Next Article in Journal
Extracellular Vesicles as Potential Bladder Cancer Biomarkers: Take It or Leave It?
Previous Article in Journal
The Role of Next-Generation Probiotics in Obesity and Obesity-Associated Disorders: Current Knowledge and Future Perspectives
Previous Article in Special Issue
Cadmium Exposure Is Associated with Behavioral Deficits and Neuroimmune Dysfunction in BTBR T+ Itpr3tf/J Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Impact of BDNF, NTRK2, NGFR, CREB1, GSK3B, AKT, MAPK1, MTOR, PTEN, ARC, and SYN1 Genetic Polymorphisms in Antidepressant Treatment Response Phenotypes

1
Centro de Investigação em Saúde e Ambiente (CISA), Escola Superior de Saúde, Instituto Politécnico do Porto, 4200-072 Porto, Portugal
2
Molecular Oncology & Viral Pathology, IPO-Porto Research Center (CI-IPOP), Portuguese Institute of Oncology, 4200-072 Porto, Portugal
3
Experimental Pathology and Therapeutics Group, IPO-Porto Research Center (CI-IPOP), Portuguese Institute of Oncology, 4200-072 Porto, Portugal
4
Hospital de Magalhães Lemos, 4149-003 Porto, Portugal
5
Instituto Universitário de Ciências da Saúde, 4585-116 Gandra, Portugal
6
Clínica Médico-Psiquiátrica da Ordem, 4000-270 Porto, Portugal
7
Trás-os-Montes e Alto Douro Hospital Centre, 5000-508 Vila Real, Portugal
8
REQUIMTE/LAQV, Instituto Superior de Engenharia do Instituto Politécnico do Porto, 4200-072 Porto, Portugal
9
Research Department, Portuguese League Against Cancer (Norte), 4200-172 Porto, Portugal
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6758; https://doi.org/10.3390/ijms24076758
Submission received: 3 March 2023 / Revised: 30 March 2023 / Accepted: 30 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue Advances in Neurogenetics and Neuroimmunology)

Abstract

:
This study aimed to investigate the influence of genetic variants in neuroplasticity-related genes on antidepressant treatment phenotypes. The BDNF-TrkB signaling pathway, as well as the downstream kinases Akt and ERK and the mTOR pathway, have been implicated in depression and neuroplasticity. However, clinicians still struggle with the unpredictability of antidepressant responses in depressed patients. We genotyped 26 polymorphisms in BDNF, NTRK2, NGFR, CREB1, GSK3B, AKT, MAPK1, MTOR, PTEN, ARC, and SYN1 in 80 patients with major depressive disorder treated according to the Texas Medical Algorithm for 27 months at Hospital Magalhães Lemos, Porto, Portugal. Our results showed that BDNF rs6265, PTEN rs12569998, and SYN1 rs1142636 SNP were associated with treatment-resistant depression (TRD). Additionally, MAPK1 rs6928 and GSK3B rs6438552 gene polymorphisms were associated with relapse. Moreover, we found a link between the rs6928 MAPK1 polymorphism and time to relapse. These findings suggest that the BDNF, PTEN, and SYN1 genes may play a role in the development of TRD, while MAPK1 and GSK3B may be associated with relapse. GO analysis revealed enrichment in synaptic and trans-synaptic transmission pathways and glutamate receptor activity with TRD-associated genes. Genetic variants in these genes could potentially be incorporated into predictive models of antidepressant response.

1. Introduction

A growing body of research suggests that neuroplasticity pathways play a significant role in the pathophysiology of depression and in the therapeutic mechanisms of antidepressant drugs (AD) [1,2,3]. The BDNF-NTRK2-CREB1 pathway is one such pathway that has received substantial attention due to its involvement in regulating neuroplasticity and synaptic function, which are disrupted in depression [4,5,6]. BDNF is one of the most investigated genes regarding depression and AD response [7,8,9]. In fact, the rs6265 (Val66Met) has been demonstrated to alter pro-BDNF processing and, consequently, the secretion of BDNF [10], as well as the hippocampal volume [11].
Activation of the TrkB neurotrophin receptor (encoded by the NTRK2 gene) has been shown to mediate most of the plasticity-enhancing effects of BDNF. NTRK2 gene polymorphism has been associated with an increased risk of developing Treatment Resistant Depression (TRD) [12], with a reduction of HAM-D21 score [13]. In addition to TrkB, BDNF can bind to the p75 neurotrophin receptor (p75NTR, encoded by the NGFR gene). Contrary to TrkB, little is known about the potential role of signaling through p75NTR. The activation of this receptor by pro-BDNF was demonstrated to induce neuronal atrophy and apoptosis, dendritic pruning, and the induction of long-term depression (LTD), opposed effects of the activation of TrkB by BDNF [14,15].
The TrkB activation induces downstream pathways, including the PI3K/Akt pathway, which is linked to the Wnt/β-catenin and to the mTOR pathway [4,16]. The PI3K/Akt pathway has intrinsically an important role in promoting the proliferation of the adult hippocampal cell and also in the inhibition of cell differentiation [15,17]. AD also produces increased Akt levels in several brain structures, including the hippocampus [18] and prefrontal cortex [15]. Alterations in several elements of this pathway, such as the cAMP response element-binding protein (CREB) transcription factor, have been described in peripheral cells and the postmortem brain of patients with affective disorders, both untreated or after AD therapy [15]. CREB is a transcription factor that regulates the expression of several genes involved in neuroplasticity (including BDNF) and cell survival and has also been widely involved in the pathophysiology of depression and in AD treatments. Moreover, a chromosomal region of 2q33–35, which included this gene, has been associated with mood disorders [19].
Furthermore, Mitogen-activated protein kinase 1 (MAPK1) signaling pathway plays a critical role in synaptic and structural plasticity [20]. Likewise, much interest has been given to the glycogen synthase kinase-3β (GSK3β), a key regulator of neuronal function [21]. In neurons, GSK3β plays an important role in the BDNF pathway and is thereby involved in mechanisms of synaptic plasticity, neurogenesis, and resilience to neuronal injury. Additionally, treatment with AD was found to inhibit GSK3β activity in mouse brains [22], and genetic variations in GSK3B have been associated with response to selective serotonin reuptake inhibitor AD in Chinese MDD patients [23].
mTOR has been recently studied in the central nervous system (CNS) physiology and diseases [15,24]. mTOR-signalling pathway has been described to be involved in synaptic plasticity, memory retention, neuroendocrine regulation, and the modulation of neuronal repair upon injury [15,24]. Activation of mTOR has been functionally linked with local protein synthesis, such as the presynaptically as synapsin I (SYN1), as well as cytoskeletal proteins, such as the activity-regulated cytoskeletal-associated protein (ARC) [15,24,25].
Although previous studies have suggested that genetic variants may play a key role in the mechanism of TRD and relapse, attempts to identify risk polymorphisms within genes with putative interest in AD response had limited success. Taking this into consideration, we aimed to evaluate the role of BDNF, CREB1, NTRK2, NGFR, ARC, GSK3B, AKT, MAPK1, MTOR, PTEN, and SYN1 genetic polymorphisms in AD treatment phenotypes in a cohort of Portuguese MDD patients.

2. Results

2.1. Relapse Phenotype

As shown in Table 1, statistically significant differences were found between relapsed and non-relapsed participants for the MAPK1 gene polymorphism rs6928. There was a higher frequency of patients carrying the C allele (GC and CC genotypes) in the non-relapse group compared with the relapse group. C allele carriers of rs6928 polymorphism presented a reduced risk of relapse (OR: 0.303; 95% CI: (0.090–1.020); p = 0.048). Moreover, Kaplan–Meier analysis revealed that patients carrying the C allele relapse later than the ones presenting the GG genotype (Figure 1; 100 vs. 82 weeks, p = 0.022, Log-rank test). A statistically significant difference was also found between relapsed and non-relapsed participants for GSK3B (rs6438552). It was observed a higher frequency of patients carrying the GG genotype in the relapse group, compared with the non-relapse group (OR: 6.667; 95% CI: (1.121–39.660); p = 0.042), as well as for the ones carrying the G allele (OR: 5.600; 95% CI: (1.137–27.571); p = 0.033). No statistically significant differences were found in genotype frequencies between relapsed and non-relapsed patients for the other evaluated SNPs (Table S1—Supplementary Data), as well as time to relapse.

2.2. Treatment Resistant Depression Phenotype

As observed in Table 1, statistically significant differences were found in genotype frequencies between TRD and non-TRD participants for the BDNF gene polymorphism rs6265, for the PTEN polymorphism rs12569998, and for the SYN1 polymorphism rs1142636. There was a significant association between BDNF rs6265 and TRD, with a higher percentage of patients with CT genotype in the TRD group, compared with the non-TRD group, which represents a three-fold higher risk of TRD development for patients carrying CT genotype (OR: 3.2; 95% CI: (0.975–10.501); p = 0.049). Regarding PTEN polymorphism, there was a statistically significant association between rs12569998 and TRD, with an overrepresentation of patients carrying TG genotype in the TRD group, which corresponds to an approximately four-fold increased risk to the development of TRD (OR: 4.231; 95% CI: (1.173–15.261); p = 0.020). Additionally, there was a statistically significant association between rs1142636 of the SYN1 gene and TRD. Carriers of the GG genotype had a six-fold increased risk of presenting a TRD phenotype (OR: 6.00; 95% CI: (1.543–23.333); p = 0.006), as well as G allele (AG and GG genotypes) carriers presented a 3-fold increased risk (OR: 3.120; 95% CI: (1.045–9.314); p = 0.037). No statistically significant differences were found between TRD and non-TRD patients for the remaining evaluated SNPs (Table S1—Supplementary Data). Moreover, no differences were found regarding time to remission for any of the evaluated SNPs.

2.3. Gene Functional Enrichment Analysis

The results of the genes associated with TRD were identified with automated gene references into function provided by Geneshot (AutoRIF). From the three genes statistically associated with TRD in our sample (BDNF, SYN1, and PTEN), BDNF and SYN1 were included in this list, bringing the total number of genes to 136, which are listed in Table S2.
To further investigate the functional implications of the TRD-associated genes, we conducted a gene ontology (GO) analysis. The results were categorized into three groups: biological process, cellular component, and molecular function. The biological process category showed that TRD-associated genes were enriched in chemical synaptic transmission (GO:0007268), anterograde trans-synaptic signaling (GO:0098916), modulation of chemical synaptic transmission (GO:0050804), and the glutamate receptor signaling pathway (GO:0007215), as evidenced by the top enrichment terms (Table 2). In the molecular function category, receptor-ligand activity (GO:0048018), glutamate receptor activity (GO:0008066, GO:0022849, GO:0004970), and hormone activity (GO:0005179) were identified as the main terms in the enrichment analysis (Table 2). Finally, in the cellular component category, neuron projections (GO:0043005), dendrites (GO:00340425), and ionotropic glutamate receptor complexes (GO:0008328) were the top enriched terms (Table 2).
KEGG analysis displayed that these genes were associated with neuroactive ligand-receptor interaction, cocaine addiction, amphetamine addiction, alcoholism, serotonergic synapse, glutamatergic synapse, cAMP signaling pathway, dopaminergic synapse, pathways of degeneration, and the calcium signaling pathway (Figure 2).

3. Discussion

The clinical AD response develops slowly over the first weeks of treatment, and one explanation for this delayed response may be the need for physical growth and reorganization in the brain, responses that are mediated by BDNF signaling and downstream pathway events [2]. In addition to the lag time between AD and clinical improvement, there are still about 30% of MDD patients who do not present improvements upon AD treatment, displaying a TRD phenotype [26]. Peculiarly, a great majority of these TRD patients improve upon ketamine administration [27]. This drug produces AD response by rapidly inducing synaptic plasticity in neuronal networks [28,29,30]. Considering this evidence, it may be reasonable to speculate that the TRD phenotype may correlate with neuroplasticity molecule alterations.
Therefore, in the present study, we have evaluated genetic polymorphisms of the genes BDNF, CREB1, NTRK2, NGFR, ARC, AKT, GSK3B, MAPK1, MTOR, PTEN, and SYN1 involved in several neurotrophic/neuroplasticity pathways with relevance to the mechanism of action of AD.
Regarding the relapse phenotype, both MAPK1 gene polymorphism rs6928 and GSK3B (rs6438552) were found to be associated with this phenotype. C allele carriers of rs6928 MAPK1 polymorphism present nearly a 70% reduced risk of relapse (OR: 0.303), and these patients also tend to relapse later than the ones presenting the GG genotype. This polymorphism has been previously studied in MDD patients and was not associated with treatment resistance, response, and remission, but its relation with relapse was not determined [20]. It is known that MAPK1 activation is altered after stress [31] and corticosterone exposure [32] and that MAPK signaling was proposed to regulate AD modulation of glial cell line-derived neurotrophic factor [33]. Therefore, it is possible that genetic polymorphism in MAPK1 may contribute to the risk of relapse. In fact, depression was previously associated with an aberrant MAPK1 signaling pathway [34].
A statistically significant difference was also found between relapsed and non-relapsed participants for GSK3B (rs6438552), patients carrying the G allele have approximately a six-fold increased risk of relapse. In vitro data have demonstrated that GSK3B (rs6438552) intronic polymorphism regulates the selection of splice acceptor sites and thus alters GSK3β transcription [35]. Furthermore, GSK3B rs6438552 was previously associated with brain structural changes in MDD [36,37]. Additionally, the inhibition of GSK3B is thought to be a key feature in the therapeutic mechanism of AD. Thus, it is likely that this polymorphism could affect the risk of relapse.
Regarding TRD, statistically significant differences were observed between TRD and non-TRD patients for PTEN polymorphism rs12569998, SYN1 polymorphism rs1142636, as well as BDNF rs6265 polymorphism. Regarding PTEN rs12569998, carriers of TG genotype and G allele carriers had a four-fold risk of developing TRD rather than TT carriers. The analysis for the homozygous GG genotype was unable to be determined due to the low frequencies of this genotype in the Caucasian population and, consequently, in our sample. Concerning SYN1 polymorphism rs114263, carriers of the GG genotype had a six-fold risk of presenting a TRD phenotype, while G allele carriers had a 3.12-fold risk compared with patients with the AA genotype. Regarding BDNF rs6265 polymorphism, CT carriers were found to have an increased 3.2-fold risk of TRD compared with CC carriers. To the best of our knowledge, this is the first study to evaluate the role of SYN1 and PTEN in AD response phenotypes.
SYN1 polymorphism has been previously evaluated in schizophrenia, and the G allele of synonymous SNP (rs1142636, Asn170Asn) in SYN1 was found to be a risk factor for schizophrenia susceptibility in Korean female patients [38]. Furthermore, mutations within this gene have been observed in a family with epilepsy [39]. The putative influence of SYN1 in TRD patients may be related to the fact that synapsins play a role during neuronal development and synapse formation [39,40]. When hippocampal neurons from SYN1 deficient mice are cultured, the axons are shorter and have fewer branches than in controls, and synapse formation is delayed [40]. It is also postulated that synapsins regulate the kinetics of neurotransmitter release during priming of synaptic vesicles at the plasma membrane [39]. Therefore, it is likely that polymorphisms in synapsin I may affect axon elongation, branching, and synaptogenesis and play an ongoing role in the distribution of neurotransmitter vesicles and neurotransmitter release, and thus contribute to a TRD phenotype [41].
Regarding PTEN, this gene has been reported as a biomarker of a high versus low suicidality state [42], although a subsequent report did not corroborate the findings [43].
Some evidence supports the influence of BDNF polymorphisms in AD response. In this respect, the most investigated variant within the BDNF gene is rs6265 [10,44,45,46,47,48]. Pharmacogenetic studies are contradictory regarding the influence of this polymorphism in AD outcomes. While some studies, such as Chi et al. [49] and Domschke et al. [50], showed a better response in the rs6265 Val/Val genotype, others found a more favorable outcome in Met allele carriers [51,52,53]. Others suggested a positive molecular heterosis effect [54,55,56]. Additionally, several negative findings were also observed, including in the large GENDEP sample [57,58,59].
To better understand the correlation between the results and TRD phenotype, we conducted a comprehensive analysis focusing on functional and pathway enrichment of the genes identified (BDNF, PTEN and SYN1) along with 135 other top genes associated with TRD. Our in-depth analysis revealed enrichment in synaptic and trans-synaptic transmission, as well as glutamate receptor signaling. The molecular mechanisms were mainly associated with glutamate receptor activity and ligand-gated channel activity, while the top enriched terms in the cellular component were neuron projections, dendrites, and ionotropic glutamate receptor complexes. These findings suggest that the molecular mechanisms of TRD-associated genes are mainly involved in synaptic and trans-synaptic transmission, affecting neuroplasticity and ionotropic glutamate receptors. This is consistent with the mechanism of action of ketamine and esketamine, two drugs that have shown efficacy in TRD patients [60]. These drugs are non-selective, non-competitive antagonists that block the NMDA receptor (an ionotropic glutamate receptor) on GABA interneurons, thereby increasing neurotrophic signaling that restores synaptic function [61]. Furthermore, the action of esketamine on AMPA receptors may ultimately improve neural plasticity and synaptogenesis through signaling pathways resulting in enhanced BDNF production [62]. These findings provide a promising avenue for the development of new treatments for TRD. Targeting the pathways and structures identified in our analysis, such as glutamate receptor activity and synaptic transmission, could potentially lead to the development of more effective therapies for this challenging phenotype of depressed patients.
Regardless of the interesting results we presented, some limitations should be considered. First, our sample is small, so it is likely that small effects exerted by single SNPs could not be detected. However, this study involved a cohort of MDD Portuguese patients with a long follow-up time. Conversely, and since we have evaluated a few SNPs in each gene, due to the target pathway design of our study, it could be possible that other SNPs could affect the outcomes. A further limitation relies on the use of drugs with different mechanisms of action, so we cannot correlate each SNP with a specific class of AD. However, patients were treated according to the Texas Medication Algorithm, and the TRD definition is independent of the drugs used.
The findings of this study provide valuable insights into the potential influence of genetic variations in the development of TRD, laying the groundwork for future research to improve the understanding of the genetic factors underlying this phenotype. In order to overcome the limitations of the present study, future research should focus on increasing the sample size and evaluating a greater number of SNPs on synaptic and trans-synaptic transmission and on the glutamate pathway to more fully elucidate its role in TRD. Moreover, exploring the relationship between specific classes of antidepressant drugs and genetic variations could yield personalized treatment options for patients with TRD.

4. Materials and Methods

4.1. Patients

The participants of this study were 80 MDD patients (followed by 27 months) from a cohort recruited at Hospital Magalhães Lemos, Oporto, Portugal, as we previously described [63,64]. Diagnostic criteria for MDD were established using the Structured Clinical Interview for DSM-IV Axis I Disorders (SCID-I), and the severity of depressive symptoms was measured using the Beck Depression Inventory (BDI). Texas Medication Algorithm [65] was used to provide medications to the patients. The BDI score was used to evaluate the clinical response to effective AD therapy (given after at least six weeks and at acceptable dosages). Time to remission (TRD), time to relapse (TR), remission, and TRD were the evaluated AD treatment response phenotypes. Patients’ clinical and socio-demographic details were previously documented [63]. Each participant provided written informed permission in accordance with “The Code of Ethics of the International Medical Association” (Declaration of Helsinki) after being briefed on the study’s purpose and methods, and this study was approved by the hospital’s ethics board.

4.2. DNA Extraction and SNP Analysis

Blood samples from MDD patients were collected in EDTA vacuum tubes. Genomic DNA was isolated from peripheral blood using a commercial kit according to standard laboratory protocols (E.Z.N.A.—Omega Bio-tek), following the manufacturer’s instructions, and stored at −20 °C until genotyping. Several polymorphisms in genes involved in several neurotrophic/neuroplasticity pathways were evaluated, namely BDNF (rs1491850, rs204946, rs6265, rs908867), CREB1 (rs11904814, rs2253206, rs6740584, rs889895), NTRK2 (rs1187323, rs1187326, rs1387926, rs1778929, rs1565445, rs1659412), NGFR (rs11466155, rs2072446, rs734194), ARC (rs10105842), AKT (rs1130233), GSK3B (rs3755557, rs6438552), MAPK1 (rs6928, rs8136867), MTOR (rs1064261), PTEN (rs12569998), and SYN1 (rs1142636). The Sequenom MassARRAY platform was used to conduct a polymorphism genotyping study (Sequenom, San Diego, CA, USA) [66].
For quality control, genotyping data were called blind to the clinical course of the disease, and 10% of the sample was repeated for the genotyping test, with the same results.

4.3. Functional Enrichment Analysis

Geneshot software (https://amp.pharm.mssm.edu/geneshot, accessed on 26 March 2023) was used to generate ranked lists of genes associated with TRD using the expression “Treatment resistant depression” and gene-publication associations encoded within an automated gene References into Function (AutoRIF) [67]. The genes ranked were then used for the enrichment analysis together with the genes identified as associated with TRD in the present study (BDNF, SYN1, and PTEN) to perform GO and KEGG Pathway analyses with Enrichr (https://maayanlab.cloud/Enrichr/ accessed on 26 march 2023), according to the protocol described by Xie, 2021 et al. [68].

4.4. Statistical Analysis

The software PAWS Statistics 18 was used for data preparation and analysis (release 18.0.0). The categorical variables were compared using the Chi-square (χ2) test at a significance threshold of 5%. As a measure of the association between genotypes and risk of developing a certain phenotype, odds ratio (OR) and 95% confidence interval (CI) were calculated. The correlation between genotypes and time to remission and relapse was evaluated using Kaplan–Meier survival curves and compared using the log-rank test. Chi-square (χ2) was used to determine the Hardy–Weinberg equilibrium of genotypic frequency.

5. Conclusions

In conclusion, our study provides valuable insights into the potential genetic factors underlying TRD and relapse predisposition. Our findings suggest that distinct molecular events contribute to these two phenotypes, with PTEN, SYN1, and BDNF polymorphisms possibly playing a role in TRD, while MAPK1 and GSK3B may be associated with relapse predisposition. The GO analysis revealed enrichment in synaptic and trans-synaptic transmission pathways and glutamate receptor activity, suggesting the involvement of neuroplasticity mechanisms in TRD pathophysiology. These polymorphisms could be included in predictive models of treatment response and help to identify targets for faster and more effective ADs to treat TRD.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24076758/s1.

Author Contributions

Conceptualization, M.S., S.C. and R.M.; data curation, L.L., J.M.-P., P.P., D.M., D.C., S.G. and R.M.; formal analysis, M.S. and L.L.; funding acquisition, S.C., A.C. and R.M.; investigation, M.S., L.L., J.M.-P., P.P., D.M., D.C. and S.G.; methodology, M.S. and L.L.; project administration, M.S. and R.M.; resources, M.S., P.P., D.M., D.C. and S.G.; supervision, S.C., A.C. and R.M.; visualization, M.S. and M.F.B.; writing—original draft, M.S.; writing—review and editing, S.C., M.F.B., A.C. and R.M. All authors have read and agreed to the published version of the manuscript.

Funding

Funding for this study was partially provided by AstraZeneca Foundation; the AstraZeneca Foundation had no further role in study design; in the collection, analysis, and interpretation of data; in the writing of the report; and in the decision to submit the paper for publication.

Institutional Review Board Statement

Each participant provided written informed permission in accordance with “The Code of Ethics of the International Medical Association” (Declaration of Helsinki) after being briefed on the study’s purpose and methods, and this study was approved by the Hospital Magalhães Lemos ethics board.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The datasets generated during the current study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors would like to thank the Genomics unit of IGC for the genotyping services.

Conflicts of Interest

The authors have no financial or proprietary interest in any material discussed in this article. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Castrén, E.; Võikar, V.; Rantamäki, T. Role of neurotrophic factors in depression. Curr. Opin. Pharmacol. 2007, 7, 18–21. [Google Scholar] [CrossRef] [PubMed]
  2. Castrén, E. Neurotrophic effects of antidepressant drugs. Curr. Opin. Pharmacol. 2004, 4, 58–64. [Google Scholar] [CrossRef] [PubMed]
  3. Duman, R.S.; Aghajanian, G.K.; Sanacora, G.; Krystal, J.H. Synaptic plasticity and depression: New insights from stress and rapid-acting antidepressants. Nat. Med. 2016, 22, 238–249. [Google Scholar] [CrossRef] [PubMed]
  4. Castren, E.; Rantamäki, T. The Role of BDNF and Its Receptors in Depression and Antidepressant Drug Action: Reactivation of Developmental Plasticity. Dev. Neurobiol. 2010, 70, 289–297. [Google Scholar] [CrossRef]
  5. Duman, R.S.; Heninger, G.R.; Nestler, E.J. A molecular and cellular theory of depression. Arch. Gen. Psychiatry 1997, 54, 597–606. [Google Scholar] [CrossRef]
  6. Monteggia, L.M.; Barrot, M.; Powell, C.M.; Berton, O.; Galanis, V.; Gemelli, T.; Meuth, S.; Nagy, A.; Greene, R.W.; Nestler, E.J. Essential role of brain-derived neurotrophic factor in adult hippocampal function. Proc. Natl. Acad. Sci. USA 2004, 101, 10827–10832. [Google Scholar] [CrossRef]
  7. Ribeiro, L.; Busnello, J.V.; Cantor, R.M.; Whelan, F.; Whittaker, P.; Deloukas, P.; Wong, M.-L.; Licinio, J. The brain-derived neurotrophic factor rs6265 (Val66Met) polymorphism and depression in Mexican-Americans. Neuroreport 2007, 18, 1291–1293. [Google Scholar] [CrossRef]
  8. Hosang, G.M.; Shiles, C.; Tansey, K.E.; McGuffin, P.; Uher, R. Interaction between stress and the BDNF Val66Met polymorphism in depression: A systematic review and meta-analysis. BMC Med. 2014, 12, 7. [Google Scholar] [CrossRef]
  9. Ozan, E.; Okur, H.; Eker, C.; Eker, O.D.; Gonul, A.S.; Akarsu, N. The effect of depression, BDNF gene val66met polymorphism and gender on serum BDNF levels. Brain Res. Bull. 2010, 81, 61–65. [Google Scholar] [CrossRef]
  10. Egan, M.F.; Kojima, M.; Callicott, J.H.; Goldberg, T.E.; Kolachana, B.S.; Bertolino, A.; Zaitsev, E.; Gold, B.; Goldman, D.; Dean, M.; et al. The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell 2003, 112, 257–269. [Google Scholar] [CrossRef]
  11. Frodl, T.; Moller, H.-J.; Meisenzahl, E. Neuroimaging genetics: New perspectives in research on major depression? Acta Psychiatry Scand. 2008, 118, 363–372. [Google Scholar] [CrossRef] [PubMed]
  12. Li, Z.; Zhang, Y.; Wang, Z.; Chen, J.; Fan, J.; Guan, Y.; Zhang, C.; Yuan, C.; Hong, W.; Wang, Y.; et al. The role of BDNF, NTRK2 gene and their interaction in development of treatment-resistant depression: Data from multicenter, prospective, longitudinal clinic practice. J. Psychiatry Res. 2013, 47, 8–14. [Google Scholar] [CrossRef] [PubMed]
  13. Dong, C.; Wong, M.-L.; Licinio, J. Sequence variations of ABCB1, SLC6A2, SLC6A3, SLC6A4, CREB1, CRHR1 and NTRK2: Association with major depression and antidepressant response in Mexican-Americans. Mol. Psychiatry 2009, 14, 1105–1118. [Google Scholar] [CrossRef] [PubMed]
  14. Zagrebelsky, M.; Holz, A.; Dechant, G.; Barde, Y.-A.; Bonhoeffer, T.; Korte, M. The p75 neurotrophin receptor negatively modulates dendrite complexity and spine density in hippocampal neurons. J. Neurosci. 2005, 25, 9989–9999. [Google Scholar] [CrossRef]
  15. Pilar-Cuéllar, F.; Vidal, R.; Díaz, A.; Castro, E.; dos Anjos, S.; Pascual-Brazo, J.; Linge, R.; Vargas, V.; Blanco, H.; Martínez-Villayandre, B.; et al. Neural plasticity and proliferation in the generation of antidepressant effects: Hippocampal implication. Neural Plast. 2013, 2013, 537265. [Google Scholar] [CrossRef]
  16. Cowen, D.S. Serotonin and neuronal growth factors—A convergence of signaling pathways. J. Neurochem. 2007, 101, 1161–1171. [Google Scholar] [CrossRef]
  17. Peltier, J.; O’Neill, A.; Schaffer, D.V. PI3K/Akt and CREB regulate adult neural hippocampal progenitor proliferation and differentiation. Dev. Neurobiol. 2007, 67, 1348–1361. [Google Scholar] [CrossRef]
  18. Mostany, R.; Valdizán, E.M.; Pazos, A. A role for nuclear β-catenin in SNRI antidepressant-induced hippocampal cell proliferation. Neuropharmacology 2008, 55, 18–26. [Google Scholar] [CrossRef]
  19. Zubenko, G.S.; Maher, B.; Hughes, H.B.; Zubenko, W.N.; Stiffler, J.S.; Kaplan, B.B.; Marazita, M.L. Genome-wide linkage survey for genetic loci that influence the development of depressive disorders in families with recurrent, early-onset, major depression. Am. J. Med. Genet. B Neuropsychiatr Genet. 2003, 123B, 1–18. [Google Scholar] [CrossRef]
  20. Calati, R.; Crisafulli, C.; Balestri, M.; Serretti, A.; Spina, E.; Calabrò, M.; Sidoti, A.; Albani, D.; Massat, I.; Höfer, P.; et al. Evaluation of the role of MAPK1 and CREB1 polymorphisms on treatment resistance, response and remission in mood disorder patients. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2013, 44, 271–278. [Google Scholar] [CrossRef]
  21. Liu, S.; Sun, N.; Xu, Y.; Yang, C.; Ren, Y.; Liu, Z.; Cao, X.; Sun, Y.; Xu, Q.; Zhang, K.; et al. Possible association of the GSK3β gene with the anxiety symptoms of major depressive disorder and P300 waveform. Genet. Test. Mol. Biomarkers 2012, 16, 1382–1389. [Google Scholar] [CrossRef] [PubMed]
  22. Li, X.; Zhu, W.; Roh, M.-S.; Friedman, A.B.; Rosborough, K.; Jope, R.S. In vivo regulation of glycogen synthase kinase-3β (GSK3β) by serotonergic activity in mouse brain. Neuropsychopharmacology 2004, 29, 1426–1431. [Google Scholar] [CrossRef] [PubMed]
  23. Tsai, S.-J.; Liou, Y.-J.; Hong, C.-J.; Yu, Y.W.-Y.; Chen, T.-J. Glycogen synthase kinase-3β gene is associated with antidepressant treatment response in Chinese major depressive disorder. Pharmacogenom. J. 2008, 8, 384–390. [Google Scholar] [CrossRef] [PubMed]
  24. Li, N.; Lee, B.; Liu, R.-J.; Banasr, M.; Dwyer, J.M.; Iwata, M.; Li, X.-Y.; Aghajanian, G.; Duman, R.S. mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science 2010, 329, 959–964. [Google Scholar] [CrossRef] [PubMed]
  25. Gong, R.; Park, C.S.; Abbassi, N.R.; Tang, S.-J. Roles of glutamate receptors and the mammalian target of rapamycin (mTOR) signaling pathway in activity-dependent dendritic protein synthesis in hippocampal neurons. J. Biol. Chem. 2006, 281, 18802–18815. [Google Scholar] [CrossRef] [PubMed]
  26. Smith, D.F. Quest for biomarkers of treatment-resistant depression: Shifting the paradigm toward risk. Front. Psychiatry 2013, 4, 57. [Google Scholar] [CrossRef]
  27. Serafini, G.; Howland, R.H.; Rovedi, F.; Girardi, P.; Amore, M. The role of ketamine in treatment-resistant depression: A systematic review. Curr. Neuropharmacol. 2014, 12, 444–461. [Google Scholar] [CrossRef]
  28. Zhou, W.; Wang, N.; Yang, C.; Li, X.-M.; Zhou, Z.-Q.; Yang, J.-J. Ketamine-induced antidepressant effects are associated with AMPA receptors-mediated upregulation of mTOR and BDNF in rat hippocampus and prefrontal cortex. Eur. Psychiatry 2014, 29, 419–423. [Google Scholar] [CrossRef]
  29. Hashimoto, K. Role of the mTOR signaling pathway in the rapid antidepressant action of ketamine. Expert Rev. Neurother. 2011, 11, 33–36. [Google Scholar] [CrossRef]
  30. Welberg, L. Psychiatric disorders: Ketamine modifies mood through mTOR. Nat. Rev. Neurosci. 2010, 11, 666. [Google Scholar] [CrossRef]
  31. Lemos, J.C.; Roth, C.A.; Chavkin, C. Signaling events initiated by kappa opioid receptor activation: Quantification and immunocolocalization using phospho-selective KOR, p38 MAPK, and K(IR) 3.1 antibodies. Methods Mol. Biol 2011, 717, 197–219. [Google Scholar] [CrossRef] [PubMed]
  32. Gourley, S.L.; Wu, F.J.; Kiraly, D.D.; PLoSki, J.E.; Kedves, A.T.; Duman, R.S.; Taylor, J.R. Regionally specific regulation of ERK MAP kinase in a model of antidepressant-sensitive chronic depression. Biol. Psychiatry 2008, 63, 353–359. [Google Scholar] [CrossRef] [PubMed]
  33. Mercier, G.; Lennon, A.M.; Renouf, B.; Dessouroux, A.; Ramaugé, M.; Courtin, F.; Pierre, M. MAP kinase activation by fluoxetine and its relation to gene expression in cultured rat astrocytes. J. Mol. Neurosci. 2004, 24, 207–216. [Google Scholar] [CrossRef] [PubMed]
  34. Dwivedi, Y.; Rizavi, H.S.; Zhang, H.; Roberts, R.C.; Conley, R.R.; Pandey, G.N. Aberrant extracellular signal-regulated kinase (ERK)1/2 signalling in suicide brain: Role of ERK kinase 1 (MEK1). Int. J. Neuropsychopharmacol. 2009, 12, 1337–1354. [Google Scholar] [CrossRef]
  35. Kwok, J.B.J.; Hallupp, M.; Loy, C.T.; Chan, D.K.Y.; Woo, J.; Mellick, G.D.; Buchanan, D.D.; Silburn, P.A.; Halliday, G.M.; Schofield, P.R. GSK3B polymorphisms alter transcription and splicing in Parkinson’s disease. Ann. Neurol. 2005, 58, 829–839. [Google Scholar] [CrossRef] [PubMed]
  36. Inkster, B.; Nichols, T.E.; Saemann, P.G.; Auer, D.P.; Holsboer, F.; Muglia, P.; Matthews, P.M. ASsociation of gsk3β polymorphisms with brain structural changes in major depressive disorder. Arch. Gen. Psychiatry 2009, 66, 721–728. [Google Scholar] [CrossRef]
  37. Liu, Z.; Guo, H.; Cao, X.; Cheng, C.; Yang, C.; Xu, C.; Zhang, A.; Sun, N.; Li, X.; Zhang, K. A combined study of GSK3β polymorphisms and brain network topological metrics in major depressive disorder. Psychiatry Res. 2014, 223, 210–217. [Google Scholar] [CrossRef]
  38. Yu, G.I.; Kim, S.K.; Park, H.J.; Kim, J.W.; Chung, J.-H.; Shin, D.H. The C allele of synonymous SNP (rs1142636, Asn170Asn) in SYN1 is a risk factor for the susceptibility of Korean female schizophrenia. Synapse 2012, 66, 979–983. [Google Scholar] [CrossRef]
  39. Garcia, C.C.; Blair, H.J.; Seager, M.; Coulthard, A.; Tennant, S.; Buddles, M.; Curtis, A.; Goodship, J.A. Identification of a mutation in synapsin I, a synaptic vesicle protein, in a family with epilepsy. J. Med. Genet. 2004, 41, 183–186. [Google Scholar] [CrossRef]
  40. Ferreira, A.; Chin, L.-S.; Li, L.; Lanier, L.M.; Kosik, K.S.; Greengard, P. Distinct roles of synapsin I and synapsin II during neuronal development. Mol. Med. 1998, 4, 22–28. [Google Scholar] [CrossRef]
  41. Rosahl, T.W.; Geppert, M.; Spillane, D.; Herz, J.; Hammer, R.E.; Malenka, R.C.; Sudhof, T.C. Short-term synaptic plasticity is altered in mice lacking synapsin I. Cell 1993, 75, 661–670. [Google Scholar] [CrossRef]
  42. Le-Niculescu, H.; Levey, D.F.; Ayalew, M.; Palmer, L.; Gavrin, L.M.; Jain, N.; Winiger, E.; Bhosrekar, S.; Shankar, G.; Radel, M.; et al. Discovery and validation of blood biomarkers for suicidality. Mol. Psychiatry 2013, 18, 1249–1264. [Google Scholar] [CrossRef]
  43. Mullins, N.; Hodgson, K.; Tansey, K.E.; Perroud, N.; Maier, W.; Mors, O.; Rietschel, M.; Hauser, J.; Henigsberg, N.; Souery, D.; et al. Investigation of blood mRNA biomarkers for suicidality in an independent sample. Transl. Psychiatry 2014, 4, e474. [Google Scholar] [CrossRef]
  44. Chen, J.; Li, X.; McGue, M. The interacting effect of the BDNF Val66Met polymorphism and stressful life events on adolescent depression is not an artifact of gene-environment correlation: Evidence from a longitudinal twin study. J. Child Psychol. Psychiatry 2013, 54, 1066–1073. [Google Scholar] [CrossRef]
  45. Chen, J.; Li, X.; McGue, M. Interacting effect of BDNF Val66Met polymorphism and stressful life events on adolescent depression. Genes Brain Behav. 2012, 11, 958–965. [Google Scholar] [CrossRef] [PubMed]
  46. Cole, J.; Weinberger, D.R.; Mattay, V.S.; Cheng, X.; Toga, A.W.; Thompson, P.M.; Powell-Smith, G.; Cohen-Woods, S.; Simmons, A.; McGuffin, P.; et al. No effect of 5HTTLPR or BDNF Val66Met polymorphism on hippocampal morphology in major depression. Genes Brain Behav. 2011, 10, 756–764. [Google Scholar] [CrossRef] [PubMed]
  47. Deng, F.-Y.; Tan, L.-J.; Shen, H.; Liu, Y.-J.; Liu, Y.-Z.; Li, J.; Zhu, X.-Z.; Chen, X.-D.; Tian, Q.; Zhao, M.; et al. SNP rs6265 regulates protein phosphorylation and osteoblast differentiation and influences BMD in humans. J. Bone Miner. Res. 2013, 28, 2498–2507. [Google Scholar] [CrossRef] [PubMed]
  48. Elfving, B.; Buttenschøn, H.N.; Foldager, L.; Poulsen, P.H.; Andersen, J.H.; Grynderup, M.B.; Hansen, A.M.; Kolstad, H.A.; Kaerlev, L.; Mikkelsen, S.; et al. Depression, the Val66Met polymorphism, age, and gender influence the serum BDNF level. J. Psychiatry Res. 2012, 46, 1118–1125. [Google Scholar] [CrossRef]
  49. Chi, M.H.; Chang, H.H.; Lee, S.Y.; Lee, I.H.; Gean, P.W.; Yang, Y.K.; Lu, R.B.; Chen, P.S. Brain derived neurotrophic factor gene polymorphism (Val66Met) and short-term antidepressant response in major depressive disorder. J. Affect. Disord. 2010, 126, 430–435. [Google Scholar] [CrossRef] [PubMed]
  50. Domschke, K.; Lawford, B.; Laje, G.; Berger, K.; Young, R.; Morris, P.; Deckert, J.; Arolt, V.; McMahon, F.J.; Baune, B.T. Brain-derived neurotrophic factor ( BDNF) gene: No major impact on antidepressant treatment response. Int. J. Neuropsychopharmacol. 2010, 13, 93–101. [Google Scholar] [CrossRef]
  51. Choi, M.-J.; Kang, R.-H.; Lim, S.-W.; Oh, K.-S.; Lee, M.-S. Brain-derived neurotrophic factor gene polymorphism (Val66Met) and citalopram response in major depressive disorder. Brain Res. 2006, 1118, 176–182. [Google Scholar] [CrossRef]
  52. Alexopoulos, G.S.; Glatt, C.E.; Hoptman, M.J.; Kanellopoulos, D.; Murphy, C.F.; Kelly, R.E., Jr.; Morimoto, S.S.; Lim, K.O.; Gunning, F.M. BDNF val66met polymorphism, white matter abnormalities and remission of geriatric depression. J. Affect. Disord. 2010, 125, 262–268. [Google Scholar] [CrossRef] [PubMed]
  53. Taylor, W.D.; McQuoid, U.R.; Ashley-Koch, A.; MacFall, J.R.; Bridgers, J.; Krishnan, R.R.; Steffens, D.C. BDNF Val66Met genotype and 6-month remission rates in late-life depression. Pharmacogenom. J. 2011, 11, 146–154. [Google Scholar] [CrossRef] [PubMed]
  54. Tsai, S.-J.; Cheng, C.-Y.; Yu, Y.W.-Y.; Chen, T.-J.; Hong, C.-J. Association study of a brain-derived neurotrophic-factor genetic polymorphism and major depressive disorders, symptomatology, and antidepressant response. Am. J. Med. Genet. B Neuropsy. Genet. 2003, 123B, 19–22. [Google Scholar] [CrossRef] [PubMed]
  55. Yoshida, K.; Higuchi, H.; Kamata, M.; Takahashi, H.; Inoue, K.; Suzuki, T.; Itoh, K.; Ozaki, N. The G196A polymorphism of the brain-derived neurotrophic factor gene and the antidepressant effect of milnacipran and fluvoxamine. J. Psychopharmacol. 2007, 21, 650–656. [Google Scholar] [CrossRef]
  56. Zou, Y.-F.; Wang, Y.; Liu, P.; Feng, X.-L.; Wang, B.-Y.; Zang, T.-H.; Yu, X.; Wei, J.; Liu, Z.-C.; Liu, Y.; et al. Association of BDNF Val66Met polymorphism with both baseline HRQOL scores and improvement in HRQOL scores in Chinese major depressive patients treated with fluoxetine. Um. Psychopharmacol. Clin. Exp. 2010, 25, 145–152. [Google Scholar] [CrossRef] [PubMed]
  57. Cattaneo, A.; Gennarelli, M.; Uher, R.; Breen, G.; Farmer, A.; Aitchison, K.J.; Craig, I.W.; Anacker, C.; Zunsztain, P.A.; McGuffin, P.; et al. Candidate genes expression profile associated with antidepressants response in the GENDEP study: Differentiating between baseline ‘predictors’ and longitudinal ‘targets’. Neuropsychopharmacology 2013, 38, 377–385. [Google Scholar] [CrossRef]
  58. Kang, R.; Chang, H.; Wong, M.-L.; Choi, M.; Park, J.; Lee, H.; Jung, I.; Joe, S.; Kim, L.; Kim, S.; et al. Brain-derived neurotrophic factor gene polymorphisms and mirtazapine responses in Koreans with major depression. J. Psychopharmacol. 2010, 24, 1755–1763. [Google Scholar] [CrossRef]
  59. Musil, R.; Zill, P.; Seemuller, F.; Bondy, B.; Obermeier, M.; Spellmann, I.; Bender, W.; Adli, M.; Heuser, I.; Zeiler, J.; et al. No influence of brain-derived neurotrophic factor (BDNF) polymorphisms on treatment response in a naturalistic sample of patients with major depression. Eur. Arch. Psychiatry Clin. Neurosci. 2013, 263, 405–412. [Google Scholar] [CrossRef]
  60. Correia-Melo, F.S.; Leal, G.C.; Vieira, F.; Jesus-Nunes, A.P.; Mello, R.P.; Magnavita, G.; Caliman-Fontes, A.T.; Echegaray, M.V.F.; Bandeira, I.D.; Silva, S.S.; et al. Efficacy and safety of adjunctive therapy using esketamine or racemic ketamine for adult treatment-resistant depression: A randomized, double-blind, non-inferiority study. J. Affect. Disord. 2020, 264, 527–534. [Google Scholar] [CrossRef]
  61. Mattson, M.P. Glutamate and Neurotrophic Factors in Neuronal Plasticity and Disease. Ann. N. Y. Acad. Sci. 2008, 1144, 97–112. [Google Scholar] [CrossRef]
  62. Salahudeen, M.S.; Wright, C.M.; Peterson, G.M. Esketamine: New hope for the treatment of treatment-resistant depression? A narrative review. Ther. Adv. Drug Saf. 2020, 11, 2042098620937899. [Google Scholar] [CrossRef]
  63. Santos, M.; Carvalho, S.; Lima, L.; Mota-Pereira, J.; Pimentel, P.; Maia, D.; Correia, D.; Gomes, S.; Cruz, A.; Medeiros, R. FAS -670A>G genetic polymorphism Is associated with Treatment Resistant Depression. J. Affect. Disord. 2015, 185, 164–169. [Google Scholar] [CrossRef] [PubMed]
  64. Santos, M.; Carvalho, S.; Lima, L.; Mota-Pereira, J.; Pimentel, P.; Maia, D.; Correia, D.; Gomes, S.; Cruz, A.; Medeiros, R. The role of IL18-607C>A and IL18-137G>C promoter polymorphisms in antidepressant treatment phenotypes: A preliminary report. Neurosci. Lett. 2016, 622, 107–112. [Google Scholar] [CrossRef] [PubMed]
  65. Trivedi, M.H.; Rush, A.J.; Crismon, M.L.; Kashner, T.M.; Toprac, M.G.; Carmody, T.J.; Key, T.; Biggs, M.M.; Shores-Wilson, K.; Witte, B.; et al. Clinical results for patients with major depressive disorder in the Texas Medication Algorithm Project. Arch. Gen. Psychiatry 2004, 61, 669–680. [Google Scholar] [CrossRef] [PubMed]
  66. Gabriel, S.; Ziaugra, L.; Tabbaa, D. SNP Genotyping Using the Sequenom MassARRAY iPLEX Platform. In Current Protocols in Human Genetics; John Wiley & Sons, Inc.: New York, NY, USA, 2001. [Google Scholar] [CrossRef]
  67. Lachmann, A.; Schilder, B.M.; Wojciechowicz, M.L.; Torre, D.; Kuleshov, M.V.; Keenan, A.B.; Ma’ayan, A. Geneshot: Search engine for ranking genes from arbitrary text queries. Nucleic Acids Res. 2019, 47, W571–W577. [Google Scholar] [CrossRef]
  68. Xie, Z.; Bailey, A.; Kuleshov, M.V.; Clarke, D.J.B.; Evangelista, J.E.; Jenkins, S.L.; Lachmann, A.; Wojciechowicz, M.L.; Kropiwnicki, E.; Jagodnik, K.M.; et al. Gene Set Knowledge Discovery with Enrichr. Curr. Protoc. 2021, 1, e90. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effect of MAPK1 rs6928 genotypes in time to relapse in MDD patients. Kaplan–Meier analysis was used to evaluate the association between time to relapse and MAPK1 rs6928 GG and GC + CC carriers. Comparison performed by Log-rank test (p = 0.022).
Figure 1. Effect of MAPK1 rs6928 genotypes in time to relapse in MDD patients. Kaplan–Meier analysis was used to evaluate the association between time to relapse and MAPK1 rs6928 GG and GC + CC carriers. Comparison performed by Log-rank test (p = 0.022).
Ijms 24 06758 g001
Figure 2. Bar chart of top enriched terms from the KEGG_2021_Human gene set library. The top 10 enriched terms for the input gene set are displayed based on the -log10 (p-value), with the actual p-value shown next to each term (*). The term at the top has the most significant overlap with the input query gene set.
Figure 2. Bar chart of top enriched terms from the KEGG_2021_Human gene set library. The top 10 enriched terms for the input gene set are displayed based on the -log10 (p-value), with the actual p-value shown next to each term (*). The term at the top has the most significant overlap with the input query gene set.
Ijms 24 06758 g002
Table 1. Association between genotypes frequencies from the most relevant polymorphisms and AD treatment outcome.
Table 1. Association between genotypes frequencies from the most relevant polymorphisms and AD treatment outcome.
Relapsed Resistant (TRD)
NoYesORCI 95%p-ValueNoYesORCI 95%p-Value
N%N%N%N%
BDNF
rs6265
CC2351.1960.01.0Referent-3253.3529.41.0Referent-
CT1737.8533.30.752[0.213–2.650]0.6572236.71164.73.200[0.975–10.501]0.049
TT511.116.70.511[0.052–5.003]1.000 *610.015.91.067[0.105–10.825]1.000 *
T carrier2248.9640.00.697[0.213–2.284]0.5502846.71270.62.743[0.860–8.750]0.081
PTEN
rs12569998
TT4191.11487.51.0Referent-5590.21368.41.0Referent-
TG48.9212.51.464[0.241–8.881]0.648 *69.8631.64.231[1.173–15.261]0.020
GG--------------
G carrier48.9212.51.464[0.241–8.881]0.648 *69.8631.64.231[1.173–15.261]0.020
SYN1
rs1142636
AA2862.2850.01.0Referent-3559.0631.61.0Referent-
AG1226.7637.51.750[0.498–6.145]0.3801829.5631.62.000[0.564–7.087]0.278
GG511.1212.51.400[0.227–8.626]0.656 *711.5736.86.000[1.543–23.333]0.006
G carrier1737.8850.01.647[0.521–5.204]0.3932541.01368.43.120[1.045–9.314]0.037
GSK3B
rs6438552
AA2044.4212.51.0Referent-2236.1947.41.0Referent-
AG1635.6850.05.000[0.929–26.913]0.074 *2439.3842.10.815[0.267–2.483]0.718
GG920.0637.56.667[1.121–39.660]0.042 *1524.6210.50.326[0.062–1.726]0.284 *
G carrier2755.61487.55.600[1.137–27.571]0.033 *3963.91052.60.627[0.221–1.775]0.377
MAPK1
rs6928
GG1840.01168.81.0Referent-2947.5526.31.0Referent-
GC1942.2318.80.258[0.062–1.080]0.066 *2236.11157.92.900[0.879–9.567]0.074
CC817.8212.40.409[0.073–2.288]0.445 *1016.4315.81.740[0.351–8.633]0.666 *
C carrier2760.0531.20.303[0.090–1.020]0.048 *3253.51473.72.538[0.813–7.919]0.102
TRD: Treatment Resistant Depression. OR: odds ratio. CI: confidence interval. Significant p values in bold. * Fisher exact test.
Table 2. Overall results of enrichment analysis using Enrichr.
Table 2. Overall results of enrichment analysis using Enrichr.
IndexNamep-ValueAdjusted
p-Value
Odds RatioCombined Score
GO Biological Process
1chemical synaptic transmission (GO:0007268)5.186 × 10−291.086 × 10−2522.001432.77
2anterograde trans-synaptic signaling (GO:0098916)3.609 × 10−253.778 × 10−2222.431262.23
3modulation of chemical synaptic transmission (GO:0050804)3.001 × 10−202.095 × 10−1733.151489.97
4glutamate receptor signaling pathway (GO:0007215)7.002 × 10−163.666 × 10−1367.142343.00
5cellular response to cytokine stimulus (GO:0071345)2.163 × 10−159.060 × 10−139.56322.96
6regulation of NMDA receptor activity (GO:2000310)8.579 × 10−152.994 × 10−1271.582318.44
7positive regulation of gene expression (GO:0010628)2.058 × 10−145.422 × 10−129.08286.14
8positive regulation of multicellular organismal process (GO:0051240)2.071 × 10−145.422 × 10−1211.01346.99
9calcium ion transmembrane import into cytosol (GO:0097553)1.848 × 10−134.224 × 10−1149.191442.14
10negative regulation of apoptotic process (GO:0043066)2.117 × 10−134.224 × 10−118.55249.46
GO Molecular Function
1receptor ligand activity (GO:0048018)7.579 × 10−192.062 × 10−1614.83618.62
2glutamate receptor activity (GO:0008066)2.072 × 10−152.818 × 10−13140.704757.03
3hormone activity (GO:0005179)5.605 × 10−155.081 × 10−1332.191056.44
4ionotropic glutamate receptor activity (GO:0004970)8.568 × 10−145.826 × 10−12137.884148.62
5glutamate-gated calcium ion channel activity (GO:0022849)1.350 × 10−117.344 × 10−1099,320.002,485,812.34
6ligand-gated channel activity (GO:0022834)1.916 × 10−118.684 × 10−1056.371391.10
7ligand-gated ion channel activity (GO:0015276)2.567 × 10−119.976 × 10−1053.921314.77
8cytokine activity (GO:0005125)1.854 × 10−106.302 × 10−913.02291.67
9NMDA glutamate receptor activity (GO:0004972)7.438 × 10−102.248 × 10−8252.685311.25
10G protein-coupled glutamate receptor activity (GO:0001640)1.664 × 10−94.116 × 10−8189.503830.60
GO Cellular Component
1neuron projection (GO:0043005)3.351 × 10−244.290 × 10−2212.87695.42
2dendrite (GO:0030425)1.012 × 10−176.476 × 10−1615.26597.32
3ionotropic glutamate receptor complex (GO:0008328)5.840 × 10−132.170 × 10−1161.131722.04
4axon (GO:0030424)6.781 × 10−132.170 × 10−1113.95391.00
5integral component of plasma membrane (GO:0005887)2.656 × 10−116.799 × 10−104.50109.69
6postsynaptic density membrane (GO:0098839)4.142 × 10−118.836 × 10−1082.861980.97
7postsynaptic specialization membrane (GO:0099634)6.178 × 10−111.130 × 10−976.941808.62
8NMDA selective glutamate receptor complex (GO:0017146)7.438 × 10−101.190 × 10−8252.685311.25
9postsynaptic density (GO:0014069)2.637 × 10−93.750 × 10−813.68270.15
10cation channel complex (GO:0034703)3.444 × 10−84.408 × 10−719.04327.14
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Santos, M.; Lima, L.; Carvalho, S.; Mota-Pereira, J.; Pimentel, P.; Maia, D.; Correia, D.; Barroso, M.F.; Gomes, S.; Cruz, A.; et al. The Impact of BDNF, NTRK2, NGFR, CREB1, GSK3B, AKT, MAPK1, MTOR, PTEN, ARC, and SYN1 Genetic Polymorphisms in Antidepressant Treatment Response Phenotypes. Int. J. Mol. Sci. 2023, 24, 6758. https://doi.org/10.3390/ijms24076758

AMA Style

Santos M, Lima L, Carvalho S, Mota-Pereira J, Pimentel P, Maia D, Correia D, Barroso MF, Gomes S, Cruz A, et al. The Impact of BDNF, NTRK2, NGFR, CREB1, GSK3B, AKT, MAPK1, MTOR, PTEN, ARC, and SYN1 Genetic Polymorphisms in Antidepressant Treatment Response Phenotypes. International Journal of Molecular Sciences. 2023; 24(7):6758. https://doi.org/10.3390/ijms24076758

Chicago/Turabian Style

Santos, Marlene, Luis Lima, Serafim Carvalho, Jorge Mota-Pereira, Paulo Pimentel, Dulce Maia, Diana Correia, M. Fátima Barroso, Sofia Gomes, Agostinho Cruz, and et al. 2023. "The Impact of BDNF, NTRK2, NGFR, CREB1, GSK3B, AKT, MAPK1, MTOR, PTEN, ARC, and SYN1 Genetic Polymorphisms in Antidepressant Treatment Response Phenotypes" International Journal of Molecular Sciences 24, no. 7: 6758. https://doi.org/10.3390/ijms24076758

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop