Next Article in Journal
Genetic Analysis Based on Mitochondrial nad2 Gene Reveals a Recent Population Expansion of the Invasive Mussel, Mytella strigata, in China
Previous Article in Journal
Unraveling the Genetic Threads of History: mtDNA HVS-I Analysis Reveals the Ancient Past of the Aburra Valley
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Association between NTRK2 Polymorphisms, Hippocampal Volumes and Treatment Resistance in Major Depressive Disorder

1
Psychiatry and Clinical Psychobiology Unit, Division of Neuroscience, IRCCS San Raffaele Scientific Institute, 20132 Milan, Italy
2
Mood Disorders Unit, IRCCS San Raffaele Scientific Institute, 20132 Milan, Italy
3
Faculty of Medicine, Vita-Salute San Raffaele University, 20132 Milan, Italy
4
Faculty of Psychology, Vita-Salute San Raffaele University, 20132 Milan, Italy
*
Author to whom correspondence should be addressed.
Genes 2023, 14(11), 2037; https://doi.org/10.3390/genes14112037
Submission received: 30 September 2023 / Revised: 28 October 2023 / Accepted: 31 October 2023 / Published: 3 November 2023
(This article belongs to the Section Human Genomics and Genetic Diseases)

Abstract

:
Despite the increasing availability of antidepressant drugs, a high rate of patients with major depression (MDD) does not respond to pharmacological treatments. Brain-derived neurotrophic factor (BDNF)-tyrosine receptor kinase B (TrkB) signaling is thought to influence antidepressant efficacy and hippocampal volumes, robust predictors of treatment resistance. We therefore hypothesized the possible role of BDNF and neurotrophic receptor tyrosine kinase 2 (NTRK2)-related polymorphisms in affecting both hippocampal volumes and treatment resistance in MDD. A total of 121 MDD inpatients underwent 3T structural MRI scanning and blood sampling to obtain genotype information. General linear models and binary logistic regressions were employed to test the effect of genetic variations related to BDNF and NTRK2 on bilateral hippocampal volumes and treatment resistance, respectively. Finally, the possible mediating role of hippocampal volumes on the relationship between genetic markers and treatment response was investigated. A significant association between one NTRK2 polymorphism with hippocampal volumes and antidepressant response was found, with significant indirect effects. Our results highlight a possible mechanistic explanation of antidepressant action, possibly contributing to the understanding of MDD pathophysiology.

1. Introduction

Major depressive disorder (MDD) is a highly prevalent psychiatric condition, associated with relatively high rates of disability and mortality [1]; antidepressants, mainly selective serotonin reuptake inhibitors (SSRIs) and serotonin and norepinephrine reuptake inhibitors (SNRIs), are currently used as the first-line pharmacological treatment of MDD [2]. However, response and remission rates to first-line treatments are modest [3], and antidepressant switch, treatment augmentation strategies or even electroconvulsive therapy are often required [4].
In recent years, it has become clear that numerous drugs have antidepressant properties without directly acting on monoamine reuptake. Additional putative antidepressant mechanisms of action include direct modulation of serotoninergic [5,6] and noradrenergic receptors [7], action on glutamatergic N-methyl-D-aspartate (NMDA) receptors [8,9] and modulation of γ-aminobutyric acid (GABA) signaling [10,11]. Furthermore, novel antidepressants acting on opioid, trace amines, cannabinoid, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and metabotropic glutamatergic receptors, as well as immuno-modulatory drugs for depression, are currently being developed [12,13].
This plethora of putative antidepressant mechanisms highlights the limitation of the old monoamine hypothesis of depression—that is, that depression stems from alterations of monoaminergic neurotransmission [14]—and suggests the existence of a common final pathway through which different pharmacological mechanisms lead to antidepressant effects.
A leading candidate for the role of a common pathway of antidepressant action is the neuroplasticity hypothesis [15]: this theory postulates a crucial role for brain-derived neurotrophic factor (BDNF) and its receptor, tyrosine receptor kinase B (TrkB), in both MDD pathophysiology and antidepressant efficacy [16]. BDNF is a member of the nerve growth factor family, which is crucial during brain development but is also expressed in adulthood, where it regulates synaptic plasticity, neuronal function and survival [17]. Lower levels of BDNF protein and mRNA have been identified in postmortem brain tissue of depressed patients, particularly in the hippocampus [18,19], as well as of people who died of suicide [20]; increased BDNF gene metilation and reduced BDNF expression has been found in peripheral blood mononuclear cells of depressed and suicidal patients [21]. Likewise, reduced levels of BDNF receptor TrkB, of its activated phosphorylated form and mRNA, have been identified in postmortem brains of depressed individuals [20,22,23]. Further, alterations of BDNF-TrkB signaling in different brain regions including the hippocampus have been shown to mediate the relation between lipopolysaccharide-induced inflammation and depression-like phenotypes in animal models.
While reductions of BDNF are not specific to depression, having been reported also in other psychiatric or neurological conditions, the BDNF-TrkB pathway seems to be crucially involved in antidepressant response [16]. Monoaminergic antidepressants, ketamine and electroconvulsive therapy (ECT) all induce increases in brain BDNF, both in mouse models and in humans [24,25,26], and BDNF signaling through TrkB appears to be crucial for antidepressant effects [27]. Deletion of both BDNF or TrkB in hippocampal neurons inhibits the effect of antidepressants in mouse models [28,29]; at the same time, only deletion of TrkB and not of BDNF in the dorsal raphe reduces antidepressant responses [30]. According to a recent suggestive hypothesis, different classes of antidepressants might exert, at least partially, their activity directly binding to the TrkB receptor, stabilizing its trans-membrane configuration and promoting BDNF-mediated signaling [16,31]; both typical monoaminergic antidepressants (selective serotonin reuptake inhibitors (SSRIs) and serotonin and norepinephrine reuptake inhibitors (SNRIs)) and ketamine have been found to have substantial binding affinity for TrkB [31].
Reduced hippocampal volumes are among the few robust structural MRI predictors of poor antidepressant response and treatment resistance [32,33]. Reduced hippocampal volumes in MDD may reflect impaired neuroplasticity and neurogenesis [14], thought to be crucial in antidepressant activity [34]; indeed, damage to hippocampal neurogenesis via X-irradiation of the mouse brain inhibits the effects of antidepressants [35]. BDNF, one of the main neurotrophic factors in the adult brain [17,36], plays a crucial role in hippocampal neurogenesis, and its levels have been repeatedly associated with hippocampal volumes [37,38,39], albeit somewhat inconsistently [40]. While the effect of BDNF polymorphisms on hippocampal volumes has been the subject of extensive research [41,42,43], the effect of genetic variation in the region coding for its receptor has been much less investigated; one study, however, reported a significant association in a general population sample [44].
The effects of BDNF genetic polymorphisms on antidepressant efficacy have been widely investigated [45]. One particular polymorphism, Val66Met, received much attention as it influences BDNF trafficking and release and affects antidepressant response in mouse models [46]; however, such effects do not translate so clearly to humans, as heterozygous subjects for Val66Met appear to have a better response to antidepressants than either Val or Met homozygous subjects [16,47].
Effects of genetic variation in the TrkB gene—NTRK2—on the other hand, have been much less studied. Neurotrophic receptor tyrosine kinase 2 (NTRK2) polymorphisms have been shown to influence response to mood-stabilizing agents in bipolar patients [48]. Concerning antidepressant response, while some studies fail to identify an association with polymorphisms mapping to the NTRK2 gene [49,50], others do report such an association [51,52]. T/T carriers in the rs1565445 NTRK2 polymorphism were reported to have higher risk of developing treatment-resistant depression compared to patients with genotype C/C and T/C [51]. Hennings and colleagues found nine NTRK2 single nucleotide polymorphisms (SNPs) to be associated with antidepressant response patterns (response at 5 weeks and remission at discharge), and for three of such SNPs, replicated the association in independent samples [52]. Polymorphisms in the NTRK2 region have also been associated with antidepressant treatment-emergent suicidal ideation, an indirect marker of poor treatment efficacy [53].
Here, our aim was to investigate the interplay between polymorphism mapping of the BDNF and NTRK2 gene region, hippocampal volumes and treatment resistance in a sample of MDD inpatients in a real-world clinical setting. We hypothesized that genetic variants of BDNF or NTRK2 genes would associate with hippocampal volume which, in turn, would affect treatment resistance.

2. Materials and Methods

2.1. Participants

Our study was performed on a sample of 126 inpatients with a diagnosis of MDD suffering from a major depressive episode without psychotic features (DSM-5 criteria) admitted to the Mood Disorders ward of San Raffaele Hospital in Milan. Patients had been referred for hospital specialized clinical treatment of depression by their general practitioners or psychiatrists in charge. Treatment was administered by staff psychiatrists upon clinical need. Prescribed pharmacotherapies can be seen in Supplementary Table S3.
Exclusion criteria were current diagnosis of any additional psychiatric disorder, including alcohol and/or substance dependence or abuse in the last 6 months, intellectual disability, pregnancy, major medical and neurological disorders.
Treatment resistance status was obtained for each patient from the clinical charts. Patients were defined as being treatment-resistant if, during the current depressive episode, they had failed to respond to at least two separate antidepressant treatments with different mechanisms of action, administered to a dose equal or superior to the minimum licensed dose for at least 4 weeks [54,55].
After a complete description of the study was given to the participants, written informed consent was obtained. All the research activities were approved by the local ethical committee. The described work has been carried out in accordance with The Code of Ethics of the World Medical Association (Declaration of Helsinki) for experiments involving humans.

2.2. Genetic Analysis

All patients underwent venous blood sampling to obtain genetic data. Blood samples were genotyped using the Infinium PsychArray 24 BeadChip (Illumina, Inc., San Diego, CA, USA), which is a cost-effective, high-density microarray developed in collaboration with the Psychiatric Genomics Consortium and containing 595,427 genetic markers associated with psychiatric disorders (https://www.illumina.com/products/by-type/microarray-kits/infinium-psycharray.html, accessed on 30 September 2023). Specifically looking at genetic variants mapping to the BDNF or NTRK2 genes regions, 46 markers for NTRK2 and 34 for BDNF are included in this array. Genotype quality control was conducted with PLINK1.9 [56]. First, markers with minor allele frequency (MAF) <5%, call rate <95% or deviant from Hardy–Weinberg equilibrium at p < 10−6 were excluded, leaving 31 markers for NTRK2 and 17 for BDNF.
Secondly, individuals with discrepant genotyped sex information, genotype rate <95% or outlying autosomal heterozygosity (Fhet > ±0.2) were removed. Finally, relatedness of participants was checked by excluding individuals with a degree of recent shared ancestry (i.e., identity by descendent, IBD) >0.1875, which is the threshold corresponding to halfway between third- and second-degree relatives [57]. Therefore, the final sample consisted of 121 MDD patients.

2.3. Brain Imaging

T1-weighted images were acquired on two 3.0 Tesla scanners: 44 patients underwent a 3T MRI scan in a Gyroscan Intera scanner, Philips, The Netherlands, employing an 8-channel SENSE head coil (T1-weighted MPRAGE sequences: TR 25.00 ms, TE 4.6 ms, field of view FOV = 230 mm, matrix = 256 × 256, in-plane resolution 0.9 × 0.9 mm, yielding 220 transversal slices with a thickness of 0.8 mm); 77 patients were scanned in a 3T Ingenia CX scanner, Philips, The Netherlands, using a 32-channel sensitivity-encoding SENSE head coil (T1-weighted MPRAGE sequence: TR 8.00 ms, TE 3.7 ms, field of view FOV = 256 mm, matrix = 256 × 256, in-plane resolution 1 × 1 mm, yielding 182 transversal slices with a thickness of 1 mm).
Images were visually inspected as part of the quality check procedure, and then processed using the Computational Anatomy Toolbox (CAT12) preprocessing pipeline [58] for SPM12 (https://www.fil.ion.ucl.ac.uk/spm/software/spm12/, accessed on 30 September 2023) in Matlab R2016b, which also allows for the extraction of ROIs tissue volumes. This included segmentation into gray matter, white matter and cerebrospinal fluid, bias regularization, non-linear modulation and normalization to MNI space using DARTEL to a 1.5 mm isotropic MNI template. Bilateral hippocampal volumes were estimated according to the Neuromorphometric Atlas and converted into percentage of total intracranial volume (TIV) with the formula: (Hippocampal Volume × 100)/TIV [59]. We will refer to the result of this calculation as normalized hippocampal volume.

2.4. Statistical Analysis

General linear models and binary logistic regression analyses were performed with StatSoft Statistica 12 (Tulsa, OK, USA); parameter estimates were obtained with least squares maximum likelihood procedures. The significance of the effects was calculated with the likelihood ratio statistic. For mediation analyses, we used Hayes’ SPSS “Process” Macro, version 4.0 with IBM SPSS statistics for Windows, Version 26.0 (IBM SPSS Statistics for Windows, Version 26.0., Armonk, NY, USA: IBM Corp.) [60].
The effect of genetic variations in BDNF and NTRK2 on bilateral hippocampal volumes were tested in the context of general linear models, entering left or right hippocampal volume as dependent variables, polymorphisms as independent variables and age, sex and MRI scan as nuisance covariates. The effect of each polymorphism was tested in a dominant (MM versus Mm + mm), recessive (MM + Mm versus mm) and over-dominant genetic model (MM + mm versus Mm) [61]. When no homozygous subject for the minor allele was identified, only the comparison between the remaining two groups was performed. This resulted in 272 comparisons, setting the Bonferroni threshold for statistical significance to p < 0.00018 (Supplementary Tables S1 and S2) [62].
The effect of genetic polymorphisms on treatment resistance status was tested with binary logistic regressions, entering treatment resistance as a dichotomous dependent variable, polymorphisms as independent variables and age and sex as nuisance covariates. Again, the dominant, recessive and over-dominant effects of polymorphisms were tested, without probing the effect of genetic models with fewer than three subjects in one group; this resulted in 128 tests, with a Bonferroni-corrected threshold of significance of p < 0.00039 (Supplementary Tables S1 and S2).
The effect of hippocampal volumes on treatment resistance status was again tested in the context of logistic regression, with resistance as a dependent variable, left or right hippocampal volumes as independent variables and age, sex and MRI scan as nuisance covariates.
After reviewing the results of the previous analyses, we tested the indirect effect of significant polymorphisms on treatment resistance through hippocampal volumes in a mediation model, entering age, sex and MRI scan as nuisance covariates. Given the high correlation between left and right hippocampal volumes (r = 0.898), a possible indirect effect of left hippocampal volume was also tested.

3. Results

Clinical and demographic characteristics of the sample can be seen in Table 1. Patients with treatment-resistant depression had significantly longer hospital stays and lower bilateral hippocampal volumes. Polymorphism distribution across subjects is presented in Supplementary Table S1.
After applying Bonferroni HWE correction for multiple comparisons, one NTRK2 polymorphism (rs1948308) was found to be significantly associated with right hippocampal volume in the over-dominant genetic model; heterozygote patients had significantly lower right hippocampal volume compared to homozygotes (p = 0.000082).
Testing the effect of all polymorphisms on treatment resistance, rs1948308 was again found to be significantly associated with treatment resistance in the over-dominant genetic model; heterozygote patients for the polymorphism had a significantly higher chance of being classified as treatment-resistant than AA or GG homozygotes (p = 0.000075).
Bilateral hippocampal volumes were significantly associated with treatment resistance (L: p = 0.000149; R: p = 0.000420).
The mediation model confirmed that rs1948308 affected bilateral hippocampal volumes (R: p < 0.001; L: p = 0.002), which in turn influenced treatment resistance status (R: p = 0.024; L: p = 0.006). Therefore, a significant indirect effect of rs1948308 heterozygosis on treatment resistance through bilateral hippocampal volumes was identified (R: 95% BCa CI [0.060, 0.853] (Figure 1); L: 95% BCa CI [0.079, 0.800]); in both cases, the direct path of the model also remained significant (R: p = 0.0031; L: p = 0.0018) (Figure 1).
No statistically significant results were identified for polymorphisms involving the BDNF gene region (see Supplementary Table S2).

4. Discussion

The main finding of the present study is the association between one polymorphism mapping to the NTRK2 region and antidepressant treatment resistance, mediated by hippocampal volumes. Reduced hippocampal volumes are a robust finding in neuroimaging studies of MDD patients [63] and are thought to reflect impaired neuroplasticity and neurogenesis processes [14], both of which are strongly affected by the BDNF-TrkB signaling pathway [64]. BDNF is a highly expressed neurotrophic factor in the adult brain, and through its high-affinity receptor—TrkB—plays a crucial role in regulating neuronal survival and synaptic plasticity [65]; furthermore, BDNF-TrkB have also been shown to play a pivotal role in promoting neurogenesis, the formation of new neurons that occurs in some regions of the adult brain, including the hippocampus [64]. Chronic stress reduces hippocampal survival and proliferation of progenitor cells, as well as dendritic complexity [66], possibly through its effect on BDNF [67]. Indeed, a model of depressive symptomatology has been proposed, postulating a stress-induced reduction in BDNF signaling, reflected in lower hippocampal volumes, and restored through successful antidepressant therapy [68].
Accordingly, genetic variations both in BDNF and in NTRK2 have been associated with hippocampal volumes, in healthy and psychiatric samples alike [42,43,44]. Lower bilateral hippocampal volumes, in turn, are one of the best replicated markers of poor antidepressant efficacy and of treatment resistance [32]. Our data suggest that NTRK2 genetic variants affect antidepressant resistance at least partially through their effect on hippocampal volumes.
In recent years, it has become apparent that immune/inflammatory alterations are involved in MDD pathophysiology [69,70] and depressive symptomatology is now thought to be a consequence of low-grade chronic inflammatory status, at least in a subset of patients [71]. Inflammation has been suggested to lead to depressive symptomatology through its effect on brain integrity including a reduction in hippocampal volumes [72,73], and has also been linked to poor antidepressant efficacy [74]. Recently, it has been shown that lipopolysaccharide-induced inflammation alters BDNF signaling in widespread brain regions including the hippocampus [75]; the BDNF-TrkB pathway, therefore, might represent the link between alterations in immune/inflammatory status and hippocampal volume, a putative neurogenesis/neuroplasticity biomarker [37,38,39] robustly linked to depressive symptomatology and treatment efficacy [32].
It has been suggested that BDNF high-affinity receptor, TrkB, encoded by the NTRK2 gene [76], plays a crucial role in biological processes leading to antidepressant response [77]. NTRK2’s rare esonic mutations impair hippocampal synaptogenesis and associate with neurobehavioral abnormalities [78]. Recently, various antidepressant compounds of different classes (including SSRIs, SNRIs and ketamine) have been shown to be able to directly bind TrkB, enhancing its signal transduction [31]; a provocative hypothesis, therefore, has been put forward, postulating a direct contribution of TrkB binding to the antidepressant (AD) mechanism of action [79]. A more conservative hypothesis postulates that AD-induced increase in monoaminergic neurotransmission may enhance brain neuroplasticity and neurogenesis, thus leading to antidepressant effects; this framework also provides an explanation for the latency of action of “classic” monoaminergic antidepressants [80]. The BDNF-TrkB signaling pathway is commonly regarded as a crucial mediator of this process [15], and growing evidence also links it to the action of fast-acting AD treatments, such as ECT and ketamine [9,81]. Regardless of the precise pathways through which ADs with various mechanism of actions exert their clinical effect—whether through a direct interaction with the TrkB receptor or an indirect increase in brain BDNF levels—BDNF-TrkB signaling emerges as a crucial molecular pathway of antidepressant action.
The polymorphism we identified was in a non-coding region, suggesting its effect on hippocampal volumes and treatment resistance are linked to genetic regulatory processes [82]. Interestingly, rs1948308, in addition to having been identified in pharmacogenetic studies of opioid use disorder [83], had previously been linked to antidepressant efficacy, although it failed to survive multiple comparison correction [52]. rs1948308 has also been robustly associated with several psychiatric phenotypes, such as neuroticism and mood swings [84,85].
We observed lower hippocampal volumes and higher rates of treatment resistance in heterozygous subjects compared to homozygous for either allele (over-dominant genetic model) [61]. Although this finding might appear counter-intuitive and certainly warrants further investigation, studies investigating the effect of BDNF Val66Met polymorphism (known to affect intracellular BDNF trafficking and release) on antidepressant efficacy also report significant effect in heterozygous subjects, with Val/Met carriers characterized by better treatment response [16,47]. However, no polymorphism in the BDNF region, including Val66Met, affected hippocampal volumes or treatment resistance in our sample. Further, although not surviving correction for multiple comparisons, we observed a general effect of the recessive allele on both TRD and hippocampal volumes (Supplementary Table S2). This may suggest a possible effect of the recessive allele that may not be visible here because of the reduced sample size but that could have a more easily interpretable biological meaning. To better understand this issue, in vitro studies could help highlight the biological effect of this genetic variation.
The observation that we found a significant result for TrkB and not BDNF genetic variants may suggest that the association between hippocampal volume and treatment resistance is not associated with the amount of BDNF released, influenced among others by the Val66Met polymorphism, but instead to its link with its receptor. TrkB is a tyrosine kinase receptor that, upon binding to its ligands—mainly BDNF but also neurotrophin 3 and 4—undergoes dimerization and starts signal transduction [86]; it is widely expressed in both peripheral and central nervous systems, including the hippocampus [87], and regulates cell growth and survival [85]. Its expression has been observed, however, in other brain cell types, such as astrocytes and microglia, with less clear biological functions [88,89]. Interestingly, a truncated form of the receptor, without the catalytic tyrosine kinase domain, is also expressed in the brain, and appears to have numerous regulatory functions [90]. Furthermore, TrkB expression on cell surface is highly regulated via numerous modulatory mechanisms, such as cAMP levels, N-glycosylation and receptor transactivation, and this strongly influences cell response to BDNF, even in short time periods [91]. All of these complex and highly integrated regulatory mechanisms of the BDNF-TrkB signaling, both on neurons and on other cell types, could be affected by the polymorphism we identified.
Our results suggest a link between BDNF-TrkB signaling, hippocampal volumes and antidepressant efficacy in MDD patients. The observation that numerous compounds and procedures, with widely different (and, at times, poorly understood) mechanisms of action, have antidepressant properties [92], has long raised the need for a comprehensive explanatory framework of MDD treatment [15]. One of the leading hypotheses postulates a common effect of antidepressants on brain neuroplasticity [93], modulated at least partially by their effect on BDNF-TrkB signaling [15]. Antidepressant treatment has been shown to increase BDNF in peripheral blood [94], in CSF [95] and in the brain [96]. BDNF in turn enhances neural plasticity in depressed brains, a process thought to be crucial in ameliorating depressive symptomatology [97]; precisely for this reason, it is considered to be a crucial mediator of antidepressant effects [98,99]. In this framework, reduced hippocampal volumes that associate with MDD and poor antidepressant response might reflect impaired neural plasticity processes [14], and antidepressant-induced hippocampal volumes increase, as was repeatedly observed [100,101], might reflect a BDNF/TrkB mediated restoration of neuroplasticity and neurogenesis.
This study has various limitations: the recruitment in a single center and the absence of a healthy control group limits the generalizability of our findings. Our sample size, albeit typical for a neuroimaging study, is smaller than those usually employed in genetic analyses. To account for this, given our research question, we a priori selected polymorphisms of interest instead of employing a genome-wide approach; however, we selected all available SNPs for BDNF and NTRK2 in our genetic array and we applied a strict Bonferroni multiple comparison correction to all our analyses. Likewise, in our MRI analysis, we did not employ a voxel-wise approach, but rather we performed a hypothesis-driven region-of-interest analysis on the hippocampus. Future better-powered multi-center studies could help to overcome these limitations.
Given the real-world nature of our study, patients were prescribed a wide variety of antidepressants drugs, whose effect we did not control for; however, we employed consensus definitions of treatment resistance widely used in real world studies [54]. Furthermore, we could not control for factors known to affect both hippocampal volumes and treatment resistance, such as adiposity and insulin resistance, inflammatory status or childhood trauma. No depression severity cut-off was used as an inclusion criteria; as a result, included patients had varied depression severity levels (Table 1). Albeit, no difference in baseline HDRS scores was identified between TRD and non-TRD patients; whether depression severity influences the relation between NTRK2 polymorphisms, hippocampal volumes and treatment resistance remains to be determined.
So far, large genome-wide association studies (GWAS) did not find genetic variation in NTRK2 or BDNF regions to associate with antidepressant efficacy [102,103]. Different study populations (both in diagnosis and severity), definitions of treatment resistance and methods employed to assess antidepressant efficacy might all justify this inconsistency. Nevertheless, this implies that our results on treatment resistance must be taken with caution. Likewise, genetic variation in the NTRK2 region was not found to be associated with hippocampal volumes in GWAS studies, which however were performed mostly on healthy individuals [104,105,106]. Alterations of the BDNF-TrkB signaling pathway might on the other hand be a specific feature of MDD pathophysiology [17], and polymorphisms affecting TrkB expression might contribute to the established reduction in hippocampal volumes in this population [63].
Despite these limitations, this study supports the notion that BDNF/TrkB genetic variations play a role in hippocampal-mediated antidepressant resistance, contributes to a better understanding of MDD pathophysiology and points to possible markers of antidepressant efficacy which, moving towards a personalized approach to treatments, might become increasingly relevant in future real-world clinical practice.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes14112037/s1, Table S1: Polymorphisms distribution in the study sample. Table S2: Effect of polymorphisms on hippocampal volumes and treatment resistance. Table S3: Details on prescribed pharmacotherapies.

Author Contributions

Conceptualization, M.P., S.P. and F.B.; methodology, M.P. and L.F.-U.; formal analysis, M.P. and L.F.-U.; investigation, C.L., S.S. and M.M.; writing—original draft preparation, M.P., S.P. and L.F.-U.; writing—review and editing, R.Z., C.C. and S.P.; supervision, S.P. and F.B. All authors have read and agreed to the published version of the manuscript.

Funding

M.P.: Italian Ministry of University, XXXVI PhD cycle. L.F.-U.: Italian Ministry of University, XXXVIII PhD cycle. This work was supported by the Italian Ministry of Health, grant RF-2018-12367249 and the Italian Ministry of University and Scientific Research, grant A_201779W93T.

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki, and approved by the local Ethics Committee. Protocol n. 10-06-SO, date of approval: 30 June 2006.

Informed Consent Statement

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

Data Availability Statement

Data available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Marx, W.; Penninx, B.W.J.H.; Solmi, M.; Furukawa, T.A.; Firth, J.; Carvalho, A.F.; Berk, M. Major depressive disorder. Nat. Rev. Dis. Primers 2023, 9, 44. [Google Scholar] [CrossRef] [PubMed]
  2. Kendrick, T.; Pilling, S.; Mavranezouli, I.; Megnin-Viggars, O.; Ruane, C.; Eadon, H.; Kapur, N. Management of depression in adults: Summary of updated NICE guidance. BMJ 2022, 378, o1557. [Google Scholar] [CrossRef] [PubMed]
  3. Sinyor, M.; Schaffer, A.; Levitt, A. The sequenced treatment alternatives to relieve depression (STAR* D) trial: A review. Can. J. Psychiatry 2010, 55, 126–135. [Google Scholar] [CrossRef] [PubMed]
  4. Voineskos, D.; Daskalakis, Z.J.; Blumberger, D.M. Management of treatment-resistant depression: Challenges and strategies. Neuropsychiatr. Dis. Treat. 2020, 16, 221–234. [Google Scholar] [CrossRef] [PubMed]
  5. Okada, M.; Fukuyama, K.; Okubo, R.; Shiroyama, T.; Ueda, Y. Lurasidone sub-chronically activates serotonergic transmission via desensitization of 5-HT1A and 5-HT7 receptors in dorsal raphe nucleus. Pharmaceuticals 2019, 12, 149. [Google Scholar] [CrossRef] [PubMed]
  6. Stahl, S.M. Modes and nodes explain the mechanism of action of vortioxetine, a multimodal agent (MMA): Actions at serotonin receptors may enhance downstream release of four pro-cognitive neurotransmitters. CNS Spectr. 2015, 20, 515–519. [Google Scholar] [CrossRef] [PubMed]
  7. Anttila, S.A.; Leinonen, E.V. A review of the pharmacological and clinical profile of mirtazapine. CNS Drug Rev. 2001, 7, 249–264. [Google Scholar] [CrossRef] [PubMed]
  8. Ostadhadi, S.; Ahangari, M.; Nikoui, V.; Norouzi-Javidan, A.; Zolfaghari, S.; Jazaeri, F.; Chamanara, M.; Akbarian, R.; Dehpour, A.-R. Pharmacological evidence for the involvement of the NMDA receptor and nitric oxide pathway in the antidepressant-like effect of lamotrigine in the mouse forced swimming test. Biomed. Pharmacother. 2016, 82, 713–721. [Google Scholar] [CrossRef]
  9. Zanos, P.; Gould, T. Mechanisms of ketamine action as an antidepressant. Mol. Psychiatry 2018, 23, 801–811. [Google Scholar] [CrossRef]
  10. Chen, S.; Gao, L.; Li, X.; Ye, Y. Allopregnanolone in mood disorders: Mechanism and therapeutic development. Pharmacol. Res. 2021, 169, 105682. [Google Scholar] [CrossRef]
  11. Althaus, A.L.; Ackley, M.A.; Belfort, G.M.; Gee, S.M.; Dai, J.; Nguyen, D.P.; Kazdoba, T.M.; Modgil, A.; Davies, P.A.; Moss, S.J.; et al. Preclinical characterization of zuranolone (SAGE-217), a selective neuroactive steroid GABAA receptor positive allosteric modulator. Neuropharmacology 2020, 181, 108333. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, Q.; Dwivedi, Y. Advances in novel molecular targets for antidepressants. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 104, 110041. [Google Scholar] [CrossRef] [PubMed]
  13. Jarończyk, M.; Walory, J. Novel molecular targets of antidepressants. Molecules 2022, 27, 533. [Google Scholar] [CrossRef]
  14. Boku, S.; Nakagawa, S.; Toda, H.; Hishimoto, A. Neural basis of major depressive disorder: Beyond monoamine hypothesis. Psychiatry Clin. Neurosci. 2018, 72, 3–12. [Google Scholar] [CrossRef] [PubMed]
  15. Harmer, C.J.; Duman, R.S.; Cowen, P.J. How do antidepressants work? New perspectives for refining future treatment approaches. Lancet Psychiatry 2017, 4, 409–418. [Google Scholar] [CrossRef] [PubMed]
  16. Castrén, E.; Monteggia, L.M. Brain-derived neurotrophic factor signaling in depression and antidepressant action. Biol. Psychiatry 2021, 90, 128–136. [Google Scholar] [CrossRef] [PubMed]
  17. Duman, R.S.; Deyama, S.; Fogaça, M.V. Role of BDNF in the pathophysiology and treatment of depression: Activity-dependent effects distinguish rapid-acting antidepressants. Eur. J. Neurosci. 2021, 53, 126–139. [Google Scholar] [CrossRef]
  18. Dunham, J.; Deakin, J.; Miyajima, F.; Payton, A.; Toro, C. Expression of hippocampal brain-derived neurotrophic factor and its receptors in Stanley consortium brains. J. Psychiatr. Res. 2009, 43, 1175–1184. [Google Scholar] [CrossRef]
  19. Thompson Ray, M.; Weickert, C.S.; Wyatt, E.; Webster, M.J. Decreased BDNF, trkB-TK+ and GAD67 mRNA expression in the hippocampus of individuals with schizophrenia and mood disorders. J. Psychiatry Neurosci. 2011, 36, 195–203. [Google Scholar]
  20. Dwivedi, Y. Brain-derived neurotrophic factor: Role in depression and suicide. Neuropsychiatr. Dis. Treat. 2009, 5, 433–449. [Google Scholar] [CrossRef]
  21. Roy, B.; Shelton, R.C.; Dwivedi, Y. DNA methylation and expression of stress related genes in PBMC of MDD patients with and without serious suicidal ideation. J. Psychiatr. Res. 2017, 89, 115–124. [Google Scholar] [CrossRef] [PubMed]
  22. Dwivedi, Y.; Rizavi, H.S.; Conley, R.R.; Roberts, R.C.; Tamminga, C.A.; Pandey, G.N. Altered gene expression of brain-derived neurotrophic factor and receptor tyrosine kinase B in postmortem brain of suicide subjects. Arch. Gen. Psychiatry 2003, 60, 804–815. [Google Scholar] [CrossRef] [PubMed]
  23. Tripp, A.; Oh, H.; Guilloux, J.-P.; Martinowich, K.; Lewis, D.A.; Sibille, E. Brain-derived neurotrophic factor signaling and subgenual anterior cingulate cortex dysfunction in major depressive disorder. Am. J. Psychiatry 2012, 169, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
  24. Autry, A.E.; Monteggia, L.M. Brain-derived neurotrophic factor and neuropsychiatric disorders. Pharmacol. Rev. 2012, 64, 238–258. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, B.; Dowlatshahi, D.; MacQueen, G.M.; Wang, J.-F.; Young, L. Increased hippocampal BDNF immunoreactivity in subjects treated with antidepressant medication. Biol. Psychiatry 2001, 50, 260–265. [Google Scholar] [CrossRef]
  26. Nibuya, M.; Morinobu, S.; Duman, R.S. Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J. Neurosci. 1995, 15, 7539–7547. [Google Scholar] [CrossRef]
  27. Saarelainen, T.; Hendolin, P.; Lucas, G.; Koponen, E.; Sairanen, M.; MacDonald, E.; Agerman, K.; Haapasalo, A.; Nawa, H.; Aloyz, R.; et al. Activation of the TrkB neurotrophin receptor is induced by antidepressant drugs and is required for antidepressant-induced behavioral effects. J. Neurosci. 2003, 23, 349–357. [Google Scholar] [CrossRef]
  28. Adachi, M.; Barrot, M.; Autry, A.E.; Theobald, D.; Monteggia, L.M. Selective loss of brain-derived neurotrophic factor in the dentate gyrus attenuates antidepressant efficacy. Biol. Psychiatry 2008, 63, 642–649. [Google Scholar] [CrossRef]
  29. Li, Y.; Luikart, B.W.; Birnbaum, S.; Chen, J.; Kwon, C.-H.; Kernie, S.G.; Bassel-Duby, R.; Parada, L.F. TrkB regulates hippocampal neurogenesis and governs sensitivity to antidepressive treatment. Neuron 2008, 59, 399–412. [Google Scholar] [CrossRef]
  30. Adachi, M.; Autry, A.E.; Mahgoub, M.; Suzuki, K.; Monteggia, L.M. TrkB signaling in dorsal raphe nucleus is essential for antidepressant efficacy and normal aggression behavior. Neuropsychopharmacology 2017, 42, 886–894. [Google Scholar] [CrossRef]
  31. Casarotto, P.C.; Girych, M.; Fred, S.M.; Kovaleva, V.; Moliner, R.; Enkavi, G.; Biojone, C.; Cannarozzo, C.; Sahu, M.P.; Kaurinkoski, K.; et al. Antidepressant drugs act by directly binding to TRKB neurotrophin receptors. Cell 2021, 184, 1299–1313.e19. [Google Scholar] [CrossRef] [PubMed]
  32. Enneking, V.; Leehr, E.J.; Dannlowski, U.; Redlich, R. Brain structural effects of treatments for depression and biomarkers of response: A systematic review of neuroimaging studies. Psychol. Med. 2020, 50, 187–209. [Google Scholar] [CrossRef] [PubMed]
  33. Paolini, M.; Harrington, Y.; Colombo, F.; Bettonagli, V.; Poletti, S.; Carminati, M.; Colombo, C.; Benedetti, F.; Zanardi, R. Hippocampal and parahippocampal volume and function predict antidepressant response in patients with major depression: A multimodal neuroimaging study. J. Psychopharmacol. 2023, 02698811231190859. [Google Scholar] [CrossRef] [PubMed]
  34. Malberg, J.E.; Hen, R.; Madsen, T.M. Adult neurogenesis and antidepressant treatment: The surprise finding by Ron Duman and the field 20 years later. Biol. Psychiatry 2021, 90, 96–101. [Google Scholar] [CrossRef] [PubMed]
  35. Santarelli, L.; Saxe, M.; Gross, C.; Surget, A.; Battaglia, F.; Dulawa, S.; Weisstaub, N.; Lee, J.; Duman, R.; Arancio, O.; et al. Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science 2003, 301, 805–809. [Google Scholar] [CrossRef] [PubMed]
  36. Bathina, S.; Das, U.N. Brain-derived neurotrophic factor and its clinical implications. Arch. Med. Sci. 2015, 11, 1164–1178. [Google Scholar] [CrossRef]
  37. Erickson, K.I.; Prakash, R.S.; Voss, M.W.; Chaddock, L.; Heo, S.; McLaren, M.; Pence, B.D.; Martin, S.A.; Vieira, V.J.; Woods, J.A.; et al. Brain-derived neurotrophic factor is associated with age-related decline in hippocampal volume. J. Neurosci. 2010, 30, 5368–5375. [Google Scholar] [CrossRef]
  38. Fujii, R.; Watanabe, K.; Okamoto, N.; Natsuyama, T.; Tesen, H.; Igata, R.; Konishi, Y.; Ikenouchi, A.; Kakeda, S.; Yoshimura, R. Hippocampal volume and plasma brain-derived neurotrophic factor levels in patients with depression and healthy controls. Front. Mol. Neurosci. 2022, 15, 857293. [Google Scholar] [CrossRef]
  39. Filimonova, E.A.; Pashkov, A.A.; Moysak, G.I.; Tropynina, A.Y.; Zhanaeva, S.Y.; Shvaikovskaya, A.A.; Akopyan, A.A.; Danilenko, K.V.; Aftanas, L.I.; Tikhonova, M.A.; et al. Brain but not serum BDNF levels are associated with structural alterations in the hippocampal regions in patients with drug-resistant mesial temporal lobe epilepsy. Front. Neurosci. 2023, 17, 1217702. [Google Scholar] [CrossRef]
  40. Puhlmann, L.; Linz, R.; Valk, S.; Vrticka, P.; de Wael, R.V.; Bernasconi, A.; Bernasconi, N.; Caldairou, B.; Papassotiriou, I.; Chrousos, G.; et al. Association between hippocampal structure and serum Brain-Derived Neurotrophic Factor (BDNF) in healthy adults: A registered report. Neuroimage 2021, 236, 118011. [Google Scholar] [CrossRef]
  41. Sanchez, M.M.; Das, D.; Taylor, J.L.; Noda, A.; Yesavage, J.A.; Salehi, A. BDNF polymorphism predicts the rate of decline in skilled task performance and hippocampal volume in healthy individuals. Transl. Psychiatry 2011, 1, e51. [Google Scholar] [CrossRef] [PubMed]
  42. Harrisberger, F.; Smieskova, R.; Schmidt, A.; Lenz, C.; Walter, A.; Wittfeld, K.; Grabe, H.; Lang, U.; Fusar-Poli, P.; Borgwardt, S. BDNF Val66Met polymorphism and hippocampal volume in neuropsychiatric disorders: A systematic review and meta-analysis. Neurosci. Biobehav. Rev. 2015, 55, 107–118. [Google Scholar] [CrossRef] [PubMed]
  43. Bueller, J.A.; Aftab, M.; Sen, S.; Gomez-Hassan, D.; Burmeister, M.; Zubieta, J.-K. BDNF Val66Met allele is associated with reduced hippocampal volume in healthy subjects. Biol. Psychiatry 2006, 59, 812–815. [Google Scholar] [CrossRef] [PubMed]
  44. de Frutos Lucas, J.; Vacher, M.; Porter, T.; Brown, B.; Laws, S. Investigating neurotrophin genetics and hippocampal volume. bioRxiv 2022. preprint. [Google Scholar] [CrossRef]
  45. Tsai, S.-J.; Hong, C.-J.; Liou, Y.-J. Effects of BDNF polymorphisms on antidepressant action. Psychiatry Investig. 2010, 7, 236. [Google Scholar] [CrossRef] [PubMed]
  46. Bath, K.G.; Jing, D.Q.; Dincheva, I.; Neeb, C.C.; Pattwell, S.S.; Chao, M.V.; Lee, F.S.; Ninan, I. BDNF Val66Met impairs fluoxetine-induced enhancement of adult hippocampus plasticity. Neuropsychopharmacology 2012, 37, 1297–1304. [Google Scholar] [CrossRef]
  47. Niitsu, T.; Fabbri, C.; Bentini, F.; Serretti, A. Pharmacogenetics in major depression: A comprehensive meta-analysis. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2013, 45, 183–194. [Google Scholar] [CrossRef]
  48. Wang, Z.; Fan, J.; Gao, K.; Li, Z.; Yi, Z.; Wang, L.; Huang, J.; Yuan, C.; Hong, W.; Yu, S.; et al. Neurotrophic tyrosine kinase receptor type 2 (NTRK2) gene associated with treatment response to mood stabilizers in patients with bipolar I disorder. J. Mol. Neurosci. 2013, 50, 305–310. [Google Scholar] [CrossRef]
  49. Deflesselle, E.; Verstuyft, C.; Petit, A.-C.; Hardy, P.; David, D.J.; Falissard, B.; Fève, B.; Rigal, L.; Becquemont, L.; Corruble, E.; et al. No impact of eight NTRK2 genetic polymorphisms on 6-month antidepressant efficacy in depressed patients. Pharmacogenomics 2017, 18, 349–357. [Google Scholar] [CrossRef]
  50. 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. [Google Scholar] [CrossRef]
  51. 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. Psychiatr. Res. 2013, 47, 8–14. [Google Scholar] [CrossRef]
  52. Hennings, J.M.; Kohli, M.A.; Czamara, D.; Giese, M.; Eckert, A.; Wolf, C.; Heck, A.; Domschke, K.; Arolt, V.; Baune, B.T.; et al. Possible associations of NTRK2 polymorphisms with antidepressant treatment outcome: Findings from an extended tag SNP approach. PLoS ONE 2013, 8, e64947. [Google Scholar] [CrossRef]
  53. Brent, D.; Melhem, N.; Turecki, G. Pharmacogenomics of suicidal events. Pharmacogenomics 2010, 11, 793–807. [Google Scholar] [CrossRef]
  54. Sforzini, L.; Worrell, C.; Kose, M.; Anderson, I.M.; Aouizerate, B.; Arolt, V.; Bauer, M.; Baune, B.T.; Blier, P.; Cleare, A.J.; et al. A Delphi-method-based consensus guideline for definition of treatment-resistant depression for clinical trials. Mol. Psychiatry 2022, 27, 1286–1299. [Google Scholar] [CrossRef]
  55. Gaynes, B.N.; Lux, L.; Gartlehner, G.; Asher, G.; Forman-Hoffman, V.; Green, J.; Boland, E.; Weber, R.P.; Randolph, C.; Bann, C.; et al. Defining treatment-resistant depression. Depress. Anxiety 2020, 37, 134–145. [Google Scholar] [CrossRef]
  56. Purcell, S.; Neale, B.; Todd-Brown, K.; Thomas, L.; Ferreira, M.A.R.; Bender, D.; Maller, J.; Sklar, P.; de Bakker, P.I.W.; Daly, M.J.; et al. PLINK: A tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 2007, 81, 559–575. [Google Scholar] [CrossRef]
  57. Anderson, C.A.; Pettersson, F.H.; Clarke, G.M.; Cardon, L.R.; Morris, A.P.; Zondervan, K.T. Data quality control in genetic case-control association studies. Nat. Protoc. 2010, 5, 1564–1573. [Google Scholar] [CrossRef] [PubMed]
  58. Gaser, C.; Dahnke, R.; Thompson, P.; Kurth, F.; Luders, E. CAT-a computational anatomy toolbox for the analysis of structural MRI data. bioRxiv 2022. preprint. [Google Scholar] [CrossRef]
  59. O’Brien, L.M.; Ziegler, D.A.; Deutsch, C.K.; Frazier, J.A.; Herbert, M.R.; Locascio, J.J. Statistical adjustments for brain size in volumetric neuroimaging studies: Some practical implications in methods. Psychiatry Res. Neuroimaging 2011, 193, 113–122. [Google Scholar] [CrossRef] [PubMed]
  60. Hayes, A.F. Introduction to Mediation, Moderation, and Conditional Process Analysis: A Regression-Based Approach; Guilford Publications: New York, NY, USA, 2017. [Google Scholar]
  61. Horita, N.; Kaneko, T. Genetic model selection for a case–control study and a meta-analysis. Meta Gene 2015, 5, 1–8. [Google Scholar] [CrossRef] [PubMed]
  62. Bender, R.; Lange, S. Adjusting for multiple testing—When and how? J. Clin. Epidemiol. 2001, 54, 343–349. [Google Scholar] [CrossRef] [PubMed]
  63. Schmaal, L.; for the ENIGMA-Major Depressive Disorder Working Group; Veltman, D.J.; van Erp, T.G.M.; Sämann, P.G.; Frodl, T.; Jahanshad, N.; Loehrer, E.; Tiemeier, H.; Hofman, A.; et al. Subcortical brain alterations in major depressive disorder: Findings from the ENIGMA Major Depressive Disorder working group. Mol. Psychiatry 2016, 21, 806–812. [Google Scholar] [CrossRef] [PubMed]
  64. Numakawa, T.; Odaka, H.; Adachi, N. Actions of brain-derived neurotrophin factor in the neurogenesis and neuronal function, and its involvement in the pathophysiology of brain diseases. Int. J. Mol. Sci. 2018, 19, 3650. [Google Scholar] [CrossRef] [PubMed]
  65. Numakawa, T.; Odaka, H.; Adachi, N. Actions of brain-derived neurotrophic factor and glucocorticoid stress in neurogenesis. Int. J. Mol. Sci. 2017, 18, 2312. [Google Scholar] [CrossRef]
  66. Krugers, H.J.; Lucassen, P.J.; Karst, H.; Joëls, M. Chronic stress effects on hippocampal structure and synaptic function: Relevance for depression and normalization by anti-glucocorticoid treatment. Front. Synaptic Neurosci. 2010, 2, 24. [Google Scholar] [CrossRef]
  67. Miranda, M.; Morici, J.F.; Zanoni, M.B.; Bekinschtein, P. Brain-derived neurotrophic factor: A key molecule for memory in the healthy and the pathological brain. Front. Cell. Neurosci. 2019, 13, 363. [Google Scholar] [CrossRef]
  68. Duman, R.S.; Monteggia, L.M. A neurotrophic model for stress-related mood disorders. Biol. Psychiatry 2006, 59, 1116–1127. [Google Scholar] [CrossRef]
  69. Beurel, E.; Toups, M.; Nemeroff, C.B. The bidirectional relationship of depression and inflammation: Double trouble. Neuron 2020, 107, 234–256. [Google Scholar] [CrossRef]
  70. Berk, M.; Williams, L.J.; Jacka, F.N.; O’Neil, A.; Pasco, J.A.; Moylan, S.; Allen, N.B.; Stuart, A.L.; Hayley, A.C.; Byrne, M.L.; et al. So depression is an inflammatory disease, but where does the inflammation come from? BMC Med. 2013, 11, 200. [Google Scholar] [CrossRef]
  71. Kiecolt-Glaser, J.K.; Derry, H.M.; Fagundes, C.P. Inflammation: Depression fans the flames and feasts on the heat. Am. J. Psychiatry 2015, 172, 1075–1091. [Google Scholar] [CrossRef]
  72. Tannous, J.; Godlewska, B.R.; Tirumalaraju, V.; Soares, J.C.; Cowen, P.J.; Selvaraj, S. Stress, inflammation and hippocampal subfields in depression: A 7 Tesla MRI Study. Transl. Psychiatry 2020, 10, 78. [Google Scholar] [CrossRef] [PubMed]
  73. Paolini, M.; Harrington, Y.; Raffaelli, L.; Poletti, S.; Zanardi, R.; Colombo, C.; Benedetti, F. Neutrophil to lymphocyte ratio and antidepressant treatment response in patients with major depressive disorder: Effect of sex and hippocampal volume. Eur. Neuropsychopharmacol. 2023, 76, 52–60. [Google Scholar] [CrossRef] [PubMed]
  74. Arteaga-Henríquez, G.; Simon, M.S.; Burger, B.; Weidinger, E.; Wijkhuijs, A.; Arolt, V.; Birkenhager, T.K.; Musil, R.; Müller, N.; Drexhage, H.A. Low-grade inflammation as a predictor of antidepressant and anti-inflammatory therapy response in MDD patients: A systematic review of the literature in combination with an analysis of experimental data collected in the EU-MOODINFLAME consortium. Front. Psychiatry 2019, 10, 458. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, J.; Wei, Y.; Kenji, H. Brain-derived neurotrophic factor (BDNF)-TrkB signaling in inflammation-related depression and potential therapeutic targets. Curr. Neuropharmacol. 2016, 14, 721–731. [Google Scholar] [CrossRef]
  76. Colucci-D’Amato, L.; Speranza, L.; Volpicelli, F. Neurotrophic factor BDNF, physiological functions and therapeutic potential in depression, neurodegeneration and brain cancer. Int. J. Mol. Sci. 2020, 21, 7777. [Google Scholar] [CrossRef]
  77. Rantamäki, T. TrkB neurotrophin receptor at the core of antidepressant effects, but how? Cell Tissue Res. 2019, 377, 115–124. [Google Scholar] [CrossRef]
  78. Sonoyama, T.; Stadler, L.K.J.; Zhu, M.; Keogh, J.M.; Henning, E.; Hisama, F.; Kirwan, P.; Jura, M.; Blaszczyk, B.K.; DeWitt, D.C.; et al. Human BDNF/TrkB variants impair hippocampal synaptogenesis and associate with neurobehavioural abnormalities. Sci. Rep. 2020, 10, 9028. [Google Scholar] [CrossRef]
  79. Casarotto, P.; Umemori, J.; Castrén, E. BDNF receptor TrkB as the mediator of the antidepressant drug action. Front. Mol. Neurosci. 2022, 15, 1032224. [Google Scholar] [CrossRef]
  80. Gourion, D. Antidepressants and their onset of action: A major clinical, methodological and pronostical issue. L’encephale 2008, 34, 73–81. [Google Scholar] [CrossRef]
  81. Psomiades, M.; Mondino, M.; Galvão, F.; Mandairon, N.; Nourredine, M.; Suaud-Chagny, M.-F.; Brunelin, J. Serum mature BDNF level is associated with remission following ECT in treatment-resistant depression. Brain Sci. 2022, 12, 126. [Google Scholar] [CrossRef]
  82. Zhang, F.; Lupski, J.R. Non-coding genetic variants in human disease. Hum. Mol. Genet. 2015, 24, R102–R110. [Google Scholar] [CrossRef] [PubMed]
  83. Crist, R.C.; Clarke, T.-K.; Berrettini, W.H. Pharmacogenetics of opioid use disorder treatment. CNS Drugs 2018, 32, 305–320. [Google Scholar] [CrossRef]
  84. Watanabe, K.; Stringer, S.; Frei, O.; Mirkov, M.U.; de Leeuw, C.; Polderman, T.J.C.; van der Sluis, S.; Andreassen, O.A.; Neale, B.M.; Posthuma, D. A global overview of pleiotropy and genetic architecture in complex traits. Nat. Genet. 2019, 51, 1339–1348. [Google Scholar] [CrossRef] [PubMed]
  85. Nagel, M.; Jansen, P.R.; Stringer, S.; Watanabe, K.; de Leeuw, C.A.; Bryois, J.; Savage, J.E.; Hammerschlag, A.R.; Skene, N.G.; Muñoz-Manchado, A.B.; et al. Meta-analysis of genome-wide association studies for neuroticism in 449,484 individuals identifies novel genetic loci and pathways. Nat. Genet. 2018, 50, 920–927. [Google Scholar] [CrossRef] [PubMed]
  86. Gupta, V.K.; You, Y.; Gupta, V.B.; Klistorner, A.; Graham, S.L. TrkB receptor signalling: Implications in neurodegenerative, psychiatric and proliferative disorders. Int. J. Mol. Sci. 2013, 14, 10122–10142. [Google Scholar] [CrossRef]
  87. Barbacid, M. The Trk family of neurotrophin receptors. J. Neurobiol. 1994, 25, 1386–1403. [Google Scholar] [CrossRef]
  88. Hu, Z.; Yu, X.; Chen, P.; Jin, K.; Zhou, J.; Wang, G.; Yu, J.; Wu, T.; Wang, Y.; Lin, F.; et al. BDNF-TrkB signaling pathway-mediated microglial activation induces neuronal KCC2 downregulation contributing to dynamic allodynia following spared nerve injury. Mol. Pain 2023, 19, 17448069231185439. [Google Scholar] [CrossRef]
  89. Fernández-García, S.; Sancho-Balsells, A.; Longueville, S.; Hervé, D.; Gruart, A.; Delgado-García, J.M.; Alberch, J.; Giralt, A. Astrocytic BDNF and TrkB regulate severity and neuronal activity in mouse models of temporal lobe epilepsy. Cell Death Dis. 2020, 11, 411. [Google Scholar] [CrossRef]
  90. Fenner, B.M. Truncated TrkB: Beyond a dominant negative receptor. Cytokine Growth Factor Rev. 2012, 23, 15–24. [Google Scholar] [CrossRef]
  91. Andreska, T.; Lüningschrör, P.; Sendtner, M. Regulation of TrkB cell surface expression—A mechanism for modulation of neuronal responsiveness to brain-derived neurotrophic factor. Cell Tissue Res. 2020, 382, 5–14. [Google Scholar] [CrossRef]
  92. Kazmi, I.; Karmakar, S.; Shaharyar, M.A.; Afzal, M.; Al-Abbasi, F.A. How Synthetic Drugs Work: Insights into Molecular Pharmacology of Classic and New Pharmaceuticals; Academic Press: Cambridge, MA, USA, 2022. [Google Scholar]
  93. Andrade, C.; Rao, N.S.K. How antidepressant drugs act: A primer on neuroplasticity as the eventual mediator of antidepressant efficacy. Indian J. Psychiatry 2010, 52, 378. [Google Scholar] [CrossRef] [PubMed]
  94. Zhou, C.; Zhong, J.; Zou, B.; Fang, L.; Chen, J.; Deng, X.; Zhang, L.; Zhao, X.; Qu, Z.; Lei, Y.; et al. Meta-analyses of comparative efficacy of antidepressant medications on peripheral BDNF concentration in patients with depression. PLoS ONE 2017, 12, e0172270. [Google Scholar] [CrossRef] [PubMed]
  95. Mindt, S.; Neumaier, M.; Hellweg, R.; Sartorius, A.; Kranaster, L. Brain-derived neurotrophic factor in the cerebrospinal fluid increases during electroconvulsive therapy in patients with depression: A preliminary report. J. ECT 2020, 36, 193–197. [Google Scholar] [CrossRef]
  96. Mannari, C.; Origlia, N.; Scatena, A.; Del Debbio, A.; Catena, M.; Dell’agnello, G.; Barraco, A.; Giovannini, L.; Dell’osso, L.; Domenici, L.; et al. BDNF level in the rat prefrontal cortex increases following chronic but not acute treatment with duloxetine, a dual acting inhibitor of noradrenaline and serotonin re-uptake. Cell. Mol. Neurobiol. 2008, 28, 457–468. [Google Scholar] [CrossRef] [PubMed]
  97. Yang, T.; Nie, Z.; Shu, H.; Kuang, Y.; Chen, X.; Cheng, J.; Yu, S.; Liu, H. The role of BDNF on neural plasticity in depression. Front. Cell. Neurosci. 2020, 14, 82. [Google Scholar] [CrossRef] [PubMed]
  98. Björkholm, C.; Monteggia, L.M. BDNF–a key transducer of antidepressant effects. Neuropharmacology 2016, 102, 72–79. [Google Scholar] [CrossRef]
  99. Cubillos, S.; Engmann, O.; Brancato, A. BDNF as a mediator of antidepressant response: Recent advances and lifestyle interactions. Int. J. Mol. Sci. 2022, 23, 14445. [Google Scholar] [CrossRef]
  100. Arnone, D.; McKie, S.; Elliott, R.; Juhasz, G.; Thomas, E.J.; Downey, D.; Williams, S.; Deakin, J.F.W.; Anderson, I.M. State-dependent changes in hippocampal grey matter in depression. Mol. Psychiatry 2013, 18, 1265–1272. [Google Scholar] [CrossRef]
  101. Schermuly, I.; Wolf, D.; Lieb, K.; Stoeter, P.; Fellgiebel, A. State dependent posterior hippocampal volume increases in patients with major depressive disorder. J. Affect. Disord. 2011, 135, 405–409. [Google Scholar] [CrossRef]
  102. Li, Q.S.; Tian, C.; Hinds, D. Genome-wide association studies of antidepressant class response and treatment-resistant depression. Transl. Psychiatry 2020, 10, 360. [Google Scholar] [CrossRef]
  103. abbri, C.; Kasper, S.; Kautzky, A.; Bartova, L.; Dold, M.; Zohar, J.; Souery, D.; Montgomery, S.; Albani, D.; Raimondi, I.; et al. Genome-wide association study of treatment-resistance in depression and meta-analysis of three independent samples. Br. J. Psychiatry 2019, 214, 36–41. [Google Scholar]
  104. Horgusluoglu-Moloch, E.; Risacher, S.L.; Crane, P.K.; Hibar, D.; Thompson, P.M.; Saykin, A.J.; Nho, K. Genome-wide association analysis of hippocampal volume identifies enrichment of neurogenesis-related pathways. Sci. Rep. 2019, 9, 14498. [Google Scholar] [CrossRef] [PubMed]
  105. Hibar, D.P.; Adams, H.H.H.; Jahanshad, N.; Chauhan, G.; Stein, J.L.; Hofer, E.; Renteria, M.E.; Bis, J.C.; Arias-Vasquez, A.; Ikram, M.K.; et al. Novel genetic loci associated with hippocampal volume. Nat. Commun. 2017, 8, 13624. [Google Scholar] [CrossRef]
  106. van der Meer, D.; Rokicki, J.; Kaufmann, T.; Córdova-Palomera, A.; Moberget, T.; Alnæs, D.; Bettella, F.; Frei, O.; Doan, N.T.; Sønderby, I.E.; et al. Brain scans from 21,297 individuals reveal the genetic architecture of hippocampal subfield volumes. Mol. Psychiatry 2020, 25, 3053–3065. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Indirect effect of NTRK2 polymorphisms on treatment resistance status through right hippocampal volume. AD: antidepressant; BCa CI: Bias-corrected and accelerated bootstrap confidence interval. * log-odds metric.
Figure 1. Indirect effect of NTRK2 polymorphisms on treatment resistance status through right hippocampal volume. AD: antidepressant; BCa CI: Bias-corrected and accelerated bootstrap confidence interval. * log-odds metric.
Genes 14 02037 g001
Table 1. Clinical and demographic characteristics of the sample.
Table 1. Clinical and demographic characteristics of the sample.
Whole Sample (n = 121)No Treatment Resistance (n = 60)Treatment Resistance (n = 61)χ2/t-Test p
Age50.08 ± 10.2650.53 ± 9.0249.64 ± 11.400.633
Sex (F/M)77/4441/1936/250.287
Education (yrs.)12.33 ± 3.8512.06 ± 3.3312.60 ± 4.320.450
Age of onset32.61 ± 12.4834.02 ± 12.0131.15 ± 12.890.214
Baseline HDRS22.02 ± 6.4122.27 ± 6.3221.80 ± 6.560.694
Duration of hospitalization (days)27.74 ± 10.4024.85 ± 7.9530.79 ± 11.800.002
BMI25.13 ± 4.3324.77 ± 4.4625.48 ± 4.200.376
R. Hippocampus as % of TIV0.221 ± 0.0170.227 ± 0.0170.216 ± 0.017<0.001
L. Hippocampus as % of TIV0.206 ± 0.0170.211 ± 0.0150.200 ± 0.016<0.001
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

Paolini, M.; Fortaner-Uyà, L.; Lorenzi, C.; Spadini, S.; Maccario, M.; Zanardi, R.; Colombo, C.; Poletti, S.; Benedetti, F. Association between NTRK2 Polymorphisms, Hippocampal Volumes and Treatment Resistance in Major Depressive Disorder. Genes 2023, 14, 2037. https://doi.org/10.3390/genes14112037

AMA Style

Paolini M, Fortaner-Uyà L, Lorenzi C, Spadini S, Maccario M, Zanardi R, Colombo C, Poletti S, Benedetti F. Association between NTRK2 Polymorphisms, Hippocampal Volumes and Treatment Resistance in Major Depressive Disorder. Genes. 2023; 14(11):2037. https://doi.org/10.3390/genes14112037

Chicago/Turabian Style

Paolini, Marco, Lidia Fortaner-Uyà, Cristina Lorenzi, Sara Spadini, Melania Maccario, Raffaella Zanardi, Cristina Colombo, Sara Poletti, and Francesco Benedetti. 2023. "Association between NTRK2 Polymorphisms, Hippocampal Volumes and Treatment Resistance in Major Depressive Disorder" Genes 14, no. 11: 2037. https://doi.org/10.3390/genes14112037

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