You are on page 1of 12

bs_bs_banner

Addiction Biology
ORIGINAL ARTICLE doi:10.1111/adb.12097

Key role of salsolinol in ethanol actions on


dopamine neuronal activity of the posterior
ventral tegmental area

Miriam Melis1, Ezio Carboni1,3,4, Pierluigi Caboni2 & Elio Acquas2,3,4


Departments of Biomedical Sciences1and Life and Environmental Sciences2, Centre of Excellence on Neurobiology of Addiction3 and INN—National Institute of
Neuroscience4, University of Cagliari, Cagliari, Italy

ABSTRACT

Ethanol excites dopamine (DA) neurons in the posterior ventral tegmental area (pVTA). This effect is responsible for
ethanol’s motivational properties and may contribute to alcoholism. Evidence indicates that catalase-mediated con-
version of ethanol into acetaldehyde in pVTA plays a critical role in this effect. Acetaldehyde, in the presence of DA,
condensates with it to generate salsolinol. Salsolinol, when administered in pVTA, excites pVTA DA cells, elicits DA
transmission in nucleus accumbens and sustains its self-administration in pVTA. Here we show, by using ex vivo
electrophysiology, that ethanol and acetaldehyde, but not salsolinol, failed to stimulate pVTA DA cell activity in mice
administered α-methyl-p-tyrosine, a DA biosynthesis inhibitor that reduces somatodendritic DA release. This effect was
specific for ethanol and acetaldehyde since morphine, similarly to salsolinol, was able to excite pVTA DA cells in
α-methyl-p-tyrosine-treated mice. However, when DA was bath applied in slices from α-methyl-p-tyrosine-treated mice,
ethanol-induced excitation of pVTA DA neurons was restored. This effect requires ethanol oxidation into acetaldehyde
given that, when H2O2-catalase system was impaired by either 3-amino-1,2,4-triazole or in vivo administration of
α-lipoic acid, ethanol did not enhance DA cell activity. Finally, high performance liquid chromatography-tandem mass
spectrometry analysis of bath medium detected salsolinol only after co-application of ethanol and DA in α-methyl-p-
tyrosine-treated mice. These results demonstrate the relationship between ethanol and salsolinol effects on pVTA DA
neurons, help to untangle the mechanism(s) of action of ethanol in this area and contribute to an exciting research
avenue prosperous of theoretical and practical consequences.

Keywords Acetaldehyde, catalase, dopamine, ethanol, salsolinol, pVTA.

Correspondence to: Elio Acquas, Department of Life and Environmental Sciences, University of Cagliari, Via Ospedale, 72—I-09124 Cagliari, Italy.
E-mail: acquas@unica.it

INTRODUCTION transmission in the nucleus accumbens shell (Howard


et al. 2008) and to affect motivation (Di Chiara 1997;
Ethanol is the main pharmacological ingredient of alco- Koob et al. 1998). These properties, in agreement with
holic drinks and one of the most known psychoactive clinical evidence that ethanol promotes DA release in
chemicals worldwide. Ethanol is also renowned for its the ventral striatum and exerts psychostimulant effects
liability to cause alcoholism, a condition that still plagues (Boileau et al. 2003), represent, in turn, potential
human civilization carrying high individual, family and grounds for development of alcoholism (Söderpalm &
societal costs, concerning and pertaining to a large scien- Ericson 2013).
tific community from bench- to bed-side. Peripheral ethanol metabolism mainly occurs by
The mesolimbic dopamine (DA) system is involved in alcohol dehydrogenase activity. Aversive peripheral
psychopharmacological effects of ethanol as a conse- effects (Escrig et al. 2012) and somatic symptoms of
quence of ethanol’s ability to excite DA neurons in poste- flushing reaction, but also centrally mediated effects
rior ventral tegmental area (pVTA) (Gessa et al. 1985; (Correa et al. 2012) are ascribed to acetaldehyde result-
Brodie, Shefner & Dunwiddie 1990; Foddai et al. 2004; ing from ethanol oxidation. In the brain, ethanol is
Okamoto, Harnett & Morikawa 2006), to stimulate DA converted into acetaldehyde by catalase (Cohen, Sinet &

© 2013 Society for the Study of Addiction Addiction Biology


2 Miriam Melis et al.

Heikkila 1980; Aragon, Spivak & Amit 1991; Aragon, by rats (Rodd et al. 2008) and, when microinjected into
Rogan & Amit 1992; Gill et al. 1992; Zimatkin et al. this brain area, increases locomotor activity (Hipólito
2006), whose inhibition prevents a number of ethanol- et al. 2010) and elicits DA release in nucleus accumbens
mediated behavioral responses ranging from voluntary shell (Hipólito et al. 2011; Deehan et al. 2013). Further-
consumption (Aragon & Amit 1992), to locomotion more, salsolinol stimulates firing rate of DA neurons ex
(Sanchis-Segura et al. 1999, 2005) and acquisition of vivo (Xie & Ye 2012; Xie et al. 2012).
conditioned place preference (Font, Miquel & Aragon Hence, provided that ethanol has to be converted into
2008). Indeed, acetaldehyde has long been suggested to acetaldehyde in order to excite DA neurons in pVTA
account for ethanol’s psychopharmacological actions (Melis et al. 2007), and that acetaldehyde in presence of
(Davis & Walsh 1970; Davis, Walsh & Yamanaka 1970; DA may condensate with it to produce salsolinol
Brown, Amit & Rockman 1979). However, this issue (Yamanaka et al. 1970; Chen et al. 2011), we hypoth-
remained controversial (Deitrich 2004) mainly because esized that salsolinol might be involved in ethanol
peripherally produced acetaldehyde displays high stimulatory action on pVTA DA cells. The present study
reactivity (Davis & Walsh 1970; Myers, Ng & Singer was designed to challenge this hypothesis by recording
1982) and conditional ability to cross blood brain barrier pVTA DA cell firing in acute brain slices ex vivo in which
(Eriksson & Sippel 1977; Tabakoff, Anderson & Ritzman somatodendritic DA release was reduced (Cheramy, Leviel
1976; but see also Correa et al. 2012 for a throughout & Glowinski 1981) by in vivo administration of α-methyl-
discussion on this issue). Nevertheless, since brain p-tyrosine (αMpT) or reserpine.
catalase-mediated ethanol metabolism was proven
(Aragon et al. 1991, 1992), pre-clinical studies have
MATERIALS AND METHODS
shown that acetaldehyde is a neuroactive molecule with
its own psychopharmacological properties (Correa et al. In vivo drug administration
2012). In particular, in vivo (Foddai et al. 2004) and in
This study was carried out in accordance with the Italian
vitro (Melis et al. 2007) evidence demonstrated that acet-
law D.L. 116/1992 which allows experiments on labora-
aldehyde stimulates the firing of pVTA DA neurons,
tory animals only after submission and approval of a
sustains its self-administration (Myers et al. 1982;
research project to the competent authorities, and with
Rodd-Henricks et al. 2002; Rodd et al. 2005; Peana et al.
the guidelines approved by the European Commission
2011), elicits conditioned place preference (Melis et al.
(n°2007/526/CE). Any possible effort was made to mini-
2007) and promotes in nucleus accumbens shell DA
mize animal pain and discomfort and to reduce the
release (Melis et al. 2007) and extracellular signal regu-
number of experimental subjects. Male CD-1 mice
lated kinase phosphorylation (Vinci et al. 2010). Remark-
(Harlan Italy SpA, Udine, Italy) used in this study were
ably, acetaldehyde has a very short half-life (Myers et al.
administered either αMpT [300 mg/kg, intraperitoneal
1982; Correa et al. 2012), and is a highly reactive
(i.p.)] (Sigma-Aldrich, Milan, Italy) or vehicle, at postna-
molecule, which may condensate, either spontaneously
tal day 21, 2 hours before sacrifice or reserpine (5 mg/kg,
or enzymatically (Chen et al. 2011), with nucleo-
i.p.) (Sigma-Aldrich) or vehicle, at postnatal day 20, 24
philic compounds, such as monoamines, to produce
hours before sacrifice. α-Lipoic acid (αLA) (100 mg/kg,
tetrahydroisoquinolines. When condensates with DA,
i.p., dissolved in 0.3 M Tris, pH 7.4) (Sigma-Aldrich) was
acetaldehyde generates 1-methyl-6,7-dihydroxy-1,2,3,4-
administered 30 minutes before sacrifice (Ledesma &
tetrahydroisoquinoline (salsolinol) (Yamanaka, Walsh &
Aragon 2012).
Davis 1970; Jamal et al. 2003).
Early studies suggested that salsolinol might have a
Assessment of DA in striatal tissue
role in psychopharmacological effects of ethanol (Davis &
Walsh 1970; Davis et al. 1970) and the relationship The striatum from both brain sides was dissected on a
between these two compounds has long been debated saline-iced block by using a Zeiss 2000 stereo microscope
(Lee et al. 2010; Correa et al. 2012; Hipólito et al. 2012; (Carl Zeiss Instruments, Milan, Italy). Tissue was imme-
Deehan, Brodie & Rodd 2013). Indeed, brain salsolinol diately weighted, suspended in 250 μl of ice-cold
quantification following in vivo ethanol administration is perchloric acid (200 mM) and sonicated with eight
still an unresolved issue (Lee et al. 2010). Accordingly, 5-second pulses followed by 3-second pauses (Sonics
salsolinol also found in chocolate (Melzig et al. 2000), Vibra Cell, Newtown, CT, USA). The suspension was then
bananas (Riggin, McCarthy & Kissinger 1976) and alco- centrifuged at 9391 × g for 10 minutes at 4°C. The super-
holic beverages (Duncan & Smythe 1982), shows poor natant was filtered through Spin-X 2 ml vials (Nylon
ability to cross blood brain barrier (Eriksson & Sippel 0.22 μm filter) (Corning-Costar, Sigma-Aldrich) and the
1977; Origitano, Hannigan & Collins 1981; Correa et al. filtrate was diluted 1:25 with distilled water. Samples
2012). Conversely, salsolinol is self-administered in pVTA of 20 μl were injected into a high performance liquid

© 2013 Society for the Study of Addiction Addiction Biology


From ethanol to salsolinol 3

chromatography (HPLC) apparatus, equipped with a medium. Ethanol, acetaldehyde, (±)-salsolinol, acetate,
15 cm LC—18 DB reverse-phase column with 5 μm par- DA (Sigma-Aldrich) and morphine (Salars, Milan, Italy)
ticle size (Supelco, Milan, Italy) and coupled with a were dissolved in distilled water, whereas 3-amino-1,2,4-
coulometric detector (ESA Coulochem II, Sunnyvale, CA, triazole (3-AT; Sigma-Aldrich) was dissolved in dimethyl
USA) to quantitate DA. Electrodes were set at +150 mV sulfoxide (DMSO) (Sigma-Aldrich). The final concentra-
(oxidation) and −200 mV (reduction). The mobile phase tion of DMSO was < 0.01%. Ethanol, acetaldehyde and
(CH3COONa: 120 mM; Citric acid: 20 mM; Na2EDTA: salsolinol concentrations were selected on the basis of
100 mg/ml; pH: 4.93; MeOH: 5 % v/v) was pumped at previous studies (Okamoto et al. 2006; Melis et al. 2007;
the flow rate of 1 ml/minute by a HPLC pump [Jasco Xie & Ye 2012; Xie et al. 2012). Action potential fre-
Europe, Cremella (LC), Italy]. The assay sensitivity for DA quency was analyzed off-line with MiniAnalysis software
was 5 fmol/injection. Data, given as mean ± standard (Synaptosoft Inc., Decatur, GA, USA). The averaged action
error (SEM), were compared and analyzed by utilizing potential frequency 5 minutes immediately before the
one-way analysis of variance (ANOVA) and, whereby drug administration was taken as baseline, and the aver-
allowed by significant main effects, followed by Tukey’s aged frequency for the third minute period centered on
post hoc comparisons. peak response was taken for drug effect. All numerical
data are given as mean ± SEM. Data were compared and
Ex vivo electrophysiology
analyzed by utilizing two-way ANOVA for repeated meas-
Whole-cell patch clamp recordings from CD-1 mouse ures (treatment × time), or one-way ANOVA for repeated
pVTA DA cells were similar to those previously described measures, when appropriate. The significance level
(Melis et al. 2010). Briefly, pups were deeply anesthetized was established at P < 0.05 and, whereby allowed by sig-
with halothane and sacrificed. A block of tissue contain- nificant main effects, ANOVA was followed by either
ing the midbrain was rapidly dissected and sliced in the Dunnett’s or Bonferroni’s multiple comparisons.
horizontal plane (230 μm) with a vibratome (VT1000S,
Leica, Wetzlar, Germany) in ice-cold low-Ca2+ solution Sample collection and sample preparation procedure for
containing (in mM): 126 NaCl, 1.6 KCl, 1.2 NaH2PO4, 1.2 high performance liquid chromatography-tandem mass
MgCl2, 0.625 CaCl2, 18 NaHCO3 and 11 glucose. Slices spectrometry (HPLC-MS/MS) analysis
were transferred to a holding chamber with artificial
Samples of aCSF (400 μl) were collected at second minute
cerebro-spinal fluid (aCSF; 37°C) saturated with 95% O2
of ethanol application to a control slice (Fig. 1A), to a slice
and 5% CO2 containing (in mM): 126 NaCl, 1.6 KCl, 1.2
from a αMpT-treated mouse (b in Fig. 4A) and at second
NaH2PO4, 1.2 MgCl2, 2.4 CaCl2, 18 NaHCO3 and 11
minute of co-application of ethanol and DA (d in Fig. 4A).
glucose. Slices (two per animal) were allowed to recover
Samples were collected from the recording chamber under
for at least 1 hour before being placed (as hemislices) in
constant perfusion flow rate of 2.5 ml/minute. Samples
the recording chamber and superfused with the aCSF
were placed into 2.0 ml glass vials and evaporated under
saturated with 95% O2 and 5% CO2. Cells were visualized
a gentle nitrogen stream. The pellet was redissolved
with an upright microscope with infrared illumination
with 100 μl of methanol containing 0.1% formic acid and
(Axioskop FS 2 plus, Zeiss, Jena, Germany), and whole-cell
subjected to HPLC-MS/MS analysis.
current-clamp recordings (one per hemislice) were
made by using an Axopatch 200B amplifier (Molecular
HPLC—Electrospray injection (ESI)—MS/MS analysis
Devices, Sunnyvale, CA, USA). Current-clamp experi-
ments were made with electrodes filled with a solution A Varian 1200 L triple-quadrupole tandem mass spec-
containing the following (in mM): 144 KCl, 10 HEPES, trometer (Varian Inc., Palo Alto, CA, USA) coupled with
3.45 BAPTA, 1 CaCl2, 2.5 Mg2ATP, and 0.25 Mg2GTP a ProStar 410 autosampler, two ProStar 210 (Varian
(pH 7.2–7.4, 275–285 mOsm). Data were filtered at Inc.) pumps and a 1200 L triple-quadrupole mass
2 kHz, digitized at 10 kHz and collected on-line with spectrometer (Varian Inc.) was used with an ESI source.
acquisition software (pClamp 8.2, Molecular Devices). DA The Varian MS workstation version 6.7 software was used
neurons were identified according to the already pub- for data acquisition and processing. Chromatographic
lished criteria (Melis et al. 2010): cell morphology and separation was performed on a Phenomenex Column
anatomical location (that is medial to the medial terminal Synergi MAX-RP 80A (4.6 mm × 150 mm i.d., 4 μm)
nucleus of the accessory optic tract), large hyperpola- [Phenomenex Italy, Castel Maggiore (Bo), Italy]. The
rization activated current (Ih > 100 pA), slow pacemaker- mobile phase consisted of (1) methanol 5% (v/v) contain-
like firing rate (< 5 Hz), long action potential duration ing 0.1% formic acid and (2) double distilled water 95%
(> 2 ms). Each slice received only a single drug exposure (v/v) containing ammonium formate 1 mM. The mobile
with the exception of data represented in Fig. 5. Drugs phase, previously degassed with high-purity helium, was
were applied in known concentrations to the superfusion pumped at a flow rate of 0.2 ml/minute, the injection

© 2013 Society for the Study of Addiction Addiction Biology


4 Miriam Melis et al.

Figure 1 Ethanol, acetaldehyde and


salsolinol excite pVTA DA neuronal firing
rate ex vivo.
Time course graphs illustrating the aver-
aged effects of ethanol (100 mM) (a), acet-
aldehyde (10 nM) (b) and salsolinol (10 nM)
(c) on firing rate of DA cells of the pVTA.
Data are normalized to the baseline. The
black bars represent the time of drug
application. n = 6 for all groups. Filled circles
represent time points significant when
compared to baseline as revealed by
Dunnett’s test. Representative traces are
shown in the insets. Calibration bar: 15 mV,
125 ms. (d) Magnification of the onset of
effects represented in panels a, b and c.
Time 0 and arrow represent the time
drugs reached the recording chamber.
(e) Concentration-response relationships
for percentage increase in firing rate pro-
duced by acetaldehyde (circles) and salso-
linol (squares). Each point shows the
mean ± SEM of responses (n = 6)

volume was 10 μl and total run time 12 minutes. ESI was sion energy used for quantitation of salsolinol were
operated in the positive ion mode. The electrospray capil- 180.1 > 115.1 (−30 V), 180.1 > 117.1 (−22 V), 180.1 >
lary potential was set at 65 V, the needle at 5850 V and 145.1 (−18 V), 180.1 > 163.1 (−14 V). The scan time
the shield at 750 V. Nitrogen, at 48 mTorr and 375°C, was was 1 second, and the detector multiplier voltage was
used as a drying gas for solvent evaporation. Full-scan set to 2000 V, with an isolation width of m/z 1.2 for
spectrum was obtained in the ranges of 100–1000 quadrupole 1 and m/z 2.0 for quadrupole 3.
atomic mass unit (amu) for salsolinol, scan time of 0.75
amu, scan width of 0.70 amu and detector at 1450 V. For RESULTS
ESI, the atmospheric pressure ionization housing was
Acetaldehyde-induced excitation of DA neuronal firing
kept at 50°C. Parent compound eluting at 8.94 minutes
rate requires DA
was subjected to collision-induced dissociation using
argon at 2.40 mTorr in the multiple reaction monitoring In agreement with previous in vitro studies, we found that
positive mode. The observed mass transitions and colli- ethanol (100 mM) (Brodie et al. 1990; Okamoto et al.

© 2013 Society for the Study of Addiction Addiction Biology


From ethanol to salsolinol 5

Tukey’s test, P < 0.05). In αMpT-treated mice, we


found that acetaldehyde was unable to stimulate DA cell
firing rate (Fig. 3A).Two-way ANOVA yielded a significant
main effect of treatment (F1,180 = 17.22, P = 0.002) dis-
closing that acetaldehyde requires the presence of
somatodendritically released DA (Cheramy et al. 1981) to
excite DA cells. Furthermore, in order to rule out the pos-
sibility that the lack of effects of acetaldehyde on pVTA DA
cells could be due to a non-specific effect of αMpT, we next
tested whether or not αMpT administration would prevent
salsolinol and morphine effects. As shown in Fig. 3B & C,
salsolinol (100 nM) and morphine (1 μM) significantly
stimulated DA cell firing rate in slices from αMpT-treated
mice similarly to control slices (two-way ANOVA F1,160 =
Figure 2 Acetate does not affect pVTA DA neuronal firing rate. 0,03, P = 0.87; F1,112 = 0,42, P = 0.53 for salsolinol and
Time course graph illustrating the averaged effects of acetate morphine, respectively). This indicates that αMpT treat-
(10 nM) on firing rate of DA cells of the pVTA (n = 5). Data are
ment did neither affect basal firing rate of these cells
normalized to the baseline.The black bar represents the time of drug
application
(αMpT 3.1 ± 0.6 and control 3.2 ± 0.5 Hz, n = 43 and 35,
respectively; P = 0.3) nor the ability of drugs, such as
salsolinol itself (Xie & Ye 2012; Xie et al. 2012) and mor-
2006; Melis et al. 2007), acetaldehyde (10 nM) (Melis phine, to excite them.
et al. 2007) and salsolinol (10 nM) (Xie & Ye 2012; Xie
et al. 2012) significantly stimulate pVTA DA cell firing
Ethanol-induced excitation of DA neuronal firing rate
rate by 142 ± 7% (one-way ANOVA for repeated meas-
requires its conversion into acetaldehyde in the
ures followed by Dunnett’s multiple comparison test:
presence of DA
F5,80 = 6.07, P = 0.0001), 252 ± 48% (one-way ANOVA
for repeated measures followed by Dunnett’s multiple Since acetaldehyde requires DA to stimulate pVTA DA cell
comparison test: F5,90 = 4.12, P = 0.0003) and 171 ± firing, we hypothesized that it might condensate with DA
26% (one-way ANOVA for repeated measures followed and generate salsolinol. Hence, we next applied ethanol
by Dunnett’s multiple comparison test: F5,76 = 6.24, (100 mM) to slices obtained from mice administered
P < 0.0001), respectively (Fig. 1A–C). Particularly, the either αMpT (300 mg/kg i.p., 2 hours before slice prepa-
onset of the effects of ethanol, acetaldehyde and ration) or reserpine (5 mg/kg i.p., 24 hours before slice
salsolinol was similar (Fig. 1D) and, in addition, acetalde- preparation). Both treatments fully prevented ethanol
hyde and salsolinol revealed overlapping dose-response ability to excite DA neurons (Fig. 4) (αMpT: two-way
curves in the range between 1 and 100 nM (two-way ANOVA F1,240 = 8,14, P = 0.012; reserpine: two-way
ANOVA F1,18 = 0.29, P = 0.6) (Fig. 1E). Notably, acetate, ANOVA F1,208 = 10,25, P = 0.006). Furthermore, we
at a concentration similar to effective acetaldehyde and tested the contribution of DA to the effect of ethanol
salsolinol one (that is 10 nM), did not affect pVTA DA (100 mM). As shown in Fig. 5A (left), ethanol failed to
cell activity (one way ANOVA F18,112 = 0.33, P = 0.99) stimulate the firing rate of DA cells; however, in the pres-
(Fig. 2). ence of bath applied DA (10 nM) its ability to enhance
We previously reported that acetaldehyde is essential pVTA DA cell activity was fully restored (one-way ANOVA
for ethanol-induced excitation of pVTA DA neurons (Melis followed by Bonferroni’s multiple comparisons test
et al. 2007). However, whether or not acetaldehyde’s F34, 175 = 2,34, P = 0.0002) [Fig. 5A (right)].
effects on pVTA DA neurons are mediated by its condensa- Since ethanol requires both acetaldehyde and DA to
tion with DA is unknown yet. Hence, we applied acetalde- increase pVTA DA neuronal activity, we next tested
hyde (10 nM) to slices obtained from mice administered whether or not application of catalase inhibitor, 3-AT
αMpT (300 mg/kg i.p., 2 hours before slice preparation), a (1 mM) would prevent DA to restore ethanol effects in
drug that inhibits DA synthesis and release (Cheramy et al. αMpT-treated mice. Remarkably [Fig. 5B (right)], 3-AT
1981) and depletes striatal DA content. Accordingly, fully abolished DA ability to recover the stimulating prop-
striatal tissue DA content was reduced following adminis- erties of ethanol on DA cell firing rate in αMpT-treated
tration of either αMpT (300 mg/kg i.p.) or reserpine mice (one-way ANOVA followed by Bonferroni’s multiple
(5 mg/kg i.p.), a drug able to deplete monoamine vesi- comparison test F31, 155 = 2.68, P = 0.0001). Accordingly,
cular stores, by irreversibly inhibiting vesicular uptake, in vivo administration of αLA (100 mg/kg i.p., 30
(two-way ANOVA F2,33 = 169, P = 0.0001 followed by minutes before slice preparation), a hydrogen peroxide

© 2013 Society for the Study of Addiction Addiction Biology


6 Miriam Melis et al.

Figure 3 Acetaldehyde-induced excita-


tion of DA neuronal firing rate requires the
presence of DA.
(a) Time course graph illustrating the aver-
aged effects of acetaldehyde (10 nM) on
firing rate of DA cells of the pVTA in mice
administered αMpT. (b) Time course graph
illustrating the averaged effects of salsolinol
(100 nM) on firing rate of DA cells of the
pVTA in mice administered αMpT. (c) Time
course graph illustrating the averaged
effects of morphine (1 μM) on firing rate of
the pVTA DA cells in mice administered
αMpT. Data are normalized to the baseline.
n = 6 for panels a and b, and n = 5 for panel
c.The black bars represent the time of drug
application. Grey areas represent the mean
response observed in control mice. Repre-
sentative traces of the spontaneous activity
of a DA neuron during baseline (a), acet-
aldehyde, salsolinol or morphine (b) and
wash-out (c) of an αMpT-treated mouse
are shown in the insets. Calibration: 15 mV,
125 ms

scavenger (Ledesma & Aragon 2012), fully prevented P = 0.66), even in the presence of bath applied DA
ethanol-induced increase in firing rate of pVTA DA cells (Fig. 5C). Lastly, 3-AT did not prevent acetaldehyde
in both naïve (one-way ANOVA F19,76 = 1.55, P = 0.09) actions on pVTA DA cells (one-way ANOVA F3,19 = 16.08,
and αMpT-treated mice (two-way ANOVA F1,52 = 0.2, P < 0.0001) (Fig. 6).

Figure 4 Ethanol-induced excitation of DA neuronal firing rate requires the presence of DA.
Time course graph illustrating the averaged effects of ethanol (100 mM) on firing rate of DA cells of the pVTA in mice pretreated with either
αMpT (white circles) or reserpine (grey circles). Data are normalized to the baseline. n = 6 for all groups. The black bar represents the time of drug
application. Grey area represents the mean response observed in control mice. Representative traces of the spontaneous activity of a DA neuron
during baseline (a), ethanol (b) and wash-out (c) of an αMpT- or reserpine-treated mouse are shown in the inset. Calibration: 15 mV, 125 ms

© 2013 Society for the Study of Addiction Addiction Biology


From ethanol to salsolinol 7

Figure 5 Ethanol-induced excitation of DA neuronal firing rate requires its oxidation into acetaldehyde and the presence of DA.
(a)Time course graph illustrating the averaged effects of ethanol (100 mM) on firing rate of DA cells of the pVTA in mice pretreated with αMpT.
Note that the effect of ethanol is restored in the presence of exogenous DA (10 nM). On the right-hand of the panel, representative traces of
the spontaneous activity of a DA neuron during baseline (a), ethanol (b), DA (c) and ethanol co-applied with DA (d) of an αMpT-treated mouse.
Calibration: 15 mV, 125 ms. (B) Time course graph illustrating the averaged effects of ethanol (100 mM) on firing rate of DA cells of the pVTA
in mice pretreated with αMpT.The effect of ethanol is abolished in the presence of DA (10 nM) when acetaldehyde formation is prevented
by 3-AT application (1 mM). On the right-hand of the panel, representative traces of the spontaneous activity of a DA neuron during baseline
(a), ethanol (b), DA and 3-AT (c), and ethanol co-applied with DA and 3-AT (d) of a αMpT-treated mouse. Calibration: 15 mV, 125 ms. Data are
normalized to the baseline. n = 6 for all groups. (C)Time course graph illustrating the averaged effects of ethanol (100 mM) on firing rate of DA
cells of the pVTA in naïve and αMpT-treated mice following systemic administration of αLA (100 mg/kg i.p., 30 minutes before slice preparation).
The effect of ethanol is abolished when acetaldehyde formation is prevented by αLA, even in the presence of DA (10 nM) applied to slices from
αMpT-treated mice. On the right-hand panel, representative traces of the spontaneous activity of a DA neuron during baseline (a), vehicle or
DA (b), and ethanol co-applied with DA (c) of naïve and αMpT-treated mice following αLA administration. Calibration: 15 mV, 125 ms. Data are
normalized to the baseline. n = 5 for both groups.The black bars represent the time of drug application

HPLC-MS/MS detects salsolinol in aCSF after following its oxidation into acetaldehyde and subsequent
co-application of ethanol and DA to slices from condensation of acetaldehyde with DA to generate
αMpT-treated mice salsolinol. If salsolinol is, indeed, the molecule responsi-
ble of the effects observed on DA firing rate, one would
Overall, these results demonstrate that ethanol stimu- expect to detect it in the aCSF under the same conditions
lates the spontaneous activity of pVTA DA neurons (see above). Hence, we analyzed by HPLC-MS/MS the

© 2013 Society for the Study of Addiction Addiction Biology


8 Miriam Melis et al.

Figure 6 Acetaldehyde effects on pVTA DA cells are independent of catalase inhibition.


(a) Time course graph illustrating the averaged effects of acetaldehyde (10 nM) in the presence of 3-AT (1 mM, 5 minutes) on firing rate of
pVTA DA cells (n = 5) in naïve mice. Data are normalized to the baseline. The black bar represents the time of drug application. Grey area
represents the mean response observed in control mice. (b) Bar graph summarizing the effects observed in (a). Acetaldehyde (ACD)
stimulates the firing rate of pVTA DA cells also in the presence of 3-AT, which per se is ineffective on DA cell spontaneous activity (one-way
ANOVA followed by Dunnett’s multiple comparison test, F3,19 = 16.18, P < 0.0001).

Figure 7 HPLC-MS/MS detects salsolinol in aCSF after co-application of ethanol and DA to slices from αMpT-treated mice.
Representative HPLC-MS/MS ion chromatograms (mass transition: 180.1 > 117.1; collision energy −22 V) of (a) standard of salsolinol (56 nM)
in aCSF; (b) aCSF from recording chamber at second-minute application of ethanol (100 mM) in control mouse; (c) aCSF from recording
chamber at second-minute application of ethanol (100 mM) in αMpT-treated mouse; (d) aCSF from recording chamber at second-minute
co-application of ethanol (100 mM) and DA (10 nM) in αMpT-treated mouse.

aCSF after application of ethanol (as in Fig. 5A, left) and (not shown) and αMpT-treated mice (D). Particularly,
after co-application of ethanol and DA (as in Fig. 5A, under conditions of ethanol and DA co-application
right). As shown in Fig. 7, we detected salsolinol when (n = 4), salsolinol was determined at a higher concentra-
ethanol was applied in slices from naïve mice (B), and tion (24 ± 3 nM) than we expected (≤10 nM) on the basis
when ethanol and DA were co-applied in slices from naïve of the stoichiometry of its formation (1:1 from the

© 2013 Society for the Study of Addiction Addiction Biology


From ethanol to salsolinol 9

ethanol-derived acetaldehyde and the added DA). One inhibition of ethanol metabolism prevents ethanol’s
possible explanation might be that undetermined, and ability to excite DA neurons in slices from αMpT-treated
not necessarily overlapping, kinetics of acetaldehyde and mice. Accordingly, in vivo administration of the H2O2
salsolinol production may take place in the recording scavenger, αLA, abolishes ethanol actions on pVTA DA
chamber. In addition, sample collection for this analysis cells in both naïve and αMpT-treated mice, thus support-
might have been carried out (second minute of ethanol ing that ethanol oxidation is a limiting step for ethanol
application) when salsolinol formation had already taken behavioral actions (Ledesma & Aragon 2012; Ledesma
place at a possibly faster rate than that of removal of DA et al. 2013; Peana et al. 2013) even in the presence of
by the constant flow rate of perfusion in the recording DA. Accordingly, 3-AT does neither affect spontaneous
chamber. activity of pVTA DA cells nor their response to acetalde-
hyde in naïve mice. In contrast, one must notice that,
while basal DA cell spontaneous activity is not affected
DISCUSSION
in αMpT-treated mice, a dysfunctional H2O2-catalase
We here provide evidence for ethanol’s two-step mecha- system alters pVTA DA cell spontaneous activity and
nism of action for its excitatory effects on pVTA DA their responses to ethanol in αMpT-treated mice. How-
neurons. This mechanism consists of its former catalase- ever, at this stage, we have no explanation for this phe-
mediated oxidation, and subsequent condensation of nomenon. One possibility might be that, since H2O2 is
ethanol-derived acetaldehyde with endogenous DA to produced by DA neurons in an activity-dependent
generate salsolinol, which excites pVTA DA neurons (Xie fashion (Avshalumov et al. 2005), impairment of H2O2-
& Ye 2012; Xie et al. 2012). catalase system might result in compensatory meta-
Several observations support our conclusions. First, bolic changes in DA cells of αMpT-treated mice. Finally,
somatodendritically released DA is required for ethanol acetaldehyde-induced excitation of pVTA DA neurons is
effects, since these were absent in mice administered absent in αMpT-treated mice, thus substantiating our
either αMpT or reserpine. Notably, ethanol effects were current hypothesis that ethanol-derived acetaldehyde
restored when DA was applied stoichiometrically with requires DA to excite DA neurons.
respect to the concentration at which acetaldehyde Third, salsolinol was detected in the aCSF only when
increases DA cell firing rate (present results and Melis ethanol and exogenous DA were co-applied in αMpT-
et al. 2007). Thus, bath application of exogenous DA, at a treated slices (Fig. 7). Notably, we applied ethanol at a
concentration that per se does not affect DA neuronal concentration that was previously reported to enhance
activity, fully recovered ethanol’s stimulatory effects on pVTA DA cell activity under similar experimental condi-
DA firing rate. While this observation appears in contrast tions (Okamoto et al. 2006; Melis et al. 2007; Xie et al.
with Xie et al. (2012), one must consider that the known 2012). Although we acknowledge that this concentra-
slow action of reserpine (Callaway, Kuczenski & Segal tion is likely higher than the one resulting from a
1989) and different experimental conditions (that is mice behaviorally relevant dose of ethanol in vivo, direct com-
systemically administered 24 hours before experiment parisons between concentrations of ethanol contingently
versus acute bath application to slices) might account self-administered in pVTA in vivo (Rodd et al. 2004,
for this discrepancy. In addition, reserpine is known to 2005) and those examined ex vivo in acute brain slices
inhibit the monoamine vesicular carrier and, transiently, (Okamoto et al. 2006; Melis et al. 2007) could not be
increase cytoplasmic DA concentration. Thus, it is possi- attempted. Nonetheless, rats self-administer ethanol
ble that in the early phases of its bath application, such as directly in pVTA at concentrations up to 300 mg%,
reported by Xie et al. (2012), reserpine may not only have which are equivalent to 66 mM. Furthermore, we applied
failed to reduce somatodendritic DA release, but also—if concentrations of ethanol and acetaldehyde with a
any—have increased it by reverting DA transporter activ- 1000:1 ratio (that is 100 mM versus 100 nM, respec-
ity (Sulzer, Maidment & Rayport 1993). Notably, both tively), which comply the ratio of those systemically
salsolinol and morphine were still able to excite pVTA DA administered (that is g/kg versus mg/kg). Additionally,
neurons in αMpT-treated mice. Collectively, these results we cannot rule out that ethanol acute actions (at
indicate that somatodendritically released DA is required 100 mM) on spontaneous activity of DA neurons might
for ethanol and acetaldehyde to stimulate pVTA DA cell be the net effect resulting from complex synaptic changes
activity. at both inhibitory (M. Melis and A. Bonci, unpublished
Second, ethanol-derived acetaldehyde (Cohen et al. observations) and excitatory inputs integrated with cell
1980; Aragon et al. 1991, 1992; Gill et al. 1992; membrane properties (Okamoto et al. 2006; Tateno &
Zimatkin et al. 2006) requires the presence of Robinson 2011). Conversely, acetaldehyde (at nM con-
extracellular DA to enhance DA cell firing rate. In fact, centrations) only acts by changing the intrinsic excitabil-
even in the presence of exogenous DA, 3-AT-mediated ity of DA neurons (Melis et al. 2007). Notably, the

© 2013 Society for the Study of Addiction Addiction Biology


10 Miriam Melis et al.

differences in terms of absolute values of acetaldehyde Acknowledgements


between in vivo and ex vivo studies may appear not so
This study was supported by the grant from Regione
critical. In fact, similar absolute amounts (number of
Autonoma della Sardegna (RAS), Italy, L.R. 7/2007,
moles) of acetaldehyde result in pVTA after its in vivo
CRP_537-CUP F71J090006200002 to E.A. The helpful
self-administration (6–90 μM in 100 nl/active lever
comments of P. J. Mackenzie and G. Zernig, and kind
press, that is between 6 × 10–13 and 9 × 10–12 moles) as in
assistance on HPLC-MS/MS analyses of A. Murgia and I.
Rodd et al. (2005) and after its bath application (10–
Ibba are gratefully acknowledged.
100 nM in 2.5 ml at the constant flow rate of 2.5 ml/
minute, that is between 1.25 × 10–11 and 1.25 × 10−10
Authors’ Contributions
moles/pVTA) as in the present ex vivo experiments. Fur-
thermore, in vivo sequential microinjections in pVTA of MM contributed to design the experiments, performed
6.9 × 10–11 moles of acetaldehyde over 10 minutes, result the electrophysiological experiments, analyzed data and
in large increases of DA efflux in nucleus accum- wrote the manuscript. EC performed the quantification of
bens shell (Deehan et al. 2012). dopamine content in the tissue by HPLC-ECD, analyzed
The stimulatory action of salsolinol on firing rate of data and contributed to write the manuscript. PC per-
DA cells has been ascribed to activation of μ-opioid recep- formed the identification of salsolinol by HPLC-MS/MS
tors and concurrent stimulation of excitatory afferents in the aCSF collected from the recording chamber, and
(Xie & Ye 2012). Thus, provided that the aim of our contributed to write the manuscript. EA conceived and
study was to demonstrate whether ethanol acts as designed the study, performed in vivo administrations and
prodrug of salsolinol, we did not further address this wrote the manuscript. All authors have critically
issue. Finally, ethanol-derived acetaldehyde via hypotha- reviewed content and approved final version submitted
lamic β-endorphins (Chronwall 1985) might be also for publication.
involved in the effects on VTA DA neurons and their
behavioral output (Spanagel, Herz & Shippenberg
1992; Sanchis-Segura et al. 2005; Pastor & Aragon References
2008). Notably, acetate, the second by-product of ethanol Ahlenius S, Carlsson A, Engel J, Svensson T, Södersten P (1973)
metabolism, has no effect on DA’s spontaneous neuronal Antagonism by alpha methyltyrosine of the ethanol-induced
activity, thus supporting the behavioral evidence that stimulation and euphoria in man. Clin Pharmacol Ther
acetate is not involved in DA-dependent effects of ethanol 14:586–591.
Aragon CM, Amit Z (1992) The effect of 3-amino-1,2,4-triazole
(Pardo et al. 2013).
on voluntary ethanol consumption: evidence for brain
The ability of ethanol to stimulate the firing rate of DA catalase involvement in the mechanism of action. Neurophar-
neurons in the pVTA has long been considered a critical macology 31:709–712.
step for its psychomotor and motivational effects (Di Aragon CM, Spivak K, Amit Z (1991) Effect of 3-amino-1,2
Chiara 1997; Koob et al. 1998; Boileau et al. 2003) and ,4-triazole on ethanol-induced narcosis, lethality and hypo-
thermia in rats. Pharmacol Biochem Behav 39:55–59.
for its ability to initiate and maintain ethanol taking
Aragon CM, Rogan F, Amit Z (1992) Ethanol metabolism in rat
behaviors (Di Chiara 1997; Koob et al. 1998; Söderpalm brain homogenates by a catalase H2O2 system. Biochem
& Ericson 2013). Remarkably, the present results are in Pharmacol 44:93–98.
line with an early clinical report suggesting that admin- Avshalumov MV, Chen BT, Koós T, Tepper JM, Rice ME (2005)
istration of αMpT prevents ethanol ability to elicit eupho- Endogenous hydrogen peroxide regulates the excitability of
ria (Ahlenius et al. 1973). An immediate implication of midbrain dopamine neurons via ATP-sensitive potassium
channels. J Neurosci 25:4222–4231.
this study would involve the potential role of salsolinol in
Boileau I, Assaad JM, Pihl RO, Benkelfat C, Leyton M, Diksic M,
stimulating firing rate of pVTA DA cells, and in exerting Tremblay RE, Dagher A (2003) Alcohol promotes dopamine
motivational effects when ingested (through foods and release in the human nucleus accumbens. Synapse 49:226–
beverages). However, the questioned ability of salsolinol 231.
to cross blood brain barrier (Origitano et al. 1981; Correa Brodie MS, Shefner SA, Dunwiddie TV (1990) Ethanol increases
the firing rate of dopamine neurons of the rat ventral
et al. 2012) greatly reduces the impact of this implica-
tegmental area in vitro. Brain Res 508:65–69.
tion. Nevertheless, given that ethanol is the main ingre- Brown ZW, Amit Z, Rockman GE (1979) Intraventricular
dient of alcoholic drinks, the present results not only self-administration of acetaldehyde, but not ethanol, in naive
provide a significant contribution to the understanding of laboratory rats. Psychopharmacology 64:271–276.
ethanol’s mechanism of action on DA cells, but also make Callaway CW, Kuczenski R, Segal DS (1989) Reserpine enhances
amphetamine stereotypies without increasing amphetamine-
a significant advance in the field of alcohol research.
induced changes in striatal dialysate dopamine. Brain Res
Finally, they provide new insights to look into the 505:83–90.
neurobiological basis of alcoholism and suggest exciting Chen A, Arshad A, Qing H, Wang R, Lu J, Deng Y (2011)
avenues of future research. Enzymatic condensation of dopamine and acetaldehyde:

© 2013 Society for the Study of Addiction Addiction Biology


From ethanol to salsolinol 11

a salsolinol synthase from rat brain. Biologia (Bratisl) Howard EC, Schier CJ, Wetzel JS, Duvauchelle CL, Gonzales RA
66:1183–1188. (2008) The shell of the nucleus accumbens has a higher dopa-
Cheramy A, Leviel V, Glowinski J (1981) Dendritic release of mine response compared with the core after non-contingent
dopamine in the substantia nigra. Nature 289:537–542. intravenous ethanol administration. Neuroscience 154:
Chronwall BM (1985) Anatomy and physiology of the 1042–1053.
neuroendocrine arcuate nucleus. Peptides 6 (Suppl. 2):1–11. Jamal M, Ameno K, Ameno S, Okada N, Ijiri I (2003) In vivo
Cohen G, Sinet PM, Heikkila R (1980) Ethanol oxidation by rat study of salsolinol produced by a high concentration of acet-
brain in vivo. Alcohol Clin Exp Res 4:366–370. aldehyde in the striatum and nucleus accumbens of free-
Correa M, Salamone JD, Segovia KN, Pardo M, Longoni R, Spina moving rats. Alcohol Clin Exp Res 27 (8 Suppl. ):79S–84S.
L, Peana AT, Vinci S, Acquas E (2012) Piecing together the Koob GF, Roberts AJ, Schulteis G, Parsons LH, Heyser CJ,
puzzle of acetaldehyde as a neuroactive agent. Neurosci Hyytiä P, Merlo-Pich E, Weiss F (1998) Neurocircuitry targets
Biobehav Rev 36:404–430. in ethanol reward and dependence. Alcohol Clin Exp Res
Davis VE, Walsh MJ (1970) Alcohol, amines, and alkaloids: a 22:3–9.
possible biochemical basis for alcohol addiction. Science Ledesma JC, Aragon CM (2012) Alpha-lipoic acid, a scavenging
167:1005–1007. agent for H2O2, reduces ethanol-stimulated locomotion in
Davis VE, Walsh MJ, Yamanaka Y (1970) Augmentation of mice. Psychopharmacol 219:171–180.
alkaloid formation from dopamine by alcohol and acetalde- Ledesma JC, Font L, Baliño P, Aragon CM (2013) Modulation
hyde in vitro. J Pharmacol Exp Ther 174:401–412. of ethanol-induced conditioned place preference in mice
Deehan GA Jr., Engleman EA, Ding ZM, McBride WJ, Rodd ZA by 3-amino-1,2,4-triazole and d-penicillamine depends on
(2012) Microinjections of acetaldehyde or salsolinol into the ethanol dose and number of conditioning trials. Psycho-
posterior ventral tegmental area increase dopamine release in pharmacology 226:673–685.
the nucleus accumbens shell. Alcohol Clin Exp Res 37:722– Lee J, Ramchandani VA, Hamazaki K, Engleman EA, McBride
729. WJ, Li TK, Kim HY (2010) A critical evaluation of influence of
Deehan GA Jr, Brodie MS, Rodd ZA (2013) What is in that drink: ethanol and diet on salsolinol enantiomers in humans and
the biological actions of ethanol, acetaldehyde, and salsolinol. rats. Alcohol Clin Exp Res 34:242–250.
Curr Top Behav Neurosci 13:163–194. Melis M, Enrico P, Peana AT, Diana M (2007) Acetaldehyde
Deitrich RA (2004) Acetaldehyde: déjà vu du jour. J Stud mediates alcohol activation of the mesolimbic dopamine
Alcohol 65:557–572. system. Eur J Neurosci 26:2824–2833.
Di Chiara G (1997) Alcohol and dopamine. Alcohol Health Res Melis M, Carta S, Fattore L, Tolu S, Yasar S, Goldberg
World 21:108–114. SR, Fratta W, Maskos U, Pistis M (2010) Peroxisome
Duncan MW, Smythe GA (1982) Salsolinol and dopamine in proliferator-activated receptors-alpha modulate dopamine cell
alcoholic beverages. Lancet 1:904–906. activity through nicotinic receptors. Biol Psychiatry 68:256–
Eriksson CJ, Sippel HW (1977) The distribution and metabolism 264.
of acetaldehyde in rats during ethanol oxidation-l. The distri- Melzig MF, Putscher I, Henklein P, Haber H (2000) In vitro phar-
bution of acetaldehyde in liver, brain, blood and breath. macological activity of the tetrahydroisoquinoline salsolinol
Biochem Pharmacol 26:241–247. present in products from Theobroma cacao L. like cocoa and
Escrig MA, Pardo M, Aragon CM, Correa M (2012) Anxio- chocolate. J Ethnopharmacol 73:153–159.
genic and stress-inducing effects of peripherally administered Myers WD, Ng KT, Singer G (1982) Intravenous self-
acetaldehyde in mice: similarities with the disulfiram-ethanol administration of acetaldehyde in the rat as a function of
reaction. Pharmacol Biochem Behav 100:404–412. schedule, food deprivation and photoperiod. Pharmacol
Foddai M, Dosia G, Spiga S, Diana M (2004) Acetaldehyde Biochem Behav 17:807–811.
increases dopaminergic neuronal activity in the VTA. Okamoto T, Harnett MT, Morikawa H (2006) Hyperpolarization
Neuropsychopharmacology 29:530–536. activated cation current (Ih) is an ethanol target in midbrain
Font L, Miquel M, Aragon CM (2008) Involvement of brain dopamine neurons of mice. J Neurophysiol 95:619–626.
catalase activity in the acquisition of ethanol-induced condi- Origitano T, Hannigan J, Collins MA (1981) Rat brain salsolinol
tioned place preference. Physiol Behav 93:733–741. and blood-brain barrier. Brain Res 224:446–451.
Gessa GL, Muntoni F, Vargiu L, Mereu G (1985) Low doses Pardo M, Betz AJ, San Miguel N, López-Cruz L, Salamone JD,
of ethanol activate dopaminergic neurons in the ventral Correa M (2013) Acetate as an active metabolite of ethanol:
tegmental area. Brain Res 348:201–203. studies of locomotion, loss of righting reflex, and anxiety in
Gill K, Menez JF, Lucas D, Deitrich RA (1992) Enzymatic produc- rodents. Front Behav Neurosci 7:81.
tion of acetaldehyde from ethanol in rat brain tissue. Alcohol Pastor R, Aragon CM (2008) Ethanol injected into the hypotha-
Clin Exp Res 16:910–915. lamic arcuate nucleus induces behavioral stimulation in rats:
Hipólito L, Sánchez-Catalán MJ, Granero L, Polache A (2010) an effect prevented by catalase inhibition and naltrexone.
Locomotor stimulant effects of acute and repeated intra- Behav Pharmacol 19:698–705.
tegmental injections of salsolinol in rats: role of mu-opioid Peana AT, Muggironi G, Fois GR, Zinellu M, Vinci S, Acquas E
receptors. Psychopharmacol 209:1–11. (2011) Effect of opioid receptor blockade on acetaldehyde self-
Hipólito L, Martí-Prats L, Sánchez-Catalán MJ, Polache A, administration and ERK phosphorylation in the rat nucleus
Granero L (2011) Induction of conditioned place preference accumbens. Alcohol 45:773–783.
and dopamine release by salsolinol in posterior VTA of rats: in- Peana AT, Muggironi G, Fois G, Diana M (2013) Alpha-lipoic
volvement of μ-opioid receptors. Neurochem Int 59:559–562. acid reduces ethanol self-administration in rats. Alcohol Clin
Hipólito L, Sánchez-Catalán MJ, Martí-Prats L, Granero L, Exp Res. doi: 10.1111/acer.12169.
Polache A (2012) Revisiting the controversial role of Riggin RM, McCarthy MJ, Kissinger PT (1976) Identification of
salsolinol in the neurobiological effects of ethanol: old and salsolinol as a major dopamine metabolite in the banana.
new vistas. Neurosci Biobehav Rev 36:362–378. J Agric Food Chem 24:189–191.

© 2013 Society for the Study of Addiction Addiction Biology


12 Miriam Melis et al.

Rodd ZA, Melendez RI, Bell RL, Kuc KA, Zhang Y, Murphy JM, Spanagel R, Herz A, Shippenberg TS (1992) Opposing tonically
McBride WJ (2004) Intracranial self-administration of active endogenous opioid systems modulate the mesolimbic
ethanol within the ventral tegmental area of male Wistar rats: dopaminergic pathway. Proc Natl Acad Sci U S A 89:2046–
evidence for involvement of dopamine neurons. J Neurosci 2050.
24:1050–1057. Sulzer D, Maidment NT, Rayport S (1993) Amphetamine and
Rodd ZA, Bell RL, Zhang Y, Murphy JM, Goldstein A, Zaffaroni A, other weak bases act to promote reverse transport of dopamine
Li TK, McBride WJ (2005) Regional heterogeneity for the in ventral midbrain neurons. J Neurochem 60:527–535.
intracranial self-administration of ethanol and acetaldehyde Tabakoff B, Anderson RA, Ritzman RF (1976) Brain acetalde-
within the ventral tegmental area of alcohol-preferring (P) hyde after ethanol administration. Biochem Pharmacol
rats: involvement of dopamine and serotonin. Neuropsycho- 25:1305–1309.
pharmacology 30:330–338. Tateno T, Robinson HP (2011) The mechanism of ethanol
Rodd ZA, Oster SM, Ding ZM, Toalston JE, Deehan G, Bell RL, Li action on midbrain dopaminergic neuron firing: a dynamic-
TK, McBride WJ (2008) The reinforcing properties of clamp study of the role of I(h) and GABAergic synaptic inte-
salsolinol in the ventral tegmental area: evidence for regional gration. J Neurophysiol 106:1901–1922.
heterogeneity and the involvement of serotonin and dopa- Vinci S, Ibba F, Longoni R, Spina L, Spiga S, Acquas E (2010)
mine. Alcohol Clin Exp Res 32:230–239. Acetaldehyde elicits ERK phosphorylation in the rat nucleus
Rodd-Henricks ZA, Melendez RI, Zaffaroni A, Goldstein A, accumbens and extended amygdala. Synapse 64:916–927.
McBride WJ, Li TK (2002) The reinforcing effects of acetalde- Xie G, Ye JH (2012) Salsolinol facilitates glutamatergic transmis-
hyde in the posterior ventral tegmental area of alcohol- sion to dopamine neurons in the posterior ventral tegmental
preferring rats. Pharmacol Biochem Behav 72:55–64. area of rats. PLoS ONE 7:e36716.
Sanchis-Segura C, Miquel M, Correa M, Aragon CM (1999) Xie G, Hipólito L, Zuo W, Polache A, Granero A, Krnjevic K,
Cyanamide reduces brain catalase and ethanol-induced Ye JH (2012) Salsolinol stimulates dopamine neurons
locomotor activity: is there a functional link? Psychophar- in slices of posterior ventral tegmental area indirectly by
macology 144:83–89. activating μ-opioid receptors. J Pharmacol Exp Ther 341:43–
Sanchis-Segura C, Correa M, Miquel M, Aragon CM (2005) 50.
Catalase inhibition in the arcuate nucleus blocks ethanol Yamanaka Y, Walsh MJ, Davis VE (1970) Salsolinol, an alkaloid
effects on the locomotor activity of rats. Neurosci Lett derivative of dopamine formed in vitro during alcohol metabo-
376:66–70. lism. Nature 227:1143–1144.
Söderpalm B, Ericson M (2013) Neurocircuitry involved in the Zimatkin SM, Pronko SP, Vasiliou V, Gonzalez FJ, Deitrich RA
development of alcohol addiction: the dopamine system and (2006) Enzymatic mechanisms of ethanol oxidation in the
its access points. Curr Top Behav Neurosci 13:127–161. brain. Alcohol Clin Exp Res 30:1500–1505.

© 2013 Society for the Study of Addiction Addiction Biology

You might also like