Skip to main content
Erschienen in: Molecular Brain 1/2014

Open Access 01.12.2014 | Research

Histamine impairs midbrain dopaminergic development in vivo by activating histamine type 1 receptors

verfasst von: Itzel Escobedo-Avila, Fernanda Vargas-Romero, Anayansi Molina-Hernández, Rodrigo López-González, Daniel Cortés, Juan A De Carlos, Iván Velasco

Erschienen in: Molecular Brain | Ausgabe 1/2014

Abstract

Background

Histamine (HA) regulates the sleep-wake cycle, synaptic plasticity and memory in adult mammals. Dopaminergic specification in the embryonic ventral midbrain (VM) coincides with increased HA brain levels. To study the effect of HA receptor stimulation on dopamine neuron generation, we administered HA to dopamine progenitors, both in vitro and in vivo.

Results

Cultured embryonic day 12 (E12) VM neural stem/progenitor cells expressed transcripts for HA receptors H1R, H2R and H3R. These undifferentiated progenitors increased intracellular calcium upon HA addition. In HA-treated cultures, dopamine neurons significantly decreased after activation of H1R. We performed intrauterine injections in the developing VM to investigate HA effects in vivo. HA administration to E12 rat embryos notably reduced VM Tyrosine Hydroxylase (TH) staining 2 days later, without affecting GABA neurons in the midbrain, or serotonin neurons in the mid-hindbrain boundary. qRT-PCR and Western blot analyses confirmed that several markers important for the generation and maintenance of dopaminergic lineage such as TH, Lmx1a and Lmx1b were significantly diminished. To identify the cell type susceptible to HA action, we injected embryos of different developmental stages, and found that neural progenitors (E10 and E12) were responsive, whereas differentiated dopaminergic neurons (E14 and E16) were not susceptible to HA actions. Proliferation was significantly diminished, whereas neuronal death was not increased in the VM after HA administration. We injected H1R or H2R antagonists to identify the receptor responsible for the detrimental effect of HA on dopaminergic lineage and found that activation of H1R was required.

Conclusion

These results reveal a novel action of HA affecting dopaminergic lineage during VM development.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​s13041-014-0058-x) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare no competing financial interests.

Authors’ contributions

IEA wrote the manuscript, analyzed the data and performed most of the experiments with help from FVR. AMH contributed to the design of experiments and discussion of results. RLG helped to perform intrauterine injections, analysis and discussion of results. DC and JADC contributed to set up techniques reviewed the manuscript and took part in the discussion of results. All authors read and approved the final manuscript. IV contributed to experiment design/supervision, reviewed the manuscript and obtained funding.

Background

Histamine (HA) is a neurotransmitter that participates in the sleep cycle, motor activity, synaptic plasticity and memory [1]–[5]. The adult mammalian brain contains somas of HA-producing neurons in the hypothalamic tuberomamillary nucleus, and its projections reach most of the CNS [3],[4],[6]. HA exerts its actions through the activation of four G-protein-coupled receptors: H1R-H4R. H1R, H2R and H3R are widely distributed throughout the CNS [6]–[8], but the presence of H4R in the brain remains unclear [9]–[11].
During embryonic development, HA is present at higher concentrations than those observed in the adult brain [12]–[14]. HA increases its content at embryonic days (E) 14–16, and then decreases perinatally [13],[15]–[17]. Histidine decarboxylase (HDC), which synthetizes HA, and the mRNAs for H1R, H2R and H3R [18]–[21] are also expressed at these developmental stages [20],[22],[23]. The transitory expression of HDC in the midbrain, together with elevated HA levels, and the expression of HRs, coincides with periods of dopaminergic specification [24], suggesting that HA might influence dopamine (DA) neuron generation in the ventral midbrain (VM). In rats, differentiated dopaminergic neurons are first identified at E12 by expression of Tyrosine Hydroxylase (TH), the rate-limiting enzyme for DA synthesis. The peak of DA neurogenesis is at E13. By E16, all DA cells are post-mitotic and their axons are en route to reach forebrain structures such as the striatum and the cerebral cortex [25],[26].
On cortical neural stem/progenitor cell (NSPC) cultures, the effect of activating H1 and H2 receptors has been reported [27]. HA increases cell proliferation by H2R activation and has a neuronal-differentiating action mediated by H1R stimulation, probably by rising prospero 1 and neurogenin 1 expression [28] and favoring differentiation to FOXP2-positive neurons [29]. Furthermore, HA stimulates neuronal differentiation of adult subventricular zone NSPC [30]. The HA concentrations required to observe these effects are 100 μM to 1 mM.
Elucidating the developmental pathways that control neuronal specification in the VM is of great relevance to increase our knowledge in differentiation of dopaminergic cells. In this work, we aimed to study the effect of HA on dopaminergic development. We found that HA is detrimental to dopaminergic differentiation of NSPC in vitro, and demonstrate that HA administration in vivo at early developmental stages decreases dopaminergic induction in the VM through H1R stimulation. This study establishes the inhibitory relationship of HA to DA neuron generation during development, and provides a novel mechanism for the future treatment of Parkinson’s disease.

Results

Midbrain NSPC cultures are multipotent and express histaminergic receptors

NSPC have the capacity to self-renew, and the potential to differentiate into neurons, astrocytes and oligodendrocytes. To characterize the expression of NSPC markers, as well as their capacity to differentiate to neuronal and glial populations, we cultured VM NSPC isolated from E12 rat embryos. We maintained these cells in proliferation during 4 days in the presence of the mitogen Fibroblast growth factor (FGF)-2, and then induced differentiation for 6 days after removal of FGF-2. We found that a very high proportion of these cells express Sox2, Vimentin and Nestin, which are markers widely used to identify undifferentiated NSPC (Figure 1A). After removing FGF-2 from cultures, cells readily differentiate into neurons (MAP2- and β-III Tubulin-positive), astrocytes (Glial Fibrillary Acidic Protein, GFAP-positive) and oligodendrocytes (O4-positive) (Figure 1B), confirming that our cultures are indeed NSPC.
Previous work has demonstrated that mRNAs for histaminergic receptors are present in rodent embryos from E14 onwards [18],[20],[21], but no information has been reported regarding the expression of these receptors at earlier stages of embryogenesis neither in the VM tissue, nor in midbrain NSPC in vitro. Therefore, using primers verified to amplify HA receptors [27], we examined whether these receptors were present in cultured VM NSPC, before and after differentiation. We found by RT-PCR that proliferative and differentiated VM NSPC cultures, isolated from E12 rat embryos, express transcripts of H1R, H2R, and three isoforms reported for the H3R, which are generated by alternative splicing [19],[21],[31],[32] (Figure 2A). The VM tissue without culture also had mRNAs for H1, H2 and H3 receptors (data not shown). We also performed Western blot analyses and found that the corresponding proteins for H1R and H2R were present in cultured NSPC (Figure 2B) and in VM tissue (data not shown). Even when the level of mRNA for H1 and H2 receptors was low, a single band was clearly detected, showing that H1R and H2R proteins are present in E12 cultured VM NSPC.

HA increases intracellular calcium in VM NSPC

Calcium (Ca2+) is an important regulator of various neuronal functions. During the formation of CNS, Ca2+ signaling is involved in the regulation of processes such as neural induction and differentiation of neural progenitors into neurons and glia, in a defined spatiotemporal pattern [33],[34]. Since the activation of H1R leads to the accumulation of Ca2+ as a second messenger, we tested if undifferentiated VM NSPC were responsive to HA stimulation by elevating intracellular Ca2+ concentration ([Ca2+]i). To address this hypothesis, we loaded cultured Nestin + VM NSPC with the ratiometric calcium probe Fura-2 AM, and measured the fluorescence changes resulting from alternating excitation of Fura-2 between 340 nm and 380 nm (R = F340/F380). We found a significant 3-fold increase in [Ca2+]i after the addition of 10 μM HA to VM NSPC, compared to the basal levels observed in control conditions (Figure 3), indicating that undifferentiated VM NSPC are responsive to HA. These results are consistent with our previous work on cerebrocortical NSPC [29], where 100 μM HA also caused calcium increases in VM cells through H1R activation.

HA increases apoptotic cell death of cultured undifferentiated midbrain NSPC without affecting cell proliferation

In order to examine the effects of HA on VM NSPC proliferation and apoptotic cell death, we treated cultures with increasing concentrations of HA (from 1 μM to 1 mM) in the presence of FGF-2 for 4 days. We used BrdU incorporation assays to analyze proliferation and the TUNEL assay to analyze apoptotic cell death. We calculated the percentage of BrdU- and TUNEL-positive cells relative to the total number of cells in control and HA-treated cultures and found that HA did not affect BrdU incorporation (Figure 4A and B). In contrast, 1 mM HA significantly increased the proportion of TUNEL-positive cells and this effect was prevented by co-incubation with the H1R antagonist chlorpheniramine (Figure 4C and D).

HA decreases the number of dopaminergic neurons in differentiated VM NSPC

To analyze whether HA has an effect on the differentiation of VM NSPC, cultures were treated for 6 days in the absence of FGF-2. Then, we evaluated by immunocytochemistry the expression of the neuronal marker β-III Tubulin. We found that 10 μM HA significantly increased the percentage of β-III Tubulin -positive cells from 21% to 37%, compared to control (Figure 5A and B). Additionally, we analyzed whether HA could have a role on dopaminergic differentiation by detecting the expression of the rate limiting enzyme for dopamine synthesis, TH, normalized by the number of neurons positive to β-III Tubulin. Figure panels 5C and D show that 1 mM HA significantly reduced the proportion of dopaminergic neurons when compared to controls. To establish if the decrease on TH-positive neurons was mediated by H1R activation, we treated cultures with chlorpheniramine, alone or in combination with 1 mM HA. We observed a significant recovery in the number of TH + cells when HA plus chlorpheniramine were added, compared to HA alone (Figure 5E and F).

In vivo experiments

Intrauterine administration of HA does not modify the morphology of the VM

In order to test if the effects of HA observed in vitro were also present in the developing brain in vivo, we performed in utero microinjections of HA dihydrochloride, which was used without neutralization. In these experiments, vehicle, HA or its receptor’s antagonists were injected directly into the ventricular lumen of E12 rat embryos. This stage was selected because it precedes the peak of neurogenesis in the midbrain. We initially injected 25 μg of HA and did not observe any change in neuronal differentiation relative to vehicle-injected embryos. We then administered 50 μg of HA, and analyzed the brains at E14. To evaluate if HA injection reached the VM region, we co-injected HA with a fluorescent tracer (Cell tracker), and found that the injected volume was enough to cover consistently the whole midbrain neuroepithelium (Figure 6A). There were no morphological differences between vehicle- and HA-injected embryos, assessed by hematoxylin-eosin staining in both coronal (Figure 6B) and sagittal (Figure 6C) sections of VM. To assess a general alteration of the VM, where TH-positive neurons are generated, the thickness of this region was measured bilaterally. No significant differences on the average VM thickness were found between vehicle- and HA-injected embryos (Figure 6D).

HA injection decreases dopaminergic induction in VM

Since we found that HA induced a decrease in TH + neurons after differentiation of cultured NSPC, we inquired whether this effect is also present in vivo. For this purpose, E12 rat embryos were injected in the cerebroventricular lumen either with vehicle or HA. Embryos were allowed to develop and then euthanized at E14. HA injection produced no decrease in TH staining without affecting β-III Tubulin-positive neurons (Figure 7A). These results suggest that HA is not interfering with neuronal differentiation since the normal pattern of β-III Tubulin-positive neurons was unaffected. In order to more quantitatively assess neuronal differentiation and the decline in dopaminergic neurons, β-III Tubulin and TH mRNA expression was evaluated by quantitative RT-PCR in vehicle- and HA-injected embryos. HA caused a significant decline in TH transcripts without affecting β-III Tubulin expression (Figure 7B). In addition, we analyzed by Western blot β-III Tubulin, TH and other markers specific for VM dopaminergic phenotype. HA induced a marked reduction in TH, Lmx1a, Lmx1b and Pitx3 protein levels relative to vehicle (Figure 7C). These experiments strongly support the notion that HA interferes with dopaminergic neuronal generation without modifying the proportion of neurons in the VM.

Injection of HA does not alter GABA or serotonin neurons

We also analyzed other neuronal phenotypes present in the developing brain in embryos injected at E12. GABAergic neurogenesis in the ventral and the dorsal regions of the midbrain begins simultaneously with dopaminergic neurogenesis [35]. GABAergic neurons express GAD65 and/or GAD67; GAD enzymes decarboxylate glutamate to produce GABA. Labeling of GAD65/67 with an antibody that recognizes both isoforms did not change after HA administration, when evaluated by immunohistochemistry (Figure 8A) and immunoblot (Figure 8B). The GAD staining pattern is restricted to the dorso-lateral regions of the midbrain compared to the ventral-most localization of dopaminergic neurons. The presence of serotoninergic neurons was identified by immunostaining against serotonin. The distribution of these neurons with or without HA was found to be in their normal arrangement in the most posterior part of the midbrain, close to the hindbrain, the cerebral regions where most serotoninergic neurons are generated (Figure 8C).

HA interferes with dopaminergic development only when administered at early stages

In order to establish whether embryos of different developmental stages were equally susceptible to HA actions as we observed at E14, we performed injections at E10, E12, E14 or E16 and analyzed TH expression in VM two days later in every case. We found that when HA was injected at E10 and E12, a marked decrease in TH staining was observed. In contrast, when HA was administered at E14 or E16, the TH expression pattern was indistinguishable from vehicle-injected embryos (Figure 9), strongly suggesting that HA is acting on neural precursors, and not on differentiated neurons.

HA decreases cell proliferation without altering cell death

After we found that HA induced a decrease in TH + neurons in vivo at early developmental stages, we asked whether this effect was produced by means of decreased proliferation or increased cell death. We injected embryos at E12 and coronal sections of vehicle- and HA-injected E14 embryos were stained with anti-phosphorylated histone H3 (pHH3), a marker for cells in M-phase. The pHH3 staining pattern of vehicle-injected embryos is consistent with previous reports [36]. However, the injection of HA caused significantly less pHH3-positive cells in the VM compared with vehicle-injected embryos (Figure 10A). We assessed apoptotic cell death in vivo and found no significant differences caused by HA administration (data not shown). We then decided to use Fluoro Jade staining to detect neuronal death, and found that HA injection did not modify the proportion of degenerating cells (Figure 10B). Thus, our results show that HA is affecting proliferation in vivo without modifying cell death in the VM.

The effect of HA on dopaminergic differentiation is mediated by H1R

In order to identify which receptor was responsible for the observed HA effects, injections of H1R or H2R antagonists, alone or in combination with HA, were performed. We analyzed the TH staining pattern in VM coronal sections of E14 embryos injected 2 days before. The H1R antagonist chlorpheniramine, injected at high doses (25 μg or 50 μg), markedly affected the normal development of the embryos, causing non-specific effects. When we injected 15 μg of chlorpheniramine, we observed a normal morphology of the treated embryos (Figure 11A). This latter dose of chlorpheniramine did not affect the expression of TH (Figure 11B); however, when co-injected with HA (Figure 12C), chlorpheniramine blocked HA-induced decrease in TH (Figure 12B), resulting in a TH staining similar to vehicle-injected embryos (Figure 12A). Cimetidine, a specific H2R antagonist, at the tested dose (50 μg), did not modify the normal embryo development nor TH staining, when injected without HA (Figure 11B). Co-injection of cimetidine with HA did not antagonize the effect triggered by HA (Figure 12D). These results demonstrate that the decrease of TH + cells is mediated through the activation of H1R.

Discussion

In this study we report that: i) midbrain NSPC express HA receptors; ii) HA stimulus increases intracellular calcium in proliferative VM NSPC; iii) one millimolar HA promotes apoptotic death of cultured VM NSPC by activating H1R; iv) Ten μM HA increases neuronal differentiation, but 1 mM HA decreases the number of differentiated dopaminergic neurons, in a H1R-dependent fashion, on VM NSPC cultures; v) HA injection at E12 decreases dopaminergic induction in VM without altering GABA or serotonin neurons in vivo; vi) HA interferes with dopaminergic development only when administered at early developmental stages; vii) HA impaired cell proliferation in the VM in vivo and viii) the effect of HA on dopaminergic differentiation is mediated by activation of H1R.
We found that H1 and H2 receptors are present at both mRNA and protein levels in proliferating and differentiated cultured VM NSPC. We showed that upon HA stimulation, undifferentiated cultures of VM NSPC significantly increased the cytoplasmic Ca2+ concentration. This data strongly suggest the presence of functional H1R coupled to the canonical calcium response in VM NSPC, consistent with the calcium responses evoked by 100 μM HA in cortical NSPC, which is blocked by chlorpheniramine [29], supporting the notion that H1R activation in undifferentiated NSPC occurs in both the forebrain and midbrain. It is well established that intracellular Ca2+ regulates neurogenesis as well as neurite/nerve growth and axonal pathfinding [34],[37],[38]. In cortical NSPC, H1R was responsible for the increased neuronal differentiation caused by HA. We explored the effect of HA on proliferation, cell death and differentiation of cultured midbrain NSPC. No significant differences on cell proliferation were found in HA-treated VM NSPC, when compared to control conditions. Regarding cell death, we observed an increasing effect of HA on apoptosis with 1 mM HA, which was clearly prevented by chlorpheniramine. We have previously reported that HA, through H1R activation, increases differentiation of E14 cortical NSPC to neurons that are positive for FOXP2, a marker of deep cortical layers [27],[29]. In contrast, in this study we found that in E12 VM NSPC, 10 μM HA increased neurogenesis and 1 mM HA affected DA neuron generation. These results highlight that the effects of HA vary depending on the studied brain region and the concentration of HA used. In VM NSPC, the decrease in TH + cells was mediated by H1R.
Previous studies have shown the expression patterns of HA-immunoreactive neurons and HRs at different stages of development [15]. HA-immunoreactive neurons are first detected on E13 in the regions of mesencephalon and metencephalon, this expression pattern remains until E17 and gradually decreases towards birth [15]. The expression patterns of HRs were reported at later stages of development (from E14 onwards) by in situ hybridization [18],[20]. We now show by RT-PCR and Western blot the presence of H1 and H2 receptors in E12 cultured VM NSPC.
Ultrasound guided-injection is a valuable technique that allows visualization and manipulation of embryos at specific developmental stages [39]–[41]. We did not find morphological differences between the vehicle- and HA-injected embryos, assessed by hematoxylin-eosin staining and the VM thickness, which allowed us to confidently proceed to analyze the effect of HA on midbrain development. We corroborated, by co-injections with Cell tracker, that HA reached all cells in the VM. When we evaluated neuronal differentiation by immunostaining, qRT-PCR and immunoblot, HA did not modify β-III Tubulin expression, suggesting that HA is not precluding neuronal differentiation. However, consistent with the data in VM NSPC cultures, when the limiting enzyme for DA synthesis, TH, was evaluated by immunohistochemistry, qRT-PCR and Western blot, we found significant decreases after HA application at E12, pointing out to a detrimental effect on dopaminergic neurons in vivo. These results suggest that HA might be affecting either DA specification and/or differentiation. To study if DA specification was being affected, we analyzed the expression of the Lim homeodomain genes Lmx1a and Lmx1b, which are specific markers for midbrain dopaminergic precursors. Lmx1a is an early marker of dividing mesencephalic precursors in the floor plate ventricular cells at early stages of development, instructing their differentiation into the midbrain dopaminergic phenotype, while Lmx1b is essential for the generation of properly differentiated midbrain dopaminergic neurons and its regional specification [42],[43]. HA significantly decreased the expression levels of both markers. Pitx3, a protein important for maintenance of dopaminergic identity, also decreased in HA-injected embryos, relative to vehicle-treated animals. To further support that HA is acting on DA precursors, E10 and E12 embryos were injected either with vehicle or HA. We demonstrated that the impairing effect of HA over the dopaminergic lineage is present in both developmental stages. In sharp contrast, DA neurons were unaffected when HA was applied at E14 or E16, when most cells already differentiated to dopaminergic neurons. These results strongly support that detrimental action of HA is on VM dopaminergic neural precursors rather than early differentiated neurons.
We found that HA did not modify BrdU incorporation in cultured VM NSPC, but significantly decreased the number of ventricular pHH3-positive cells in vivo. This different effect of HA on proliferation might be explained by the fact that BrdU was added to cultures for 3 h, and is incorporated by cells in the S phase of cell cycle, whereas pHH3 is present in cells undergoing mitoses. An additional factor to consider is that 100 μM HA caused a significant increase of BrdU incorporation in cultured cortical NSPC in the presence of daily-added 10 ng/ml FGF-2 [27], the same conditions used here for VM NSPC. A potentiation of thymidine incorporation has also been reported for tracheal smooth muscle human cells incubated with 10 ng/ml Epidermal Growth Factor and 10 μM HA, compared to the growth factor-only condition; this increased proliferation was associated to activation of Akt with participation of Gq/11 receptors [44]. It is possible that in cultured VM cells there are subpopulations that differentially respond to HA: some might decrease (i.e. dopaminergic precursors), while others (non-dopaminergic cells) could increase BrdU incorporation, resulting in no overall changes after HA addition. This possibility is supported by the fact that HA caused a significant decrease in the number of differentiated dopaminergic neurons both in vitro and in vivo.
We also analyzed the differentiation of other neuronal phenotypes present in midbrain-surrounding regions. Both serotoninergic and dopaminergic neurons arise from the ventral-most neuroepithelial progenitors, although at different rostro-caudal levels [45]. GABAergic cells are present in the substantia nigra pars reticulata and to some extent within the substantia nigra pars compacta and also in the ventral tegmental area [35]. Here, we demonstrate that the HA effect on midbrain development was selective for dopaminergic phenotype without affecting the generation of GABA neurons. In line with these findings, serotonin neurons, which are formed in the hindbrain very close to the midbrain-hindbrain boundary, were unaffected by HA.
In order to identify the cellular mechanism by which HA was exerting its action on decreasing TH + dopaminergic neurons, we co-injected HA with H1R or H2R antagonists. We found that only chlorpheniramine, a selective H1R antagonist, completely prevented the impairing action of HA over DA neurons. Our data support the hypothesis that HA action is on precursor dopaminergic cells by activation of H1R. We suggest that HA might be acting as a detrimental signal during early stages of DA specification. As we showed in vitro, incubation with HA leads to the elevation of Ca2+ in Nestin-positive VM NSPC, which might suggest that the mechanistic basis of HA effect over NSPC population affecting dopaminergic differentiation/specification involves intracellular Ca2+. Interestingly, Ca2+ signaling appears to control NSPC proliferation, regulating the switch from proliferation to neuronal differentiation [46]. For instance, in human NSPC, mobilization of IP3-dependent Ca2+ stores lengthens the cell cycle and increases the number of intermediate neural progenitors [34]. In this model, Ca2+ controls the duration of cell cycle by increasing the level of p53 protein, a known regulator of the cyclin-dependent kinase inhibitor p21 [47]. We have previously shown in cortical NSPC cultures the effect of HA on cell division patterns, evaluating asymmetric and symmetric divisions [28]. In the developing VM, proliferating cells are present in close to the ventricular zone [36]; we show that the number of mitotic cells is decreased by HA injection in vivo, and this could be a contributing mechanism for the decreased number of dopaminergic neurons.
The effects of artificially administering HA during VM development are clear, but we did not directly assess the participation of endogenous HA on dopaminergic development for example by decreasing its levels by inhibiting Histidine decarboxylase activity; nonetheless, the injection of H1 or H2 receptor antagonists alone did not modify TH staining, suggesting that the endogenous HA levels are not sufficient to affect dopaminergic development. Interestingly, the participation of HA in the pathogenesis of Parkinson’s disease in adult organisms has been suggested by previous work. For example, there is an increase in brain HA content in post-mortem putamen and substantia nigra[48] and HA blood levels are increased in parkinsonian patients [49]. Also, the number of histaminergic fibers innervating the substantia nigra is increased in individuals suffering Parkinson’s disease relative to controls [48],[50]. In addition to these suggestive data, there is experimental evidence showing that HA can affect differentiated dopaminergic neurons in the adult brain. Direct infusion of HA in the rodent substantia nigra leads to a selective damage of DA neurons [51]. One of the most widely used animal model for Parkinson’s disease is the intracerebral injection of the DA neurotoxin 6-hydroxydopamine in one hemisphere. When endogenous brain HA levels were increased or decreased, 6-hydroxydopamine-induced damage was potentiated or inhibited, respectively, assessed by TH immunostaining and by the rotational test after apomorphine administration. Supporting the participation of HA in the damage induced by this neurotoxin, the H1R antagonist pyrilamine, significantly, albeit transiently, decreased DA neuron degeneration; the H2R antagonist cimetidine, did not modify the 6-hydroxydopamine-induced damage [52].
Alternative mechanisms for the deleterious effects of HA on DA cells can include activation of microglial cells. In the adult substantia nigra injection of HA generates a microglial reaction, promoting dopaminergic cell death [51]. The possibility that microglia might contribute to DA degeneration in our model needs to be tested. There is evidence that microglia are present at this developmental stages [53], and that activated microglial cells reduce the number of neural precursor cells by phagocytosis in the developing cerebral cortex [54].

Conclusion

Our findings reveal an unexpected role of HA receptor activation during embryogenesis in a specific neuronal phenotype. These effects are relevant for its significance as a possible therapeutic target in CNS pathologies that involve the interaction between both histamine and dopamine systems. In conclusion, our study contributes to the knowledge of factors participating on the generation and degeneration of DA neurons and reveals novel actions of HA during midbrain development, which might lead to design new and more specific protocols for the generation of dopaminergic neurons and future therapies for Parkinson’s disease.

Methods

Cell culture

Multipotent NSPC were isolated from E12-E12.5 embryos obtained from time-pregnant Wistar rats as described [27]. For simplicity, we only use E12 to designate these embryos and the same criterion applies for other embryonic stages. VM was dissected in Krebs solution then mechanically dissociated. Cells were recovered by centrifugation. Isolated cells were cultured on plastic ware treated with 15 μg/ml poly-L-ornithine (Sigma) and 1 μg/mL human Fibronectin (Invitrogen) in N2 medium (DMEM/F12 1:1, supplemented with 25 μg/L of human insulin, 30 nM sodium selenite, 100 μM putrescine, 20 nM progesterone and 100 mg/L of apotransferrin) containing 10 ng/mL of basic fibroblast growth factor (FGF-2; Peprotech). Cultures were maintained at 37°C in humidified 5% CO2/95% air atmosphere. These cultures were considered as passage (P) 0. When cultures reached 80% confluence, cells were detached with 0.1 mM EDTA in phosphate buffered saline (PBS) and re-plated to P1. Cells were maintained under proliferation conditions for 4 days in control (N2 medium + 10 ng/ml FGF-2) and experimental conditions that vary concentrations of HA (from 1 μM to 1 mM) with or without HA receptor antagonists. Differentiation was promoted by removal of FGF-2 during 6 additional days in control and experimental conditions. HA antagonists for H1R (1 μM chlorpheniramine, Sigma), H2R (30 μM cimetidine, Sigma) were added daily to HA-treated cells during proliferation and differentiation stages. For immunocytochemistry analysis, BrdU and TUNEL assays, cells were seeded at 1.5×104/well in 24-well plates (Corning). For RNA extraction, cells were seeded at 3×105/well in 6-well plates (Corning). HA dihydrochloride was used without neutralization.

RNA extraction, RT-PCR and Real-Time RT-PCR

Total RNA was isolated from proliferating cultures, differentiated cultures, and midbrain tissue from ultrasound injected and control embryos using TRIZOL reagent (Invitrogen). One μg of RNA was reverse transcribed using dNTPs, 100 U of Superscript III Reverse Transcriptase, 50 μM of random hexamers and 40 U of Recombinant RNase Inhibitor (Invitrogen). PCR reactions were performed using 100 ng of cDNA, 2 U of recombinant Taq DNA polymerase (Invitrogen), 20 pmol of each oligonucleotide, 500 μM dNTPs (Sigma) and 1.5 mM MgCl2. PCR products were analyzed by electrophoresis to determine their molecular weights after ethidium bromide staining. The sequences of the PCR products were verified and correspond to the indicated HA receptor. Reactions with RNA in the absence of retrotranscription were included as negative controls. RNA extracted from adult rat cerebral cortex was used as a positive control for HA receptors. The forward (F) and reverse (R) primers listed in Table 1 were used.
Table 1
Primer sequences for detection of histaminergic receptors by end-point RT-PCR
Gene
Sense primer
Antisense primer
H1R
F: 5′-CTTCTACCTCCCCACTTTgCT-3′
R: 5′-TTCCCTTTCCCCCTCTTg-3′
H2R
F: 5′-TTCTTggACTCCTggTgCTgC-3′
R: 5′-CATgCCCCCTCTggTCCC-3′
H3R
F: 5′-CCAgAACCCCCACCAgATg-3′
R: 5′-CCAgCAgAgCCCAAAgATg-3′
These primers have been used previously by our group [27] and after retrotranscription and amplification, PCR products were extracted from the gel and sequenced, confirming that these sequences correspond to H1, H2 and the 3 isoforms of H3 receptors.
For qRT-PCR [55], we used 100 ng of the synthesized cDNA, 2.5 U of recombinant Taq DNA polymerase (Invitrogen), 0.4 mM of each oligonucleotide and dNTPs, 1.5 mM MgCl2 and 5 μM Syto 9 (Invitrogen). Quantitative analysis of cDNA amplification was performed using the Rotor Gene 6000 System (Corbett Life Science). Melting curves were performed in every reaction to ensure that only one product was present. GAPDH amplification was used as loading control for all reactions and every transcript value was normalized to GAPDH expression. The 2exp (−ΔΔCt) method was used to determine the relative expression of each gene compared to control condition. The primers listed in Table 2 were used.
Table 2
Primer sequences for detection of transcripts by qRT-PCR
Gene
Sense primer
Antisense primer
β-III Tubulin
F: 5′-ggCCTTTggACACCTATTCA-3′
R: 5′-TgCAggCAgTCACAATTCTC-3′
TH
F: 5′-CCACTggAggCTgTggTATT-3′
R: 5′-CCgggTCTCTAAgTggTgAA-3′
GAPDH
F: 5′-gTggACCTCATggCCTACAT-3′
R: 5′-ggATggAATTgTgAgggAgA-3′

Western blot

Cells were cultured and seeded at 3.5×104/well in 6-well plates (Corning) in control and experimental conditions, following standard protocols [27],[56]. Briefly, cells were homogenized in lysis buffer (25 mM Tris–HCl pH 7, 1% IGEPAL, 100 mM NaCl) supplemented with protease inhibitors (Complete mini, Roche). Proteins were obtained by centrifugation at 13 800 g at 4°C for 15 min, and quantified by Bradford assay (BioRad). Proteins (30 μg) were resolved on 8% sodium dodecyl sulfate–polyacrylamide gel electrophoresis and transferred to nitrocellulose membranes (Amersham Bioscience). The membranes were blocked with 5% non-fat milk and incubated overnight with primary antibodies. Pre-stained markers (Invitrogen) were included for size determination. The following antibodies were used: rabbit anti-rat H1R polyclonal antibody (1:1500, Santa Cruz Biotechnology); goat anti-rat H2R polyclonal antibody (1:1500, Santa Cruz Biotechnology); rabbit anti-Lmx1a polyclonal antibody (1:100, Abcam); rabbit anti-Lmx1b polyclonal antibody (1:100, Abcam); rabbit polyclonal anti-tyrosine hydroxylase (TH, 1:1000; Pel-freez) and rabbit anti-Pitx3 polyclonal antibody (1:500, Zymed). Membranes were washed and incubated with corresponding horseradish peroxidase-coupled secondary antibodies (1:10 000, Santa Cruz Biotechnology) for 2 hours. Immunoreactive bands were detected using enhanced chemiluminescence method (Amersham Bioscience) and developed on photographic film (Kodak).

Immunocytochemistry

Cells were fixed with 4% paraformaldehyde in PBS, pH 7.4 at the end of proliferation and differentiation stages. Immunocytochemistry assays were performed following standard protocols [57]. Briefly, cells were permeabilized and blocked for 1 hour with 0.3% Triton X-100 and 10% normal goat serum (NGS) in PBS. Cells were incubated overnight at 4°C with the following primary antibodies diluted in PBS containing 10% NGS: rabbit polyclonal anti-β-III Tubulin (1:2000, Covance); rabbit polyclonal anti-glial fibrillary acidic protein (GFAP; 1:2000, DAKO); mouse monoclonal antibody anti-microtubule associated protein 2 (MAP2; 1:500, Chemicon); mouse monoclonal anti-Nestin (1:100, Developmental Studies Hybridoma Bank); rabbit anti-Nestin (1:1000, Covance); rabbit polyclonal anti-TH (1:1000; Pel-freez); rabbit polyclonal anti-Sox2 (1:200; Millipore). Appropriate Alexa-Fluor secondary antibodies were used diluted in PBS (1:1000; Molecular Probes) containing 10% NGS. Nuclei were stained with Hoechst 33258 (1 ng/mL; Sigma). Immunostainings were analyzed with an epifluorescence microscope (Nikon, Eclipse TE2000-U) and photographed with a Nikon digital camera (DMX1200 F). Negative controls were performed in the absence of primary antibodies and showed no unspecific staining.

BrdU incorporation assay

Cells treated with FGF-2 were incubated for 3 hours with 10 μM 5-bromo-2-deoxyuridine (BrdU, Roche), washed with medium and fixed with 4% paraformaldehyde in PBS 21 h later [27]. After fixation, cells were incubated with 1 N HCl for 30 min at 25°C, and neutralized by washing with 0.1 M borate buffer, pH 8.5. Preparations were blocked for 1 h with 10% NGS (Invitrogen) and 0.3% Triton X-100 (Sigma) in PBS. Cells were incubated overnight at 4°C with monoclonal rat anti-BrdU antibody (1:100; Accurate) followed by three washes with 1% bovine serum albumin in PBS. Alexa-Fluor 488 goat anti-rat IgG was added (1:1000, Invitrogen) in blocking solution for 1 h at RT and washed three times with PBS. Nuclei were stained with Hoechst 33258 (1 ng/mL; Sigma).

TUNEL assay

Proliferating and differentiated cells were fixed with 4% paraformaldehyde in PBS, and subsequently washed three times with PBS, pH = 7.4 [27]. For detection of apoptotic cells, the Terminal deoxytransferase-mediated deoxyuridine triphosphate nick-end labeling kit (TUNEL, InSitu Cell Death Detection Kit, Roche Diagnostics) was used to evaluate the effect of HA on cell death. Apoptotic cells were visualized by fluorescence microscopy.

Fluoro Jade staining

Fluoro Jade was used to label and quantify degenerating neurons in the VM. The sections were mounted on slides and air-dried. The slides were dipped in absolute ethanol for 3 min followed by 1 min in 70% ethanol and washed in deionized water for 1 min. Tissue was then oxidized with a 0.06% KMnO4 solution diluted in PBS for 15 min, followed by three rinses with deionized water. From this point onwards, the slides were protected from light. Tissue sections were stained with a 0.001% solution of Fluoro-Jade in 0.1% acetic acid for 30 min. Slides were rinsed with water and air-dried. The slides were cleared in xylene and coverslipped with DPX mounting medium [58],[59].

Cell counting

Cell counts from BrdU, immunocytochemistry, TUNEL, phosphorylated histone H3 (pHH3) and Fluoro Jade experiments were performed from digital pictures. Quantification of cells was performed by using the ImageJ software considering the total number of Hoechst-stained nuclei (total cells), in at least 10 random fields from a minimum of 3 independent experiments.

Intracellular Ca2+ imaging

Proliferating VM NSPC in suspension were loaded with the Ca2+-sensitive fluorescence indicator Fura-2 AM (Invitrogen) at a final concentration of 5 μM at 37°C for 45 min [29]. Loaded cells were centrifuged and seeded at 1×105 cells on 3.5 cm glass bottom dishes (MatTek Corp) treated with 15 mg/mL poly-L-ornithine (Sigma) and 1 μg/ml human Fibronectin (Invitrogen) in N2 medium. After 2 hours, attached cells were incubated in Hank’s solution (137 mM NaCl, 5 mM KCl, 0.3 mM NaH2PO4, 0.8 mM MgSO4, 1 mM MgCl2, 5 mM glucose, 2 mM glutamine, 10 mM Tris–HCl without CaCl2; pH 7.4) at 37°C for 10 minutes. Subsequently, Hank’s solution was removed and new solution with 5 mM CaCl2 was added. UV excitation was set to alternate between 340 and 380 nm and the emission was monitored at 510 nm in a Carl Zeiss microscope. Images and the emission ratio generated between both excitations wavelengths (R = E 340/E 380) were obtained and analyzed using AxioVision software.

Ultrasound guided injections

We used timed-pregnant Wistar rats with embryos at gestational ages E10, E12, E14 and E16 to perform ultrasound-guided injections. All animal procedures were approved by the Instituto de Fisiología Celular Animal Use and Care Committee and conformed to Mexican and NIH guidelines. All injections were performed by duplicate or triplicate in different embryos from the same dam, and repeated at least in 3 different pregnant rats. Control and experimental animals were included in the same pregnant rat, but to make sure that the different conditions do not mix, each group was evaluated in a single uterine horn. The results were reproducible when we analyzed embryos from different dams. To perform ultrasound guided injections [40], we used an MHF-1 Ultraview UltraSound System (E-Technologies) with a focal distance of 7 mm for imaging. Rats were deeply anesthetized in a chamber and then maintained with a facemask to administer inhaled sevoflurane (5% for induction and 1% for maintenance). We performed midline laparotomy, uterine horns were carefully exposed and the number of embryos was recorded. Intraventricular injections were performed through the uterine wall using pulled microcapillaries (borosilicate glass, Sutter Instruments). The glass needles were visualized in the ultrasound imaging system and the verification of the site of injection in the lateral ventricles was made during the procedure by observing the liquid going out of the needle. HA, chlorpheniramine (H1R antagonist) or cimetidine (H2R antagonist) were injected on the ventricular lumen at different doses in a final volume of 2 μl aided by an automatic injector (Quintessential injector, Stoelting). The injected amount was of HA dihydrochloride, which was used without neutralization. DMSO and Cell Tracker (Invitrogen) injections were performed as controls. Once the desired number of embryos was injected, the uterus was repositioned in its physiological site. The incisions of the abdominal muscle and skin were stitched with separate sutures. Finally we applied Buprenorphine as analgesic (0.1 mg/kg, Pisa Laboratories) and animals were monitored and left to recover. Two days after the injections, pregnant rats were euthanized by overdose of the anesthetic Sodium pentobarbital (Pfizer), and identified embryos were recovered for analysis. We performed the following number of successfully injected E12 embryos: vehicle: 48; HA: 12; chlorpheniramine: 9; cimetidine: 9; HA plus chlorpheniramine: 9; HA plus cimetidine: 9. The number of dams used was 18.

Immunohistochemistry

Embryos were removed and transferred to PBS. Subsequently, the embryos were fixed by immersion overnight in 4% PFA in PBS at 4°C. Thereafter, the tissues were cryoprotected overnight in 30% (v/v) sucrose, embedded in Tissue Freezing Medium (Tissue-Tek, Zakura Finetek) and frozen. Sagittal and coronal sections (20 μm) were obtained on a cryostat (Leica). For immunohistochemistry, slides were rehydrated in PBS. Tissue sections were permeabilized and blocked for 1 hour at RT with 0.3% Triton X-100 and 10% NGS in PBS. Slides were incubated overnight at 4°C with the following primary antibodies diluted in PBS containing 10% NGS: rabbit polyclonal anti-pHH3 (1:250, Cell Signaling), rabbit polyclonal anti-β-III Tubulin (1:1000, Babco Covance); mouse monoclonal anti-Nestin (1:100, Developmental Studies Hybridoma Bank); rabbit polyclonal anti-tyrosine hydroxylase (TH; 1:1000, Pel-freez); rabbit polyclonal anti-GAD65/67 (1:500, Millipore); rabbit polyclonal anti-serotonin (1:500, Sigma). Subsequently, the sections were washed three times for 5 minutes in PBS and incubated with the secondary antibodies Alexa-Fluor 488 anti-rabbit IgG and Alexa-Fluor 568 anti-mouse IgG (1:500; Molecular Probes) in PBS containing 10% NGS for 2 hours. Nuclei were stained with Hoechst 33258 (1 ng/mL). After three further washes in PBS, the sections were mounted in Aqua Poly/Mount (Polysciences, Inc.) Immunostainings were analyzed with an epifluorescence microscope (Nikon, Eclipse TE2000-U) and photographed with a Nikon digital camera (DMX1200 F). Some images were acquired with a FV1000 Olympus confocal microscope. Negative controls were performed in the absence of primary antibodies and showed no unspecific staining.

Statistical analysis

We performed 3–5 independent experiments, defined as duplicate or triplicate experiments from a different pregnant rat. All data are presented as mean ± standard error of mean (S.E.M). One-way ANOVA was performed for statistical analysis and multiple comparisons between treated and control groups were made using the post-hoc Student-Newman-Keuls test. Differences were considered statistically significant at P ≤ 0.05. Graphs were generated using GraphPad Prism software.

Acknowledgements

This work was supported by grants from Dirección General de Asuntos del Personal Académico, Universidad Nacional Autónoma de México (Papiit IN208713), Consejo Nacional de Ciencia y Tecnología (CONACyT 131281) to I.V., and by the Ministerio de Ciencia e Innovación de España (Grant BFU2010-21377) to J.A.D.C. Itzel Escobedo-Avila, Rodrigo López-González and Daniel Cortés received a graduate fellowship from CONACyT. Fernanda Vargas-Romero received an undergraduate fellowship from DGAPA (IN208713). We thank Drs. Rosa E. Navarro and Jesús Santa-Olalla for continuous discussion and advice. The technical assistance of the microscopy unit of Instituto de Fisiología Celular is appreciated. The Nestin monoclonal antibody developed by S. Hockfield was obtained from the Developmental Studies Hybridoma Bank developed under the auspices of the NICHD and maintained at The University of Iowa, Department of Biology, Iowa City, IA 52242, USA. We thank Dr. Daniel Hoeppner for critical reading of the manuscript.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made.
The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
The Creative Commons Public Domain Dedication waiver (https://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Competing interests

The authors declare no competing financial interests.

Authors’ contributions

IEA wrote the manuscript, analyzed the data and performed most of the experiments with help from FVR. AMH contributed to the design of experiments and discussion of results. RLG helped to perform intrauterine injections, analysis and discussion of results. DC and JADC contributed to set up techniques reviewed the manuscript and took part in the discussion of results. All authors read and approved the final manuscript. IV contributed to experiment design/supervision, reviewed the manuscript and obtained funding.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Neurologie & Psychiatrie

Kombi-Abonnement

Mit e.Med Neurologie & Psychiatrie erhalten Sie Zugang zu CME-Fortbildungen der Fachgebiete, den Premium-Inhalten der dazugehörigen Fachzeitschriften, inklusive einer gedruckten Zeitschrift Ihrer Wahl.

e.Med Neurologie

Kombi-Abonnement

Mit e.Med Neurologie erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes, den Premium-Inhalten der neurologischen Fachzeitschriften, inklusive einer gedruckten Neurologie-Zeitschrift Ihrer Wahl.

Literatur
1.
Zurück zum Zitat Wada H, Inagaki N, Yamatodani A, Watanabe T: Is the histaminergic neuron system a regulatory center for whole-brain activity?. Trends Neurosci. 1991, 14 (9): 415-418. 10.1016/0166-2236(91)90034-R. SepPubMedCrossRef Wada H, Inagaki N, Yamatodani A, Watanabe T: Is the histaminergic neuron system a regulatory center for whole-brain activity?. Trends Neurosci. 1991, 14 (9): 415-418. 10.1016/0166-2236(91)90034-R. SepPubMedCrossRef
2.
Zurück zum Zitat Onodera K, Yamatodani A, Watanabe T, Wada H: Neuropharmacology of the histaminergic neuron system in the brain and its relationship with behavioral disorders. Prog Neurobiol. 1994, 42 (6): 685-702. 10.1016/0301-0082(94)90017-5. AprPubMedCrossRef Onodera K, Yamatodani A, Watanabe T, Wada H: Neuropharmacology of the histaminergic neuron system in the brain and its relationship with behavioral disorders. Prog Neurobiol. 1994, 42 (6): 685-702. 10.1016/0301-0082(94)90017-5. AprPubMedCrossRef
3.
Zurück zum Zitat Haas H, Panula P: The role of histamine and the tuberomamillary nucleus in the nervous system. Nat Rev Neurosci. 2003, 4 (2): 121-130. 10.1038/nrn1034. FebPubMedCrossRef Haas H, Panula P: The role of histamine and the tuberomamillary nucleus in the nervous system. Nat Rev Neurosci. 2003, 4 (2): 121-130. 10.1038/nrn1034. FebPubMedCrossRef
4.
Zurück zum Zitat Haas HL, Sergeeva OA, Selbach O: Histamine in the nervous system. Physiol Rev. 2008, 88 (3): 1183-1241. 10.1152/physrev.00043.2007. JulPubMedCrossRef Haas HL, Sergeeva OA, Selbach O: Histamine in the nervous system. Physiol Rev. 2008, 88 (3): 1183-1241. 10.1152/physrev.00043.2007. JulPubMedCrossRef
5.
Zurück zum Zitat Panula P, Nuutinen S: The histaminergic network in the brain: basic organization and role in disease. Nat Rev Neurosci. 2013, 14 (7): 472-487. 10.1038/nrn3526. JunPubMedCrossRef Panula P, Nuutinen S: The histaminergic network in the brain: basic organization and role in disease. Nat Rev Neurosci. 2013, 14 (7): 472-487. 10.1038/nrn3526. JunPubMedCrossRef
6.
Zurück zum Zitat Watanabe T, Yanai K: Studies on functional roles of the histaminergic neuron system by using pharmacological agents, knockout mice and positron emission tomography. Tohoku J Exp Med. 2001, 195 (4): 197-217. 10.1620/tjem.195.197. DecPubMedCrossRef Watanabe T, Yanai K: Studies on functional roles of the histaminergic neuron system by using pharmacological agents, knockout mice and positron emission tomography. Tohoku J Exp Med. 2001, 195 (4): 197-217. 10.1620/tjem.195.197. DecPubMedCrossRef
7.
Zurück zum Zitat Leurs R, Smit MJ, Timmerman H: Molecular pharmacological aspects of histamine receptors. Pharmacol Ther. 1995, 66 (3): 413-463. 10.1016/0163-7258(95)00006-3. JunPubMedCrossRef Leurs R, Smit MJ, Timmerman H: Molecular pharmacological aspects of histamine receptors. Pharmacol Ther. 1995, 66 (3): 413-463. 10.1016/0163-7258(95)00006-3. JunPubMedCrossRef
8.
Zurück zum Zitat Hill SJ, Ganellin CR, Timmerman H, Schwartz JC, Shankley NP, Young JM, Schunack W, Levi R, Haas HL: International Union of Pharmacology, XIII. Classification of histamine receptors. Pharmacol Rev. 1997, 49 (3): 253-278. SepPubMed Hill SJ, Ganellin CR, Timmerman H, Schwartz JC, Shankley NP, Young JM, Schunack W, Levi R, Haas HL: International Union of Pharmacology, XIII. Classification of histamine receptors. Pharmacol Rev. 1997, 49 (3): 253-278. SepPubMed
9.
Zurück zum Zitat Nakamura T, Itadani H, Hidaka Y, Ohta M, Tanaka K: Molecular cloning and characterization of a new human histamine receptor, HH4R. Biochem Biophys Res Commun. 2000, 279 (2): 615-620. 10.1006/bbrc.2000.4008. Dec 20PubMedCrossRef Nakamura T, Itadani H, Hidaka Y, Ohta M, Tanaka K: Molecular cloning and characterization of a new human histamine receptor, HH4R. Biochem Biophys Res Commun. 2000, 279 (2): 615-620. 10.1006/bbrc.2000.4008. Dec 20PubMedCrossRef
10.
Zurück zum Zitat Liu C, Ma X, Jiang X, Wilson SJ, Hofstra CL, Blevitt J, Pyati J, Li X, Chai W, Carruthers N, Lovenberg TW: Cloning and pharmacological characterization of a fourth histamine receptor (H (4)) expressed in bone marrow. Mol Pharmacol. 2001, 59 (3): 420-426. MarPubMed Liu C, Ma X, Jiang X, Wilson SJ, Hofstra CL, Blevitt J, Pyati J, Li X, Chai W, Carruthers N, Lovenberg TW: Cloning and pharmacological characterization of a fourth histamine receptor (H (4)) expressed in bone marrow. Mol Pharmacol. 2001, 59 (3): 420-426. MarPubMed
11.
Zurück zum Zitat Lethbridge NL, Chazot PL: Immunological identification of the mouse H4 histamine receptor on spinal cord motor neurons using a novel anti-mouse H4R antibody. Inflamm Res. 2010, 59 (Suppl 2): S197-198. 10.1007/s00011-009-0127-2. MarPubMedCrossRef Lethbridge NL, Chazot PL: Immunological identification of the mouse H4 histamine receptor on spinal cord motor neurons using a novel anti-mouse H4R antibody. Inflamm Res. 2010, 59 (Suppl 2): S197-198. 10.1007/s00011-009-0127-2. MarPubMedCrossRef
12.
Zurück zum Zitat Pearce LA, Schanberg SM: Histamine and spermidine content in brain during development. Science. 1969, 166 (3910): 1301-1303. 10.1126/science.166.3910.1301. Dec 5PubMedCrossRef Pearce LA, Schanberg SM: Histamine and spermidine content in brain during development. Science. 1969, 166 (3910): 1301-1303. 10.1126/science.166.3910.1301. Dec 5PubMedCrossRef
13.
Zurück zum Zitat Vanhala A, Yamatodani A, Panula P: Distribution of histamine-, 5-hydroxytryptamine-, and tyrosine hydroxylase-immunoreactive neurons and nerve fibers in developing rat brain. J Comp Neurol. 1994, 347 (1): 101-114. 10.1002/cne.903470108. Sep 1PubMedCrossRef Vanhala A, Yamatodani A, Panula P: Distribution of histamine-, 5-hydroxytryptamine-, and tyrosine hydroxylase-immunoreactive neurons and nerve fibers in developing rat brain. J Comp Neurol. 1994, 347 (1): 101-114. 10.1002/cne.903470108. Sep 1PubMedCrossRef
14.
Zurück zum Zitat Panula P, Lintunen M, Karlstedt K: Histamine in brain development and tumors. Semin Cancer Biol. 2000, 10 (1): 11-14. 10.1006/scbi.2000.0302. FebPubMedCrossRef Panula P, Lintunen M, Karlstedt K: Histamine in brain development and tumors. Semin Cancer Biol. 2000, 10 (1): 11-14. 10.1006/scbi.2000.0302. FebPubMedCrossRef
15.
Zurück zum Zitat Auvinen S, Panula P: Development of histamine-immunoreactive neurons in the rat brain. J Comp Neurol. 1988, 276 (2): 289-303. 10.1002/cne.902760211. Oct 8PubMedCrossRef Auvinen S, Panula P: Development of histamine-immunoreactive neurons in the rat brain. J Comp Neurol. 1988, 276 (2): 289-303. 10.1002/cne.902760211. Oct 8PubMedCrossRef
16.
Zurück zum Zitat Nissinen MJ, Panula P: Developmental patterns of histamine-like immunoreactivity in the mouse. J Histochem Cytochem. 1995, 43 (2): 211-227. 10.1177/43.2.7822777. FebPubMedCrossRef Nissinen MJ, Panula P: Developmental patterns of histamine-like immunoreactivity in the mouse. J Histochem Cytochem. 1995, 43 (2): 211-227. 10.1177/43.2.7822777. FebPubMedCrossRef
17.
Zurück zum Zitat Panula P, Sundvik M, Karlstedt K: Developmental roles of brain histamine. Trends Neurosci. 2014, 37 (3): 159-168. 10.1016/j.tins.2014.01.001. MarPubMedCrossRef Panula P, Sundvik M, Karlstedt K: Developmental roles of brain histamine. Trends Neurosci. 2014, 37 (3): 159-168. 10.1016/j.tins.2014.01.001. MarPubMedCrossRef
18.
Zurück zum Zitat Kinnunen A, Lintunen M, Karlstedt K, Fukui H, Panula P: In situ detection of H1-receptor mRNA and absence of apoptosis in the transient histamine system of the embryonic rat brain. J Comp Neurol. 1998, 394 (1): 127-137. 10.1002/(SICI)1096-9861(19980427)394:1<127::AID-CNE10>3.0.CO;2-L. Apr 27PubMedCrossRef Kinnunen A, Lintunen M, Karlstedt K, Fukui H, Panula P: In situ detection of H1-receptor mRNA and absence of apoptosis in the transient histamine system of the embryonic rat brain. J Comp Neurol. 1998, 394 (1): 127-137. 10.1002/(SICI)1096-9861(19980427)394:1<127::AID-CNE10>3.0.CO;2-L. Apr 27PubMedCrossRef
19.
Zurück zum Zitat Heron A, Rouleau A, Cochois V, Pillot C, Schwartz JC, Arrang JM: Expression analysis of the histamine H (3) receptor in developing rat tissues. Mech Dev. 2001, 105 (1–2): 167-173. 10.1016/S0925-4773(01)00389-6. JulPubMedCrossRef Heron A, Rouleau A, Cochois V, Pillot C, Schwartz JC, Arrang JM: Expression analysis of the histamine H (3) receptor in developing rat tissues. Mech Dev. 2001, 105 (1–2): 167-173. 10.1016/S0925-4773(01)00389-6. JulPubMedCrossRef
20.
Zurück zum Zitat Karlstedt K, Nissinen M, Michelsen KA, Panula P: Multiple sites of L-histidine decarboxylase expression in mouse suggest novel developmental functions for histamine. Dev Dyn. 2001, 221 (1): 81-91. 10.1002/dvdy.1127. MayPubMedCrossRef Karlstedt K, Nissinen M, Michelsen KA, Panula P: Multiple sites of L-histidine decarboxylase expression in mouse suggest novel developmental functions for histamine. Dev Dyn. 2001, 221 (1): 81-91. 10.1002/dvdy.1127. MayPubMedCrossRef
21.
Zurück zum Zitat Karlstedt K, Ahman MJ, Anichtchik OV, Soinila S, Panula P: Expression of the H3 receptor in the developing CNS and brown fat suggests novel roles for histamine. Mol Cell Neurosci. 2003, 24 (3): 614-622. 10.1016/S1044-7431(03)00208-2. NovPubMedCrossRef Karlstedt K, Ahman MJ, Anichtchik OV, Soinila S, Panula P: Expression of the H3 receptor in the developing CNS and brown fat suggests novel roles for histamine. Mol Cell Neurosci. 2003, 24 (3): 614-622. 10.1016/S1044-7431(03)00208-2. NovPubMedCrossRef
22.
Zurück zum Zitat Nissinen MJ, Karlstedt K, Castren E, Panula P: Expression of histidine decarboxylase and cellular histamine-like immunoreactivity in rat embryogenesis. J Histochem Cytochem. 1995, 43 (12): 1241-1252. 10.1177/43.12.8537641. DecPubMedCrossRef Nissinen MJ, Karlstedt K, Castren E, Panula P: Expression of histidine decarboxylase and cellular histamine-like immunoreactivity in rat embryogenesis. J Histochem Cytochem. 1995, 43 (12): 1241-1252. 10.1177/43.12.8537641. DecPubMedCrossRef
23.
Zurück zum Zitat Wiener Z, Toth S, Gocza E, Kobolak J, Falus A: Mouse embryonic stem cells express histidine decarboxylase and histamine H1 receptors. Inflamm Res. 2003, 52 (Suppl 1): S53-54. 10.1007/s000110300052. AprPubMedCrossRef Wiener Z, Toth S, Gocza E, Kobolak J, Falus A: Mouse embryonic stem cells express histidine decarboxylase and histamine H1 receptors. Inflamm Res. 2003, 52 (Suppl 1): S53-54. 10.1007/s000110300052. AprPubMedCrossRef
24.
25.
Zurück zum Zitat Specht LA, Pickel VM, Joh TH, Reis DJ: Light-microscopic immunocytochemical localization of tyrosine hydroxylase in prenatal rat brain.I. Early ontogeny. J Comp Neurol. 1981, 199 (2): 233-253. 10.1002/cne.901990207. Jun 20PubMedCrossRef Specht LA, Pickel VM, Joh TH, Reis DJ: Light-microscopic immunocytochemical localization of tyrosine hydroxylase in prenatal rat brain.I. Early ontogeny. J Comp Neurol. 1981, 199 (2): 233-253. 10.1002/cne.901990207. Jun 20PubMedCrossRef
26.
Zurück zum Zitat Vitalis T, Cases O, Parnavelas JG: Development of the dopaminergic neurons in the rodent brainstem. Exp Neurol. 2005, 191 (Suppl 1): S104-112. 10.1016/j.expneurol.2004.05.044. FebPubMedCrossRef Vitalis T, Cases O, Parnavelas JG: Development of the dopaminergic neurons in the rodent brainstem. Exp Neurol. 2005, 191 (Suppl 1): S104-112. 10.1016/j.expneurol.2004.05.044. FebPubMedCrossRef
27.
Zurück zum Zitat Molina-Hernandez A, Velasco I: Histamine induces neural stem cell proliferation and neuronal differentiation by activation of distinct histamine receptors. J Neurochem. 2008, 106 (2): 706-717. 10.1111/j.1471-4159.2008.05424.x. JulPubMedCrossRef Molina-Hernandez A, Velasco I: Histamine induces neural stem cell proliferation and neuronal differentiation by activation of distinct histamine receptors. J Neurochem. 2008, 106 (2): 706-717. 10.1111/j.1471-4159.2008.05424.x. JulPubMedCrossRef
28.
Zurück zum Zitat Rodriguez-Martinez G, Velasco I, Garcia-Lopez G, Solis KH, Flores-Herrera H, Diaz NF, Molina-Hernandez A: Histamine is required during neural stem cell proliferation to increase neuron differentiation. Neuroscience. 2012, 216: 10-17. 10.1016/j.neuroscience.2012.04.049. Aug 2PubMedCrossRef Rodriguez-Martinez G, Velasco I, Garcia-Lopez G, Solis KH, Flores-Herrera H, Diaz NF, Molina-Hernandez A: Histamine is required during neural stem cell proliferation to increase neuron differentiation. Neuroscience. 2012, 216: 10-17. 10.1016/j.neuroscience.2012.04.049. Aug 2PubMedCrossRef
29.
Zurück zum Zitat Molina-Hernandez A, Rodriguez-Martinez G, Escobedo-Avila I, Velasco I: Histamine up-regulates fibroblast growth factor receptor 1 and increases FOXP2 neurons in cultured neural precursors by histamine type 1 receptor activation: conceivable role of histamine in neurogenesis during cortical development in vivo. Neural Dev. 2013, 8: 4-10.1186/1749-8104-8-4.PubMedPubMedCentralCrossRef Molina-Hernandez A, Rodriguez-Martinez G, Escobedo-Avila I, Velasco I: Histamine up-regulates fibroblast growth factor receptor 1 and increases FOXP2 neurons in cultured neural precursors by histamine type 1 receptor activation: conceivable role of histamine in neurogenesis during cortical development in vivo. Neural Dev. 2013, 8: 4-10.1186/1749-8104-8-4.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Bernardino L, Eiriz MF, Santos T, Xapelli S, Grade S, Rosa AI, Cortes L, Ferreira R, Braganca J, Agasse F, Ferreira L, Malva JO: Histamine stimulates neurogenesis in the rodent subventricular zone. Stem Cells. 2012, 30 (4): 773-784. 10.1002/stem.1042. AprPubMedCrossRef Bernardino L, Eiriz MF, Santos T, Xapelli S, Grade S, Rosa AI, Cortes L, Ferreira R, Braganca J, Agasse F, Ferreira L, Malva JO: Histamine stimulates neurogenesis in the rodent subventricular zone. Stem Cells. 2012, 30 (4): 773-784. 10.1002/stem.1042. AprPubMedCrossRef
31.
Zurück zum Zitat Smit MJ, Timmerman H, Alewijnse AE, Punin M, van den Nieuwenhof I, Blauw J, van Minnen J, Leurs R: Visualization of agonist-induced internalization of histamine H2 receptors. Biochem Biophys Res Commun. 1995, 214 (3): 1138-1145. 10.1006/bbrc.1995.2404. Sep 25PubMedCrossRef Smit MJ, Timmerman H, Alewijnse AE, Punin M, van den Nieuwenhof I, Blauw J, van Minnen J, Leurs R: Visualization of agonist-induced internalization of histamine H2 receptors. Biochem Biophys Res Commun. 1995, 214 (3): 1138-1145. 10.1006/bbrc.1995.2404. Sep 25PubMedCrossRef
32.
Zurück zum Zitat Matsuda N, Jesmin S, Takahashi Y, Hatta E, Kobayashi M, Matsuyama K, Kawakami N, Sakuma I, Gando S, Fukui H, Hattori Y, Levi R: Histamine H1 and H2 receptor gene and protein levels are differentially expressed in the hearts of rodents and humans. J Pharmacol Exp Ther. 2004, 309 (2): 786-795. 10.1124/jpet.103.063065. MayPubMedCrossRef Matsuda N, Jesmin S, Takahashi Y, Hatta E, Kobayashi M, Matsuyama K, Kawakami N, Sakuma I, Gando S, Fukui H, Hattori Y, Levi R: Histamine H1 and H2 receptor gene and protein levels are differentially expressed in the hearts of rodents and humans. J Pharmacol Exp Ther. 2004, 309 (2): 786-795. 10.1124/jpet.103.063065. MayPubMedCrossRef
33.
Zurück zum Zitat Ermakov A, Pells S, Freile P, Ganeva VV, Wildenhain J, Bradley M, Pawson A, Millar R, De Sousa PA: A role for intracellular calcium downstream of G-protein signaling in undifferentiated human embryonic stem cell culture. Stem Cell Res. 2012, 9 (3): 171-184. 10.1016/j.scr.2012.06.007. NovPubMedCrossRef Ermakov A, Pells S, Freile P, Ganeva VV, Wildenhain J, Bradley M, Pawson A, Millar R, De Sousa PA: A role for intracellular calcium downstream of G-protein signaling in undifferentiated human embryonic stem cell culture. Stem Cell Res. 2012, 9 (3): 171-184. 10.1016/j.scr.2012.06.007. NovPubMedCrossRef
34.
Zurück zum Zitat Leclerc C, Neant I, Moreau M: The calcium: an early signal that initiates the formation of the nervous system during embryogenesis. Front Mol Neurosci. 2012, 5: 3-10.3389/fnmol.2012.00064.PubMedPubMedCentralCrossRef Leclerc C, Neant I, Moreau M: The calcium: an early signal that initiates the formation of the nervous system during embryogenesis. Front Mol Neurosci. 2012, 5: 3-10.3389/fnmol.2012.00064.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Lahti L, Achim K, Partanen J: Molecular regulation of GABAergic neuron differentiation and diversity in the developing midbrain. Acta Physiol (Oxf). 2013, 207 (4): 616-627. 10.1111/apha.12062. AprCrossRef Lahti L, Achim K, Partanen J: Molecular regulation of GABAergic neuron differentiation and diversity in the developing midbrain. Acta Physiol (Oxf). 2013, 207 (4): 616-627. 10.1111/apha.12062. AprCrossRef
36.
Zurück zum Zitat Andersson ER, Salto C, Villaescusa JC, Cajanek L, Yang S, Bryjova L, Nagy II, Vainio SJ, Ramirez C, Bryja V, Arenas E: Wnt5a cooperates with canonical Wnts to generate midbrain dopaminergic neurons in vivo and in stem cells. Proc Natl Acad Sci U S A. 2013, 110 (7): E602-610. 10.1073/pnas.1208524110. Feb 12PubMedPubMedCentralCrossRef Andersson ER, Salto C, Villaescusa JC, Cajanek L, Yang S, Bryjova L, Nagy II, Vainio SJ, Ramirez C, Bryja V, Arenas E: Wnt5a cooperates with canonical Wnts to generate midbrain dopaminergic neurons in vivo and in stem cells. Proc Natl Acad Sci U S A. 2013, 110 (7): E602-610. 10.1073/pnas.1208524110. Feb 12PubMedPubMedCentralCrossRef
37.
Zurück zum Zitat Meier E, Hertz L, Schousboe A: Neurotransmitters as developmental signals. Neurochem Int. 1991, 19: 1-15. 10.1016/0197-0186(91)90113-R.CrossRef Meier E, Hertz L, Schousboe A: Neurotransmitters as developmental signals. Neurochem Int. 1991, 19: 1-15. 10.1016/0197-0186(91)90113-R.CrossRef
38.
Zurück zum Zitat Arie Y, Iketani M, Takamatsu K, Mikoshiba K, Goshima Y, Takei K: Developmental changes in the regulation of calcium-dependent neurite outgrowth. Biochem Biophys Res Commun. 2009, 379 (1): 11-15. 10.1016/j.bbrc.2008.11.128. Jan 30PubMedCrossRef Arie Y, Iketani M, Takamatsu K, Mikoshiba K, Goshima Y, Takei K: Developmental changes in the regulation of calcium-dependent neurite outgrowth. Biochem Biophys Res Commun. 2009, 379 (1): 11-15. 10.1016/j.bbrc.2008.11.128. Jan 30PubMedCrossRef
39.
Zurück zum Zitat Slevin JC, Byers L, Gertsenstein M, Qu D, Mu J, Sunn N, Kingdom JC, Rossant J, Adamson SL: High resolution ultrasound-guided microinjection for interventional studies of early embryonic and placental development in vivo in mice. BMC Dev Biol. 2006, 6: 10-10.1186/1471-213X-6-10.PubMedPubMedCentralCrossRef Slevin JC, Byers L, Gertsenstein M, Qu D, Mu J, Sunn N, Kingdom JC, Rossant J, Adamson SL: High resolution ultrasound-guided microinjection for interventional studies of early embryonic and placental development in vivo in mice. BMC Dev Biol. 2006, 6: 10-10.1186/1471-213X-6-10.PubMedPubMedCentralCrossRef
40.
Zurück zum Zitat Garcia-Moreno F, Pedraza M, Di Giovannantonio LG, Di Salvio M, Lopez-Mascaraque L, Simeone A, De Carlos JA: A neuronal migratory pathway crossing from diencephalon to telencephalon populates amygdala nuclei. Nat Neurosci. 2010, 13 (6): 680-689. 10.1038/nn.2556. JunPubMedCrossRef Garcia-Moreno F, Pedraza M, Di Giovannantonio LG, Di Salvio M, Lopez-Mascaraque L, Simeone A, De Carlos JA: A neuronal migratory pathway crossing from diencephalon to telencephalon populates amygdala nuclei. Nat Neurosci. 2010, 13 (6): 680-689. 10.1038/nn.2556. JunPubMedCrossRef
41.
Zurück zum Zitat Pierfelice TJ, Gaiano N: Ultrasound-guided microinjection into the mouse forebrain in utero at E9.5. J Vis Exp. 2010, 45: 2047-Nov 13PubMed Pierfelice TJ, Gaiano N: Ultrasound-guided microinjection into the mouse forebrain in utero at E9.5. J Vis Exp. 2010, 45: 2047-Nov 13PubMed
42.
Zurück zum Zitat Abeliovich A, Hammond R: Midbrain dopamine neuron differentiation: factors and fates. Dev Biol. 2007, 304 (2): 447-454. 10.1016/j.ydbio.2007.01.032. Apr 15PubMedCrossRef Abeliovich A, Hammond R: Midbrain dopamine neuron differentiation: factors and fates. Dev Biol. 2007, 304 (2): 447-454. 10.1016/j.ydbio.2007.01.032. Apr 15PubMedCrossRef
43.
Zurück zum Zitat Smidt MP, Burbach JP: How to make a mesodiencephalic dopaminergic neuron. Nat Rev Neurosci. 2007, 8 (1): 21-32. 10.1038/nrn2039. JanPubMedCrossRef Smidt MP, Burbach JP: How to make a mesodiencephalic dopaminergic neuron. Nat Rev Neurosci. 2007, 8 (1): 21-32. 10.1038/nrn2039. JanPubMedCrossRef
44.
Zurück zum Zitat Kong KC, Billington CK, Gandhi U, Panettieri RA, Penn RB: Cooperative mitogenic signaling by G protein-coupled receptors and growth factors is dependent on G (q/11). FASEB J. 2006, 20 (9): 1558-1560. 10.1096/fj.05-5622fje. JulPubMedCrossRef Kong KC, Billington CK, Gandhi U, Panettieri RA, Penn RB: Cooperative mitogenic signaling by G protein-coupled receptors and growth factors is dependent on G (q/11). FASEB J. 2006, 20 (9): 1558-1560. 10.1096/fj.05-5622fje. JulPubMedCrossRef
45.
Zurück zum Zitat Goridis C, Rohrer H: Specification of catecholaminergic and serotonergic neurons. Nat Rev Neurosci. 2002, 3 (7): 531-541. 10.1038/nrn871. JulPubMedCrossRef Goridis C, Rohrer H: Specification of catecholaminergic and serotonergic neurons. Nat Rev Neurosci. 2002, 3 (7): 531-541. 10.1038/nrn871. JulPubMedCrossRef
46.
Zurück zum Zitat Weissman TA, Riquelme PA, Ivic L, Flint AC, Kriegstein AR: Calcium waves propagate through radial glial cells and modulate proliferation in the developing neocortex. Neuron. 2004, 43 (5): 647-661. 10.1016/j.neuron.2004.08.015. Sep 2PubMedCrossRef Weissman TA, Riquelme PA, Ivic L, Flint AC, Kriegstein AR: Calcium waves propagate through radial glial cells and modulate proliferation in the developing neocortex. Neuron. 2004, 43 (5): 647-661. 10.1016/j.neuron.2004.08.015. Sep 2PubMedCrossRef
47.
Zurück zum Zitat Garcia-Garcia E, Pino-Barrio MJ, Lopez-Medina L, Martinez-Serrano A: Intermediate progenitors are increased by lengthening of the cell cycle through calcium signaling and p53 expression in human neural progenitors. Mol Biol Cell. 2012, 23 (7): 1167-1180. 10.1091/mbc.E11-06-0524. AprPubMedPubMedCentralCrossRef Garcia-Garcia E, Pino-Barrio MJ, Lopez-Medina L, Martinez-Serrano A: Intermediate progenitors are increased by lengthening of the cell cycle through calcium signaling and p53 expression in human neural progenitors. Mol Biol Cell. 2012, 23 (7): 1167-1180. 10.1091/mbc.E11-06-0524. AprPubMedPubMedCentralCrossRef
48.
Zurück zum Zitat Rinne JO, Anichtchik OV, Eriksson KS, Kaslin J, Tuomisto L, Kalimo H, Roytta M, Panula P: Increased brain histamine levels in Parkinson's disease but not in multiple system atrophy. J Neurochem. 2002, 81 (5): 954-960. 10.1046/j.1471-4159.2002.00871.x. JunPubMedCrossRef Rinne JO, Anichtchik OV, Eriksson KS, Kaslin J, Tuomisto L, Kalimo H, Roytta M, Panula P: Increased brain histamine levels in Parkinson's disease but not in multiple system atrophy. J Neurochem. 2002, 81 (5): 954-960. 10.1046/j.1471-4159.2002.00871.x. JunPubMedCrossRef
49.
Zurück zum Zitat Coelho MH, Silva IJ, Azevedo MS, Manso CF: Decrease in blood histamine in drug-treated parkinsonian patients. Mol Chem Neuropathol. 1991, 14 (2): 77-85. 10.1007/BF03159928. AprPubMedCrossRef Coelho MH, Silva IJ, Azevedo MS, Manso CF: Decrease in blood histamine in drug-treated parkinsonian patients. Mol Chem Neuropathol. 1991, 14 (2): 77-85. 10.1007/BF03159928. AprPubMedCrossRef
50.
Zurück zum Zitat Anichtchik OV, Rinne JO, Kalimo H, Panula P: An altered histaminergic innervation of the substantia nigra in Parkinson's disease. Exp Neurol. 2000, 163 (1): 20-30. 10.1006/exnr.2000.7362. MayPubMedCrossRef Anichtchik OV, Rinne JO, Kalimo H, Panula P: An altered histaminergic innervation of the substantia nigra in Parkinson's disease. Exp Neurol. 2000, 163 (1): 20-30. 10.1006/exnr.2000.7362. MayPubMedCrossRef
51.
Zurück zum Zitat Vizuete ML, Merino M, Venero JL, Santiago M, Cano J, Machado A: Histamine infusion induces a selective dopaminergic neuronal death along with an inflammatory reaction in rat substantia nigra. J Neurochem. 2000, 75 (2): 540-552. 10.1046/j.1471-4159.2000.0750540.x. AugPubMedCrossRef Vizuete ML, Merino M, Venero JL, Santiago M, Cano J, Machado A: Histamine infusion induces a selective dopaminergic neuronal death along with an inflammatory reaction in rat substantia nigra. J Neurochem. 2000, 75 (2): 540-552. 10.1046/j.1471-4159.2000.0750540.x. AugPubMedCrossRef
52.
Zurück zum Zitat Liu CQ, Chen Z, Liu FX, Hu DN, Luo JH: Involvement of brain endogenous histamine in the degeneration of dopaminergic neurons in 6-hydroxydopamine-lesioned rats. Neuropharmacology. 2007, 53 (7): 832-841. 10.1016/j.neuropharm.2007.08.014. DecPubMedCrossRef Liu CQ, Chen Z, Liu FX, Hu DN, Luo JH: Involvement of brain endogenous histamine in the degeneration of dopaminergic neurons in 6-hydroxydopamine-lesioned rats. Neuropharmacology. 2007, 53 (7): 832-841. 10.1016/j.neuropharm.2007.08.014. DecPubMedCrossRef
53.
Zurück zum Zitat Chan WY, Kohsaka S, Rezaie P: The origin and cell lineage of microglia: new concepts. Brain Res Rev. 2007, 53 (2): 344-354. 10.1016/j.brainresrev.2006.11.002. FebPubMedCrossRef Chan WY, Kohsaka S, Rezaie P: The origin and cell lineage of microglia: new concepts. Brain Res Rev. 2007, 53 (2): 344-354. 10.1016/j.brainresrev.2006.11.002. FebPubMedCrossRef
54.
Zurück zum Zitat Cunningham CL, Martinez-Cerdeno V, Noctor SC: Microglia regulate the number of neural precursor cells in the developing cerebral cortex. J Neurosci. 2013, 33 (10): 4216-4233. 10.1523/JNEUROSCI.3441-12.2013. Mar 6PubMedPubMedCentralCrossRef Cunningham CL, Martinez-Cerdeno V, Noctor SC: Microglia regulate the number of neural precursor cells in the developing cerebral cortex. J Neurosci. 2013, 33 (10): 4216-4233. 10.1523/JNEUROSCI.3441-12.2013. Mar 6PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Lopez-Gonzalez R, Camacho-Arroyo I, Velasco I: Progesterone and 17beta-estradiol increase differentiation of mouse embryonic stem cells to motor neurons. IUBMB Life. 2011, 63 (10): 930-939. 10.1002/iub.560. OctPubMedCrossRef Lopez-Gonzalez R, Camacho-Arroyo I, Velasco I: Progesterone and 17beta-estradiol increase differentiation of mouse embryonic stem cells to motor neurons. IUBMB Life. 2011, 63 (10): 930-939. 10.1002/iub.560. OctPubMedCrossRef
56.
Zurück zum Zitat Rodriguez-Rivera NS, Molina-Hernandez A, Sanchez-Cruz E, Escalante-Alcalde D, Velasco I: Activated Notch1 is a stronger astrocytic stimulus than leukemia inhibitory factor for rat neural stem cells. Int J Dev Biol. 2009, 53 (7): 947-953.PubMedCrossRef Rodriguez-Rivera NS, Molina-Hernandez A, Sanchez-Cruz E, Escalante-Alcalde D, Velasco I: Activated Notch1 is a stronger astrocytic stimulus than leukemia inhibitory factor for rat neural stem cells. Int J Dev Biol. 2009, 53 (7): 947-953.PubMedCrossRef
57.
Zurück zum Zitat Diaz NF, Guerra-Arraiza C, Diaz-Martinez NE, Salazar P, Molina-Hernandez A, Camacho-Arroyo I, Velasco I: Changes in the content of estrogen alpha and progesterone receptors during differentiation of mouse embryonic stem cells to dopamine neurons. Brain Res Bull. 2007, 73 (1–3): 75-80. 10.1016/j.brainresbull.2007.02.002. Jun 15PubMedPubMedCentralCrossRef Diaz NF, Guerra-Arraiza C, Diaz-Martinez NE, Salazar P, Molina-Hernandez A, Camacho-Arroyo I, Velasco I: Changes in the content of estrogen alpha and progesterone receptors during differentiation of mouse embryonic stem cells to dopamine neurons. Brain Res Bull. 2007, 73 (1–3): 75-80. 10.1016/j.brainresbull.2007.02.002. Jun 15PubMedPubMedCentralCrossRef
58.
Zurück zum Zitat Ajmo CT, Vernon DO, Collier L, Hall AA, Garbuzova-Davis S, Willing A, Pennypacker KR: The spleen contributes to stroke-induced neurodegeneration. J Neurosci Res. 2008, 86 (10): 2227-2234. 10.1002/jnr.21661.PubMedPubMedCentralCrossRef Ajmo CT, Vernon DO, Collier L, Hall AA, Garbuzova-Davis S, Willing A, Pennypacker KR: The spleen contributes to stroke-induced neurodegeneration. J Neurosci Res. 2008, 86 (10): 2227-2234. 10.1002/jnr.21661.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Schmued LC, Albertson C, Slikker W: Fluoro-Jade: a novel fluorochrome for the sensitive and reliable histochemical localization of neuronal degeneration. Brain Res. 1997, 751 (1): 37-46. Mar 14PubMedCrossRef Schmued LC, Albertson C, Slikker W: Fluoro-Jade: a novel fluorochrome for the sensitive and reliable histochemical localization of neuronal degeneration. Brain Res. 1997, 751 (1): 37-46. Mar 14PubMedCrossRef
Metadaten
Titel
Histamine impairs midbrain dopaminergic development in vivo by activating histamine type 1 receptors
verfasst von
Itzel Escobedo-Avila
Fernanda Vargas-Romero
Anayansi Molina-Hernández
Rodrigo López-González
Daniel Cortés
Juan A De Carlos
Iván Velasco
Publikationsdatum
01.12.2014
Verlag
BioMed Central
Erschienen in
Molecular Brain / Ausgabe 1/2014
Elektronische ISSN: 1756-6606
DOI
https://doi.org/10.1186/s13041-014-0058-x

Weitere Artikel der Ausgabe 1/2014

Molecular Brain 1/2014 Zur Ausgabe

Leitlinien kompakt für die Neurologie

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Update Neurologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.