Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Apr 29.
Published in final edited form as: CNS Drugs. 2020 Apr;34(4):389–414. doi: 10.1007/s40263-020-00702-y

Genetic Influence on Efficacy of Pharmacotherapy for Pediatric Attention-Deficit/Hyperactivity Disorder: Overview and Current Status of Research

Nada A Elsayed 1,2, Kaila M Yamamoto 1, Tanya E Froehlich 1,3,*
PMCID: PMC8083895  NIHMSID: NIHMS1689953  PMID: 32133580

Abstract

Multiple stimulant and non-stimulant medications are approved for the treatment of Attention-Deficit/Hyperactivity Disorder (ADHD), one of the most prevalent childhood neurodevelopmental disorders. Choosing among the available agents and determining the most effective ADHD medication for a given child can be a time-consuming process due to the high inter-individual variability in treatment efficacy. As a result, there is growing interest in identifying predictors of ADHD medication response in children through the burgeoning field of pharmacogenomics. This article reviews childhood ADHD pharmacogenomic efficacy studies published during the last decade (2009–2019), which have largely focused on pharmacodynamic candidate gene investigations of methylphenidate and atomoxetine response, with a lesser number investigating pharmacokinetic candidate genes and genome-wide approaches. Findings from studies which have advanced the field of ADHD pharmacogenomics through investigation of meta-analytic approaches and gene*gene interactions are also overviewed. Despite this progress, no one genetic variant or currently available pharmacogenomics test has demonstrated clinical utility in pinpointing the optimal ADHD medication for a given individual patient, highlighting the need for further investigation.

1. Introduction

Attention-Deficit/Hyperactivity Disorder (ADHD)--a common childhood neurodevelopmental disorder affecting 3–7% of children worldwide[1, 2]--is associated with marked impairment in academic, social, and family functioning.[3, 4] Medications for ADHD that effectively ameliorate symptoms of inattention, hyperactivity, and impulsivity are available,[5] with stimulant medications (methylphenidate and amphetamine/dextroamphetamine) considered to be first-line treatment and non-stimulants (atomoxetine and alpha-2-agonists) considered to be second-line.[3] Although data showing that ADHD medications can enhance long-term outcomes are sparse, evidence is emerging that treatment may improve academic outcomes[6, 7] and decrease risk for motor vehicle accidents,[8, 9] injuries,[10] depression,[11] and criminality.[12]

Despite the above evidence of ADHD medication efficacy and improved outcomes in groups of individuals with ADHD, finding an effective medication and dosage for a given child with ADHD can be a complex process. Currently, no phenotypic or patient factors have been shown to consistently predict ADHD medication response.[13] As a result, pediatric clinical practice guidelines continue to recommend empirical determination of each child’s ADHD treatment regimen through a process of gradual upward dose titration and a trial-and-error approach to trying different preparations.[3] This trial-and-error method often yields marked delays in achieving optimal symptom remission, as well as premature termination of treatment due to provider or family frustration with the titration process.[14]

In other branches of medicine, progress in predicting medication response has been made through pharmacogenomics (PGx), which utilizes knowledge of an individual’s genes to predict medication efficacy, dosing, and adverse effects. In fact, evidence has accrued to the point where the Clinical Pharmacogenetics Implementation Consortium (CPIC) has established universal guidelines for genotype-based drug prescription[15] for more than 20 drug-gene pairs.[16] Clinical implementation of these guidelines began as early as 2011 with the launch of the Pharmacogenomics Research Network Translational Pharmacogenomics Program at 8 different medical centers, including the University of Maryland. There, an anti-platelet pharmacogenetics initiative has shown much success in prescribing heart catherization patients anti-platelet drugs based on CYP2C19 genotype.[17] Unfortunately, ADHD PGx has not fully realized its promise to facilitate more expeditious identification of the optimal medication regimen for each child with ADHD. Nonetheless, numerous pediatric ADHD pharmacogenomics studies have been conducted in recent years and warrant further examination to guide future research and clinical care efforts. Hence, this article provides an overview of pediatric ADHD PGx studies published from 2009–2019 to provide an update to our prior review.[18]

This review is limited to ADHD PGx studies enrolling participants who were <18 years old which used 1) standardized measures to assess improvements in symptoms or global functioning with medication treatment (e.g., ADHD rating scales or Clinical Global Impression [CGI] ratings) and/or 2) direct observations of inattention or activity level. ADHD PGx investigations of academic or neuropsychological outcomes are beyond the scope of this review. A comprehensive literature search was conducted using the PubMed (http://www.ncbi.nlm.nih.gov/sites) and PsycINFO (http://www.apa.org/psycinfo/) databases to identify relevant English-language articles. Search terms included the following: ‘pharmacogen*’ AND ‘ADHD’ OR ‘attention-deficit’ OR ‘attention’ OR ‘hyperactiv*’ OR ‘methylphenidate’ OR ‘amphetamine’ OR ‘atomoxetine’ OR ‘stimulant’ OR ‘psychostimulant’ OR ‘clonidine’ OR ‘guanfacine’ or ‘α-agonist’. References of identified papers were also screened to pinpoint additional relevant literature.

Due to the sparsity of pharmacogenomics investigations evaluating response to amphetamines, guanfacine, and clonidine, this review focuses primarily on methylphenidate (MPH) and atomoxetine (ATX) studies. The MPH and ATX sections each open with a review of findings for PGx studies of the gene encoding their respective metabolizing enzymes. A summary of findings for pharmacodynamic candidate genes that have been significantly associated with MPH or ATX response in at least one prior study follows in the text as well as in Tables 1–6, with additional information about each study’s methods (medication trial duration, dosage used, and primary outcome measures) summarized in Supplemental Tables 16. Minor allele frequencies for relevant polymorphisms (extracted from the Ensembl Genome Browser[19]) are also presented in the text. The section for each genetic variant closes with an integration paragraph in cases where this polymorphism has been the focus of a meta-analysis or significant prior literature. Additional genetic variants that did not show significant effects in MPH or ATX PGx studies from 2009–2019 are included in Tables 1–6 (with further study information available in Supplemental Tables 16) rather than being discussed in the text.

Table 1.

Pharmacokinetic Candidate Gene Studies of Methylphenidate Efficacy Published from 2009–2019

Gene Polymorphism(s) Authors Design / Sample Findings
Carboxylesterase 1 (CES1)
  • Gly143Glu

Nemoda et al 2009 [23]
  • POL

  • N=122

  • Hungary

  • No effect.

  • rs3815589

  • rs2287194

  • rs2244613

  • rs2002577

  • rs2307244

  • rs2307240

  • rs12443580

Johnson et al 2013 [27]
  • POL

  • N=77

  • Ireland

  • No effects.

POL=prospective, open-label

2. Methylphenidate Pharmacogenomic Studies

MPH’s main mechanism of action involves increase in synaptic dopamine and norepinephrine via blockade of the dopamine transporter and norepinephrine transporter.[20] Additional MPH actions which lead to an increase in extracellular dopamine and norepinephrine include its agonist activity at the serotonin 1A receptor and its redistribution of vesicular monoamine transporter 2, which regulates release of monoamines from vesicular storage.[21]

2.1. Methylphenidate Pharmacokinetic Genetic Variants

2.1.1. Carboxylesterase 1 (CES1) [Table 1].

The hepatic enzyme CES1 is the main metabolizer of MPH, de-esterifying it to form inactive ritalinic acid.[22]

Prior to 2009, little was known regarding the impact of CES1 variants on MPH efficacy. Since then, two prospective, open-label studies have investigated the link between CES1 polymorphisms and MPH efficacy. Nemoda et al.[23] explored the association between the CES1 Gly143Glu (rs71647871, minor allele frequency [MAF]=0.01–0.04 [T/glutamate]) functional polymorphism and response to MPH. This polymorphism is of particular interest since the glutamate substitution has been associated with a lower level of CES1 enzyme activity and altered MPH metabolism in several pharmacokinetics studies.[2426] Nemoda et al.’s study, which included five glutamate allele carriers, did not find a link between the Gly143Glu polymorphism and MPH responder status or reduction in hyperactivity/impulsivity or inattention scores. However, they did observe that MPH responders possessing the 143Glu allele required a lower MPH dose for symptom reduction compared to responders who were Gly/Gly homozygotes. The second study, by Johnson et al.[27], evaluated seven CES1 polymorphisms (but not Gly143Glu) and found no association between any of these single nucleotide polymorphisms (SNPs) and MPH dose response or efficacy.[27]

2.2. Methylphenidate Pharmacodynamic Genetic Variants

2.2.1. Adrenergic α2A –Receptor (ADRA2A) [Table 2].

Table 2.

Neurotransmitter Receptor Candidate Gene Studies of Methylphenidate Efficacy Published from 2009–2019

Gene Polymorphism(s) Authors Design / Sample Findings
Adrenergic α2A –Receptor (ADRA2A)
  • MspI

Cheon et al 2009 [38]
  • POL

  • N=114

  • South Korea

  • Improved MPH response with MspI G allele.

  • MspI

Froehlich et al 2011 [33]
  • DBPC

  • N=89

  • U.S.A.

  • No effect after multiple comparison correction.

  • MspI

  • DraI

Kim et al 2011 [34]
Hong et al 2012 [35]
Park et al 2013 [36]
  • POL

  • N=102 for Kim et al

  • N=103 for Hong et al

  • N=115 for Park et al

  • South Korea

Kim et al, Hong et al, and Park et al:
  • No main effects in whole sample analyses.

Park et al:
  • Improved MPH response with DraI C/C homozygosity in analyses limited to combined type participants.

Hong et al:
  • Interaction effects between the DraI SNP and the DRD4 VNTR, NET G1287A and NET −3081(A/T) polymorphisms.

  • MspI

Unal et al 2016 [40]
  • POL

  • N=108

  • Turkey

  • Decreased MPH response with MspI GG genotype.

  • MspI

  • DraI

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effects.

  • MspI

Huang et al 2018 [39]
  • POL

  • N=59

  • Taiwan

  • Improved MPH response with MspI GG genotype.

Dopamine Receptor D1 (DRD1)
  • rs4867798

  • rs251937

  • rs11749676

  • rs835540

  • rs835616

  • rs835541

  • rs863126

  • rs265977

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effects.

Dopamine Receptor D2 (DRD2)
  • rs1800497

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • Faster MPH response with C allele.

  • rs4630328

  • rs7131056

  • rs4245146

  • rs17529477

  • rs2002453

  • rs12363125

  • rs2283265

  • rs2242592

  • rs1554929

  • rs2234689

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effect on MPH efficacy for any variant.

  • Trend for rs2283265 T allele carriers to require a higher MPH dose.

Dopamine Receptor D3 (DRD3)
  • rs6280

  • rs9825563

  • rs1800828

  • rs10934256

  • rs167770

  • rs167771

  • rs324035

  • rs9880168

  • rs2134655

  • rs3732790

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • Decreased MPH response nominally associated with rs2134655 G allele.

  • Decreased MPH response with GG haplotype at rs2134655-rs1800828 in whole sample analyses, with effects accentuated in children with prenatal smoke exposure.

  • No effects for other variants.

  • rs6280

Fageera et al 2018 [62]
  • DBPC

  • N=327 boys

  • Canada

  • Improved MPH response with C/C homozygosity according to teacher but not parent ratings.

Dopamine Receptor D4 (DRD4)
  • Exon 3 VNTR

  • 120 BP promoter duplication

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • No effects.

  • Exon 3 VNTR

Froehlich et al 2011 [33]
  • DBPC

  • N=89

  • U.S.A.

  • Decreased response on hyperactive-impulsive symptoms across MPH doses with absence of 4-repeat (short) allele.

  • Exon 3 VNTR

Hong et al 2012 [35]
  • POL

  • N=103

  • South Korea

  • No independent effect.

  • Interaction effect with the ADRA2A DraI SNP.

  • Exon 3 VNTR

Ji et al 2013 [67]
  • POL

  • N=114

  • South Korea

  • No effect.

  • Exon 3 VNTR

  • 120 BP promoter duplication

  • rs3758653

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • Decreased MPH response with homozygous short allele promoter duplication.

  • No effects for other variants.

  • Exon 3 VNTR

Naumova et al 2017 [66]
  • DBPC

  • N=374

  • Canada

  • Improved MPH response with 7-repeat (long) allele homozygosity based on parent but not teacher ratings.

  • rs936465

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effect.

Dopamine Receptor D5 (DRD5)
  • rs10033951

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effect.

5-Hydroxytryptamine (Serotonin) Receptor 2A (HTR2A)
  • rs7322347

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

5-Hydroxytryptamine (Serotonin) Receptor 2C (HTR2C)
  • rs6318

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Metabotropic Glutamate Receptor Subtype 7 (mGluR7/GRM7)
  • rs3792452

Park et al 2014b [82]
  • POL

  • N=175

  • South Korea

  • Improved MPH response with G/A genotype.

  • rs3792452

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

N-methyl-D-aspartate (NMDA) Receptor Subunit 2A (GRIN2A)
  • rs2229193

Kim et al 2017 [101]
  • POL

  • N=75

  • South Korea

  • No effect.

N-methyl-D-aspartate (NMDA) Receptor Subunit 2B (GRIN2B).
  • rs2284411

Kim et al 2017 [101]
  • POL

  • N=75

  • South Korea

  • Improved MPH response with C/C genotype.

BP=base pair

DBPC=double-blind, placebo-controlled

POL=prospective, open-label

SNP=single nucleotide polymorphism

VNTR=Variable nucleotide tandem repeat

ADRA2A encodes a norepinephrine autoreceptor whose activation limits norepinephrine release.[28] Prior animal studies suggest ADRA2A may mediate MPH effects, since α2-adrenoceptor antagonists have been shown to block MPH’s beneficial effects.[29, 30]

MspI (−1291 C>G; rs1800544)

Since 2009, one placebo-controlled and five open-label MPH PGx cohorts have assessed ADRA2A’s −1291 C>G SNP (rs1800544, MAF=0.46 [C]), which results in an MspI restriction site within a putative transcription factor binding site [31] in the gene promoter region.[32] The randomized controlled trial (RCT) reported a main effect such that G allele homozygosity was linked to higher hyperactive-impulsive scores on placebo and across MPH doses, but MspI genotype did not influence MPH dose response.[33] Among the open-label cohorts, two found no effects[3437], two linked the G allele with improved MPH response[38, 39] and one [40] linked the G allele to diminished MPH response.

Meta-analysis Findings.

Although the findings described above do not depict a clear impact of the MspI variant on MPH efficacy, a recent meta-analysis by Myer et al.[41] integrated data from four studies, including two described above[33, 42], and identified an association between the MspI G allele and improved MPH response (Odds Ratio [OR]=1.69, 95% Confidence Interval (CI):1.12–2.55, p=0.01; I2 =80%, indicating considerable heterogeneity in individual study results).

DraI (rs553668)

An A to G substitution in the ADRA2A 3’ Untranslated Region (UTR) (rs553668, MAF=0.33 [A]) creates a DraI site that has also been the subject of ADHD PGx investigation. From 2009–2019, this polymorphism was assessed in two cohorts, and none identified significant independent associations with MPH response in whole sample analyses.[3437] However, when Park et al. excluded inattentive type participants and limited analysis to children with ADHD-combined type, they found a link between C/C homozygosity and improved MPH response.[36] Significant pairwise interactions of the ADRA2A DraI SNP with the DRD4 exon 3 VNTR, the DraI SNP with the norepinephrine transporter (NET) G1287A SNP, and the DraI SNP with NET −3081(A/T) SNP on MPH treatment response were also reported within the same cohort.[35]

2.2.2. Brain-derived Neurotrophic Factor (BDNF) [Table 3].

Table 3.

Neuronal/Synaptic Plasticity and Synaptic Effector Candidate Gene Studies of Methylphenidate Efficacy Published from 2009–2018

Gene Polymorphism(s)
Authors Design / Sample Findings
Brain-derived Neurotrophic Factor (BDNF)
  • Val66Met

Kim et al 2011 [34]
  • POL

  • N=102

  • South Korea

  • Improved MPH response with Val/Val homozygosity.

Cadherin 13 (CDH13)
  • rs6565113

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Fatty Acid Desaturase 2 (FADS2)
  • rs498793

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Glucose-fructose Oxidoreductase Domain Containing 1 (GFOD1)
  • rs552655

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Latrophilin 3 / Adhesion G Protein-coupled Receptor L3 (LPHN3/ADGRL3) Arcos-Burgos et al:
  • rs6551665

  • rs6551665, rs1947275, rs9683662 haplotype

Jain et al:
  • rs6551665

  • 11q haplotype block1

Arcos-Burgos et al 2010 [75]
Jain et al 2012 [80]
  • POL

  • N=240 for Arcos-Burgos et al

  • N=82 for Jain et al

  • U.S.A.

Arcos-Burgos et al:
  • Improved MPH response on inattentive symptoms with rs6551665 G allele in both single marker and rs6551665, rs1947275, rs9683662 haplotype analysis.

Jain et al:
  • Improved stimulant response linked to having rs6551665 G/G homozygosity and two copies of the 11q ADHD susceptibility haplotype.

  • rs2122643

  • rs1868790

  • rs6551665

  • rs1947274

  • rs6858066

  • rs2345039

Choudhry et al 2012 [79]
  • DBPC

  • N=380

  • Canada

  • Improved MPH response with rs6551665 G allele.

  • Improved MPH response with rs1868790 T allele.

  • Improved MPH response for the rs6551665, rs1947274 and rs6858066 GCA haplotype in children whose mothers did not experience prenatal stress.

  • No effects for other variants.

  • rs1868790

  • rs1947274

  • rs2122643

  • rs2345039

  • rs6551665

  • rs6858066

Labbe et al 2012 [78]
  • DBPC

  • N=416

  • Canada

  • Decreased MPH response linked to rs1947274 C allele, rs2345039 G allele, rs6551665 G allele, and rs6858066 A allele.

  • No effects for other variants.

  • rs1947274

  • rs2345039

  • rs6551665

  • rs6551665 and rs1947274 haplotype

Song et al 2014 [76]
  • POL

  • N=139

  • South Korea

  • No independent or haplotype effects.

  • rs6813183, rs1355368, rs734644 haplotype

  • rs6551665 and rs1947275 haplotype

Bruxel et al 2015 [77]
  • POL

  • N=172

  • Brazil

  • Faster MPH treatment response for rs6813183, rs1355368, and rs734644 haplotype CGC homozygotes.

  • No effects for the rs6551665 and rs1947275 GT haplotype after multiple comparison correction.

  • rs1397548

  • rs1868790

  • rs2305339

  • rs6551655

  • rs6813183

  • rs6858066

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • Decreased MPH response for rs1868790 AA homozygotes.

  • No effects for other variants.

Neurotrophin-3 (NTF-3)
  • rs6332

  • rs6489630

Kim et al 2011 [34]
  • POL

  • N=102

  • South Korea

  • No effects.

Steroid Sulfatase (STS)
  • rs12861247

  • rs17268988

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effects.

1

spans all or part of the genes encoding neural cell adhesion molecule 1 (NCAM1), encompasses the tetratricopeptide repeat domain 12 (TTC12) and the ankyrin repeat and kinase domain containing 1 (ANKK1), and is adjacent to the 5’ untranslated region of the dopamine receptor D2 gene (DRD2)

DBPC=double-blind, placebo-controlled

POL=prospective, open-label

BDNF plays an important role in neuronal development and maturation, particularly for dopaminergic neurons.[43]

We identified one pediatric study which evaluated the link between a BDNF polymorphism and MPH efficacy. Kim et al.’s open-label study investigated the BDNF Val66Met (rs6265, MAF = 0.20 [T]) polymorphism and found that valine/valine (Val/Val) homozygotes had a better MPH response (in terms of both symptom improvement and a response definition which required both improvement in symptoms and CGI ratings) than methionine (Met) carriers.[34] Several lines of evidence suggest the Met allele may be linked to adverse changes in brain anatomy and function. For example, decreased prefrontal cortex, temporal lobe, and occipital lobe volumes have been documented for Met carriers compared to Val/Val homozygotes.[44, 45] An additional study showed that the neurons of Val/Val homozygotes have increased activity-dependent BDNF secretion compared to Met/Met homozygotes.[46]

2.2.3. Catechol-O-methyltransferase (COMT) [Table 4].

Table 4.

Neurotransmitter Synthesis and Degradation Candidate Gene Studies of Methylphenidate Efficacy Published from 2009–2019

Gene Polymorphism(s) Authors Design / Sample Findings
Catechol-O-methyltransferase (COMT)
  • Val158Met

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • No effect.

  • Val158Met

Froehlich et al 2011 [33]
  • DBPC

  • N=89

  • U.S.A.

  • Trend toward greater improvement on hyperactive-impulsive symptoms with increasing MPH dose for Val/Val homozygotes.

  • Val158Met

Salatino-Oliveira et al 2011 [53]
  • POL

  • N=251

  • Brazil

  • No independent effect on ADHD symptoms.

  • Met allele was linked to improved MPH effects on oppositional symptoms over time.

  • Val158Met

Yatsuga et al 2014 [52]
  • POL

  • N=50 boys

  • Japan

  • No effect.

  • Val158Met

Park et al 2014a [50]
  • POL

  • N=120

  • South Korea

  • Improved MPH response on hyperactive-impulsive symptoms with Val/Val homozygosity (not significant after multiple comparison correction).

  • Val158Met

Unal et al 2016 [40]
  • POL

  • N=108

  • Turkey

  • No effect.

  • Val158Met

  • rs4818

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effects.

  • Val158Met

  • rs2020917

  • rs933271

  • rs1544325

  • rs740603

  • rs740601

  • rs4646316

  • rs165774

  • rs9332377

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effects.

DOPA Decarboxylase (DDC)
  • rs6592961

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Dopamine Beta Hydroxylase (DBH)
  • rs2007153

  • rs2797851

  • rs1548364

  • rs2797855

  • rs1541332

  • rs2519154

  • rs2797853

  • rs6479643

  • rs77905

  • rs2073833

  • rs1611131

  • rs129883

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • MPH non-response nominally associated with rs2073833 C allele.

  • Decreased MPH response with TC genotype at rs1541332–rs2073833 in whole sample analyses, with effects accentuated in children with prenatal smoke exposure.

Tyrosine Hydroxylase (TH)
  • rs10770140

  • rs6356

  • rs2070762

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effect for rs10770140 or rs6356.

  • Decreased MPH response with rs2070762 C/C homozygosity, with effects accentuated in children with prenatal smoke exposure.

DBPC=double-blind, placebo-controlled

POL=prospective, open-label

COMT degrades dopamine and norepinephrine in the synaptic cleft and plays an important role in the catabolism of dopamine in the prefrontal cortex due to the relative paucity of dopamine transporters in this region.[47]

A SNP at COMT’s codon 158 (Val158Met, rs4680, MAF=0.37 [A]) results in a substitution of methionine for valine which changes enzyme activity levels: Met/Met homozygotes exhibit four times lower enzymatic activity than those with wild type Val/Val homozygosity.[48] From 2009 onward, we identified two RCTs and six naturalistic studies which evaluated effects of the Val158Met polymorphism on MPH efficacy. Both RCTs evaluated genotype*dose effects: McGough et al. did not find significant COMT*MPH dose effects on ADHD symptom response,[49] while Froehlich et al. identified a trend (p=0.09) toward COMT*MPH dose effects suggesting improved response on hyperactive-impulsive symptoms with increasing dose for Val/Val homozygotes.[33] Park et al.’s open-label study also suggested a link between Val/Val homozygosity and an improved MPH response, but this finding was not significant after multiple comparison correction.[50] An additional five open-label studies reported null effects of the Val158Met genotype on MPH treatment of ADHD symptoms.[37, 40, 5153]

Meta-analysis Findings.

Although studies from the past decade do not conclusively show a role for this variant in MPH response, aggregation of the available COMT PGx data in a recent meta-analysis by Myer et al.[41] does shed light on the impact of the Val158Met polymorphism. This meta-analysis, which included data from five of the studies detailed above,[33, 49, 50, 52, 53] found that the Val/Val genotype was associated with improved MPH treatment response compared with Met carriers (OR=1.4, 95%CI:1.04–1.87; p=0.02; I2 =70%, may represent substantial heterogeneity in individual study findings).

2.2.4. Dopamine Beta Hydroxylase (DBH) [Table 4].

The enzyme DBH converts dopamine to norepinephrine.[54]

We identified only one study which evaluated the link between DBH variants and MPH response in children. This naturalistic study by Pagerols et al.[51] found a nominal association between DBH rs2073833 (MAF=0.43 [C]) and MPH treatment response that did not meet the Bonferroni-corrected threshold for statistical significance. However, a DBH two-marker haplotype was identified as a significant predictor of MPH response: the TC allelic combination of a haplotype involving rs1541332 (MAF=0.36 [G]) and rs2073833 was overrepresented among treatment-resistant patients (OR=3.43, 95%CI 1.59–7.4; padjusted=0.049), with risk for treatment failure further accentuated in children whose mothers smoked during pregnancy.[51] The functional significances of the rs1541332 and rs2073833 variants are currently unclear, although both SNPs are intronic.[55]

2.2.5. Dopamine Receptor D2 (DRD2) [Table 2].

DRD2 encodes the D2 dopamine autoreceptor, a feedback regulator critical for dopamine transmission and release, modulation of working memory, and regulation of the reward pathway.[56]

From 2009–2019, two naturalistic studies focusing on different polymorphisms assessed DRD2’s role in MPH efficacy. Gomez-Sanchez et al.[37] found an association between faster MPH response and the C allele of rs1800497 (MAF=0.33 [A]), which has previously been linked to increased DRD2 expression and availability.[57, 58] Pagerols et al. reported a trend for T allele carriers of rs2283265 (MAF=0.23 [T]) to require a higher MPH dose than G/G homozygotes (p=0.06), but did not find differences between the rs2283265 genotypes in terms of treatment response as assessed by CGI ratings.[51] One prior study of the intronic rs2283265 SNP found that T allele carriers expressed lower levels of the short isoform DRD2 (presynaptic) receptor and were more likely to have impairments on working memory and attentional tasks despite higher prefrontal cortex activity levels on functional magnetic resonance imaging (fMRI).[59]

2.2.6. Dopamine Receptor D3 (DRD3) [Table 2].

DRD3 receptors are involved in the modulation of dopamine synthesis and release, and are known to play a role in reward processing, addictive behaviors, incentive-based learning, and inhibition of motor responses.[60]

rs6280

The C allele of rs6280 (MAF=0.49 [C]) has been shown to confer higher dopamine binding affinity.[61] One placebo-controlled and one naturalistic MPH PGx study have investigated effects of rs6280 variants. In their RCT, Fageera et al. observed greater behavioral improvement with MPH treatment for rs6280 C allele homozygotes according to teacher but not parent ratings.[62] However, Pagerols et al.’s naturalistic study did not observe a significant impact of rs6280 on MPH efficacy.[51]

Additional Polymorphisms and Haplotypes

After correcting for multiple comparisons, Pagerols et al. did not find a significant main effect of nine other DRD3 SNPs (see Table 1) on MPH efficacy, although they did find that children with the GG haplotype at rs2134655-rs1800828 were more likely to be MPH non-responders.[51] Both rs2134655 (MAF=0.20 [T]) and rs1800828 (MAF=0.22 [G]) are intronic variants whose exact functions remain unknown.[55]

2.2.7. Dopamine Receptor D4 (DRD4) [Table 2].

DRD4 encodes dopamine receptor D4, a G-protein coupled receptor whose roles include inhibition of adenylyl cyclase.[63, 64] DRD4 has many important functions in the central nervous system, including regulation of corticostriatal neurotransmission by controlling glutamate receptor activity and by influencing dopamine synthesis and release.[60, 64]

Exon 3 VNTR

The 48 base pair DRD4 exon 3 VNTR has been widely evaluated in ADHD pharmacogenomics, with evidence suggesting the seven-repeat (long) allele confers decreased DRD4 activity.[65] From 2009–2019, three placebo-controlled and three naturalistic MPH PGx studies have focused on this polymorphism. The three RCTs produced varying results. McGough et al. failed to find an association between the exon 3 VNTR and MPH effects on ADHD symptoms.[49] Froehlich et al. observed that those lacking the four-repeat (short) allele among 89 American children had a decreased treatment response on hyperactive-impulsive symptoms.[33] Naumova et al. reported an association between the seven-repeat allele and improved treatment response according to parent but not teacher ratings.[66] All three naturalistic studies failed to identify an association between the exon 3 VNTR polymorphism and MPH efficacy.[35, 37, 67] However, Hong et al. identified an interaction between the DRD4 exon 3 VNTR and ADRA2A MspI genotype and MPH response.[35]

Meta-analysis Findings.

In a meta-analysis of six studies evaluating the 4-repeat VNTR (including four discussed above[33, 35, 49, 67]) and five studies evaluating the 7-repeat VNTR, Myer et al. documented a significant association between 4-repeat homozygosity and improved MPH response (OR=1.66, 95%CI:1.16–2.37, p=0.005; I2 =39%, may represent moderate heterogeneity in study findings), and a trend toward a link between the 7-repeat allele and diminished MPH response (OR=0.68, 95%CI:0.47–1.00, p=0.05; I2 =65%, may represent substantial heterogeneity in study findings).[41]

Promoter Region Polymorphism

One RCT and one naturalistic MPH PGx study have investigated the role of a 120-base pair repeat sequence located in DRD4’s promoter region which may be involved in regulating DRD4 gene expression.[68] Only Gomez-Sanchez et al.’s naturalistic study reported an association between short allele promoter duplication homozygosity and decreased MPH response,[37] while the RCT documented null effects on the outcome of ADHD symptomatic improvement.[49]

2.2.8. Dopamine Transporter (DAT/SLC6A3) [Table 5].

Table 5.

Neurotransmitter Reuptake and Release Candidate Gene Studies of Methylphenidate Efficacy Published from 2009–2019

Gene Polymorphism(s) Authors Design / Sample Findings
Dopamine Transporter (DAT/SLC6A3)
  • 3’UTR VNTR

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • No effect.

  • 3’UTR VNTR

Froehlich et al 2011 [33]
  • DBPC

  • N=89

  • U.S.A.

  • Diminished MPH response on hyperactive-impulsive symptoms with increasing MPH dose for 10-repeat carriers.

  • 3’UTR VNTR

Hong et al 2012 [35]
  • POL

  • N=103

  • South Korea

  • No independent effect.

  • Trend toward an interaction with NET −3081 (A/T).

  • 3’UTR VNTR

Stein et al 2014 [71]
  • DBPC

  • N=56

  • U.S.A

  • Diminished MPH response over time for 9-repeat homozygotes.

  • 3’UTR VNTR

  • Intron 8 VNTR

  • rs2550948

  • rs2652511

  • rs11564750

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • Improved MPH response with rs2550948 G allele.

  • Faster MPH response over time with intron 8 VNTR 6/− or −/− genotypes.

  • No effects for other variants.

  • rs460700

  • rs37020

  • rs13161905

  • rs27048

  • rs6347

  • rs11133767

  • rs40184

Pagerols et al 2017 [51]
  • POL

  • N=107

  • Spain

  • No effects.

Norepinephrine Transporter (NET/SLC6A2)
  • G1287A

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • No effect.

  • G1287A

  • −3081(A/T)

Kim et al 2010 [91]
  • POL

  • N=112

  • South Korea

  • Trend toward improved MPH response for T allele of −3081(A/T).

  • No effect for G1287A.

  • G1287A

  • −3081(A/T)

Kim et al 2011 [34]
Hong et al 2012 [35]
  • POL

  • N=102 for Kim et al

  • N=103 for Hong et al

  • South Korea

Kim et al:
  • No effects.

Hong et al:
  • No independent effects.

  • NET G1287A*ADRA2A DraI interaction effect (improved response with G allele of G1287A).

NET −3081(A/T)*ADRA2A DraI interaction effect (improved response with T allele of −3081(A/T)).
  • G1287A

  • −3081(A/T)

  • rs5568

  • rs998424

  • rs1616905

  • rs2242446

  • rs2242446 and rs1610905 haplotype

  • rs5569 and rs998424 haplotype

Lee et al 2011 [92]
  • POL

  • N=137

  • South Korea

  • No single marker or haplotype effects.

  • G1287A

Song et al 2011 [94]
  • POL

  • N=114

  • South Korea

  • Improved MPH response with G1287A G/G homozygosity.

  • G1287A

  • −3081(A/T)

Park et al 2012 [95]
  • POL

  • N=37

  • South Korea

  • Improved MPH response with G1287A G/G homozygosity based on clinician but not parent ratings.

  • No effect for −3081(A/T).

  • 30 tag SNPs

  • G1287A

Thakur et al 2012 [97]
  • DBPC

  • N=475

  • Canada

  • Improved MPH response with rs36021 T allele in children with prenatal smoke exposure.

  • Improved MPH response with rs3785152 C allele in children who lacked prenatal smoke exposure.

  • rs192303

  • rs3785143

Song et al 2014 [76]
  • POL

  • N=139

  • South Korea

  • Improved MPH response with rs192303 C/C homozygosity (not significant after multiple comparison correction).

  • No effect for rs3785143.

  • G1287A

  • −3081(A/T)

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effects.

  • G1287A

  • −3081(A/T)

  • rs2242446

  • rs3785143

  • rs3785157

  • rs7194256

Angyal et al 2018 [93]
  • POL

  • N=122

  • Hungary

  • Improved MPH response on hyperactive-impulsive symptoms for −3081(A/T) T allele carriers.

  • No independent effects for other variants.

  • Improved MPH response on inattention symptoms for the TCT haplotype of −3081(A/T), rs2242446, and rs3785143 (compared to the ATC haplotype).

Serotonin Transporter (SLC6A4/5-HTT)
  • 5HTTLPR

  • Intron 2 VNTR

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • Decreased MPH response with Intron 2 VNTR −12/−12 homozygosity.

  • No effect for 5HTTLPR.

  • 5HTTLPR

Thakur et al 2010 [105]
  • DBPC

  • N=157

  • Canada

  • Improved MPH response with long allele homozygosity.

  • 5HTTLPR

Park et al 2015 [106]
  • POL

  • N=114

  • South Korea

  • No effect.

  • 5HTTLPR

  • Intron 2 VNTR

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effects.

Solute Carrier Family 9 (SLC9A9)
  • rs9810857

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • No effect.

Synaptosomal-Associated Protein 25
(SNAP25)
  • T1065G

  • T1069C

McGough et al 2009 [49]
  • DBPC

  • N=82

  • U.S.A

  • No effects.

  • T1065G

Song et al 2014 [76]
  • POL

  • N=139

  • South Korea

  • Improved MPH response with T allele (not significant after multiple comparison adjustment).

  • T1065G

Gomez-Sanchez et al 2017 [37]
  • POL

  • N=208

  • Spain

  • Improved MPH response with G allele (no adjustment for multiple comparisons).

DBPC=double-blind, placebo-controlled

POL=prospective, open-label

SNP=single nucleotide polymorphism

VNTR=Variable nucleotide tandem repeat

DAT encodes a presynaptic protein responsible for reuptake of synaptic dopamine. One of MPH’s main mechanisms of action is to increase the concentration of dopamine in the synapse by inhibiting DAT.[69] Early ADHD PGx studies focused on DAT,[18] making it among the most well-studied candidate genes, although the rate of investigation has slowed over the past decade.

3’ UTR VNTR

A VNTR in DAT’s 3’ UTR has been linked to functional differences, as neuroimaging studies have demonstrated that individuals with the 10-repeat allele exhibit ~50% greater DAT densities compared to other genotypes.[70] From 2009–2019, three RCTs and two naturalistic studies have evaluated the effects of DAT 3’ UTR VNTR variants on MPH efficacy. The RCTs found inconsistent results: Froehlich et al.[33] observed a greater reduction in hyperactive-impulsive symptoms with increasing MPH dose for children who were not 10-repeat carriers, Stein et al.[71] found an improved MPH response in those who were 10-repeat carriers, and McGough et al.[49] did not observe a significant impact of variation in the 3’ UTR VNTR on MPH response. On the other hand, the prospective open-label MPH PGx studies were uniform in their documentation of null effects for the DAT 3’ UTR VNTR.[35, 37]

Meta-analyses Findings.

During the past decade, three meta-analyses have examined the impact of DAT 3’ UTR VNTR polymorphisms on MPH efficacy. Kambeitz et al.[72] aggregated data from 13 pediatric (none of which appear above due to their publication prior to 2009) and three adult studies, and reported no effect on ADHD total symptom scores or the specific dimensions of hyperactivity-impulsivity or inattention. Effects remained non-significant when analyses were limited to pediatric studies. However, when analyses were limited to naturalistic studies, 10-repeat homozygosity was associated with a poor response to MPH (d=−0.63, 95%CI:−1.1 to −0.15, p=0.009, I2=91%, indicating considerable heterogeneity in study findings).[72] Echoing Kambeitz et al.’s findings for naturalistic studies, Myer et al.[41] analyzed data from 16 pediatric studies (three of which were reviewed above[33, 35, 49]) and documented reduced MPH efficacy for 10-repeat homozygotes (OR=0.74, 95%CI:0.60–0.90, p=0.004; I2 =68%, may represent substantial heterogeneity in study findings). Similarly, Soleimani et al.[73] linked improved MPH treatment response with lack of 10-repeat homozygosity (10/R/10R) in a pooled analysis of eight naturalistic pediatric studies (pooled Cohen’s d=0.44; 95%CI: 0.12–0.75, p<0.01, I2 statistic not available), all of which were published prior to 2009, but effects of the 10R/10R genotype were non-significant in pooled analyses of seven clinical trials, including one reviewed above.[35]

Additional Polymorphisms

A recent study by Gomez-Sanchez et al.[37] investigated additional DAT loci (see Table 1). They observed only two significant associations: a link between improved MPH response and the G allele of rs2550948 (MAF=0.33 [T]) and an association between faster MPH response over time and lack of 6-repeat homozygosity at the intron 8 VNTR. The functional significance of the intronic rs2550948 SNP is currently unknown, while intron 8 VNTR increased repeat number has been correlated with increased DAT expression.[74]

2.2.9. Latrophilin 3/Adhesion G Protein-coupled Receptor L3 (LPHN3/ADGRL3) [Table 3].

LPHN3 encodes a brain-specific G-protein coupled receptor (Adhesion G protein-coupled Receptor L3) that is important for synaptic function and neurotransmitter exocytosis.[75]

rs1947274

From 2009–2019, four MPH PGx studies evaluated the rs1947274 (MAF=0.40 [C]) intronic SNP and reported inconsistent results. Two naturalistic studies by Song et al.[76] and Bruxel et al.[77] found no independent effects of rs1947274 variants on MPH efficacy. Song et al.[76] also reported that a haplotype block containing rs1947274 and rs6551665 (MAF=0.40 [G]) had no effect on treatment response. In contrast, in their RCT, Labbe et al. found that the rs1947274 C allele predicted MPH non-response, as assessed using an outcome measure combining direct observation of child behavior and actigraphy data.[78] On the other hand, an RCT by Choudhry et al. observed no significant main effects of rs1947274 variants; however, a haplotype analysis involving rs6551665, rs1947274, and rs6858066 (MAF=0.50 [G]) found that the GCA haplotype was associated with improved response to MPH (as rated on the CGI), but only in children whose mothers did not experience stress during pregnancy.[79] Of note, effect modification by prenatal stress may help to explain why Labbe et al. (who did not test for this interaction) and Choudhry et al. (who did) found that different rs1947274 alleles portended poor MPH response.

Meta-analysis Findings.

When Myer et al. performed a meta-analysis which included three of the studies discussed above[76, 78, 79], they did not observe a significant impact of rs1947274 variants on MPH response (OR=0.95, 95%CI:0.71–1.26, p= 0.70; I2 =81%, indicating considerable heterogeneity in study findings).[41]

rs6551665

The six MPH PGx studies evaluating the rs6551665 (MAF=0.40 [G]) intronic SNP documented mixed effects. Three naturalistic studies by Song et al.[76], Bruxel et al.[77], and Gomez-Sanchez et al.[37] reported null effects. However, Labbe et al.’s RCT linked the rs6551665 G allele to MPH non-response,[78] while a naturalistic study by Arcos-Burgos et al.[75] and an RCT by Choudhry et al.[79] observed improved MPH response with the G allele in single marker analyses. As above, Choudhry et al. also found a link between improved MPH response and the rs6551665-rs1947274-rs6858066 GCA haplotype, but only in children whose mothers did not experience prenatal stress.[79] Furthermore, in addition to the independent effects reported above, Arcos-Burgos evaluated an LPHN3 haplotype block containing rs6551665, rs1947275, and rs9683662 (MAF=0.18 [T]) and found that the haplotype containing the rs6551665 G allele conferred better treatment response. [75] An analysis by Jain et al.[80] which utilized a subset of Arcos-Burgos et al.’s[75] sample investigated the interaction between LPHN3’s rs6551665 and a haplotype block on chromosome 11q containing all or part of the genes encoding neural cell adhesion molecule 1 (NCAM1), the tetratricopeptide repeat domain 12 (TTC12), the ankyrin repeat and kinase domain containing 1 (ANKK1), and situated adjacent to the 5’ UTR of DRD2. They found that being a G/G homozygote at rs6551665 and having two copies of the 11q ADHD susceptibility haplotype to be correlated with good response to stimulant medication treatment.[80]

Meta-analysis Findings.

In Myer et al.’s meta-analysis, which analyzed data from four of studies detailed above[7679], effects of the rs6551665 variant on MPH response were non-significant (OR=1.07, 95%CI:0.84–1.37, p=0.59; I2 =77%, indicating considerable heterogeneity in study findings).[41]

rs1868790

Effects of rs1868790 (MAF=0.48 [A]) on MPH efficacy were investigated in two RCT and one naturalistic study: two reported null effects[37, 78], while Choudhry et al.’s RCT linked the T allele to improved MPH response.[79]

rs2345039

Effects of rs2345039 (MAF=0.38 [C]) on MPH efficacy were investigated in three studies. In their RCT, Labbe et al.[78] found that the G allele predicted MPH non-response, while a naturalistic study by Song et al.[76] and an RCT by Choudhry et al.[79] reported null effects.

rs6858066

Three studies evaluated the impact of rs6858066 (MAF=0.50 [G]) variants on MPH efficacy. Labbe et al.’s RCT linked the rs6858066 A allele to diminished MPH response,[78] while Gomez-Sanchez et al.’s naturalistic study[37] and Choudhry et al.’s RCT[79] found no significant main effects for rs6858066. However, Choudhry et al. reported a link between improved MPH response and the rs6551665-rs1947274-rs6858066 GCA haplotype in children whose mothers did not experience prenatal stress.[79] Of note, Choudhry et al.’s haplotype findings and Labbe et al.’s main effect findings suggest opposite effects of the A allele on MPH efficacy.

Additional Polymorphisms and Haplotypes

Although additional LPHN3 SNPs (see Table 1) have been investigated in one MPH PGx study each, none had significant main effects on MPH efficacy. However, a haplotype analysis by Bruxel et al. documented a faster MPH treatment response for CGC homozygotes of a haplotype involving rs6813183 (MAF=0.39 [G]), rs1355368 (MAF=0.48 [A]), and rs734644 (MAF=0.32 [T]).[77]

2.2.10. Metabotropic Glutamate Receptor Subtype 7 (mGluR7/GRM7) [Table 2].

The glutamate receptor subtype 7 (mGluR7/GRM7) is activated by the excitatory neurotransmitter L-glutamate and is an important presynaptic neurotransmission regulator.[81]

rs3792452

Since 2009, rs3792452’s (MAF=0.14 [A]) association with MPH efficacy has been evaluated in two open-label studies. Park et al. found that children with the G/A genotype had a better response to MPH than G/G homozygotes,[82] while Gomez-Sanchez et al. reported null effects.[37]

Genome-wide Association Study (GWAS) Findings and Integration with the Prior Literature.

Consistent with Park et al.’s findings, a 2008 GWAS by Mick et al. reported an association between rs3792452’s G allele and MPH treatment response.[83] A 2018 GWAS by Pagerols et al. did not replicate this rs3792452 finding, although they did identify a nominal association between GRM7 rs17047590 (MAF=0.16 [T]) and treatment response.[84] The functional significance has not been elucidated for either rs3792452 or rs17047590, which are both intronic.[85]

2.2.11. Norepinephrine Transporter (NET/SLC6A2) [Table 5].

NET plays a role in presynaptic reuptake of both norepinephrine and dopamine.[86, 87] It is thought to be more important for dopamine reuptake in the prefrontal cortex than the dopamine transporter due to the low density of dopamine transporters in this region and dopamine’s higher affinity for the NET compared to DAT.[88] Inhibition of NET is a key MPH mechanism of action.[89]

G1287A (rs5569)

The most well-studied NET variant in MPH PGx investigations is G1287A (rs5569, MAF=0.23 [A]). While it is routinely regarded as a silent polymorphism, G allele homozygosity has been associated with elevated levels of 3-methoxy-4-hydroxyphenylglycol, the major brain metabolite of norepinephrine.[90] From 2009–2019, G1287A was assessed in one RCT and seven prospective open-label cohorts. In their RCT, McGough did not observe significant effects of G1287A variants on MPH efficacy.[49] The open-label G1287A MPH studies also had null findings by and large, with five studies finding no significant effects,[34, 37, 9193] and only three[35, 94, 95] suggesting a role for this SNP. Both Song et al. and Park et al. documented an improved MPH response for G1287A G/G homozygotes.[94, 95] Hong et al. (who examined the same cohort as Kim et al.[34]) did not observe significant independent effects of G1287A variants, but did document significant NET G1287A*ADRA2A DraI interaction effects on MPH response, such that having the G1287 G allele portended an improved MPH response.[35]

Meta-analyses Findings:

Although the results of individual studies failed to provide conclusive evidence of G1287A’s impact on MPH efficacy, aggregation of data from seven studies (including five discussed above)[35, 49, 91, 92, 94] in Myer et al.’s meta-analysis[41] provides evidence of a link between G/G homozygosity and improved MPH response (OR versus A carriers=1.73, 95%CI:1.26–2.37, p=0.0007; I2=0%, indicating little heterogeneity in study findings). On the other hand, a meta-analysis by Angyal et al.[93] which combined data from five studies[35, 9194] did not show significant effects of the G1287A polymorphism on MPH response (MPH positive response OR for A carriers versus G homozygotes=0.92, 95%CI:0.51–1.74, p=0.76; heterogeneity statistic not provided) despite the substantial overlap between studies included in their and Myer et al.’s[41] meta-analyses.

−3081(A/T) (rs28386840)

An additional NET polymorphism of particular interest is the promoter region −3081 A to T SNP (rs28386840, MAF=0.39 [T]) whose T-allele has been linked to lower NET expression.[96] From 2009–2019, this polymorphism was investigated in six open-label cohorts. Although four of these studies failed to document significant effects of the −3081(A/T) SNP,[34, 37, 92, 95] Kim et al.[91] reported a trend toward improved MPH response with the T allele and Angyal et al.[93] found improved MPH effects on hyperactive-impulsive symptoms for T allele carriers. While Hong et al.[35] (who evaluated the same sample as Kim et al. [34]) did not observe independent effects of −3081(A/T) variants, they did find significant NET −3081(A/T)*ADRA2A DraI interaction effects on MPH response such that having the T allele of −3081(A/T) predicted improved MPH effects.

Meta-analyses Findings:

Two meta-analyses evaluated effects of the −3081(A/T) SNP on MPH efficacy. Myer et al.[41] combined data from three studies, including two which were reviewed above[35, 91] and documented improved MPH response for T allele carriers (OR versus A/A homozygotes=2.93, 95%CI:1.76–4.90, p<0.0001; I2=8%, indicating little heterogeneity in study findings). An additional meta-analysis by Angyal et al.[93], which aggregated data from the same studies as Myer et al. also showed a link between the T allele and improved MPH response (OR versus A/A homozygotes=2.70, 95%CI:1.27–5.73, p=0.01; heterogeneity statistic not provided).

Additional Polymorphisms and Haplotypes.

Additional NET polymorphisms and haplotypes (see Table 1) have been evaluated in one RCT[97] and three open-label studies[76, 92, 93]. Angyal et al.[93] observed improved MPH effects on inattention symptoms for the TCT haplotype of −3081(A/T), rs2242446 (MAF=0.25 [C]), and rs3785143 (MAF=0.10 [T]) (compared to that of the most frequent haplotype [ATC]). In an RCT, Thakur et al. found the effects of two NET SNPs were moderated by exposure to prenatal smoking: the T allele of rs36021 (MAF=0.40 [T]) was linked to improved MPH response in children whose mothers smoked during pregnancy, while the C allele of rs3785152 (MAF=0.15 [T]) was associated with improved MPH response in children whose mothers did not smoke during pregnancy.[97] The functional significance of the intronic SNPs rs36021 and rs3785152 has not been fully elucidated.[97]

2.2.12. N-methyl-D-aspartate (NMDA) Receptor Subunit 2B (GRIN2B) [Table 2].

GRIN2B encodes subunit 2B of a glutamate receptor known as the NMDA receptor.[98] Since the glutamatergic system is regulated by the dopamine and norepinephrine systems and inhibition of NMDA receptors has been shown to interfere with MPH response in animal models, it is plausible that GRIN2B polymorphisms may impact ADHD pharmacological response.[99]

We found one study which investigated the link between variants in rs2284411 (MAF=0.28 [T]), an intronic GRIN2B SNP [100] with unknown functional significance, and MPH efficacy. In an open-label trial, Kim et al. found that rs2284411 C/C homozygotes showed improved treatment effects compared to T carriers in terms of inattentive symptoms scores and overall CGI ratings, while effects on hyperactive and total symptom scores were nominally significant.[101]

2.2.13. Serotonin Transporter (SLC6A4/5-HTT) [Table 5].

SLC6A4 encodes the serotonin reuptake transporter 5-HTT. Previous evidence has linked 5-HTT dysfunction with hyperactivity and inattentiveness,[102] indicating that it may play a role in progression of ADHD and ADHD treatment response.

5HTTLPR

Over the past decade, two MPH PGx RCT and two prospective, open-label studies investigated effects of a 44 base pair insertion/deletion region in SLC6A4’s promoter region (5HTTLPR) which results in long and short variant alleles. The 5HTTLPR short allele is associated with reduced SLC6A4 expression and serotonin reuptake, while the long allele is linked to higher 5-HTT expression and more rapid serotonin uptake.[103, 104] In their RCT, Thakur et al. reported a link between long-allele homozygosity and improved MPH response.[105] McGough et al.’s RCT did not find an effect of 5HTTLPR variants on MPH response in terms of ADHD symptomatic improvement, although long-allele carriers were significantly less responsive to higher doses of MPH on a math accuracy measure.[49] Neither open-label trial identified significant effects of 5HTTLPR variants on MPH efficacy.[37, 106]

Integration with the prior literature.

Similar to findings over the past decade, the literature prior to 2009 contains scant evidence that 5HTTLPR variants impact MPH efficacy: two prior naturalistic studies found no association between 5HTTLPR variants and MPH response[107, 108], while a naturalistic study by Seeger et al. found a decreased response to MPH for children who were both 5HTTLPR long-allele homozygotes and DRD4 7-repeat carriers.[109]

Intron 2 VNTR

One placebo-controlled and one open-label study evaluated effects of a VNTR in SLC6A4’s intron 2. McGough et al.’s RCT documented a decreased MPH response (in terms of symptomatic improvement) for −12/−12 homozygotes at SLC6A4 intron 2,[49] while effects of this VNTR on MPH response were not observed in Gomez-Sanchez et al.’s naturalistic study.[37]

2.2.14. Synaptosomal-associated protein 25 (SNAP25) [Table 5].

SNAP-25 is a highly localized presynaptic neuronal protein that plays a key role in exocytosis of catecholamine-containing vesicles.[110]

T1065G (rs3746544)

During the 2009–2019 period, we identified one RCT and two naturalistic studies evaluating effects of T1065G (rs3746544, MAF=0.28 [G]) variants on MPH efficacy. McGough et al.’s RCT[49] found no association between T1065G variants and MPH response,[49] while the open-label studies documented opposing effects. Song et al. found a link between the T allele of T1065G and improved MPH response (on an outcome which combined CGI and ADHD symptom ratings) that was no longer significant when adjusted for multiple comparisons.[76] In contrast, Gomez-Sanchez et al. reported that carriers for the G allele of T1065G showed a better MPH response than T/T homozygotes, although this finding was not adjusted for multiple comparisons.[37]

Integration with the Prior Literature:

Only one prior study by McGough et al.[111] investigated effects of T1065G variants on MPH response: they found that the T allele of T1065G was linked to better MPH dose response. Given the sparsity of the prior literature and recent null or mixed effects, there does not appear to be conclusive evidence that T1065G variants impact MPH efficacy.

2.2.15. Tyrosine Hydroxylase (TH) [Table 4].

TH is the rate-limiting enzyme in the synthesis of catecholamines, including dopamine, noradrenaline and adrenaline, as it converts tyrosine into l-dopa.[112, 113]

During the past decade, one naturalistic pediatric study by Pagerols et al. investigated effects of three TH variants on MPH efficacy (see Table 1).[51] They reported that C/C homozygotes for rs2070762 (MAF=0.42 [C]) were less responsive to MPH than T allele carriers in whole sample analyses, and that this adverse impact of C/C homozygosity on MPH response was accentuated in children who had in utero tobacco exposure.[51] In vitro assays indicate that the rs2070762 C allele might be a functional enhancer element regulating TH gene expression.[114]

2.3. Methylphenidate Genome-Wide Approaches

One pediatric MPH PGx GWAS was identified by our 2009–2019 literature search. This GWAS did not identify any variants which had a significant association with MPH efficacy (as assessed via CGI ratings) after multiple comparison correction in a discovery cohort of 173 Spanish children.[84] However, a meta-analysis of this pediatric cohort and a second cohort of 189 adults utilized a polygenic risk score model and found a significant association between MPH response and variants in the PEBP4 (rs17685420, MAF=0.16 [T]) gene, as well as suggestive effects of markers within the ALDH1L1 (rs2886059, MAF=0.24 [A]), CDH23 (rs17712523, MAF=0.13 [A]), and ARSA (rs2071421, MAF=0.22 [C]) genes.[84] PEBP4, a serine protease important for cellular development, has been implicated in Alzheimer’s disease.[115] Both ALDH1L1, which encodes an aldehyde dehydrogenase, and CDH23, which encodes an important cell adhesion molecule, have previously been associated with ADHD.[116, 117] Additionally, deficiencies in ARSA, a sulfatase, have been associated with poor school performance and cerebellar ataxia.[118]

3. Amphetamine Pharmacogenomic Studies

Like MPH, amphetamines act by blocking reuptake of dopamine and norepinephrine through dopamine transporter and NET inhibition.[21] In addition, amphetamines increase synaptic dopamine by inhibiting the vesicular monoamine transporter 2, which releases dopamine from vesicular storage, as well as by reverse transport at DAT.[21, 119]. The hepatic enzyme cytochrome P450 2D6 (CYP2D6), which also breaks down ATX, has been implicated in amphetamine metabolism.[120] However, our literature search did not identify any amphetamine PGx studies investigating the link between ADHD symptom response after amphetamine treatment and variants in pharmacokinetic genes.

3.1. Amphetamine Pharmacodynamic Genetic Variants

3.1.1. Dopamine Transporter (DAT/SLC6A3).

We identified only one PGx study from 2009–2019 that evaluated pharmacodynamics genetic predictors of amphetamine effects on ADHD symptoms. This RCT by Stein et al.[71] investigated whether polymorphisms in DAT’s 3’ UTR VNTR moderated the dose-response effects of both dexmethylphenidate (d-MPH) and mixed amphetamine salts (MAS). They found a reduced response to both medications for 9-repeat homozygotes: 10–20 mg doses of either d-MPH or MAS had little to no effect on hyperactivity-impulsivity or ADHD total symptom scores in children with the 9/9 genotype, but low to moderate stimulant doses did lead to improvement in 10-repeat carriers.

4. Atomoxetine Pharmacogenomic Studies

ATX was the first non-stimulant medication approved for use in ADHD. ATX acts to increase availability of synaptic norepinephrine through selective inhibition of NET.[121, 122] In addition, since NET is thought to play a large role in prefrontal cortex dopamine reuptake given the low density of DAT in this region,[88] ATX blockade of NET may also increase synaptic dopamine in the prefrontal cortex .

4.1. Atomoxetine Pharmacokinetic Genetic Variants

4.1.1. Cytochrome P450 2D6 (CYP2D6) [Table 6].

Table 6.

Candidate Gene Studies of Atomoxetine Efficacy Published from 2009–2019

Gene Polymorphism(s) Authors Design / Sample Findings
Adrenergic α1AReceptor (ADRA1A)
  • rs17426222

  • rs573514

  • rs3808585

Yang et al 2013 [132]
  • POL

  • N=111

  • China

  • No effects.

Adrenergic α2AReceptor (ADRA2A)
  • MspI

  • DraI

Yang et al 2013 [132]
  • POL

  • N=111

  • China

  • Non-remission status linked to GG haplotype of MspI and DraI (not significant after correction for multiple comparisons).

Cytochrome P-450 2D6 (CYP2D6)
  • *3, *4, *5, *6, *7, *8 alleles1

Ramoz et al 2009 [126] Included two cohorts:
Multinational
  • POL

  • N=160

U.S.A.
  • DB crossover with methylphenidate

  • N=105

  • No effects.

Dopamine Beta Hydroxylase (DBH)
  • rs1076150

  • rs1611115

  • rs1108580

  • rs2873804

  • rs1548364

  • rs2519154

  • rs2073837

  • rs129882

Fang et al 2015 [130]
  • POL

  • N=87

  • China

  • Decreased ATX response with C allele of rs2519154.

  • No effects for other SNPs or for haplotypes from 2 linkage disequilibrium blocks.

Norepinephrine Transporter (NET/SLC6A2)
  • 108 SNPs

Ramoz et al 2009 [126] Included two cohorts:
Multinational
  • POL

  • N=160

U.S.A.
  • DB crossover with methylphenidate

  • N=105

  • No effects for any SNP after multiple comparison correction.

  • ATX response linked to a linkage disequilibrium (LD) block covering exons 4–9, including rs3785143.

  • rs3785143

  • rs3785152

  • rs2279805

  • rs5569

  • rs36009

  • rs2242447

Yang et al 2013 [132]
  • POL

  • N=111

  • China

  • ATX non-response linked to rs3785143 T allele.

  • No effects for other SNPs after multiple comparison correction.

1

Participants who were homozygous or heterozygous for alleles *3, *4, *5, *6, *7, and/or *8 were considered poor metabolizers.

ATX=atomoxetine

DB=double-blind

POL=prospective, open-label

SNP=single nucleotide polymorphism

ATX catabolism is primarily mediated by the liver enzyme CYP2D6[123], which converts ATX to 4-hydroxyatomoxetine, an active metabolite.[124] CYP2D6 is a highly polymorphic enzyme; the link between several functional CYP2D6 polymorphisms and altered metabolizer status has been well-documented. Poor metabolizers (PM) possess one or more mutations in CYP2D6 conferring reduced enzyme activity while extensive metabolizers (EM) exhibit normal/wild type enzyme activity.[125]

Since 2009, the impact of CYP2D6 variants on ATX efficacy was investigated in two samples (one from the U.S. and one multinational) via an RCT design. Findings from both samples were reported by Ramoz et al.[126] They did not observe differences between individuals with CYP2D6 PM versus EM genotypes in ATX efficacy in either cohort, when analyses combined both cohorts, or when analyses were limited to Caucasian participants.

Integration with the Prior Literature:

We previously reported[18] on a 2007 pooled analysis conducted by Michelson et al.[127] which found that CYP2D6 PMs had greater overall symptom reduction scores compared to EMs. Of note, half of the Ramoz et al. cohort [126] was included in the Michelson et al. pooled analysis; therefore, the lack of effect for CYP2D6 metabolizer status observed by Ramoz et al. may be due to small sample size and attendant loss of power. Indeed, in a separate pooled analysis of 1326 children conducted by Trzepacz et al. in 2008,[128] CYP2D6 PMs were found to have significantly greater improvement in inattention symptom scores than EMs. Ultimately, these efficacy findings, combined with evidence that PMs experience increased adverse effects (including greater changes in heart rate and blood pressure) compared to EMs[127, 128] as well repeated demonstration of CYP2D6 genotype effects on ATX pharmacokinetics (reviewed in Brown et al.’s Supplement[129] and by Yu et al.[125]), were deemed strong enough to form the basis for a newly issued CPIC guideline providing therapeutic recommendations for ATX based on CYP2D6 genotype.[129]

4.2. Atomxetine Pharmacodynamic Genetic Variants

Since ATX selectively impacts NET, several studies have explored the effect of variability in noradrenergic-related genes on ATX response in children with ADHD.

4.2.1. Dopamine Beta Hydroxylase (DBH) [Table 6].

DBH plays a key role in the synthesis of norepinephrine from dopamine.[54]

Only one ATX PGx efficacy study[130] during the 2009–2019 period evaluated DBH variants. In their naturalistic study, Fang et al. investigated the link between ATX response and eight DBH SNPs as well as haplotypes from two linkage disequilibrium blocks. Although nominal associations with ATX efficacy were found for multiple DBH SNPs and haplotypes, only one SNP, rs2519154 (MAF=0.43 [T]), was significantly linked to ATX response after multiple comparison correction, with the C allele being associated with non-responder status. Rs2519154 is intronic, and is thought to play a role in DBH gene regulation.[130]

4.2.2. Norepinephrine Transporter (NET/SLC6A2) [Table 6].

Given that ATX’s mechanism of action involves selective inhibition of NET, which leads to increased synaptic norepinephrine in multiple brain regions as well as increased synaptic dopamine in the prefrontal cortex,[131] it is not surprising that ATX PGx studies have also focused on NET variants.

From 2009–2019, two naturalistic studies investigated the link between NET variants and ATX efficacy. Ramoz et al.[126] investigated the association between NET polymorphisms and ATX response in two independent cohorts (one from the U.S.A and one multinational). Of 108 sequenced NET polymorphisms, 20 SNPs were nominally associated with ATX response in at least one of the two cohorts, although none were statistically significant after multiple comparison correction. However, Ramoz et al. did identify a linkage disequilibrium (LD) block covering NET’s exons 4–9 (where 36 SNPs were genotyped) which was significantly associated with treatment response such that a carrier of this allele sequence would have twice the chance of responding to ATX compared to a non-carrier (OR=1.83, p<0.01 in the combined cohort). Yang et al.[132] investigated the link between six NET SNPs and ATX efficacy. They only found one SNP to be linked to ATX efficacy after correction for multiple comparisons: the rs3785143 T allele was significantly associated with ATX non-responder status and the C allele with being an ATX responder. Yang et al. considered their results to be consistent with prior findings: when Ramoz et al.[126] evaluated the LD block covering NET’s exons 4–9, they found that the haplotype containing the rs3785143 C allele was associated with ATX responder status. Although the exact function of rs3785143 has not been fully elucidated, it is located in intron 1, a region important for NET transcriptional regulation.[133]

5. Alpha-agonist Pharmacogenomic Studies

In recent years, the non-stimulant medications guanfacine and clonidine have been FDA-approved for the treatment of ADHD. Guanfacine and clonidine both selectively stimulate α2-adrenergic receptors in the prefrontal cortex, modulating norepinephrine activity and decreasing impulsivity and hyperactivity.[134] Guanfacine is metabolized by CYP3A4.[134] While clonidine metabolism is not well-understood, evidence indicates that its breakdown is primarily mediated by CYP2D6 and to a lesser extent by CYP3A4/5 and CYP1A1/2.[135] Unfortunately, our search of the 2009–2019 literature did not identify any guanfacine or clonidine PGx pharmacodynamic or pharmacokinetic studies.

6. Current Challenges and Future Directions

The past decade of pediatric ADHD medication efficacy PGx research has been marked by further accumulation of candidate gene studies focusing mostly on associations with MPH response. A lesser number of ADHD PGx studies have addressed ATX and little to none have evaluated amphetamine, guanfacine, and clonidine. Unfortunately, even when studies of a given genetic variant and ADHD medication efficacy are available, our understanding of their findings is complicated by numerous factors.

Specifically, interpretation of ADHD PGx study results and comparison of findings across studies continues to be constrained considerably by heterogeneity and limitations in study design.[18] For example, many studies are open-label: their inability to account for placebo effects makes it difficult to identify true medication response and its association with genetic variability. Indeed, Polanczyk et al. found that study design (open studies versus RCTs) was associated with heterogeneity of study results in ADHD pharmacogenetics studies.[136] In addition, variability in dosing regimens also presents a barrier to interpretation, as the use of only lower doses in some studies may bias against finding significant treatment effects. Current studies are also limited by their outcome measures, as dichotomous measures (e.g., medication responder versus non-responder status) have less power to detect effects than quantitative measures. The most commonly used measures, including ADHD symptom rating scales and Clinician Global Impression ratings, are also subjective, and relationships between genetic variants and medication response have sometimes varied in the same sample depending on who completed the ratings (e.g., parents versus teachers).[66] In addition, ADHD PGx studies have focused on predictors of short-term effects, with few assessing long-term outcomes, even though responder status to the same medication and dose has been shown to change over time.[37, 49]

The heterogeneity of ADHD PGx study samples--in terms of racial/ethnic and clinical characteristics—also presents a barrier to understanding study findings. Genetic factors predicting medication response may vary by racial/ethnic group, suggesting that the polymorphisms being studied may not be themselves influencing medication response but may be in linkage disequilibrium with the actual functional genetic variants.[137] Unique risk and protective factors which influence medication efficacy may also be present in different ethnic groups.[41] Given prior evidence that some polymorphisms may influence MPH response on inattentive or hyperactive-impulsive symptoms but not both domains[33, 101], the ADHD presentation (e.g., inattentive, hyperactive-impulsive, or combined) profile of the sample may impact PGx study findings.[36] Differences in the comorbidity profile of study samples may also influence study results: children with varied coexisting conditions may have unique underlying neurophysiologies such that different genetic variants may be operative in ADHD medication response. In addition, there is emerging evidence that the effect of a given genetic variant on medication response may be moderated by other genetic polymorphisms,[35, 37] yet few ADHD PGx investigations have considered these interactions. Care must be taken, however, to ensure that studies of interactions or specific subgroups are independent and hypothesis-driven, rather than being conducted for the sake of documenting “statistically significant” findings obtained by chance after performing multiple non-adjusted comparisons.

ADHD PGx studies have also been limited by their generally small sample sizes, which have limited power to detect modest genetic effects. Therefore, the ADHD PGx knowledge base has been considerably advanced by recent meta-analyses[41, 72, 93] which aggregated data from multiple studies to identify a significant impact on MPH efficacy for variants in DAT, DRD4, ADRA2A, COMT, and NET. However, in discussing their meta-analysis findings, Myer et al. note that the OR for each variant is low; therefore, knowing an individual’s genotype for any one variant is unlikely to reliably predict ADHD treatment choice.[41] Rather, multivariable models that incorporate the impact of multiple genes and/or investigation of polygenic risk scores may one day yield clinically actionable MPH efficacy information. Furthermore, to more fully capture the inputs which drive family choice of treatment, ADHD PGx studies and multivariable prediction models need to incorporate not just measures of medication efficacy, but also side effects. Of note, there is a burgeoning ADHD PGx adverse effects literature[138], although its review is beyond the scope of the present article.

Our ability to determine the most important ADHD pharmacogenomic predictors is also likely limited by the field’s focus on pharmacodynamic genes and its dependence on the candidate gene approach. In many branches of medicine where PGx tools are gaining traction, the approach has centered on elucidating the role of pharmacokinetic gene polymorphisms.[139] Indeed, we have made progress in understanding how CYP2D6 polymorphisms impact ATX response such that modifications to ATX dosing based on CYP2D6 genotyping are included in ATX’s FDA labeling. In addition, an ATX-CYP2D6 CPIC guideline has recently been published.[129] However, few prior publications have examined the association between pharmacokinetic genes and MPH, amphetamine, guanfacine, or clonidine efficacy. Furthermore, given our imperfect understanding of the neurobiology underlying ADHD medication response, it is unlikely that we can develop candidate polymorphism hypotheses that cover all the salient pharmacodynamic inputs. To address this limitation, it is likely that more hypothesis-free efforts, such as genome-wide approaches, are needed. However, the need for large sample sizes, which entail considerable expense, presents a significant barrier to conducting PGx GWAS. Investigators targeting other diseases/disorders have addressed this barrier by creating large databases/registries which bank standardized clinical data and biological samples (including DNA) academic institution-wide or across multiple collaborating institutions in a study consortium (e.g., the Children’s Oncology Group).[140] Collection of DNA samples as a standard for all ADHD medication randomized controlled trials would also considerably increase the available data for future ADHD PGx efforts. The resultant sizable databases could be used not just for GWAS, but could also serve as a convenient repository facilitating replication of candidate gene findings from discovery samples (thereby acting as a check on the release of false positive findings).

Increasingly, personalized predictive efforts in other branches of medicine are also expanding beyond standard pharmacogenomic approaches focusing on the code of individual genes. Indeed, there is growing recognition that regulation of gene expression through epigenetic mechanisms may play an important role in drug response, and there is increasing evidence that epigenetic patterns in blood samples can predict and/or reflect response to psychiatric medications, including antidepressants[141148] and antipsychotics.[146, 149153] It is also being proposed that pharmacometabolomic approaches may be able to capture the impact of environmental, gene*environment, gene*gene, and epigenetic influences. The metabolome is the repertoire of small molecules–including amine neurotransmitters, lipid signaling molecules, and many other classes of compounds--found within a biological system and representing the final product of interactions between gene expression, protein expression, and the cellular environment.[154] Intriguingly, pharmacometabolomic studies of schizophrenia[155] and major depression[156] have furthered understanding of drug response and provided groundwork for identifying predictive metabolic markers for psychiatric medications.

7. Conclusions

Despite the proliferation of numerous additional PGx publications over the past decade, the era of personalized predictive ADHD clinical care has not yet dawned. To make progress, adequately powered studies which pay careful attention to the study design and sample considerations discussed above are needed. ADHD pharmacogenomic efforts which incorporate pharmacokinetic as well as pharmacodynamic gene investigations, genome-wide approaches, and gene*gene interactions are also likely necessary to drive the field forward. Efforts which go beyond investigation of gene coding sequences, possibly by incorporating epigenetic and metabolomic approaches, may also be crucial to the development of clinically salient ADHD medication response prediction tools.

Supplementary Material

Supplemental Table 1
Supplemental Table 2
Supplemental Table 3
Supplemental Table 4
Supplemental Table 5
Supplemental Table 6

Key Points.

  • Over the past decade, pediatric ADHD medication efficacy pharmacogenomics research has largely focused on pharmacodynamic candidate gene studies of methylphenidate response. Although results of individual studies have often lacked consistency, recent meta-analyses have reported a small but significant impact on methylphenidate efficacy of variants in the dopamine transporter (DAT/SLC6A3), dopamine receptor D4 (DRD4), adrenergic α2A-receptor (ADRA2A), catechol-O-methyltrasferase (COMT), and norepinephrine transporter (NET/SLC6A2) genes.

  • Progress has also been made in understanding how genetic polymorphisms in the atomoxetine-metabolizing enzyme CYP2D6 impact atomoxetine response and pharmacokinetics. In fact, a recently published Clinical Pharmacogenetics Implementation Consortium (CPIC) guideline outlines some recommended variations in atomoxetine dosing based on CYP2D6 genotype.

  • Notwithstanding these advances, the field to date has been unable to translate the plethora of available ADHD pharmacogenomics study findings into a tool that can predict with precision which ADHD medication will be most effective for each individual patient. There are myriad factors underlying this impasse--including the small effects of individual genetic variants, study design weaknesses and inconsistencies, lack of confirmatory studies in different racial/ethnic populations, and failure to consider gene-gene interactions--which should be addressed when undertaking future ADHD pharmacogenomics research.

Funding:

No funding was received for the preparation of this manuscript.

Footnotes

Conflicts of Interest:

TF is the principal investigator for a National Institutes of Mental Health (NIMH) award relating to methylphenidate. This NIMH funding is awarded to TF’s institution (so she does not receive any funding directly for this grant), and NIMH is a U.S. government agency that does not have any financial stake in methylphenidate. NE and KY have no potential conflicts of interest to report.

REFERENCES

  • 1.Thomas R, Sanders S, Doust J, Beller E, Glasziou P. Prevalence of attention-deficit/hyperactivity disorder: a systematic review and meta-analysis. Pediatrics. 2015. April;135(4):e994–1001. [DOI] [PubMed] [Google Scholar]
  • 2.Polanczyk GV, Salum GA, Sugaya LS, Caye A, Rohde LA. Annual research review: A meta-analysis of the worldwide prevalence of mental disorders in children and adolescents. J Child Psychol Psychiatry. 2015. Mar;56(3):345–65. [DOI] [PubMed] [Google Scholar]
  • 3.Wolraich ML, Hagan JF Jr., Allan C, Chan E, Davison D, Earls M, et al. Clinical Practice Guideline for the Diagnosis, Evaluation, and Treatment of Attention-Deficit/Hyperactivity Disorder in Children and Adolescents. Pediatrics. 2019. October;144(4). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Diagnostic and Statistical Manual of Mental Disorders (DSM-5). 5th ed. Washington, DC: American Psychiatric Association; 2013. [Google Scholar]
  • 5.Cortese S, Adamo N, Del Giovane C, Mohr-Jensen C, Hayes AJ, Carucci S, et al. Comparative efficacy and tolerability of medications for attention-deficit hyperactivity disorder in children, adolescents, and adults: a systematic review and network meta-analysis. Lancet Psychiatry. 2018. September;5(9):727–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Barbaresi WJ, Katusic SK, Colligan RC, Weaver AL, Jacobsen SJ. Modifiers of long-term school outcomes for children with attention-deficit/hyperactivity disorder: does treatment with stimulant medication make a difference? Results from a population-based study. Journal of Developmental and Behavioral Pediatrics. 2007. August;28(4):274–86. [DOI] [PubMed] [Google Scholar]
  • 7.Scheffler RM, Brown TT, Fulton BD, Hinshaw SP, Levine P, Stone S. Positive association between attention-deficit/ hyperactivity disorder medication use and academic achievement during elementary school. Pediatrics. 2009. May;123(5):1273–9. [DOI] [PubMed] [Google Scholar]
  • 8.Chang Z, Lichtenstein P, D’Onofrio BM, Sjolander A, Larsson H. Serious transport accidents in adults with attention-deficit/hyperactivity disorder and the effect of medication: a population-based study. JAMA Psychiatry. 2014. March;71(3):319–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Chang Z, Quinn PD, Hur K, Gibbons RD, Sjolander A, Larsson H, et al. Association Between Medication Use for Attention-Deficit/Hyperactivity Disorder and Risk of Motor Vehicle Crashes. JAMA Psychiatry. 2017. June 1;74(6):597–603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Dalsgaard S, Leckman JF, Mortensen PB, Nielsen HS, Simonsen M. Effect of drugs on the risk of injuries in children with attention deficit hyperactivity disorder: a prospective cohort study. Lancet Psychiatry. 2015. August;2(8):702–9. [DOI] [PubMed] [Google Scholar]
  • 11.Chang Z, D’Onofrio BM, Quinn PD, Lichtenstein P, Larsson H. Medication for Attention-Deficit/Hyperactivity Disorder and Risk for Depression: A Nationwide Longitudinal Cohort Study. Biol Psychiatry. 2016. December 15;80(12):916–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lichtenstein P, Halldner L, Zetterqvist J, Sjolander A, Serlachius E, Fazel S, et al. Medication for attention deficit-hyperactivity disorder and criminality. The New England journal of medicine. 2012. November 22;367(21):2006–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Lowe N, Barry E, Gill M, Hawi Z. An Overview of the Pharmacogenetics and Molecular Genetics of ADHD. Current Pharmacogenomics. 2006;4:231–43. [Google Scholar]
  • 14.Stein MA, McGough JJ. The pharmacogenomic era: promise for personalizing attention deficit hyperactivity disorder therapy. Child and adolescent psychiatric clinics of North America. 2008. April;17(2):475–90, xi-xii. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Relling MV, Klein TE. CPIC: Clinical Pharmacogenetics Implementation Consortium of the Pharmacogenomics Research Network. Clin Pharmacol Ther. 2011. March;89(3):464–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Clinical Pharmacogenetics Implementation Consortium. Guidelines. 2019. [cited 2019, April 10; Available from: https://cpicpgx.org/guidelines
  • 17.Shuldiner AR, Palmer K, Pakyz RE, Alestock TD, Maloney KA, O’Neill C, et al. Implementation of pharmacogenetics: the University of Maryland Personalized Anti-platelet Pharmacogenetics Program. Am J Med Genet C Semin Med Genet. 2014. March;166C(1):76–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Froehlich TE, McGough JJ, Stein MA. Progress and promise of attention-deficit hyperactivity disorder pharmacogenetics. CNS drugs. 2010. February 1;24(2):99–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Zerbino DR, Achuthan P, Akanni W, Amode MR, Barrell D, Bhai J, et al. Ensembl 2018. Nucleic Acids Res. 2018. January 4;46(D1):D754–D61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Solanto MV. Neuropsychopharmacological mechanisms of stimulant drug action in attention-deficit hyperactivity disorder: a review and integration. Behavioural brain research. 1998. July;94(1):127–52. [DOI] [PubMed] [Google Scholar]
  • 21.Faraone SV. The pharmacology of amphetamine and methylphenidate: Relevance to the neurobiology of attention-deficit/hyperactivity disorder and other psychiatric comorbidities. Neurosci Biobehav Rev. 2018. April;87:255–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Sun Z, Murry DJ, Sanghani SP, Davis WI, Kedishvili NY, Zou Q, et al. Methylphenidate is stereoselectively hydrolyzed by human carboxylesterase CES1A1. J Pharmacol Exp Ther. 2004. August;310(2):469–76. [DOI] [PubMed] [Google Scholar]
  • 23.Nemoda Z, Angyal N, Tarnok Z, Gadoros J, Sasvari-Szekely M. Carboxylesterase 1 gene polymorphism and methylphenidate response in ADHD. Neuropharmacology. 2009. December;57(7–8):731–3. [DOI] [PubMed] [Google Scholar]
  • 24.Zhu HJ, Patrick KS, Yuan HJ, Wang JS, Donovan JL, DeVane CL, et al. Two CES1 gene mutations lead to dysfunctional carboxylesterase 1 activity in man: clinical significance and molecular basis. American journal of human genetics. 2008. June;82(6):1241–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Stage C, Jurgens G, Guski LS, Thomsen R, Bjerre D, Ferrero-Miliani L, et al. The impact of CES1 genotypes on the pharmacokinetics of methylphenidate in healthy Danish subjects. Br J Clin Pharmacol. 2017. July;83(7):1506–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Kaddurah-Daouk R, Hankemeier T, Scholl EH, Baillie R, Harms A, Stage C, et al. Pharmacometabolomics Informs About Pharmacokinetic Profile of Methylphenidate. CPT Pharmacometrics Syst Pharmacol. 2018. August;7(8):525–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Johnson KA, Barry E, Lambert D, Fitzgerald M, McNicholas F, Kirley A, et al. Methylphenidate side effect profile is influenced by genetic variation in the attention-deficit/hyperactivity disorder-associated CES1 gene. J Child Adolesc Psychopharmacol. 2013. December;23(10):655–64. [DOI] [PubMed] [Google Scholar]
  • 28.Nestler EJ, Hyman SE, Malenka RC. Molecular Neuropharmacology: A Foundation for Clinical Neuroscience. New York: The McGraw-Hill Companies, Inc.; 2001. [Google Scholar]
  • 29.Arnsten AF, Dudley AG. Methylphenidate improves prefrontal cortical cognitive function through alpha2 adrenoceptor and dopamine D1 receptor actions: Relevance to therapeutic effects in Attention Deficit Hyperactivity Disorder. Behav Brain Funct. 2005. April 22;1(1):2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Andrews GD, Lavin A. Methylphenidate increases cortical excitability via activation of alpha-2 noradrenergic receptors. Neuropsychopharmacology. 2006. March;31(3):594–601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Hegvik TA, Jacobsen KK, Fredriksen M, Zayats T, Haavik J. A candidate gene investigation of methylphenidate response in adult attention-deficit/hyperactivity disorder patients: results from a naturalistic study. J Neural Transm (Vienna). 2016. August;123(8):859–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Lario S, Calls J, Cases A, Oriola J, Torras A, Rivera F. MspI identifies a biallelic polymorphism in the promoter region of the alpha 2A-adrenergic receptor gene. Clin Genet. 1997. February;51(2):129–30. [PubMed] [Google Scholar]
  • 33.Froehlich TE, Epstein JN, Nick TG, Melguizo Castro MS, Stein MA, Brinkman WB, et al. Pharmacogenetic predictors of methylphenidate dose-response in attention-deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry. 2011. November;50(11):1129–39 e2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Kim BN, Cummins TD, Kim JW, Bellgrove MA, Hong SB, Song SH, et al. Val/Val genotype of brain-derived neurotrophic factor (BDNF) Val(6)(6)Met polymorphism is associated with a better response to OROS-MPH in Korean ADHD children. Int J Neuropsychopharmacol. 2011. November;14(10):1399–410. [DOI] [PubMed] [Google Scholar]
  • 35.Hong SB, Kim JW, Cho SC, Shin MS, Kim BN, Yoo HJ. Dopaminergic and noradrenergic gene polymorphisms and response to methylphenidate in korean children with attention-deficit/hyperactivity disorder: is there an interaction? J Child Adolesc Psychopharmacol. 2012. October;22(5):343–52. [DOI] [PubMed] [Google Scholar]
  • 36.Park S, Kim JW, Kim BN, Hong SB, Shin MS, Yoo HJ, et al. No significant association between the alpha-2A-adrenergic receptor gene and treatment response in combined or inattentive subtypes of attention-deficit hyperactivity disorder. Pharmacopsychiatry. 2013. July;46(5):169–74. [DOI] [PubMed] [Google Scholar]
  • 37.Gomez-Sanchez CI, Carballo JJ, Riveiro-Alvarez R, Soto-Insuga V, Rodrigo M, Mahillo-Fernandez I, et al. Pharmacogenetics of methylphenidate in childhood attention-deficit/hyperactivity disorder: long-term effects. Sci Rep. 2017. September 4;7(1):10391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Cheon KA, Cho DY, Koo MS, Song DH, Namkoong K. Association between homozygosity of a G allele of the alpha-2a-adrenergic receptor gene and methylphenidate response in Korean children and adolescents with attention-deficit/hyperactivity disorder. Biol Psychiatry. 2009. April 1;65(7):564–70. [DOI] [PubMed] [Google Scholar]
  • 39.Huang HC, Wu LS, Yu SC, Wu BJ, Lua AC, Lee SM, et al. The Alpha-2A Adrenergic Receptor Gene −1291C/G Single Nucleotide Polymorphism is Associated with the Efficacy of Methylphenidate in Treating Taiwanese Children and Adolescents with Attention-Deficit Hyperactivity Disorder. Psychiatry Investig. 2018. March;15(3):306–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Unal D, Unal MF, Alikasifoglu M, Cetinkaya A. Genetic Variations in Attention Deficit Hyperactivity Disorder Subtypes and Treatment Resistant Cases. Psychiatry Investig. 2016. July;13(4):427–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Myer NM, Boland JR, Faraone SV. Pharmacogenetics predictors of methylphenidate efficacy in childhood ADHD. Mol Psychiatry. 2018. September;23(9):1929–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Kim BN, Kim JW, Cummins TD, Bellgrove MA, Hawi Z, Hong SB, et al. Norepinephrine genes predict response time variability and methylphenidate-induced changes in neuropsychological function in attention deficit hyperactivity disorder. J Clin Psychopharmacol. 2013. June;33(3):356–62. [DOI] [PubMed] [Google Scholar]
  • 43.Trzaska KA, King CC, Li KY, Kuzhikandathil EV, Nowycky MC, Ye JH, et al. Brain-derived neurotrophic factor facilitates maturation of mesenchymal stem cell-derived dopamine progenitors to functional neurons. J Neurochem. 2009. August;110(3):1058–69. [DOI] [PubMed] [Google Scholar]
  • 44.Pezawas L, Verchinski BA, Mattay VS, Callicott JH, Kolachana BS, Straub RE, et al. The brain-derived neurotrophic factor val66met polymorphism and variation in human cortical morphology. J Neurosci. 2004. November 10;24(45):10099–102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ho BC, Milev P, O’Leary DS, Librant A, Andreasen NC, Wassink TH. Cognitive and magnetic resonance imaging brain morphometric correlates of brain-derived neurotrophic factor Val66Met gene polymorphism in patients with schizophrenia and healthy volunteers. Arch Gen Psychiatry. 2006. July;63(7):731–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Flanagin BA, Cook EH Jr., de Wit H. An association study of the brain-derived neurotrophic factor Val66Met polymorphism and amphetamine response. Am J Med Genet B Neuropsychiatr Genet. 2006. September 5;141B(6):576–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Napolitano A, Cesura AM, Da Prada M. The role of monoamine oxidase and catechol O-methyltransferase in dopaminergic neurotransmission. J Neural Transm Suppl. 1995;45:35–45. [PubMed] [Google Scholar]
  • 48.Lachman HM, Papolos DF, Saito T, Yu YM, Szumlanski CL, Weinshilboum RM. Human catechol-O-methyltransferase pharmacogenetics: description of a functional polymorphism and its potential application to neuropsychiatric disorders. Pharmacogenetics. 1996. June;6(3):243–50. [DOI] [PubMed] [Google Scholar]
  • 49.McGough JJ, McCracken JT, Loo SK, Manganiello M, Leung MC, Tietjens JR, et al. A candidate gene analysis of methylphenidate response in attention-deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry. 2009. December;48(12):1155–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Park S, Kim JW, Kim BN, Shin MS, Yoo HJ, Cho SC. Catechol-O-methyltransferase Val158-Met polymorphism and a response of hyperactive-impulsive symptoms to methylphenidate: A replication study from South Korea. J Psychopharmacol. 2014. July;28(7):671–6. [DOI] [PubMed] [Google Scholar]
  • 51.Pagerols M, Richarte V, Sanchez-Mora C, Garcia-Martinez I, Corrales M, Corominas M, et al. Pharmacogenetics of methylphenidate response and tolerability in attention-deficit/hyperactivity disorder. Pharmacogenomics J. 2017. January;17(1):98–104. [DOI] [PubMed] [Google Scholar]
  • 52.Yatsuga C, Toyohisa D, Fujisawa TX, Nishitani S, Shinohara K, Matsuura N, et al. No association between catechol-O-methyltransferase (COMT) genotype and attention deficit hyperactivity disorder (ADHD) in Japanese children. Brain Dev. 2014. August;36(7):620–5. [DOI] [PubMed] [Google Scholar]
  • 53.Salatino-Oliveira A, Genro JP, Zeni C, Polanczyk GV, Chazan R, Guimaraes AP, et al. Catechol-O-methyltransferase valine158methionine polymorphism moderates methylphenidate effects on oppositional symptoms in boys with attention-deficit/hyperactivity disorder. Biol Psychiatry. 2011. August 1;70(3):216–21. [DOI] [PubMed] [Google Scholar]
  • 54.Kaufman S, Friedman S. Dopamine-Beta-Hydroxylase. Pharmacol Rev. 1965. June;17:71–100. [PubMed] [Google Scholar]
  • 55.Ribases M, Ramos-Quiroga JA, Hervas A, Sanchez-Mora C, Bosch R, Bielsa A, et al. Candidate system analysis in ADHD: evaluation of nine genes involved in dopaminergic neurotransmission identifies association with DRD1. World J Biol Psychiatry. 2012. April;13(4):281–92. [DOI] [PubMed] [Google Scholar]
  • 56.Ford CP. The role of D2-autoreceptors in regulating dopamine neuron activity and transmission. Neuroscience. 2014. December 12;282:13–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Thompson J, Thomas N, Singleton A, Piggott M, Lloyd S, Perry EK, et al. D2 dopamine receptor gene (DRD2) Taq1 A polymorphism: reduced dopamine D2 receptor binding in the human striatum associated with the A1 allele. Pharmacogenetics. 1997. December;7(6):479–84. [DOI] [PubMed] [Google Scholar]
  • 58.Pohjalainen T, Rinne JO, Nagren K, Lehikoinen P, Anttila K, Syvalahti EK, et al. The A1 allele of the human D2 dopamine receptor gene predicts low D2 receptor availability in healthy volunteers. Mol Psychiatry. 1998. May;3(3):256–60. [DOI] [PubMed] [Google Scholar]
  • 59.Zhang Y, Bertolino A, Fazio L, Blasi G, Rampino A, Romano R, et al. Polymorphisms in human dopamine D2 receptor gene affect gene expression, splicing, and neuronal activity during working memory. Proc Natl Acad Sci U S A. 2007. December 18;104(51):20552–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Wu J, Xiao H, Sun H, Zou L, Zhu LQ. Role of dopamine receptors in ADHD: a systematic meta-analysis. Mol Neurobiol. 2012. June;45(3):605–20. [DOI] [PubMed] [Google Scholar]
  • 61.Jeanneteau F, Funalot B, Jankovic J, Deng H, Lagarde JP, Lucotte G, et al. A functional variant of the dopamine D3 receptor is associated with risk and age-at-onset of essential tremor. Proc Natl Acad Sci U S A. 2006. July 11;103(28):10753–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Fageera W, Sengupta SM, Labbe A, Grizenko N, Joober R. DRD3 Gene and ADHD: A Pharmaco-Behavioural Genetic Study. Neuromolecular Med. 2018. December;20(4):515–24. [DOI] [PubMed] [Google Scholar]
  • 63.Beaulieu JM, Gainetdinov RR. The physiology, signaling, and pharmacology of dopamine receptors. Pharmacol Rev. 2011. March;63(1):182–217. [DOI] [PubMed] [Google Scholar]
  • 64.Wong AH, Van Tol HH. The dopamine D4 receptors and mechanisms of antipsychotic atypicality. Prog Neuropsychopharmacol Biol Psychiatry. 2003. October;27(7):1091–9. [DOI] [PubMed] [Google Scholar]
  • 65.Pappa I, Mileva-Seitz VR, Bakermans-Kranenburg MJ, Tiemeier H, van IMH. The magnificent seven: A quantitative review of dopamine receptor d4 and its association with child behavior. Neurosci Biobehav Rev. 2015. October;57:175–86. [DOI] [PubMed] [Google Scholar]
  • 66.Naumova D, Grizenko N, Sengupta SM, Joober R. DRD4 exon 3 genotype and ADHD: Randomised pharmacodynamic investigation of treatment response to methylphenidate. World J Biol Psychiatry. 2017. December 15:1–10. [DOI] [PubMed] [Google Scholar]
  • 67.Ji HS, Paik KC, Park WS, Lim MH. No Association between the Response to Methylphenidate and DRD4 Gene Polymorphism in Korean Attention Deficit Hyperactivity Disorder: A Case Control Study. Clin Psychopharmacol Neurosci. 2013. April;11(1):13–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.D’Souza UM, Russ C, Tahir E, Mill J, McGuffin P, Asherson PJ, et al. Functional effects of a tandem duplication polymorphism in the 5’flanking region of the DRD4 gene. Biol Psychiatry. 2004. November 1;56(9):691–7. [DOI] [PubMed] [Google Scholar]
  • 69.Volkow ND, Fowler JS, Wang G, Ding Y, Gatley SJ. Mechanism of action of methylphenidate: insights from PET imaging studies. J Atten Disord. 2002;6 Suppl 1:S31–43. [DOI] [PubMed] [Google Scholar]
  • 70.Winsberg BG, Comings DE. Association of the dopamine transporter gene (DAT1) with poor methylphenidate response. J Am Acad Child Adolesc Psychiatry. 1999. December;38(12):1474–7. [DOI] [PubMed] [Google Scholar]
  • 71.Stein MA, Waldman I, Newcorn J, Bishop J, Kittles R, Cook EH Jr. Dopamine transporter genotype and stimulant dose-response in youth with attention-deficit/hyperactivity disorder. J Child Adolesc Psychopharmacol. 2014. June;24(5):238–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kambeitz J, Romanos M, Ettinger U. Meta-analysis of the association between dopamine transporter genotype and response to methylphenidate treatment in ADHD. Pharmacogenomics J. 2013. April 16. [DOI] [PubMed] [Google Scholar]
  • 73.Soleimani R, Salehi Z, Soltanipour S, Hasandokht T, Jalali MM. SLC6A3 polymorphism and response to methylphenidate in children with ADHD: A systematic review and meta-analysis. Am J Med Genet B Neuropsychiatr Genet. 2018. April;177(3):287–300. [DOI] [PubMed] [Google Scholar]
  • 74.Brookes KJ, Neale BM, Sugden K, Khan N, Asherson P, D’Souza UM. Relationship between VNTR polymorphisms of the human dopamine transporter gene and expression in post-mortem midbrain tissue. Am J Med Genet B Neuropsychiatr Genet. 2007. December 5;144B(8):1070–8. [DOI] [PubMed] [Google Scholar]
  • 75.Arcos-Burgos M, Jain M, Acosta MT, Shively S, Stanescu H, Wallis D, et al. A common variant of the latrophilin 3 gene, LPHN3, confers susceptibility to ADHD and predicts effectiveness of stimulant medication. Mol Psychiatry. 2010. November;15(11):1053–66. [DOI] [PubMed] [Google Scholar]
  • 76.Song J, Kim SW, Hong HJ, Lee MG, Lee BW, Choi TK, et al. Association of SNAP-25, SLC6A2, and LPHN3 with OROS methylphenidate treatment response in attention-deficit/hyperactivity disorder. Clin Neuropharmacol. 2014. Sep-Oct;37(5):136–41. [DOI] [PubMed] [Google Scholar]
  • 77.Bruxel EM, Salatino-Oliveira A, Akutagava-Martins GC, Tovo-Rodrigues L, Genro JP, Zeni CP, et al. LPHN3 and attention-deficit/hyperactivity disorder: a susceptibility and pharmacogenetic study. Genes Brain Behav. 2015. June;14(5):419–27. [DOI] [PubMed] [Google Scholar]
  • 78.Labbe A, Liu A, Atherton J, Gizenko N, Fortier ME, Sengupta SM, et al. Refining psychiatric phenotypes for response to treatment: contribution of LPHN3 in ADHD. Am J Med Genet B Neuropsychiatr Genet. 2012. October;159B(7):776–85. [DOI] [PubMed] [Google Scholar]
  • 79.Choudhry Z, Sengupta SM, Grizenko N, Fortier ME, Thakur GA, Bellingham J, et al. LPHN3 and attention-deficit/hyperactivity disorder: interaction with maternal stress during pregnancy. J Child Psychol Psychiatry. 2012. August;53(8):892–902. [DOI] [PubMed] [Google Scholar]
  • 80.Jain M, Velez JI, Acosta MT, Palacio LG, Balog J, Roessler E, et al. A cooperative interaction between LPHN3 and 11q doubles the risk for ADHD. Mol Psychiatry. 2012. July;17(7):741–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Fisher NM, Seto M, Lindsley CW, Niswender CM. Metabotropic Glutamate Receptor 7: A New Therapeutic Target in Neurodevelopmental Disorders. Front Mol Neurosci. 2018;11:387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Park S, Kim BN, Cho SC, Kim JW, Kim JI, Shin MS, et al. The metabotropic glutamate receptor subtype 7 rs3792452 polymorphism is associated with the response to methylphenidate in children with attention-deficit/hyperactivity disorder. J Child Adolesc Psychopharmacol. 2014. May;24(4):223–7. [DOI] [PubMed] [Google Scholar]
  • 83.Mick E, Neale B, Middleton FA, McGough JJ, Faraone SV. Genome-wide association study of response to methylphenidate in 187 children with attention-deficit/hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet. 2008. December 5;147B(8):1412–8. [DOI] [PubMed] [Google Scholar]
  • 84.Pagerols M, Richarte V, Sanchez-Mora C, Rovira P, Soler Artigas M, Garcia-Martinez I, et al. Integrative genomic analysis of methylphenidate response in attention-deficit/hyperactivity disorder. Sci Rep. 2018. January 30;8(1):1881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.National Center for Biotechnology Information. Database of Single Nucleotide Polymorphisms (dbSNP) U.S. National Library of Medicine.
  • 86.Xu F, Gainetdinov RR, Wetsel WC, Jones SR, Bohn LM, Miller GW, et al. Mice lacking the norepinephrine transporter are supersensitive to psychostimulants. Nat Neurosci. 2000. May;3(5):465–71. [DOI] [PubMed] [Google Scholar]
  • 87.Carboni E, Silvagni A, Vacca C, Di Chiara G. Cumulative effect of norepinephrine and dopamine carrier blockade on extracellular dopamine increase in the nucleus accumbens shell, bed nucleus of stria terminalis and prefrontal cortex. J Neurochem. 2006. January;96(2):473–81. [DOI] [PubMed] [Google Scholar]
  • 88.Madras BK, Miller GM, Fischman AJ. The dopamine transporter and attention-deficit/hyperactivity disorder. Biol Psychiatry. 2005. June 1;57(11):1397–409. [DOI] [PubMed] [Google Scholar]
  • 89.Berridge CW, Devilbiss DM, Andrzejewski ME, Arnsten AF, Kelley AE, Schmeichel B, et al. Methylphenidate preferentially increases catecholamine neurotransmission within the prefrontal cortex at low doses that enhance cognitive function. Biol Psychiatry. 2006. November 15;60(10):1111–20. [DOI] [PubMed] [Google Scholar]
  • 90.Jonsson EG, Nothen MM, Gustavsson JP, Neidt H, Bunzel R, Propping P, et al. Polymorphisms in the dopamine, serotonin, and norepinephrine transporter genes and their relationships to monoamine metabolite concentrations in CSF of healthy volunteers. Psychiatry Res. 1998. June 2;79(1):1–9. [DOI] [PubMed] [Google Scholar]
  • 91.Kim BN, Kim JW, Hong SB, Cho SC, Shin MS, Yoo HJ. Possible association of norepinephrine transporter −3081(A/T) polymorphism with methylphenidate response in attention deficit hyperactivity disorder. Behav Brain Funct. 2010. October 7;6:57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Lee SH, Kim SW, Lee MG, Yook KH, Greenhill LL, Fradin KN, et al. Lack of association between response of OROS-methylphenidate and norepinephrine transporter (SLC6A2) polymorphism in Korean ADHD. Psychiatry Res. 2011. April 30;186(2–3):338–44. [DOI] [PubMed] [Google Scholar]
  • 93.Angyal N, Horvath EZ, Tarnok Z, Richman MJ, Bognar E, Lakatos K, et al. Association analysis of norepinephrine transporter polymorphisms and methylphenidate response in ADHD patients. Prog Neuropsychopharmacol Biol Psychiatry. 2018. June 8;84(Pt A):122–8. [DOI] [PubMed] [Google Scholar]
  • 94.Song J, Song DH, Jhung K, Cheon KA. Norepinephrine transporter gene (SLC6A2) is involved with methylphenidate response in Korean children with attention deficit hyperactivity disorder. Int Clin Psychopharmacol. 2011. March;26(2):107–13. [DOI] [PubMed] [Google Scholar]
  • 95.Park MH, Kim JW, Yang YH, Hong SB, Park S, Kang H, et al. Regional brain perfusion before and after treatment with methylphenidate may be associated with the G1287A polymorphism of the norepinephrine transporter gene in children with attention-deficit/hyperactivity disorder. Neurosci Lett. 2012. April 18;514(2):159–63. [DOI] [PubMed] [Google Scholar]
  • 96.Sigurdardottir HL, Kranz GS, Rami-Mark C, James GM, Vanicek T, Gryglewski G, et al. Effects of norepinephrine transporter gene variants on NET binding in ADHD and healthy controls investigated by PET. Hum Brain Mapp. 2016. March;37(3):884–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Thakur GA, Sengupta SM, Grizenko N, Choudhry Z, Joober R. Comprehensive phenotype/genotype analyses of the norepinephrine transporter gene (SLC6A2) in ADHD: relation to maternal smoking during pregnancy. PloS one. 2012;7(11):e49616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Hu C, Chen W, Myers SJ, Yuan H, Traynelis SF. Human GRIN2B variants in neurodevelopmental disorders. J Pharmacol Sci. 2016. October;132(2):115–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Wanchoo SJ, Swann AC, Dafny N. Descending glutamatergic pathways of PFC are involved in acute and chronic action of methylphenidate. Brain Res. 2009. December 8;1301:68–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.dbSNP Short Genetic Variations rs2284411 Reference SNP Report. [cited 2019 August 15]; Available from: https://www.ncbi.nlm.nih.gov/snp/rs2284411#publications
  • 101.Kim JI, Kim JW, Park JE, Park S, Hong SB, Han DH, et al. Association of the GRIN2B rs2284411 polymorphism with methylphenidate response in attention-deficit/hyperactivity disorder. J Psychopharmacol. 2017. August;31(8):1070–7. [DOI] [PubMed] [Google Scholar]
  • 102.Fishbein DH, Lozovsky D, Jaffe JH. Impulsivity, aggression, and neuroendocrine responses to serotonergic stimulation in substance abusers. Biol Psychiatry. 1989. April 15;25(8):1049–66. [DOI] [PubMed] [Google Scholar]
  • 103.Lesch KP, Bengel D, Heils A, Sabol SZ, Greenberg BD, Petri S, et al. Association of anxiety-related traits with a polymorphism in the serotonin transporter gene regulatory region. Science. 1996. November 29;274(5292):1527–31. [DOI] [PubMed] [Google Scholar]
  • 104.Hanna GL, Himle JA, Curtis GC, Koram DQ, Veenstra-VanderWeele J, Leventhal BL, et al. Serotonin transporter and seasonal variation in blood serotonin in families with obsessive-compulsive disorder. Neuropsychopharmacology. 1998. February;18(2):102–11. [DOI] [PubMed] [Google Scholar]
  • 105.Thakur GA, Grizenko N, Sengupta SM, Schmitz N, Joober R. The 5-HTTLPR polymorphism of the serotonin transporter gene and short term behavioral response to methylphenidate in children with ADHD. BMC psychiatry. 2010;10:50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Park SY, Kim EJ, Cheon KA. Association Between 5-HTTLPR Polymorphism and Tics after Treatment with Methylphenidate in Korean Children with Attention-Deficit/Hyperactivity Disorder. J Child Adolesc Psychopharmacol. 2015. October;25(8):633–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Zeni CP, Guimaraes AP, Polanczyk GV, Genro JP, Roman T, Hutz MH, et al. No significant association between response to methylphenidate and genes of the dopaminergic and serotonergic systems in a sample of Brazilian children with attention-deficit/hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet. 2007. April 5;144(3):391–4. [DOI] [PubMed] [Google Scholar]
  • 108.Tharoor H, Lobos EA, Todd RD, Reiersen AM. Association of dopamine, serotonin, and nicotinic gene polymorphisms with methylphenidate response in ADHD. Am J Med Genet B Neuropsychiatr Genet. 2008. June 5;147B(4):527–30. [DOI] [PubMed] [Google Scholar]
  • 109.Seeger G, Schloss P, Schmidt MH. Marker gene polymorphisms in hyperkinetic disorder--predictors of clinical response to treatment with methylphenidate? Neurosci Lett. 2001. November 2;313(1–2):45–8. [DOI] [PubMed] [Google Scholar]
  • 110.Schiavo G, Stenbeck G, Rothman JE, Sollner TH. Binding of the synaptic vesicle v-SNARE, synaptotagmin, to the plasma membrane t-SNARE, SNAP-25, can explain docked vesicles at neurotoxin-treated synapses. Proc Natl Acad Sci U S A. 1997. February 4;94(3):997–1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.McGough J, McCracken J, Swanson J, Riddle M, Kollins S, Greenhill L, et al. Pharmacogenetics of methylphenidate response in preschoolers with ADHD. J Am Acad Child Adolesc Psychiatry. 2006. November;45(11):1314–22. [DOI] [PubMed] [Google Scholar]
  • 112.Nagatsu T, Levitt M, Udenfriend S. Conversion of L-tyrosine to 3,4-dihydroxyphenylalanine by cell-free preparations of brain and sympathetically innervated tissues. Biochem Biophys Res Commun. 1964;14:543–9. [DOI] [PubMed] [Google Scholar]
  • 113.Nagatsu T, Levitt M, Udenfriend S. Tyrosine Hydroxylase. The Initial Step in Norepinephrine Biosynthesis. J Biol Chem. 1964. September;239:2910–7. [PubMed] [Google Scholar]
  • 114.Wang L, Li B, Lu X, Zhao Q, Li Y, Ge D, et al. A functional intronic variant in the tyrosine hydroxylase (TH) gene confers risk of essential hypertension in the Northern Chinese Han population. Clin Sci (Lond). 2008. September;115(5):151–8. [DOI] [PubMed] [Google Scholar]
  • 115.Hohman TJ, Bush WS, Jiang L, Brown-Gentry KD, Torstenson ES, Dudek SM, et al. Discovery of gene-gene interactions across multiple independent data sets of late onset Alzheimer disease from the Alzheimer Disease Genetics Consortium. Neurobiol Aging. 2016. February;38:141–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Lesch KP, Timmesfeld N, Renner TJ, Halperin R, Roser C, Nguyen TT, et al. Molecular genetics of adult ADHD: converging evidence from genome-wide association and extended pedigree linkage studies. J Neural Transm (Vienna). 2008. November;115(11):1573–85. [DOI] [PubMed] [Google Scholar]
  • 117.Lyon GJ, Jiang T, Van Wijk R, Wang W, Bodily PM, Xing J, et al. Exome sequencing and unrelated findings in the context of complex disease research: ethical and clinical implications. Discov Med. 2011. July;12(62):41–55. [PMC free article] [PubMed] [Google Scholar]
  • 118.Lugowska A, Mierzewska H, Bekiesinska-Figatowska M, Szczepanik E, Goszczanska-Ciuchta A, Bednarska-Makaruk M. A homozygote for the c.459+1G>A mutation in the ARSA gene presents with cerebellar ataxia as the only first clinical sign of metachromatic leukodystrophy. J Neurol Sci. 2014. March 15;338(1–2):214–7. [DOI] [PubMed] [Google Scholar]
  • 119.Steingard R, Taskiran S, Connor DF, Markowitz JS, Stein MA. New Formulations of Stimulants: An Update for Clinicians. J Child Adolesc Psychopharmacol. 2019. June;29(5):324–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Wu D, Otton SV, Inaba T, Kalow W, Sellers EM. Interactions of amphetamine analogs with human liver CYP2D6. Biochem Pharmacol. 1997. June 1;53(11):1605–12. [DOI] [PubMed] [Google Scholar]
  • 121.Gehlert DR, Gackenheimer SL, Robertson DW. Localization of rat brain binding sites for [3H]tomoxetine, an enantiomerically pure ligand for norepinephrine reuptake sites. Neurosci Lett. 1993. July 23;157(2):203–6. [DOI] [PubMed] [Google Scholar]
  • 122.Wong DT, Threlkeld PG, Best KL, Bymaster FP. A new inhibitor of norepinephrine uptake devoid of affinity for receptors in rat brain. J Pharmacol Exp Ther. 1982. July;222(1):61–5. [PubMed] [Google Scholar]
  • 123.Zhou SF, Liu JP, Chowbay B. Polymorphism of human cytochrome P450 enzymes and its clinical impact. Drug Metab Rev. 2009;41(2):89–295. [DOI] [PubMed] [Google Scholar]
  • 124.Ring BJ, Gillespie JS, Eckstein JA, Wrighton SA. Identification of the human cytochromes P450 responsible for atomoxetine metabolism. Drug Metab Dispos. 2002. March;30(3):319–23. [DOI] [PubMed] [Google Scholar]
  • 125.Yu G, Li GF, Markowitz JS. Atomoxetine: A Review of Its Pharmacokinetics and Pharmacogenomics Relative to Drug Disposition. J Child Adolesc Psychopharmacol. 2016. May;26(4):314–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Ramoz N, Boni C, Downing AM, Close SL, Peters SL, Prokop AM, et al. A haplotype of the norepinephrine transporter (Net) gene Slc6a2 is associated with clinical response to atomoxetine in attention-deficit hyperactivity disorder (ADHD). Neuropsychopharmacology. 2009. August;34(9):2135–42. [DOI] [PubMed] [Google Scholar]
  • 127.Michelson D, Read HA, Ruff DD, Witcher J, Zhang S, McCracken J. CYP2D6 and clinical response to atomoxetine in children and adolescents with ADHD. J Am Acad Child Adolesc Psychiatry. 2007. February;46(2):242–51. [DOI] [PubMed] [Google Scholar]
  • 128.Trzepacz PT, Williams DW, Feldman PD, Wrishko RE, Witcher JW, Buitelaar JK. CYP2D6 metabolizer status and atomoxetine dosing in children and adolescents with ADHD. Eur Neuropsychopharmacol. 2008. February;18(2):79–86. [DOI] [PubMed] [Google Scholar]
  • 129.Brown JT, Bishop JR, Sangkuhl K, Nurmi EL, Mueller DJ, Dinh JC, et al. Clinical Pharmacogenetics Implementation Consortium Guideline for Cytochrome P450 (CYP)2D6 Genotype and Atomoxetine Therapy. Clin Pharmacol Ther. 2019. July;106(1):94–102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Fang Y, Ji N, Cao Q, Su Y, Chen M, Wang Y, et al. Variants of Dopamine Beta Hydroxylase Gene Moderate Atomoxetine Response in Children with Attention-Deficit/Hyperactivity Disorder. J Child Adolesc Psychopharmacol. 2015. October;25(8):625–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Easton N, Steward C, Marshall F, Fone K, Marsden C. Effects of amphetamine isomers, methylphenidate and atomoxetine on synaptosomal and synaptic vesicle accumulation and release of dopamine and noradrenaline in vitro in the rat brain. Neuropharmacology. 2007. February;52(2):405–14. [DOI] [PubMed] [Google Scholar]
  • 132.Yang L, Qian Q, Liu L, Li H, Faraone SV, Wang Y. Adrenergic neurotransmitter system transporter and receptor genes associated with atomoxetine response in attention-deficit hyperactivity disorder children. J Neural Transm (Vienna). 2013. July;120(7):1127–33. [DOI] [PubMed] [Google Scholar]
  • 133.Kim CH, Kim HS, Cubells JF, Kim KS. A previously undescribed intron and extensive 5’ upstream sequence, but not Phox2a-mediated transactivation, are necessary for high level cell type-specific expression of the human norepinephrine transporter gene. J Biol Chem. 1999. March 5;274(10):6507–18. [DOI] [PubMed] [Google Scholar]
  • 134.Briars L, Todd T. A Review of Pharmacological Management of Attention-Deficit/Hyperactivity Disorder. J Pediatr Pharmacol Ther. 2016. May-Jun;21(3):192–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Claessens AJ, Risler LJ, Eyal S, Shen DD, Easterling TR, Hebert MF. CYP2D6 mediates 4-hydroxylation of clonidine in vitro: implication for pregnancy-induced changes in clonidine clearance. Drug Metab Dispos. 2010. September;38(9):1393–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Polanczyk G, Faraone SV, Bau CH, Victor MM, Becker K, Pelz R, et al. The impact of individual and methodological factors in the variability of response to methylphenidate in ADHD pharmacogenetic studies from four different continents. Am J Med Genet B Neuropsychiatr Genet. 2008. December 5;147B(8):1419–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Rosenberg NA, Huang L, Jewett EM, Szpiech ZA, Jankovic I, Boehnke M. Genome-wide association studies in diverse populations. Nat Rev Genet. 2010. May;11(5):356–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Bruxel EM, Akutagava-Martins GC, Salatino-Oliveira A, Contini V, Kieling C, Hutz MH, et al. ADHD pharmacogenetics across the life cycle: New findings and perspectives. Am J Med Genet B Neuropsychiatr Genet. 2014. June;165(4):263–82. [DOI] [PubMed] [Google Scholar]
  • 139.Wehry AM, Ramsey L, Dulemba SE, Mossman SA, Strawn JR. Pharmacogenomic Testing in Child and Adolescent Psychiatry: An Evidence-Based Review. Curr Probl Pediatr Adolesc Health Care. 2018. February;48(2):40–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Children’s Oncology Group. 2019. [cited 2019 December 10]; Available from: https://childrensoncologygroup.org/
  • 141.Duclot F, Kabbaj M. Epigenetic mechanisms underlying the role of brain-derived neurotrophic factor in depression and response to antidepressants. J Exp Biol. 2015. January 1;218(Pt 1):21–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Menke A, Binder EB. Epigenetic alterations in depression and antidepressant treatment. Dialogues Clin Neurosci. 2014. September;16(3):395–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tadic A, Muller-Engling L, Schlicht KF, Kotsiari A, Dreimuller N, Kleimann A, et al. Methylation of the promoter of brain-derived neurotrophic factor exon IV and antidepressant response in major depression. Mol Psychiatry. 2014. March;19(3):281–3. [DOI] [PubMed] [Google Scholar]
  • 144.Lopez JP, Mamdani F, Labonte B, Beaulieu MM, Yang JP, Berlim MT, et al. Epigenetic regulation of BDNF expression according to antidepressant response. Mol Psychiatry. 2013. April;18(4):398–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Yehuda R, Daskalakis NP, Desarnaud F, Makotkine I, Lehrner AL, Koch E, et al. Epigenetic Biomarkers as Predictors and Correlates of Symptom Improvement Following Psychotherapy in Combat Veterans with PTSD. Front Psychiatry. 2013;4:118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Tang H, Dalton CF, Srisawat U, Zhang ZJ, Reynolds GP. Methylation at a transcription factor-binding site on the 5-HT1A receptor gene correlates with negative symptom treatment response in first episode schizophrenia. Int J Neuropsychopharmacol. 2014. April;17(4):645–9. [DOI] [PubMed] [Google Scholar]
  • 147.Powell TR, Smith RG, Hackinger S, Schalkwyk LC, Uher R, McGuffin P, et al. DNA methylation in interleukin-11 predicts clinical response to antidepressants in GENDEP. Transl Psychiatry. 2013. September 3;3:e300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Domschke K, Tidow N, Schwarte K, Deckert J, Lesch KP, Arolt V, et al. Serotonin transporter gene hypomethylation predicts impaired antidepressant treatment response. Int J Neuropsychopharmacol. 2014. August;17(8):1167–76. [DOI] [PubMed] [Google Scholar]
  • 149.Burghardt KJ, Goodrich JM, Dolinoy DC, Ellingrod VL. Gene-specific DNA methylation may mediate atypical antipsychotic-induced insulin resistance. Bipolar Disord. 2016. August;18(5):423–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Ovenden ES, McGregor NW, Emsley RA, Warnich L. DNA methylation and antipsychotic treatment mechanisms in schizophrenia: Progress and future directions. Prog Neuropsychopharmacol Biol Psychiatry. 2018. February 2;81:38–49. [DOI] [PubMed] [Google Scholar]
  • 151.Sugawara H, Bundo M, Ishigooka J, Iwamoto K, Kato T. Epigenetic regulation of serotonin transporter in psychiatric disorders. J Genet Genomics. 2013. July 20;40(7):325–9. [DOI] [PubMed] [Google Scholar]
  • 152.Abdolmaleky HM, Pajouhanfar S, Faghankhani M, Joghataei MT, Mostafavi A, Thiagalingam S. Antipsychotic drugs attenuate aberrant DNA methylation of DTNBP1 (dysbindin) promoter in saliva and post-mortem brain of patients with schizophrenia and Psychotic bipolar disorder. Am J Med Genet B Neuropsychiatr Genet. 2015. December;168(8):687–96. [DOI] [PubMed] [Google Scholar]
  • 153.Rukova B, Staneva R, Hadjidekova S, Stamenov G, Milanova V, Toncheva D. Whole genome methylation analyses of schizophrenia patients before and after treatment. Biotechnol Biotechnol Equip. 2014. May 4;28(3):518–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Quinones MP, Kaddurah-Daouk R. Metabolomics tools for identifying biomarkers for neuropsychiatric diseases. Neurobiology of disease. 2009. August;35(2):165–76. [DOI] [PubMed] [Google Scholar]
  • 155.Kaddurah-Daouk R, McEvoy J, Baillie RA, Lee D, Yao JK, Doraiswamy PM, et al. Metabolomic mapping of atypical antipsychotic effects in schizophrenia. Mol Psychiatry. 2007. October;12(10):934–45. [DOI] [PubMed] [Google Scholar]
  • 156.Abo R, Hebbring S, Ji Y, Zhu H, Zeng ZB, Batzler A, et al. Merging pharmacometabolomics with pharmacogenomics using ‘1000 Genomes’ single-nucleotide polymorphism imputation: selective serotonin reuptake inhibitor response pharmacogenomics. Pharmacogenetics and genomics. 2012. April;22(4):247–53. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental Table 1
Supplemental Table 2
Supplemental Table 3
Supplemental Table 4
Supplemental Table 5
Supplemental Table 6

RESOURCES