the pathology of neurode-
the anti-apoptotic protein Bcl-2, while the expression of pro- generative disorders.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


16 K. Tulpule and R. Dringen

Conflict of interest in rat nasal mucosal homogenates. Biochem. Pharmacol. 33, 1137
1142.
The authors have no conict of interest to declare. Cheng X. and Zhang X. (2007) Structural dynamics of protein lysine
methylation and demethylation. Mutat. Res. 618, 102115.
Cloos P. A., Christensen J., Agger K. and Helin K. (2008) Erasing the
References methyl mark: histone demethylases at the center of cellular
differentiation and disease. Genes Dev. 22, 11151140.
Ahmed A. E. and Ahmed M. W. (1978) Metabolism of dihalomethanes Cole S. P. C. and Deeley R. G. (2006) Transport of glutathione and
to formaldehyde and inorganic halide II. Studies on the mechanism glutathione conjugates by MRP1. Trends Pharmacol. Sci. 27, 438
of the reaction. Biochem. Pharmacol. 27, 20212025. 446.
Airas L., Mikkola J., Vainio J. M., Elovaara I. and Smith D. J. (2006) Cook R. J., Champion K. M. and Giometti C. S. (2001) Methanol
Elevated serum soluble vascular adhesion protein-1 (VAP-1) in toxicity and formate oxidation in NEUT2 mice. Arch. Biochem.
patients with active relapsing remitting multiple sclerosis. Biophys. 393, 192198.
J. Neuroimmunol. 177, 132135. Davis M. J., Iancu O. D., Acher F. C., Stewart B. M., Eiwaz M. A.,
Alnouti Y. and Klaassen C. D. (2008) Tissue distribution, ontogeny, and Duvoisin R. M. and Raber J. (2013) Role of mGluR4 in acquisition
regulation of aldehyde dehydrogenase (Aldh) enzymes mRNA by of fear learning and memory. Neuropharmacology 66, 365372.
prototypical microsomal enzyme inducers in mice. Toxicol. Sci. Debernardi R., Pierre K., Lengacher S., Magistretti P. J. and Pellerin L.
101, 5164. (2003) Cell-specic expression pattern of monocarboxylate
Anthony T. E. and Heintz N. (2007) The folate metabolic enzyme transporters in astrocytes and neurons observed in different
ALDH1L1 is restricted to the midline of the early CNS, suggesting mouse brain cortical cell cultures. J. Neurosci. Res. 73, 141155.
a role in human neural tube defects. J. Comp. Neurol. 500, 368383. Deeley R. G. and Cole S. P. (2006) Substrate recognition and transport
Aslan H., Songur A., Tunc A. T., Ozen O. A., Bas O., Yagmurca M., by multidrug resistance protein 1 (ABCC1). FEBS Lett. 580, 1103
Turgut M., Sarsilmaz M. and Kaplan S. (2006) Effects of 1111.
formaldehyde exposure on granule cell number and volume of Dhareshwar S. S. and Stella V. J. (2008) Your prodrug releases
dentate gyrus: a histopathological and stereological study. Brain formaldehyde: should you be concerned? No! J. Pharm. Sci. 97,
Res. 1122, 191200. 41844193.
Belkacemi A. and Ramassamy C. (2012) Time sequence of oxidative Dicker E. and Cederbaum A. I. (1984) Inhibition of the oxidation of
stress in the brain from transgenic mouse models of Alzheimers acetaldehyde and formaldehyde by hepatocytes and mitochondria
disease related to the amyloid-b cascade. Free Radic. Biol. Med. by crotonaldehyde. Arch. Biochem. Biophys. 234, 187196.
52, 593600. Dringen R. and Hamprecht B. (1998) Glutathione restoration as indicator
Bennani-Baiti I. M., Machado I., Llombart-Bosch A. and Kovar H. for cellular metabolism of astroglial cells. Dev. Neurosci. 20, 401
(2012) Lysine-specic demethylase 1 (LSD1/KDM1A/AOF2/ 407.
BHC110) is expressed and is an epigenetic drug target in Dringen R., Kranich O. and Hamprecht B. (1997) The c-glutamyl
chondrosarcoma, Ewings sarcoma, osteosarcoma, and transpeptidase inhibitor acivicin preserves glutathione released by
rhabdomyosarcoma. Hum. Pathol. 43, 13001307. astroglial cells in culture. Neurochem. Res. 22, 727733.
Berkich D. A., Ola M. S., Cole J., Sweatt A. J., Hutson S. M. and Dringen R., Hoepken H. H., Minich T. and Ruedig C. (2007) Pentose
LaNoue K. F. (2007) Mitochondrial transport proteins of the brain. phosphate pathway and NADPH metabolism, in Handbook of
J. Neurosci. Res. 85, 33673377. Neurochemistry and Molecular Neurobiology (Dienel G. and
Bola~
nos J. P., Almeida A. and Moncada S. (2010) Glycolysis: a Gibson G. S., eds), pp. 4162. Neural Energy Utilization, Springer,
bioenergetic or a survival pathway?. Trends Biochem. Sci. 35, 145 Heidelberg.
149. Fernandez-Fernandez S., Almeida A. and Bola~nos J. P. (2012)
Boomsma F., de Kam P. J., Tjeerdsma G., van den Meiracker A. H. and Antioxidant and bioenergetic coupling between neurons and
van Veldhuisen D. J. (2000) Plasma semicarbazide-sensitive amine astrocytes. Biochem. J. 443, 312.
oxidase (SSAO) is an independent prognostic marker for mortality Ferrer I., Lizcano J. M., Hernandez M. and Unzeta M. (2002)
in chronic heart failure. Eur. Heart J. 21, 18591863. Overexpression of semicarbazide sensitive amine oxidase in the
Boomsma F., Hut H., Bagghoe U., van der Houwen A. and van den cerebral blood vessels in patients with Alzheimers disease and
Meiracker A. (2005) Semicarbazide-sensitive amine oxidase cerebral autosomal dominant arteriopathy with subcortical infarcts
(SSAO): from cell to circulation. Med. Sci. Monit. 11, RA122 and leukoencephalopathy. Neurosci. Lett. 321, 2124.
RA126. Forneris F., Battaglioli E., Mattevi A. and Binda C. (2009) New roles of
Boumezbeur F., Mason G. F., de Graaf R. A., Behar K. L., Cline G. W., avoproteins in molecular cell biology: histone demethylase LSD1
Shulman G. I., Rothman D. L. and Petersen K. F. (2010) Altered and chromatin. FEBS J. 276, 43044312.
brain mitochondrial metabolism in healthy aging as assessed by Fountoulakis M., Gulesserian T. and Lubec G. (2003) Overexpression of
in vivo magnetic resonance spectroscopy. J. Cereb. Blood Flow C1-tetrahydrofolate synthase in fetal Down syndrome brain.
Metab. 30, 211221. J. Neural Transm. Suppl. 67, 8593.
Burg D., Wielinga P., Zelcer N., Saeki T., Mulder G. J. and Borst P. Franks S. J. (2005) A mathematical model for the absorption and
(2002) Inhibition of the multidrug resistance protein 1 (MRP1) by metabolism of formaldehyde vapour by humans. Toxicol. Appl.
peptidomimetic glutathione-conjugate analogs. Mol. Pharmacol. Pharmacol. 206, 309320.
62, 11601166. Friedenson B. (2011) A common environmental carcinogen unduly
Cahoy J. D., Emery B., Kaushal A. et al. (2008) A transcriptome affects carriers of cancer mutations: carriers of genetic mutations in
database for astrocytes, neurons, and oligodendrocytes: a new a specic protective response are more susceptible to an
resource for understanding brain development and function. environmental carcinogen. Med. Hypotheses 77, 791797.
J. Neurosci. 28, 264278. Fukuda T., Tokunaga A., Sakamoto R. and Yoshida N. (2011) Fbxl10/
Casanova-Schmitz M., David R. M. and Heck H. D. (1984) Oxidation of Kdm2b deciency accelerates neural progenitor cell death and
formaldehyde and acetaldehyde by NAD+-dependent dehydrogenases leads to exencephaly. Mol. Cell. Neurosci. 46, 614624.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 17

Galter D., Carmine A., Buervenich S., Duester G. and Olson L. (2003) polymorphism in the alcohol dehydrogenase 3 (ADH3) promoter.
Distribution of class I, III and IV alcohol dehydrogenase mRNAs Pharmacogenetics 11, 815824.
in the adult rat, mouse and human brain. Eur. J. Biochem. 270, Henzi H. (1984) Chronic methanol poisoning with the clinical and
13161326. pathologic-anatomical features of multiple sclerosis. Med.
Gennuso F., Fernetti C., Tirolo C. et al. (2004) Bilirubin protects Hypotheses 13, 6375.
astrocytes from its own toxicity by inducing up-regulation and Herrero-Mendez A., Almeida A., Fernandez E., Maestre C., Moncada S.
translocation of multidrug resistance-associated protein 1 (Mrp1). and Bola~nos J. P. (2009) The bioenergetic and antioxidant status of
Proc. Natl Acad. Sci. USA 101, 24702475. neurons is controlled by continuous degradation of a key glycolytic
Gr
onvall-Nordquist J. L., Backlund L. B., Garpenstrand H., Ekblom J., enzyme by APC/CCdh1. Nat. Cell Biol. 11, 747752.
Landin B., Yu P. H., Oreland L. and Rosenqvist U. (2001) Follow- van Horssen J., Schreibelt G., Drexhage J., Hazes T., Dijkshtra C. D.,
up of plasma semicarbazide-sensitive amine oxidase activity and van der Valk P. and de Vries H. E. (2008) Severe oxidative
retinopathy in type 2 diabetes mellitus. J. Diabetes Complications damage in multiple sclerosis lesions coincides with enhanced
15, 250256. antioxidant enzyme expression. Free Radic. Biol. Med. 45, 1729
de Groot A. C., Flyvholm M. A., Lensen G., Menne T. and Coenraads P. 1737.
J. (2009) Formaldehyde-releasers: Relationship to formaldehyde van Horssen J., Witte M. E., Schreibelt G. and de Vries H. E. (2011)
contact allergy. Contact allergy to formaldehyde and inventory of Radical changes in multiple sclerosis pathogenesis. Biochim.
formaldehyde-releasers. Contact Derm. 61, 6385. Biophys. Acta 1812, 141150.
de Groot A., White I. R., Flyvholm M. A., Lensen G. and Coenraads P. Hou H. and Yu H. (2010) Structural insights into histone lysine
J. (2010) Formaldehyde-releasers in cosmetics: relationship to demethylation. Curr. Opin. Struct. Biol. 20, 739748.
formaldehyde contact allergy. Part 2. Patch test relationship Iborra F. J., Renau-Piqueras J., Portoles M., Boleda M. D., Guerri C. and
to formaldehyde contact allergy, experimental provocation Pares X. (1992) Immunocytochemical and biochemical
tests, amount of formaldehyde released, and assessment of risk demonstration of formaldehyde dehydrogenase (class III alcohol
to consumers allergic to formaldehyde. Contact Derm. 62, dehydrogenase) in the nucleus. J. Histochem. Cytochem. 40, 1865
1831. 1878.
Guo J. M., Liu A. J., Zang P. et al. (2013) ALDH2 protects against Izzo A. and Schneider R. (2010) Chatting histone modications in
stroke by clearing 4-HNE. Cell Res. 2013, 116. mammals. Brief. Funct. Genomics 9, 429443.
Gurel A., Coskun O., Armutcu F., Kante M. and Ozen O. A. (2005) Jalkanen S. and Salmi M. (2001) Cell surface monoamine oxidases:
Vitamin E against oxidative damage caused by formaldehyde in enzymes in search of a function. EMBO J. 20, 38933901.
frontal cortex and hippocampus: biochemical and histological Jalkanen S. and Salmi M. (2008) VAP-1 and CD73, endothelial cell
studies. J. Chem. Neuroanat. 29, 173178. surface enzymes in leukocyte extravasation. Arterioscler. Thromb.
Haider L., Fischer M. T., Frischer J. M., Bauer J., Hoftberger R., Botond Vasc. Biol. 28, 1826.
G., Esterbauer H., Binder C. J., Witztum J. L. and Lassmann H. Janaky R., Cruz-Aguado R., Oja S. S. and Shaw C. A. (2007)
(2011) Oxidative damage in multiple sclerosis lesions. Brain 134, Glutathione in the nervous system: roles in neural function and
19141924. health and implications for neurological disease, in Handbook of
Halim N. D., McFate T., Mohyeldin A. et al. (2010) Phosphorylation Neurochemistry (Oja S. S., Schousboe A. and Saransaari P., eds),
status of pyruvate dehydrogenase distinguishes metabolic pp. 347399. Amino Acids and Peptides in the Nervous System,
phenotypes of cultured rat brain astrocytes and neurons. Glia 58, Springer, Heidelberg.
11681176. Julia P., Farres J. and Pares X. (1987) Characterization of three
Harris C., Wang S. W., Lauchu J. J. and Hansen J. M. (2003) Methanol isoenzymes of rat alcohol-dehydrogenase - tissue distribution
metabolism and embryotoxicity in rat and mouse conceptuses: and physical and enzymatic-properties. Eur. J. Biochem. 162,
comparisons of alcohol dehydrogenase (ADH1), formaldehyde 179189.
dehydrogenase (ADH3), and catalase. Reprod. Toxicol. 17, 349 Kahl P., Gullotti L., Heukamp L. C. et al. (2006) Androgen receptor
357. coactivators lysine-specic histone demethylase 1 and four and a
Hayami S., Yoshimatsu M., Veerakumarasivam A. et al. (2010) half LIM domain protein 2 predict risk of prostate cancer
Overexpression of the JmjC histone demethylase KDM5B in recurrence. Cancer Res. 66, 1134111347.
human carcinogenesis: involvement in the proliferation of cancer Kapur B. M., Vandenbroucke A. C., Adamchik Y., Lehotay D. C. and
cells through the E2F/RB pathway. Mol. Cancer 9, 59. Carlen P. L. (2007) Formic acid, a novel metabolite of chronic
Hayami S., Kelly J. D., Cho H. S. et al. (2011) Overexpression of LSD1 ethanol abuse, causes neurotoxicity, which is prevented by folic
contributes to human carcinogenesis through chromatin regulation acid. Alcohol. Clin. Exp. Res. 31, 21142120.
in various cancers. Int. J. Cancer 128, 574586. Karadi I., Meszaros Z., Csanyi A., Szombathy T., Hosszufalusi N.,
He R., Lu J. and Miao J. (2010) Formaldehyde stress. Sci. China Life Romics L. and Magyar K. (2002) Serum semicarbazide-sensitive
Sci. 53, 13991404. amine oxidase (SSAO) activity is an independent marker of carotid
Heck H. D. and Casanova M. (2004) The implausibility of leukemia atherosclerosis. Clin. Chim. Acta 323, 139146.
induction by formaldehyde: a critical review of the biological Kato S., Burke P. J., Koch T. H. and Bierbaum V. M. (2001)
evidence on distant-site toxicity. Regul. Toxicol. Pharmacol. 40, Formaldehyde in human cancer cells: detection by preconcentration-
92106. chemical ionization mass spectrometry. Anal. Chem. 73, 2992
Heck H. D., Casanova-Schmitz M., Dodd P. B., Schachter E. N., Witek 2997.
T. J. and Tosun T. (1985) Formaldehyde (CH2O) concentrations in Keppler D. (2011) Multidrug resistance proteins (MRPs, ABCCs):
the blood of humans and Fischer-344 rats exposed to CH2O under importance for pathophysiology and drug therapy. Handb. Exp.
controlled conditions. Am. Ind. Hyg. Assoc. J. 46, 13. Pharmacol. 201, 299323.
Heck H. D., Casanova M. and Starr T. B. (1990) Formaldehyde toxicity - Khokhlov A. P., Zavalishin I. A., Savchenko I. N. and Dziuba A. N.
new understanding. Crit. Rev. Toxicol. 20, 397426. (1989) Disorders of formaldehyde metabolism and its metabolic
Hedberg J. J., Backlund M., Stromberg P., Lonn S., Dahl M. L., precursors in patients with multiple sclerosis. Zh. Nevropatol.
Ingelman-Sundberg M. and Hoog J. O. (2001) Functional Psikhiatr. Im. S. S. Korsakova 89, 4548.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


18 K. Tulpule and R. Dringen

Kiernan J. A. (2000) Formaldehyde, formalin, paraformaldehyde and MacAllister S. L., Choi J., Dedina L. and OBrien P. J. (2011) Metabolic
glutaraldehyde: what they are and what they do. Microsc. Today 1, mechanisms of methanol/formaldehyde in isolated rat hepatocytes:
812. Carbonyl-metabolizing enzymes versus oxidative stress. Chem.
Kiernan M. C., Vucic S., Cheah B. C., Turner M. R., Eisen A., Hardiman Biol. Interact. 191, 308314.
O., Burrell J. R. and Zoing M. C. (2011) Amyotrophic lateral MacFarlane A. J., Perry C. A., Girnary H. H., Gao D., Allen R. H.,
sclerosis. Lancet 377, 942955. Stabler S. P., Shane B. and Stover P. J. (2009) Mthfd1 is an
Kilburn K. H., Seidman B. C. and Warshaw R. (1985a) Neurobehavioral essential gene in mice and alters biomarkers of impaired one-
and respiratory symptoms of formaldehyde and xylene exposure in carbon metabolism. J. Biol. Chem. 284, 15331539.
histology technicians. Arch. Environ. Health 40, 229233. Mahad D., Ziabreva I., Lassmann H. and Turnbull D. (2008)
Kilburn K. H., Warshaw R., Boylen C. T., Johnson S. J., Seidman B., Mitochondrial defects in acute multiple sclerosis lesions. Brain
Sinclair R. and Takaro T. Jr (1985b) Pulmonary and 131, 17221735.
neurobehavioral effects of formaldehyde exposure. Arch. Malek F. A., Moritz K. U. and Fanghanel J. (2003) A study on specic
Environ. Health 40, 254260. behavioral effects of formaldehyde in the rat. J. Exp. Anim. Sci. 42,
Krupenko S. A. (2009) FDH: an aldehyde dehydrogenase fusion enzyme 160170.
in folate metabolism. Chem. Biol. Interact. 178, 8493. del Mar Hernandez M., Esteban M., Szabo P., Boada M. and Unzeta M.
Krupenko N. I., Dubard M. E., Strickland K. C., Moxley K. M., Oleinik (2005) Human plasma semicarbazide sensitive amine oxidase
N. V. and Krupenko S. A. (2010) ALDH1L2 is the mitochondrial (SSAO), b-amyloid protein and aging. Neurosci. Lett. 384, 183187.
homolog of 10-formyltetrahydrofolate dehydrogenase. J. Biol. Martinez S. E., Vaglenova J., Sabria J., Martinez M. C., Farres J. and
Chem. 285, 2305623063. Pares X. (2001) Distribution of alcohol dehydrogenase mRNA in
Kurkijarvi R., Yegutkin G. G., Gunson B. K., Jalkanen S., Salmi M. and the rat central nervous system - consequences for brain ethanol and
Adams D. H. (2000) Circulating soluble vascular adhesion protein retinoid metabolism. Eur. J. Biochem. 268, 50455056.
1 accounts for the increased serum monoamine oxidase activity in Mason M. J., Mattsson K., Pasternack M., Voipio J. and Kaila K. (1990)
chronic liver disease. Gastroenterology 119, 10961103. Postsynaptic fall in intracellular pH and increase in surface pH
Laitinen J., Makela M., Mikkola J. and Huttu I. (2010) Fire ghting caused by efux of formate and acetate anions through GABA-
trainers exposure to carcinogenic agents in smoke diving gated channels in craysh muscle-bers. Neuroscience 34, 359
simulators. Toxicol. Lett. 192, 6165. 368.
Lee E. S., Chen H., Hardman C., Simm A. and Charlton C. (2008) Meinerz D. F., Comprasi B., Allebrandt J. et al. (2013) Sub-acute
Excessive S-adenosyl-L-methionine-dependent methylation administration of (S)-dimethyl 2-(3-(phenyltellanyl) propanamido)
increases levels of methanol, formaldehyde and formic acid in rat succinate induces toxicity and oxidative stress in mice: unexpected
brain striatal homogenates: possible role in S-adenosyl- effects of N-acetylcysteine. Springerplus 2, 182.
L-methionine-induced Parkinsons disease-like disorders. Life Meszaros Z., Szombathy T., Raimondi L., Karadi I., Romics L. and
Sci. 83, 821827. Magyar K. (1999) Elevated serum semicarbazide-sensitive amine
Lee M., Schwab C. and McGeer P. L. (2011) Astrocytes are GABAergic oxidase activity in non-insulin-dependent diabetes mellitus:
cells that modulate microglial activity. Glia 59, 152165. correlation with body mass index and serum triglyceride.
Leuner K., Muller W. E. and Reichert A. S. (2012) From mitochondrial Metabolism 48, 113117.
dysfunction to amyloid beta formation: novel insights into the Metz B., Kersten G. F., Hoogerhout P. et al. (2004) Identication of
pathogenesis of Alzheimers disease. Mol. Neurobiol. 46, 186 formaldehyde-induced modications in proteins: reactions with
193. model peptides. J. Biol. Chem. 279, 62356243.
Lewerenz J., Dargusch R. and Maher P. (2010) Lactacidosis modulates Metz B., Kersten G. F., Baart G. J., de Jong A., Meiring H., ten Hove J.,
glutathione metabolism and oxidative glutamate toxicity. van Steenbergen M. J., Hennink W. E., Crommelin D. J. and
J. Neurochem. 113, 502514. Jiskoot W. (2006) Identication of formaldehyde-induced
Li F., Liu X., Su Z. and Sun R. (2011) Acidosis leads to brain modications in proteins: reactions with insulin. Bioconjug.
dysfunctions through impairing cortical GABAergic neurons. Chem. 17, 815822.
Biochem. Biophys. Res. Commun. 410, 775779. Miao J. and He R. (2012) Chronic formaldehyde-mediated impairments
Li F., Gong Q., Dong H. and Shi J. (2012) Resveratrol, a neuroprotective and age-related dementia, in Neurodegeneration (Martin L. M. and
supplement for Alzheimers disease. Curr. Pharm. Des. 18, 2733. Loh S. H. Y., eds), pp. 5976. InTech, doi: 10.5772/34949.
Lim S., Janzer A., Becker A., Zimmer A., Schule R., Buettner R. and Monte W. C. (2010) Methanol: a chemical Trojan horse as the root of the
Kirfel J. (2010) Lysine-specic demethylase 1 (LSD1) is highly inscrutable U. Med. Hypotheses 74, 493496.
expressed in ER-negative breast cancers and a biomarker Moschen I., Broer A., Galic S., Lang F. and Broer S. (2012) Signicance
predicting aggressive biology. Carcinogenesis 31, 512520. of short chain fatty acid transport by members of the
Lu S. C. (2013) Glutathione synthesis. Biochim. Biophys. Acta 1830, monocarboxylate transporter family (MCT). Neurochem. Res. 37,
31433153. 25622568.
Lu Z., Li C. M., Qiao Y., Yan Y. and Yang X. (2008) Effect of inhaled Mukherjee S. and Manahan-Vaughan D. (2013) Role of metabotropic
formaldehyde on learning and memory of mice. Indoor Air 18, 77 glutamate receptors in persistent forms of hippocampal plasticity
83. and learning. Neuropharmacology 66, 6581.
Lu J., Li C., Su T., Liu Y. and He R. (2013) Formaldehyde induces Nazarian A., Hermannsson B. J., Muller J., Zurakowski D. and Snyder
hyperphosphorylation and polymerization of Tau protein both B. D. (2009) Effects of tissue preservation on murine bone
in vitro and in vivo. Biochim. Biophys. Acta 1830, 41024116. mechanical properties. J. Biomech. 42, 8286.
Luo F. C., Zhou J., Lv T., Qi L., Wang S. D., Nakamura H., Yodoi J. and Neves A., Costalat R. and Pellerin L. (2012) Determinants of brain
Bai J. (2012) Induction of endoplasmic reticulum stress and the cell metabolic phenotypes and energy substrate utilization unraveled
modulation of thioredoxin-1 in formaldehyde-induced with a modeling approach. PLoS Comput. Biol. 8, e1002686.
neurotoxicity. Neurotoxicology 33, 290298. Neymeyer V., Tephly T. R. and Miller M. W. (1997) Folate and 10-
Lushchak V. I. (2012) Glutathione homeostasis and functions: potential formyltetrahydrofolate dehydrogenase (FDH) expression in the
targets for medical interventions. J. Amino Acids 2012, 736837. central nervous system of the mature rat. Brain Res. 766, 195204.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 19

Nicholls P. (1975) Formate as an inhibitor of cytochrome c oxidase. layer in hippocampus and hemisphere in the rat: a stereological
Biochem. Biophys. Res. Commun. 67, 610616. study. Brain Res. 1145, 157167.
Nishimura M. and Naito S. (2006) Tissue-specic mRNA expression Sasseville D. (2004) Hypersensitivity to preservatives. Dermatol. Ther.
proles of human phase I metabolizing enzymes except for 17, 251263.
cytochrome P450 and phase II metabolizing enzymes. Drug Schad A., Fahimi H. D., Volkl A. and Baumgart E. (2003) Expression
Metab. Pharmacokinet. 21, 357374. of catalase mRNA and protein in adult rat brain: detection
Obata T. (2006) Diabetes and semicarbazide-sensitive amine oxidase by nonradioactive in situ hybridization with signal amplication
(SSAO) activity: a review. Life Sci. 79, 417422. by catalyzed reporter deposition (ISH-CARD) and
Obata T. and Yamanaka Y. (2000) Evidence for existence of immunohistochemistry (IHC)/immunouorescence (IF). J.
immobilization stress-inducible semicarbazide-sensitive amine Histochem. Cytochem. 51, 751760.
oxidase inhibitor in rat brain cytosol. Neurosci. Lett. 296, 5860. Scheiber I. F. and Dringen R. (2011) Copper accelerates glycolytic ux
Oldham M. C., Konopka G., Iwamoto K., Langfelder P., Kato T., in cultured astrocytes. Neurochem. Res. 36, 894903.
Horvath S. and Geschwind D. (2008) Functional organization of Schildhaus H. U., Riegel R., Hartmann W. et al. (2011) Lysine-specic
the transcriptome in the human brain. Nat. Neurosci. 11, 1271 demethylase 1 is highly expressed in solitary brous tumors,
1282. synovial sarcomas, rhabdomyosarcomas, desmoplastic small round
Olsen R. W. and Sieghart W. (2009) GABAA receptors: subtypes provide cell tumors, and malignant peripheral nerve sheath tumors. Hum.
diversity of function and pharmacology. Neuropharmacology 56, Pathol. 42, 16671675.
141148. Schmidt M. M. and Dringen R. (2012) GSH synthesis and metabolism,
OSullivan J., Unzeta M., Healy J., OSullivan M. I., Davey G. and in Advances in Neurobiology (Gruetter R. and Choi I. Y., eds), pp.
Tipton K. F. (2004) Semicarbazide-sensitive amine oxidases: 10291050. Neural Metabolism in vivo, Springer, New York.
enzymes with quite a lot to do. Neurotoxicology 25, 303315. Schulte J. H., Lim S., Schramm A. et al. (2009) Lysine-specic
Oyama Y., Sakai H., Arata T., Okano Y., Akaike N., Sakai K. and Noda demethylase 1 is strongly expressed in poorly differentiated
K. (2002) Cytotoxic effects of methanol, formaldehyde, and neuroblastoma: implications for therapy. Cancer Res. 69, 2065
formate on dissociated rat thymocytes: a possibility of aspartame 2071.
toxicity. Cell Biol. Toxicol. 18, 4350. Schwyzer R. U. and Henzi H. (1983) Multiple sclerosis: plaques caused
Paling D., Golay X., Wheeler-Kingshott C., Kapoor R. and Miller D. by 2-step demyelination? Med. Hypotheses 12, 129142.
(2011) Energy failure in multiple sclerosis and its investigation Skrzydlewska E. (2003) Toxicological and metabolic consequences of
using MR techniques. J. Neurol. 258, 21132127. methanol poisoning. Toxicol. Mech. Methods 13, 277293.
Parnetti L., Reboldi G. P. and Gallai V. (2000) Cerebrospinal uid Smith D. J. and Vainio P. J. (2007) Targeting vascular adhesion protein-
pyruvate levels in Alzheimers disease and vascular dementia. 1 to treat autoimmune and inammatory diseases. Ann. N. Y. Acad.
Neurology 54, 735737. Sci. 1110, 382388.
Pauwels P. J., Opperdoes F. R. and Trouet A. (1985) Effects of Sohal R. S. and Orr W. C. (2012) The redox stress hypothesis of aging.
antimycin, glucose deprivation, and serum on cultures of neurons, Free Radic. Biol. Med. 52, 539555.
astrocytes, and neuroblastoma cells. J. Neurochem. 44, 143148. Song M. S., Baker G. B., Dursun S. M. and Todd K. G. (2010) The
Pitten F. A., Kramer A., Herrmann K., Breme I. and Koch S. (2000) antidepressant phenelzine protects neurons and astrocytes
Formaldehyde neurotoxicity in animal experiments. Pathol. Res. against formaldehyde-induced toxicity. J. Neurochem. 114,
Pract. 196, 193198. 14051413.
Prasannan P., Pike S., Peng K., Shane B. and Appling D. R. (2003) Songur A., Sarsilmaz M., Ozen O. A., Sahin S., Koken R., Zararsiz I. and
Human mitochondrial C1-tetrahydrofolate synthase: gene structure, Ilhan N. (2008) The effects of inhaled formaldehyde on oxidant and
tissue distribution of the mRNA, and immunolocalization in antioxidant systems of rat cerebellum during the postnatal
Chinese hamster ovary cells. J. Biol. Chem. 278, 4317843187. development process. Toxicol. Mech. Methods 18, 569574.
Regan R. F. and Guo Y. P. (1999a) Extracellular reduced glutathione Songur A., Ozen O. A. and Sarsilmaz M. (2010) The toxic effects of
increases neuronal vulnerability to combined chemical hypoxia and formaldehyde on the nervous system. Rev. Environ. Contam.
glucose deprivation. Brain Res. 817, 145150. Toxicol. 203, 105118.
Regan R. F. and Guo Y. P. (1999b) Potentiation of excitotoxic injury by Sorg B. A., Swindell S. and Tschirgi M. L. (2004) Repeated low level
high concentrations of extracellular reduced glutathione. formaldehyde exposure produces enhanced fear conditioning to
Neuroscience 91, 463470. odor in male, but not female, rats. Brain Res. 1008, 1119.
Richard T., Pawlus A. D., Iglesias M. L., Pedrot E., Waffo-Teguo P., Staab C. A., Alander J., Morgenstern R., Grafstrom R. C. and Hoog J. O.
Merillon J. M. and Monti J. P. (2011) Neuroprotective properties (2009) The janus face of alcohol dehydrogenase 3. Chem. Biol.
of resveratrol and derivatives. Ann. N. Y. Acad. Sci. 1215, 103 Interact. 178, 2935.
108. Staub F., Peters J., Kempski O., Schneider G. H., Schurer L.
Rose C. F. (2010) Increase brain lactate in hepatic encephalopathy: cause and Baethmann A. (1993) Swelling of glial cells in lactacidosis
or consequence? Neurochem. Int. 57, 389394. and by glutamate: signicance of Cltransport. Brain Res. 610, 69
Ross J. M., Oberg J., Brene S. et al. (2010) High brain lactate is a 74.
hallmark of aging and caused by a shift in the lactate Steele M. L. and Robinson S. R. (2012) Reactive astrocytes give neurons
dehydrogenase A/B ratio. Proc. Natl Acad. Sci. USA 107, less support: implications for Alzheimers disease. Neurobiol.
2008720092. Aging 33, 423.e1423.e13.
Salthammer T., Mentese S. and Marutzky R. (2010) Formaldehyde in the Steullet P., Cabungcal J. H., Kulak A., Kraftsik R., Chen Y., Dalton T.
indoor environment. Chem. Rev. 110, 25362572. P., Cuenod M. and Do K. Q. (2010) Redox dysregulation affects
Sarsilmaz M., Kaplan S., Songur A., Colakoglu S., Aslan H., Tunc A. T., the ventral but not dorsal hippocampus: impairment of
Ozen Q. A., Turgut M. and Bas O. (2007) Effects of postnatal parvalbumin neurons, gamma oscillations, and related behaviors.
formaldehyde exposure on pyramidal cell number, volume of cell J. Neurosci. 30, 25472558.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


20 K. Tulpule and R. Dringen

Stewart M. J., Malek K. and Crabb D. W. (1996) Distribution of Turkoglu A. O., Sarsilmaz M., Ogeturk M., Kus I. and Songur A. (2008)
messenger RNAs for aldehyde dehydrogenase 1, aldehyde Benecial effects of caffeic acid phenethyl ester on formaldehyde-
dehydrogenase 2, and aldehyde dehydrogenase 5 in human induced learning and memory disabilities: a labyrinth test
tissues. J. Investig. Med. 44, 4246. performance study. Erciyes Med. J. 30, 211217.
Sullivan P. G. and Brown M. R. (2005) Mitochondrial aging and Tyihak E. and Kiraly-Veghely Z. (2008) Interaction of trans-resveratrol
dysfunction in Alzheimers disease. Prog. Neuropsychopharmacol. with endogenous formaldehyde as one basis of its diverse
Biol. Psychiatry 29, 407410. benecial biological effects. Bull. de IOIV 81, 6574.
Suzuki A., Stern S. A., Bozdagi O., Huntley G. W., Walker R. H., Unzeta M., Sole M., Boada M. and Hernandez M. (2007)
Magistretti P. J. and Alberini C. M. (2011) Astrocyte-neuron Semicarbazide-sensitive amine oxidase (SSAO) and its possible
lactate transport is required for long-term memory formation. Cell contribution to vascular damage in Alzheimers disease. J. Neural
144, 810823. Transm. 114, 857862.
Tang X., Bai Y., Duong A., Smith M. T., Li L. and Zhang L. (2009) Uotila L. (1981) Thioesters of glutathione. Methods Enzymol. 77, 424
Formaldehyde in China: production, consumption, exposure levels, 430.
and health effects. Environ. Int. 35, 12101224. Usanmaz S. E., Akarsu E. S. and Vural N. (2002) Neurotoxic effects of
Tang X. Q., Ren Y. K., Chen R. Q. et al. (2011) Formaldehyde induces acute and subacute formaldehyde exposures in mice. Environ.
neurotoxicity to PC12 cells involving inhibition of paraoxonase-1 Toxicol. Pharmacol. 11, 93100.
expression and activity. Clin. Exp. Pharmacol. Physiol. 38, 208 Valente T., Gella A., Sole M., Durany N. and Unzeta M. (2012)
214. Immunohistochemical study of semicarbazide-sensitive amine
Tang X. Q., Ren Y. N., Zhou C. F. et al. (2012) Hydrogen sulde oxidase/vascular adhesion protein-1 in the hippocampal
prevents formaldehyde-induced neurotoxicity to PC12 cells by vasculature: pathological synergy of Alzheimers disease and
attenuation of mitochondrial dysfunction and pro-apoptotic diabetes mellitus. J. Neurosci. Res. 90, 19891996.
potential. Neurochem. Int. 61, 1624. Velez-Fort M., Audinat E. and Angulo M. C. (2011) Central role of
Teng S., Beard K., Pourahmad J., Moridani M., Easson E., Poon R. and GABA in neuron-glia interactions. Neuroscientist 18, 237250.
OBrien P. J. (2001) The formaldehyde metabolic detoxication Wallace K. B., Eells J. T., Madeira V. M., Cortopassi G. and Jones D. P.
enzyme systems and molecular cytotoxic mechanism in isolated rat (1997) Mitochondria-mediated cell injury. Symposium overview.
hepatocytes. Chem. Biol. Interact. 130132, 285296. Fundam. Appl. Toxicol. 38, 2337.
Tephly T. R. (1991) The toxicity of methanol. Life Sci. 48, 1031 Wang R. S., Nakajima T., Kawamoto T. and Honma T. (2002) Effects of
1041. aldehyde dehydrogenase-2 genetic polymorphisms on metabolism
Thigpen A. E., West M. G. and Appling D. R. (1990) Rat C1- of structurally different aldehydes in human liver. Drug Metab.
tetrahydrofolate synthase. cDNA isolation, tissue-specic levels of Dispos. 30, 6973.
the mRNA, and expression of the protein in yeast. J. Biol. Chem. Wang E. Y., Gao H. F., Salter-Cid L., Zhang J., Huang L., Podar E. M.,
265, 79077913. Miller A., Zhao J. J., ORourke A. and Linnik M. D. (2006)
Thompson C. M., Ceder R. and Grafstrom R. C. (2010) Formaldehyde Design, synthesis and biological evaluation of semicarbazide-
dehydrogenase: beyond phase I metabolism. Toxicol. Lett. 193, sensitive amine oxidase (SSAO) inhibitors with anti-inammatory
13. activity. J. Med. Chem. 49, 21662173.
Tibbetts A. S. and Appling D. R. (2010) Compartmentalization of Weisskopf M. G., Morozova N., OReilly E. J., McCullough M. L.,
mammalian folate-mediated one-carbon metabolism. Annu. Rev. Calle E. E., Thun M. J. and Ascherio A. (2009) Prospective study
Nutr. 30, 5781. of chemical exposures and amyotrophic lateral sclerosis. J. Neurol.
Tong Z., Luo W., Wang Y. et al. (2010) Tumor tissue-derived Neurosurg. Psychiatry 80, 558561.
formaldehyde and acidic microenvironment synergistically induce Wolf S. S., Patchev V. K. and Obendorf M. (2007) A novel variant of the
bone cancer pain. PLoS ONE 5, e10234. putative demethylase gene, s-JMJD1C, is a coactivator of the AR.
Tong Z., Zhang J., Luo W. et al. (2011) Urine formaldehyde level is Arch. Biochem. Biophys. 460, 5666.
inversely correlated to mini mental state examination scores in Xiang Y., Zhu Z., Han G. et al. (2007) JARID1B is a histone H3 lysine
senile dementia. Neurobiol. Aging 32, 3141. 4 demethylase up-regulated in prostate cancer. Proc. Natl Acad.
Tong Z., Han C., Luo W., Wang X., Li H., Luo H., Zhou J., Qi J. and He Sci. USA 104, 1922619231.
R. (2013a) Accumulated hippocampal formaldehyde induces age- Yin J. and Zhang J. (2011) Multidrug resistance-associated protein 1
dependent memory decline. Age (Dordr.) 35, 583596. (MRP1/ABCC1) polymorphism: from discovery to clinical
Tong Z., Han C., Luo W. et al. (2013b) Aging-associated excess application. J. Cent. South Univ. 36, 927938.
formaldehyde leads to spatial memory decits. Sci. Rep. 3, 1807. Yu P. H. (2001) Involvement of cerebrovascular semicarbazide-sensitive
Tulpule K. and Dringen R. (2011) Formaldehyde stimulates Mrp1- amine oxidase in the pathogenesis of Alzheimers disease and
mediated glutathione deprivation of cultured astrocytes. vascular dementia. Med. Hypotheses 57, 175179.
J. Neurochem. 116, 626635. Yu P. H., Wright S., Fan E. H., Lun Z. R. and Gubisne-Harberle D.
Tulpule K. and Dringen R. (2012) Formate generated by cellular (2003) Physiological and pathological implications of
oxidation of formaldehyde accelerates the glycolytic ux in semicarbazide-sensitive amine oxidase. Biochim. Biophys. Acta
cultured astrocytes. Glia 60, 582593. 1647, 193199.
Tulpule K., Schmidt M. M., Boecker K., Goldbaum O., Richter- Zararsiz I., Kus I., Akpolat N., Songur A., Ogeturk M. and Sarsilmaz M.
Landsberg C. and Dringen R. (2012) Formaldehyde induces rapid (2006) Protective effects of x-3 essential fatty acids against
glutathione export from viable oligodendroglial OLN-93 cells. formaldehyde-induced neuronal damage in prefrontal cortex of
Neurochem. Int. 61, 13021313. rats. Cell Biochem. Funct. 24, 237244.
Tulpule K., Hohnholt M. C. and Dringen R. (2013) Formaldehyde Zararsiz I., Kus I., Ogeturk M., Akpolat N., Kose E., Meydan S. and
metabolism and formaldehyde-induced stimulation of lactate Sarsilmaz M. (2007) Melatonin prevents formaldehyde-induced
production and glutathione export in cultured neurons. neurotoxicity in prefrontal cortex of rats: An immunohistochemical
J. Neurochem. 125, 260272. and biochemical study. Cell Biochem. Funct. 25, 413418.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21


Formaldehyde in brain 21

Zararsiz I., Meydan S., Sarsilmaz M., Songur A., Ozen O. A. and Sogut pyramidal neurons. Biochem. Biophys. Res. Commun. 415, 224
S. (2011) Protective effects of omega-3 essential fatty acids against 228.
formaldehyde-induced cerebellar damage in rats. Toxicol. Ind. Zibetti C., Adamo A., Binda C., Forneris F., Toffolo E., Verpelli C.,
Health 27, 489495. Ginelli E., Mattevi A., Sala C. and Battaglioli E. (2010) Alternative
Zhang Y. Z., Zhang Q. H., Ye H., Zhang Y., Luo Y. M., Ji X. M. and Su splicing of the histone demethylase LSD1/KDM1 contributes to the
Y. Y. (2010) Distribution of lysine-specic demethylase 1 in the modulation of neurite morphogenesis in the mammalian nervous
brain of rat and its response in transient global cerebral ischemia. system. J. Neurosci. 30, 25212532.
Neurosci. Res. 68, 6672. Zimatkin S. M. and Lindros K. O. (1996) Distribution of catalase in rat
Zhao H., Cai Y., Yang Z., He D. and Shen B. (2011) Acidosis leads brain: aminergic neurons as possible targets for ethanol effects.
to neurological disorders through overexciting cortical Alcohol Alcohol. 31, 167174.

2013 International Society for Neurochemistry, J. Neurochem. (2013) 127, 7--21

Potrebbero piacerti anche