Dual-Specificity Phosphatases in Mental and Neurological Disorders

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Progress in Neurobiology xxx (xxxx) xxx

Contents lists available at ScienceDirect

Progress in Neurobiology
journal homepage: www.elsevier.com/locate/pneurobio

Review article

Dual-specificity phosphatases in mental and neurological disorders


Ning An a, c, Katherine Bassil a, c, Ghazi I. Al Jowf a, b, c, Harry W.M. Steinbusch a, c,
Markus Rothermel c, d, Laurence de Nijs a, c, 1, Bart P.F. Rutten a, c, *, 1
a
Department of Psychiatry and Neuropsychology, School for Mental Health and Neuroscience (MHeNs), Faculty of Health, Medicine and Life Sciences, Maastricht
University, Maastricht, the Netherlands
b
College of Applied Medical Sciences, Department of Public Health, King Faisal University, Al-Ahsa, Saudi Arabia
c
European Graduate School of Neuroscience, Maastricht University, Maastricht, the Netherlands
d
Department of Chemosensation – AG Neuromodulation, RWTH Aachen University, Aachen, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: The dual-specificity phosphatase (DUSP) family includes a heterogeneous group of protein phosphatases that
DUSP dephosphorylate both phospho-tyrosine and phospho-serine/phospho-threonine residues within a single sub­
Mental health strate. These protein phosphatases have many substrates and modulate diverse neural functions, such as neu­
Neurological diseases
rogenesis, differentiation, and apoptosis. DUSP genes have furthermore been associated with mental disorders
Phosphatase
Psychiatric disorders
such as depression and neurological disorders such as Alzheimer’s disease. Herein, we review the current
literature on the DUSP family of genes concerning mental and neurological disorders. This review i) outlines the
structure and general functions of DUSP genes, and ii) overviews the literature on DUSP genes concerning mental
and neurological disorders, including model systems, while furthermore providing perspectives for future
research.

1. Introduction mental and neurological disorders is associated with an interplay be­


tween genetic predisposition and epigenetic mechanisms that are
Mental and neurological disorders are becoming highly prevalent, influenced by environmental exposures throughout life (Klengel and
and in turn, exerting significant socioeconomic and personal burdens (; Binder, 2015; Nestler et al., 2016). Developments in the fields of psy­
Wittchen et al., 2011). For instance, mental disorders are often associ­ chiatric genetics and epigenetics have enabled the first wave of
ated with a reduced quality of life and also reduced maximal life ex­ genome-wide analyses on groups of people diagnosed with specific
pectancy (Chesney et al., 2014), possibly due to the early onset of these mental and neurological disorders compared to control populations.
disorders, the impact on social interactions and their frequent occur­ These analyses have provided critical initial insights, suggesting the
rence as a comorbidity with other physiological health issues, i.e., car­ involvement of distinct genes and biological processes in the onset and
diovascular, metabolic and immune-related disorders (Wykes et al., course of mental and neurological disorders. That being said, increasing
2015). evidence from a range of genetic and epigenetic studies have identified
Recent progress in the study of mental and neurological disorders has associations between several genes in the dual-specificity phosphatases
allowed for a better understanding of critical underlying determinants, (DUSP) family and a variety of mental and neurological disorders (Duric
as well as associated molecular and cellular mechanisms. The etiology of et al., 2010; Perez-Sen et al., 2019).

Abbreviations: AD, Alzheimer’s disease; ADF, actin-depolymerizing factor; ASDs, autism spectrum disorders; ATM, ataxia-telangiectasia mutated; Aβ, amyloid-
beta; APP, amyloid precursor protein; CDC14, cell division cycle 14 phosphatases; CUS, chronic unpredictable stress; DG, dentate gyrus; DUSPs, dual-specificity
phosphatases; ECS, electroconvulsive seizure; ERK, extracellular signal-regulated kinase; HD, Huntington’s disease; JNK, c-Jun amino-terminal kinase; LBs, lafora
bodies; LD, lafora disease; LTP, long-term potentiation; MAPK, MAP kinases; MKPs, mitogen-activated protein kinase phosphatases; PD, Parkinson’s disease; PI3K,
phosphatidylinositol 3-kinase; PRL, phosphatases of regenerating liver; PTEN, phosphatase and tensin homologs deleted on chromosome 10; PTP, protein tyrosine
phosphatase; PTSD, posttraumatic stress disorder; SSH, slingshot homolog; SNO-PTEN, S-nitrosylation of PTEN; 6-OHDA, 6-hydroxydopamine.
* Corresponding author at: Department of Psychiatry and Neuropsychology, School for Mental Health and Neuroscience (MHeNs), Faculty of Health, Medicine and
Life Sciences, Maastricht University, P.O. Box 616, 6200 MD Maastricht, the Netherlands.
E-mail address: b.rutten@maastrichtuniversity.nl (B.P.F. Rutten).
1
Equal contributions.

https://doi.org/10.1016/j.pneurobio.2020.101906
Received 21 August 2019; Received in revised form 26 August 2020; Accepted 1 September 2020
Available online 6 September 2020
0301-0082/© 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Please cite this article as: Ning An, Progress in Neurobiology, https://doi.org/10.1016/j.pneurobio.2020.101906
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

These groups of phosphatases are characterized by the removal of a Table 1


phosphorus group from both phospho-tyrosine and phospho-serine/ DUSP Protein or RNA expression in animal and human brain.
phospho-threonine residues within a single substrate, leading to Isoforms Expression Protein/ Species Ref.
conformational changes in proteins, a process that can be reversed by mRNA
kinase phosphorylation. Additionally, these protein phosphatases have a SSH1 Frontal cortex, Protein Human (Barone et al.,
variety of substrates and as such can modulate diverse cellular functions, cerebellum 2014; Zafar et al.,
such as neurogenesis, neuronal differentiation, and apoptosis via three 2017)
main signaling pathways, including MAP kinase pathways (Patterson Primary cortical Protein Mice (Barone et al.,
neurons 2014)
et al., 2009), Phosphatidylinositol 3-kinase (PI3K)/AKT (Carracedo and Dorsal root Protein Chick (Endo et al., 2003)
Pandolfi, 2008), and BDNF/ p75NTR pathways (Finelli et al., 2013; ganglion neurons
Perez-Sen et al., 2019).
While a range of studies on DUSP genes has linked these genes to SSH2 Brain, thalamus, mRNA Mice (Kousaka et al.,
cancer (Bermudez et al., 2010; Low and Zhang, 2016) and disorders of hippocampus 2008)

the immune system (Lang et al., 2006), accumulating evidence for links
SSH3 Hippocampus, mRNA Human (Wang and Wang,
between DUSP genes and mental disorders, such as depression and bi­ temporal gyrus, 2020)
polar disorder, as well as neurological disorders such as Alzheimer’s frontal gyrus
disease, is increasingly documented (Rios et al., 2014). Prefrontal cortex, mRNA Mice (van Heerden
Herein, we aim to summarize, converge, and critically review the hypothalamus et al., 2009)
Amygdala mRNA Mice (McCullough
current literature on the DUSP family of genes implicated in mental and
et al., 2018)
neurological disorders.
(Dumaual et al.,
Cerebellum mRNA Human
2. Structure, expression, and function of DUSP family members 2006)
PRL1
Neuronal lineages
mRNA Zebrafish (Lin et al., 2013)
in the embryo.
DUSPs’ primary mode of action is the dephosphorylation of tyrosine
and/or serine/threonine residues and the resulting activity regulation of
Cerebellum, (Dumaual et al.,
their substrates. The physiological outcomes of DUSPs’ functions thus cerebral cortex
mRNA Human
2006)
PRL2
hinge on their substrate specificity and phosphatase activity, however, Neuronal lineages
mRNA Zebrafish (Lin et al., 2013)
the substrates for DUSPs are not precisely defined. The archetypical in an embryo.
DUSP, DUSP1/MKP1, was initially discovered to regulate the activities
(Sebollela et al.,
of MAP kinases by dephosphorylating the TXY motif in the kinase CDC14A Cortical slices mRNA Human
2012)
domain. However, although DUSPs were discovered more than a decade (www.protein
ago, on Wittchen et al., 2011ly in the past few years have their various Cerebral cortex Protein Human
atlas.org)
functions begun to be described. DUSPs can be categorized based on the
presence or absence of a MAP kinase-interacting domain into typical hCDC14Bpar
Adult and fetal
mRNA Human
(Rosso et al.,
DUSPs and atypical DUSPs, respectively. A subset of DUSPs contains an brain 2008)

N-terminal region composed of two CDC25 homology 2 (CH2) domains


Hippocampus,
and an intervening cluster of basic amino acids known as the MAP prefrontal cortex, (Rosso et al.,
kinase-binding (MKB) motif. All DUSP proteins contain a common hCDC14B1 mRNA Human
amygdala, 2008)
phosphatase domain which consists of conserved Asp, Cys, and Arg hypothalamus
residues forming the catalytic site (Huang and Tan, 2012).
The subcategorization of DUSPs into subgroups is supported by the Hippocampus,
prefrontal cortex, (Rosso et al.,
phylogenetic tree of DUSP sequences similarities with consideration of hCDC14B2
amygdala,
mRNA Human
2008)
substrate preferences (Huang and Tan, 2012). Based on this, the DUSP hypothalamus
family members can be categorized into the following subgroups: 1)
slingshot homolog (SSH) family of phosphatases, 2) phosphatases of regen­ Hippocampus,
erating liver (PRL) family, 3) cell division cycle 14 (CDC14) phosphatases, prefrontal cortex, (Rosso et al.,
hCDC14B3 mRNA Human
amygdala, 2008)
4) phosphatase and tensin homologs deleted on chromosome 10 (PTEN), 5)
hypothalamus
myotubularins, 6) mitogen-activated protein kinase phosphatases (MKPs)
and 7) atypical DUSPs (Patterson et al., 2009). DUSP family members Adult brain and
and synonyms can be found in Table1 of the Supplementary materials, embryonic
while atypical DUSPs are presented in Supplementary Table 2. hCDC14C
forebrain,
mRNA Human
(Rosso et al.,
including the 2008)
In the following paragraphs, we will review the current knowledge
dorsal
about the structure, tissue expression and function of the DUSP family telencephalon
members. A summary of DUSP Protein or RNA expression in animal and
human brains is presented in Table 1, while their expression in non- Cerebral cortex,
(www.protein
brain tissues is shown in Supplementary Table 3. The function of cerebellum, Protein Human
atlas.org)
hippocampus
DUSP proteins is summarized in Table 2.
Anterior olfactory
nucleus, cerebral
2.1. Slingshot homolog (SSH) family of phosphatases cortex,
PTENs amygdaloid
2.1.1. Structure and homologues nucleus,
Protein Rat (Cai et al., 2009)
hippocampus,
Slingshot homolog (SSH), encoded by the SSH gene, was initially Purkinje’s cells,
recognized as a cofilin phosphatase in genetic studies performed in basal ganglia,
Drosophila (Jung et al., 2007). Cofilin is an actin-depolymerizing factor thalamus,
(ADF) that is abundant in human and other mammalian brains (Mina­ midbrain, pons.
mide et al., 2000). In mammals, three SSH homologs have been (continued on next page)

2
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Table 1 (continued ) Table 1 (continued )


Isoforms Expression Protein/ Species Ref. Isoforms Expression Protein/ Species Ref.
mRNA mRNA

Ventral prefrontal Protein Human (Karege et al., callosum,


cortex 2011) hippocampus,
Layer III temporal Protein Human (Rickle et al., hypothalamus,
cortex 2006) substantia nigra,
Frontal cortex Protein Human (Zhang et al., subthalamic
2006) nucleus, and
thalamus
Myotubularin Spinal cord and mRNA Human (Raess et al., MTMR2 Cortex, the Protein/ Rat (Lee et al., 2010)
substantia nigra 2017) hippocampus, mRNA
Cerebral cortex, Protein Human (www.protein cerebellum,
hippocampus, and atlas.org)
cerebellum MTMR5 Brain, the sciatic Protein Mice (Mammel et al.,
nerve 2019)
MKP1 Hippocampus Protein/ Human/ (Duric et al.,
mRNA Mice 2010) Laforin Cerebellum, mRN Mice (Ganesh et al.,
(DUSP1) Caudate mRNA Human (Taylor et al., (EPM2A) hippocampus, 2001)
2013) cerebral cortex,
Temporal cortex Protein Human (Du et al., 2019) and the olfactory
Cerebral cortex, Protein Human (Dwivedi et al., bulb.
cerebellum 2001)
Striatum mRNA Mice (Taylor et al., DUSP12 Hippocampus mRNA Human (Duric et al.,
2013) 2010)
Striatum, mRNA Rat (Ujike et al.,
thalamus, cortex 2002) DUSP14 Retinal ganglion mRNA Rat (Galvao et al.,
Dorsolateral mRNA Human (Renton, 2010) cells 2018)
prefrontal cortex
Neocortex, lateral Protein Rat (Gass et al., 1996)
DUSP15 Schwann cell mRNA Rat (Rodriguez-
nucleus of the bed
Molina et al.,
of the stria
2017)
terminalis
Myelinating mRNA Rat (Muth et al.,
Oligodendrocytes 2016)
MKP2 Cerebellar vermis Protein Human (Kyosseva et al.,
1999)
DUSP18 Fetal brain mRNA Human (Wu et al., 2003b)
(DUSP4) Cortex, Protein Human (Dwivedi et al.,
hippocampus, 2001)
cerebellum DUSP19 Hippocampus mRNA Human (Duric et al.,
2010)
ventral subiculum mRNA Human (Labonte et al.,
The medial mRNA Rat (Kodama et al.,
2017)
prefrontal cortex, 2005)
lateral frontal
cortex, parietal DUSP22 Hippocampus Protein Human (Sanchez-Mut
cortex, et al., 2014)
hippocampus
DUSP23 Nucleus mRNA Human (Labonte et al.,
MKP3 The medial mRNA Rat (Kodama et al., accumbens, the 2017)
(DUSP6) prefrontal cortex, 2005) Broadman area
lateral frontal
cortex, parietal DUSP26 Hippocampus Protein/ Human (Duric et al.,
cortex, mRNA 2010; Jung et al.,
hippocampus 2016)
Ventromedial PFC mRNA Human/ (Labonte et al.,
Mice 2017)
Striatum, cortex, mRNA Rat (Ujike et al., identified so far and denoted as SSH-1, SSH-2, and SSH-3 (Jung et al.,
hippocampus, 2002) 2007).
The sequence alignment of the three SSH phosphatases shows more
PAC1 Hippocampus Protein/ Mice (Martin-Segura
than 80 % sequence similarity (Jung et al., 2007). For instance, SSH
(DUSP2) mRNA et al., 2019)
Hippocampus mRNA Human (Duric et al., proteins from human, mouse, and Drosophila possess three highly
2010a) conserved domains, A, B, and P (a phosphatase domain) in the N-ter­
minal region (Ohta et al., 2003). Amino acid sequences of the P domain
Hvhr3 Brain (not mRNA Mice (Kutty et al., of the SSH family are distantly related to those of a family of MKPs and
(DUSP5) specified) 2016)
share a Dual Specific Phosphatase (DSP) active site (HCxxGxxR)
Nucleus mRNA Mice (Seleman et al.,
accumbens 2014)
conserved within both DSP and protein tyrosine phosphatases (PTP)
The medial mRNA Rat (Kuntz-Melcavage (Barford et al., 1998; Niwa et al., 2002). The short serine-rich sequence
prefrontal cortex et al., 2009) motif (S domain) is conserved only in SSH-1 and SSH-2 in mouse and
human but not in SSH-3 of the mouse, human, or Drosophila SSH protein
MKP-X Whole-brain mRNA Mice (Sokolov et al., (Fig. 1A) (Ohta et al., 2003).
(DUSP7) 2003)

Hvh5 Amygdala, caudate mRNA Human (Martell et al.,


2.1.2. Expression and function
(DUSP8) nucleus, the corpus 1995b) At the cellular level, human SSH (hSSH)-1 is primarily expressed in
the plasma membrane, cytosol, and nucleoplasm, while hSSH-2 and
hSSH-3 reside in the cytoplasm [the Human Protein Atlas (www.

3
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Table 2
DUSP genes in normal and pathological brain function.
Isoforms Function in Normal Brain Function in Pathological Brain

SSH1 along with Cofilin1 promoted dynamic changes in the


cytoskeleton needed for axon engagement like growth cone SSH1 reduced and remained inactive as Cofilin1 in the frontal cortex of
collapse and neurite outgrowth, and myelination in Schwann sporadic AD in human (Barone et al., 2014).
SSH1 cells in rat (Hsieh et al., 2006; Sparrow et al., 2012).
SSH1 increases growth cone motility and extension, and the
SSH1 reduced protein level and became inactive as Cofilin1 in the brain
growth cone becomes slender and branchy in chick (Endo et al.,
of APP/PS1 mice model of AD (Barone et al., 2014).
2003).

SSH3 mRNA is differentially regulated gene in hippocampus tissue,


SSH3 temporal gyrus tissue, frontal gyrus tissue and whole blood in patients
diagnosed with AD (Wang and Wang, 2020).
SSH3 mRNA has been found to be downregulated in central amygdala
Drd2-expressing population following foot shock fear conditioning
compared to controls in mice (McCullough et al., 2018).
SSH3 mRNA is over expressed in prefrontal cortex, and hypothalamus
in mice subjected to maternal separation for 3 h per day and lasted for
14 days (van Heerden et al., 2009).

PRL3 promotes cell proliferation, migration, and invasion in glioblastoma


PRL3
cells in Human (Mu et al., 2018; Wang et al., 2016a).

CDC14A involved in cell cycle regulation of human brain


Upregulated in response to Amyloid- Oligomers in adult human cortical
CDC14A vascular endothelial cells following injury induced by high
brain slices (Sebollela et al., 2012).
glucose, free fatty acids, and hypoxia (Su et al., 2015)

Regulates RNA polymerase II and represses cell cycle


CDC14B transcription in primary mouse embryonic fibroblasts Involved in glioblastoma growth in human (Galeano et al., 2013).
(Guillamot et al., 2011).

PTEN single nucleotide polymorphisms associated with increased risk of


depression in a Chinese cohort (Liu et al., 2016). PTEN protein levels are
PTENs higher in postmortem lysates of the ventral prefrontal cortex from suicide
victims diagnosed with depression when compared to non-depressed healthy
controls (Karege et al., 2011).

PTEN moved to nucleus and promoted neuron survival in mice PTEN recruitment controls synaptic and cognitive function in AD model
(Goh et al., 2014) App/Psen1 mice (Knafo et al., 2016).
PTEN and β-catenin signaling regulates normal brain growth
PTEN mediates brain growth during development and PTEN mutation-induced
trajectory by controlling cell number, and imbalance in this
Autism-like behavior in mice (Chen et al., 2015; Clipperton-Allen and
relationship can result in abnormal brain growth in mice (Chen
Page, 2014; Fraser et al., 2008; Igarashi et al., 2018; Kwon et al., 2006).
et al., 2015).
PTEN mutations are late events in the malignant progression of glioma, and
the occurrence of PTEN mutations are significantly correlated to patients’
short-term survival (Yang et al., 2010).
PTEN mutation coexisted with autism behavior and macrocephaly in human
(Vanderver et al., 2014).

Depression mice model showed increased hippocampal MKP1 expression,


MKP1 controls axon branching induced by BDNF signaling via
MKP1 (DUSP1) which can be normalized by antidepressant, while mice lacking MKP1 are
mediating JNK deactivation in mice (Jeanneteau et al., 2010)
resilient to stress (Duric et al., 2010)
MKP-1 functions in light-dependent and time-of-day-dependent
MKP1 mRNA increased in the DG and CA1 of human diagnosed with
manners in the mice central clock structure-the suprachiasmatic
depression (Duric et al., 2010).
nucleus (Doi et al., 2007)
Increased MKP-1 expression levels could be the cause of the high resistance
to conventional chemotherapeutics in human glioblastoma multiforme in
human (Yu et al., 2012)
DUSP1 protein levels increased in the hippocampus (Iio et al., 2011) and the
ventrolateral orbital cortex (Zhao et al., 2017) of stressed rats when
compared to controls.

MKP2 promotes neuroplasticity and memory, and its deletion


MKP2 were significantly decreased in cerebellar vermis from schizophrenic
MKP2 (DUSP4) impairs synaptic plasticity and hippocampal-dependent memory
patients compared to control subjects (Kyosseva et al., 1999).
in mice (Abdul Rahman et al., 2016)
MKP2 was increased in BAs 8, 9, 10, and hippocampus, without any change
in the cerebellum of depressed suicide subjects compared with control
subjects (Dwivedi et al., 2001).

DUSP6 mRNA was differentially expressed in post-mortem tissue ventromedial


DHA-enriched fish-oil induced MKP3 that enhance GFAP in
MKP3 (DUSP6) PFC of patients with depressive disorder in a sex-specific
developing rat brain astrocytes (Tripathi et al., 2017).
manner (Labonte et al., 2017).

Regulate the signaling of pressure-dependent myogenic cerebral DUSP5 served as transcriptional target of tumor suppressor p53 in glioblastoma
Hvhr3 (DUSP5)
arterial constriction in rat (Wickramasekera et al., 2013). in human (Ueda et al., 2003).
(continued on next page)

4
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Table 2 (continued )
Isoforms Function in Normal Brain Function in Pathological Brain

MKP-X (DUSP7) The expression of DUSP7 was mediated by ERK1/2 activity both DUSP7 mRNA expression was reduced in the whole brain after chronic
in resting and LPS-stimulated microglia in rat (Ham et al., 2015) amphetamine injections in mice (Sokolov et al., 2003).

Hvh5 (DUSP8) Abundant in human and mice brain and inactivate mitogen- Hvh5 mRNA was induced in the nucleus accumbens, caudate putamen,
activated protein kinase (Martell et al., 1995a). frontal cortex, and hippocampus by i.p. Injection of cocaine and
fluoxetine in rat (Thiriet et al., 1998).

MK-STYX induced neurite extensions through the Rho signaling


MK-STYX missense mutation was identified in intellectual disability,
pathway and forms synapse in PC-12 cells. MK-STYX altered
accompanied by seizures and behavioral problems in
their morphology in primary hippocampal neurons in rat (Banks
human (Isrie et al., 2015)
MK-STYX (STYX-L1) et al., 2017; Flowers et al., 2014)
MK-STYX was differentially expressed in the blood of depression patients
(Guilloux et al., 2015; Miyata et al., 2016). or depressive-like behavior
in mice (Miyata et al., 2016).

MTMR2 contributes to the maintenance of excitatory synapses Loss of MTMR2 in Schwann cells causes CMT4B1 neuropathy, which is
by inhibiting excessive endosome formation and destructive characterized by a dysmyelinating neuropathy with myelin outfoldings
endosomal traffic to lysosomes in rat (Lee et al., 2010). in mice (Bolis et al., 2005).
MTMR2
MTMR2 interacted with the neurofilament light chain protein,
Mutation in the MTMR2 gene is a causative mutation in patients with
NF-L, in both Schwann cells and neurons in rat and human (
Charcot-Marie-Tooth Disease Type 4B1 (Abdalla-Moady et al., 2017).
Previtali et al., 2003).

MTMR4 was differentially expressed between grade II-III gliomas and


MTMR4 .
(grade IV) glioblastomas in human (Bourgonje et al., 2016).

MTMR5 suppress neurite growth in hippocampal neurons in rat.


SBF1 mutations may cause a syndromic form of autosomal recessive axonal
MTMR5 (SBF1) Overexpression of MTMR5 reduce hippocampal neurite
neuropathy (AR-CMT2) in addition to CMT4B3 in Human (Manole et al., 2017).
outgrowth (Buchser et al., 2010).

An intronic variant was identified in the genetic locus of MTMR7 linked to


MTMR7 variant Creutzfeldt-Jakob disease susceptibility in human
(Sanchez-Juan et al., 2012).
MTMR7 was differentially expressed in the substantia nigra of patients with
PD in the GWAS study (Hook et al., 2018).

The copy number loss for MTMR8 was identified in 65 % of the glioblastoma
MTMR8
multiforme patient sample (Waugh, 2016).

MTMR10 was affected via 15q13.3 microdeletion and displayed strong


phenotypes related to autism-like behavior (Forsingdal et al., 2016;
MTMR10 Kogan et al., 2015) and autism like behavior in mice
MTMR10 was affected via 15q13.3 deletion and was present in ADHD patients
(Valbonesi et al., 2015).

MTMR11 was frameshift mutated in children with epileptic encephalopathies


MTMR11
(Veeramah et al., 2013).

Mtmr13− /− mice show both the initial dysmyelination and later degenerative
pathology of CMT4B2 (Ng et al., 2013; Robinson et al., 2008). Mutation in the
MTMR13
MTMR13 gene associated with a classical Charcot-Marie-Tooth 4B2 phenotype
in human (Chen et al., 2014; Negrão et al., 2014).

The phosphatase activity of laforin is dispensable to rescue Epm2a − /− mice from


Lafora disease (Gayarre et al., 2014). EPM2A gene was expressed lower to control
Laforin (EPM2A) cells in fibroblasts from Lafora disease in human (Garcia-Gimeno et al., 2018).
EPM2A gene has 11 kinds of mutations in the patients of Lafora disease in human
(Gomez-Garre et al., 2000)

DUSP11 mRNA was downregulated in the model of nicotine-induced seizures in


DUSP11
mice (Kedmi and Orr-Urtreger, 2007).

DUSP13A interacts with the N-terminal domain of Apoptosis signal-regulating


kinase 1 in an oxidative stress-independent manner in brain neuroblastoma.
DUSP13A
The knock-down of DUSP13A decreased the phosphorylation and activation of
apoptosis signal-regulating kinase 1 (Park et al., 2010).

DUSP14 is a direct negative-feedback mechanism of 3, 4-meth­


DUSP14 mRNA was upregulated in the frontal cortex in response to
ylenedioxymethamphetamine (MDMA)- induced ERK signaling
Environmental chronic mild stress in mice (Tordera et al., 2011).
DUSP14 in the striatum of mice (Marie-Claire et al., 2008).
DUSP14 was a gene target limiting axon growth and DUSP14 was the delayed primary response genes downregulated in DG neurons
regeneration downstream of Krűppel-like transcription factor 9 of mice experienced foot shock after 24 h in mice (Rao-Ruiz et al., 2019).
(continued on next page)

5
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Table 2 (continued )
Isoforms Function in Normal Brain Function in Pathological Brain

(KLF9)’s ability to suppress axon growth in retinal ganglion cells


of rat (Galvao et al., 2018).
DUSP14 was decreased in HD mice and can be enriched after pridopidine
treatment in mice (Geva et al., 2016).

DUSP15 is necessary for full activation of ERK1/2


phosphorylation and represses expression of several myelin The SNP (rs3746599) of DUSP15 was significantly associated with autism in
genes, including myelin basic protein, in Schwann cells of rat human (Tian et al., 2017)
DUSP15
(Rodriguez-Molina et al., 2017).
DUSP15 influences oligodendroglial differentiation and myelin
gene expression in rat (Muth et al., 2016).

The mRNA levels of DUSP19 were increased in the dentate gyrus from
DUSP19
depression postmortem tissue (Duric et al., 2010).

The promoter hypermethylation of the DUSP22 gene was identified in the


hippocampus from controls and AD patients. DUSP22 inhibits PKA activity
and thereby determines tau phosphorylation status and CREB signaling
(Sanchez-Mut et al., 2014).
DUSP22 DUSP22 gene promoter showed higher DNA methylation levels in the
famine-exposed schizophrenia patients compared to non-famine exposed
groups (Boks et al., 2018).
Increased DNA methylation within/around the DUSP22 gene was linked to
increased PTSD symptoms in human (Rutten et al., 2018)

DUSP23 affected neuronal differentiation in mice. The knock-


The mRNA expression of the gene DUSP23 was significantly lower in patients
down of DUSP23 decreased neuronal differentiation in terms of
DUSP23 that have died from the disease compared with neuroblastoma patients with
neuronal outgrowth and the expression of neuronal marker
no evidence of disease (Caren et al., 2011).
proteins (Kim et al., 2016).

DUSP26 suppresses receptor tyrosine kinases and regulates DUSP26 was differentially expressed between grade II-III gliomas and (grade IV)
neuronal development in zebrafish (Yang et al., 2017) glioblastomas in human (Bourgonje et al., 2016).
DUSP26 inhibition via NSC-87877 function in neuroblastoma, resulting in
DUSP26 decreased tumor growth and increased p53 and p38 activity in
mice (Shi et al., 2015).
DUSP26 expression and JNK activation were enhanced in the hippocampus
of AD patients (Jung et al., 2016).

The variant rs950302 of cytosolic gene DUSP27 associate with heroin addiction
DUSP27
vulnerability in African Americans (Nielsen et al., 2010).

Note: AD: Alzheimer’s disease; HD: Huntingdon’s disease; CMT4: Charcot-Marie-Tooth Neuropathy Type 4.

proteinatlas.org)]. At the tissue level, hSSH-1 proteins were expressed in addition to the number, size, and morphology of dendritic spines in
the frontal cortex and cerebellum of human post-mortem brain tissues of cortical neurons by controlling the actin cytoskeleton via ADF/cofilin
healthy elderly persons (Barone et al., 2014; Zafar et al., 2017) as well as activation. Interestingly, AMPA receptor trafficking and the morphology
in human keratinocytes (Kligys et al., 2007). of dendritic spines in cortical neurons are strongly associated with
SSH phosphatases modulate actin separation and reunion by regu­ synaptic plasticity, which underlies cognitive functions such as learning
lating the ADF /cofilin complex in vivo (Niwa et al., 2002). SSH phos­ and memory processes (Yuen et al., 2010).
phatases remove a phosphorus group from ADF/cofilin to activate this
complex (Endo et al., 2007, 2003; Kim et al., 2009; Wang et al., 2005). In 2.2. Phosphatases of regenerating liver (PRLs) family
turn, the activated ADF/cofilin complex depolymerizes and dismantles
actin filaments in order to drive the protraction of growth cones and 2.2.1. Structure and homologues
neurite extensions, as observed in rat hippocampal neurons. In contrast, While the cellular functions of this family remain yet to be uncov­
activated LIM domain kinase 1, a negative regulator of ered, what we currently know is that PRLs are oncogenes (Wei et al.,
actin-polymerization dynamics, adds a phosphorus group to ADF/cofi­ 2018). Within this subgroup, three proteins subtypes, PRL-1, PRL-2, and
lin, inhibiting the formation of this complex. This inhibition then drives PRL-3, have been identified based on their amino acid sequences
actin polymerization, which leads to a reduction in the protraction of (Campbell and Zhang, 2014), which share greater than 50 % homology
growth cones and neurite extensions (Soosairajah et al., 2005). The in humans (Rios et al., 2013).
balance between SSH and LIM domain kinase 1 is responsible for PRLs carry the CAAX motif and are the only CAAX proteins in the
modulating actin filament assembly at the tip of the growth cone and is DUSP family (Rios et al., 2013). CAAX proteins are involved in global
essential for driving the repulsive or attractive responses of growth cellular functions, such as proliferation and differentiation. A polybasic
cones and neurite extensions, which in turn leads to modifications in region localizes next to the CAAX box and mediates membrane or nu­
neuronal morphology and sprouting (Endo et al., 2007; Meberg et al., clear localization of PRLs. The catalytic or protein phosphatase (PTP)
1998; Ohashi, 2015; Wen et al., 2007). Specifically, SSH1, along with domain is responsible for enzymatic activity, requiring the P-loop resi­
Cofilin1 modulated growth cone extension in rat (Hsieh et al., 2006; dues and the WPD loop (conserved in the PTP family) residues for the
Sparrow et al., 2012) and chick (Endo et al., 2003). SSH2 reverse transfer of a phosphate group (Fig. 1B) (Rios et al., 2013).
actin-severing defects and improves actomyosin parameters in in­
terneurons of mice (Tielens et al., 2016). Furthermore, SSH phospha­ 2.2.2. Expression and function
tases play a pivotal role in the control of AMPA receptor trafficking, in At the cellular level, PRLs localize in the plasma membrane and

6
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Fig. 1. Schematic representations of the structure of the family members of the DUSP proteins.
A. Structures of SSH family phosphatases. The schematic diagrams show protein structures of mouse (m), human (h), and Drosophila (D) SSHs. The highly conserved
regions between SSH family proteins are indicated as A, B, P (protein phosphatase) and S (serine-rich) domains (adapted from (Niwa et al., 2002; Ohta et al., 2003)).
B. Simplified schematic diagram of the PRL proteins. The PTP domain is responsible for enzymatic activity, requiring the P-loop residues and the WPD loop residues
for phosphate transfer. A polybasic region localizes next to the PTP domains and mediates membrane or nuclear localization of PRLs. The CAAX protein prenylation
anchors the PRLs to the cellular membrane (simplified from (Bessette et al., 2008)).
C. Schematic of the primary structure of CDC14 in humans. The conserved domain is depicted in blue (adapted from (Gray et al., 2003; Rosso et al., 2008)).
D. The domain architecture of PTEN. PTEN is composed of four domains: the PIP2-binding domain (PBD)/PTP, C2 domain, C tail domain, and a PDZ binding domain.
The active catalytic site is HCxxGxxR (adapted from review (Hopkins et al., 2014; Worby and Dixon, 2014)).
E. Scaled representation of the protein domains of human myotubularins. All myotubularins share the PH-GRAM and phosphatase (active or dead) domains.
Additionally, myotubularins can also carry other functional domains, including the PDZ binding site, the PH (Pleckstrin homology) and FYVE (Fab1-YOTB-Va­
c1-EEA1) domains, as well as the DENN (Differentially Expressed in Normal and Neoplastic cells) domain. Except for MTMR10, all myotubularins are composed of a
coiled-coil domain. For each myotubularin, the amino acid length for the most described protein isoform is indicated (adapted from review (Begley and Dixon, 2005;
Raess et al., 2017)).
F. Classification and domain structure of the MKP family. The three subgroups are based on substrates and subcellular localization. In addition to the MAPK binding
(MKB) domain and dual-specificity phosphatase (DSP) domain, the nuclear localization signal (NLS), nuclear export signal (NES), and PEST sequences are shown in
the figure. (adapted from (Caunt and Keyse, 2013; Kondoh and Nishida, 2007)).
G. Atypical DUSPs predominantly contain the consensus DSP catalytic domain, whereas some atypical DUSPs contain the CH2 domain, the carbohydrate-binding
domain, and is an Arginine-rich or Proline-rich region (adapted from (Patterson et al., 2009)).

nucleus, whereas at the tissue level, subtypes of human PRLs (hPRLs) Trautmann and McCollum, 2002). CDC14 comprises three isoforms,
differ depending on the type and severity of the tumor in question. hPRL- CDC14A, CDC14B, and CDC14C (Bremmer et al., 2012). The isoform
1 and hPRL-2 mRNA expression patterns are widespread in healthy adult CDC14B encodes four splice variants, including CDC14Bpar, CDC14B1,
human tissues. hPRL-2 is expressed at higher levels in the brain than CDC14B2, and CDC14B3 (Rosso et al., 2008). Interestingly, the isoform
hPRL-1, especially in the granular layer of the cerebellum (Dumaual CDC14C (also known as CDC14Bretro) is produced by the CDC14B splice
et al., 2006). hPRL-3 mRNA is expressed in both skeletal muscle and the variant CDC14Bpar, by gene retroduplication, a process which occurred
heart during development (Matter et al., 2001). When compared to in hominoids (Mocciaro and Schiebel, 2010; Rosso et al., 2008).
other mammalian tissue, such as mouse tissue, similarities do exist. For CDC14A contains a core domain and a nuclear export signal, which is
instance, mouse PRL (mPRL)-1 is 100 % identical to hPRL-1 in the responsible for the translocation of CDC14A from the nucleus to the
number of amino acid subunits (Zeng et al., 1998). mPRL-2 mRNA is cytoplasm. The core domain contains two structurally similar domains,
expressed in skeletal muscle, and mPRL-3 is expressed in both skeletal A and B (Gray et al., 2003). The protein structure of human hCDC14A
muscle and heart tissue (Zeng et al., 1998). shares 65 % compatibility with hCDC14B except for the nucleolar tar­
PRLs promotes cell proliferation, migration, invasion, tumor growth, geting sequence (N-terminal 44 amino acids) that is responsible for
and metastasis via multiple signaling pathways, including extracellular localizing CDC14B to the nucleolus throughout interphase during cell
signal-regulated kinase (ERK) 1/2 pathways (Luo et al., 2009; Peng division (Gray et al., 2003). CDC14C structure is similar to hCDC14B
et al., 2009), the mechanistic target of the rapamycin (Ye et al., 2015), except for the C-terminus (Fig. 1C) (Rosso et al., 2008).
and the phosphatidylinositol 3-kinase (PI3K)/AKT pathways (Zhang
et al., 2018). It has been observed that PRL-1 and PRL-2 induce cell 2.3.2. Expression and function
invasion and motility through the activation of ERK 1/2. Moreover, At the cellular level, CDC14A localizes on the centrosomes of cells
PRL-3 stimulates both cell proliferation and epithelial-mesenchymal during interphase (Mailand et al., 2002). CDC14B1 is expressed in the
transition, a crucial developmental process, by acting upstream of nucleoli and CDC14B2 in nuclear speckles localized within the nucleus,
PI3K. Interestingly, ERK1/2 and PI3K are common signaling pathways as demonstrated in COS7-cells. Moreover, CDC14B3 and CDC14Bpar
in cell proliferation, migration, invasion, tumor growth, and metastasis exhibit co-localization with microtubules in COS7-cells (Rosso et al.,
(Rios et al., 2013). 2008). It has been also demonstrated that CDC14C co-localizes with a
marker of the endoplasmic reticulum in COS7 cells, human HeLa, or
LN229 cell lines (Rosso et al., 2008).
2.3. Cell division cycle 14 (CDC14) phosphatases
At the tissue level, hCDC14A and hCDC14B are both found in the
cerebral cortex, lymph nodes, liver, colon, kidneys, and testis (www.
2.3.1. Structure and homologues
proteinatlas.org). While hCDC14Bpar mRNA is predominantly
CDC14 phosphatases, encoded by the CDC14 gene, are the subgroup
expressed in the adult and fetal brain, hCDC14B1, hCDC14B2, and
that is mainly involved in cell cycle regulation (Clifford et al., 2008;

7
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

hCDC14B3 mRNAs are expressed in the adult brain, including the hip­ have been described, with MTM1 being the first recognized myotubu­
pocampus, prefrontal cortex, amygdala, and hypothalamus (Rosso et al., larin. Subsequently, 13 myotubularin-related proteins, labeled MTMR1
2008). hCDC14C mRNA is found in the adult brain and embryonic to MTMR13, were identified (Raess et al., 2017).
forebrain, including the dorsal telencephalon (Rosso et al., 2008). All myotubularins share the PH-GRAM (Pleckstrin Homology - Glu­
The CDC14 family is conserved within eukaryotes and plays a role in cosyltransferase, Rab-like GTPase Activator, and Myotubularins)
the inactivation of mitotic cyclin-dependent kinase via dephosphoryla­ domain and catalytically active or inactive phosphatase domains.
tion. While activation of cyclin-dependent kinase drives cells into Additionally, myotubularins can also carry other functional domains,
mitosis, their inactivation promotes mitotic exit and cytokinesis. including the PDZ binding site, the PH (Pleckstrin homology) and FYVE
Furthermore, CDC14 regulates a variety of other cellular events, such as (Fab1-YOTB-Vac1-EEA1) domains, and the DENN (Differentially
DNA recombination, telomere segregation, mitotic spindle dynamics, Expressed in Normal and Neoplastic cells) domain. Except for MTMR10,
and cytokinesis (Bremmer et al., 2012; Stegmeier and Amon, 2004). all myotubularins are composed of a coiled-coil domain (Fig. 1E) (Begley
and Dixon, 2005; Raess et al., 2017).
2.4. Phosphatase and tensin homologs deleted on chromosome 10
(PTENs) family 2.5.2. Expression and function
At the cellular level, myotubularins do not show any nuclear
2.4.1. Structure and homologues expression but are primarily localized in the cytoplasm as a richly-
The PTEN gene codes for the PTEN protein, a protein in which tumor- formed network. Additionally, myotubularins have also been shown to
suppression functions due to its phosphatase activity (Govender and localize to Rac1-inducible plasma membrane ruffles. This localization to
Chetty, 2012; Knafo and Esteban, 2017). PTEN carries two isoforms: a Rac1-induced ruffles seems to be associated with a highly conserved
CUG initiated isoform designated PTENα, and an AUG initiated isoform myotubularin domain referred to as RID (Laporte et al., 2002).
referred to as PTENβ with a smaller molecular weight than PTENα (Liang At the tissue level, myotubularin mRNA expression has been docu­
et al., 2017). The N-terminal domain of PTEN contains the catalytic mented for various human organs, including the spinal cord and sub­
N-terminal PIP2-binding domain and PTP domain. The C-terminal stantia nigra (SN) of the central nervous system, skin, lungs, and vagina
domain consists of the following subdomains: C2, C-tail, and the PDZ, (Raess et al., 2017). At the protein level, myotubularins are expressed in
with the active catalytic site being HCxxGxxR (Fig. 1D) (Hopkins et al., the brain, including the cerebral cortex, hippocampus, and cerebellum,
2014; Worby and Dixon, 2014). PTEN’s phosphatase domain carries a as well as non-brain tissues like lungs, muscles, endocrine tissue, bone
similar structure to protein phosphatases but with a more significant marrow, immune system, liver, gallbladder, and pancreas (www.protein
active site allowing it to bind other substrates such as phosphoinositide atlas.org).
(PI). Additionally, PTEN’s C2 domain has been shown to bind phos­ Myotubularins are involved in several cellular processes, including
pholipid membranes in vitro and thus aid in the steering and anchoring autophagy, apoptosis, the actin cytoskeleton, and intermediate fila­
of PTEN to the cellular membrane (Lee et al., 1999). Together this ments dynamic and PI metabolism. Active myotubularin can remove
interplay between both domains carries implications for the suppression phosphate on carbon number 3 of PtdIns3P or PtdIns(3,5)P2, and turn it
and stimulation of tumor cell growth. into PtdIns or PtdIns5P, respectively (Hnia et al., 2012). Additionally,
they are also involved in myelin formation in neurons. For instance,
2.4.2. Expression and function MTMR2 is present in the nucleus and cytoplasmic compartments of
At the cellular level, PTENα is expressed in the cytoplasm and the Schwann cells and motor neurons but not in the nucleus of sensory
mitochondrial inner membrane, while PTENβ is in the nucleolus (Liang neurons. Moreover, MTMR2 interacts with the neurofilament light chain
et al., 2017, 2014). In neurons, PTEN dynamically localized to special­ protein NF-L in both Schwann cells and neurons (Previtali et al., 2003).
ized subcellular compartments, such as the neuronal growth cone or Schwann cell/dorsal root ganglion neuron co-cultures from MTMR2
dendritic spines, as well as the nucleus (Kreis et al., 2014; LaSarge et al., knock-out mice exhibit excessive redundant myelin, also known as
2016). myelin outfolding, and MTMR2 replacement was shown to rescue this
At the tissue level, PTEN protein and mRNA expression have been myelin outfolding phenotype (Bolis et al., 2009). Interestingly, the
observed in the human pancreas (Ebert et al., 2000) and in the human deletion of MTMR2 phospholipid phosphatase in humans causes
brain, including the cerebral cortex, cerebellum, and hippocampus childhood-onset of an autosomal recessive demyelinating neuropathy,
(www.proteinatlas.org). PTEN exhibits differential distribution in the also known as Charcot–Marie–Tooth type 4B1 (Bolino et al., 2000).
rat brain, with the highest levels found in the anterior olfactory nucleus,
cerebral cortex, amygdaloid nucleus, hippocampus, Purkinje cells, and 2.6. Mitogen-activated protein kinase phosphatases (MKPs)
several nuclei in the basal ganglia, thalamus, midbrain, and pons (Cai
et al., 2009). 2.6.1. Structure and homologues
PTEN regulates cellular proliferation, survival, energy metabolism, The MKP genes encode phosphatases that dephosphorylate MAP
cellular architecture, and motility (Worby and Dixon, 2014). During kinase (MAPK) signaling elements in vivo hence deactivating them
development, PTEN modulates cell proliferation and neuronal growth (Kondoh and Nishida, 2007). In turn, this leads to the modulation of
by dephosphorylating PIP3 and antagonizing PI3K signaling. PI3K several physiological processes via a conformational change of their
signaling mediates responses to different cellular stimuli, including substrates.
hormones and growth factors (Bermudez Brito et al., 2015; Leslie and MKPs are composed of the MAPK-binding (MKB) domain in the N-
Downes, 2002; Ye et al., 2012). On the other hand, the inhibition of terminal end and the DSP domain in the C-terminal end. The N-terminal
PTEN causes axonal regeneration and neural repair (Ohtake et al., MAPK-binding domain regulates enzymatic specificity through docking
2015). PTEN plays a fundamental role in the maintenance of chromo­ interaction with MAPK. The binding of phosphorylated MAPK to the
somal stability through physical interactions with centromeres and MAPK-binding domain alters the structure of the DUSP domain. This
controls DNA repair (Shen et al., 2007). alteration in the conformation of the protein, in addition to the inter­
action of the catalytic domain with MAPK, increases MKPs’ catalytic
2.5. Myotubularins activity (Fig. 1F).

2.5.1. Structure and homologues 2.6.2. Expression and function


Myotubularins can be found in almost all eukaryotes, ranging from MKPs seem to be highly expressed in a variety of tissue types,
yeast to mammals. In humans, 14 clearly defined myotubularin paralogs including the brain, endocrine tissues, lung, digestive tract, liver,

8
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

gallbladder, kidneys, male and female reproductive tissues, adipose current antidepressants, and genetic factors have been shown to
tissue (www.proteinatlas.org). contribute to the risk of treatment-resistant depression (Willner et al.,
The MKP family can bind to substrates from MAPK signaling path­ 2013).
ways, including ERK, JNK, and p38 (Caunt and Keyse, 2013; Kondoh
and Nishida, 2007). The DSP domain of MKPs inactivates MAPK by 3.1.1. PTEN and depression
docking into phospho-MAPK (Kondoh and Nishida, 2007). MAPK Genetic studies have demonstrated an association between three
signaling converts extracellular stimuli into a variety of intracellular distinct PTEN single nucleotide polymorphisms, rs701848, rs2735343,
responses, such as proliferation, differentiation, survival, apoptosis, and and rs112025902, and increased risk of depression in a Chinese cohort
migration (Dwivedi et al., 2001; Kim and Choi, 2015; Yuan et al., 2010). (Liu et al., 2016). Other studies have provided evidence that PTEN
MAPK signaling pathways have many predominant kinases, including protein levels are higher in postmortem lysates of the ventral prefrontal
JNK, ERK, and p38, all of which can be inactivated by different MKPs cortex (Brodmann’s area 11) from suicide victims diagnosed with
(Owens and Keyse, 2007). depression when compared to non-depressed healthy controls (Karege
et al., 2011). Besides, the enzymatic activity of the kinases PI3K and
2.7. Atypical DUSPs AKT1 were found to be decreased in the ventral prefrontal cortex
(Brodmann’s area 11) of suicide victims diagnosed with depression,
2.7.1. Structure and homologues while their protein levels did not differ. Conversely, PTEN protein levels
Atypical DUSP genes have multiple nomenclatures and remain in the ventral prefrontal cortex were observed to be increased in patients
poorly characterized (Huang and Tan, 2012; Rios et al., 2014). The with a depressive disorder. This attenuation of PI3K and AKT1 activity in
HUGO (Human Genome Organization) Gene Nomenclature Committee suicide victims with a depressive disorder may be related to elevated
includes 16 atypical DUSPs genes (Cain and Beeser, 2013), Patterson levels of PTEN, which in turn may result in insufficient phosphorylation
et al., identify 20 members of atypical DUSPs (Patterson et al., 2009), of second lipid messengers PI3-phosphate, PIP2 and PIP3 (Karege et al.,
and Huang and Tan 15 atypical DUSP members (Huang and Tan, 2012). 2011).
Moreover, phylogeny analysis has shown that atypical DUSPs appear not PI3K and AKT are also involved in mediating depressive-like
to be derived from a common proximal ancestor (Huang and Tan, 2012; behavior in mice induced by stress (Deng et al., 2019; Wu et al.,
Patterson et al., 2009). Atypical DUSPs encode proteins with a molecular 2018), and inhibitors of PI3K/ AKT have been shown to prevent
weight of less than 27 kDa (Patterson et al., 2009; Rios et al., 2014). All antidepressant-like effects (characterized by decreased immobility time)
the members of atypical DUSPs are listed in Supplementary Table 2. induced by creatine in mice following the stress-inducing tail suspension
Atypical DUSPs predominantly contain the consensus DSP catalytic test (Cunha et al., 2016; Ludka et al., 2016). Together, these results
domain. Some atypical DUSPs contain a CH2 domain, carbohydrate- highlight that the phosphorylation of AKT and the downstream effects
binding domain, and an Arginine-rich or Proline-rich region (Fig. 1G) might be of interest as a potential treatment of depression. Previous
(Patterson et al., 2009). work has indicated that deficits in vital cellular processes such as cell
survival and neuroplasticity are observed in major depression (Pittenger
2.7.2. Expression and function and Duman, 2008). Therefore, looking into the enzymatic activity of
At the cellular level, the majority of atypical DUSPs are localized in PTEN and PI3K and their association with abnormalities in neurotrophic
the cytoplasm, with some being in the nucleus, and another subset in the signaling is pertinent (Karege et al., 2011).
mitochondria and Golgi in various cell types (Patterson et al., 2009).
At the tissue level, the brain expression of atypical DUSPs is pre­ 3.1.2. DUSP1 and depression
sented in Table 1, while their localization in other tissues s described in DUSP1 has also been documented to play a role in the pathophysi­
Supplementary Table 3. ology of depression in human subjects (Akbarian and Davis, 2010). For
Some atypical DUSPs regulate MAPK, playing a role in cell prolif­ instance, DUSP1 mRNA expression was shown to be increased in the
eration and apoptosis. In addition to MAPK protein substrates, several hippocampus of depressed patients when compared to healthy controls
substrates of atypical DUSPs include nucleic acids (such as RNA), and (Duric et al., 2010). As negative regulators of DUSP1, MAPK and its
phosphorylated carbohydrates (such as amylopectin and glycogen). downstream kinases were decreased in the prefrontal cortical areas and
Nonetheless, the physiological substrates of many atypical DUSPs the hippocampus of suicide subjects with depression (Dwivedi et al.,
remain unknown (Bayón and Alonso, 2010; Patterson et al., 2009). 2006, 2003; Dwivedi et al., 2001). Recent evidence from animal studies
suggests the involvement of MKPs in depression-like behavior. For
3. DUSP genes and mental disorders instance, in one study, DUSP1 protein levels were observed to be
increased in the hippocampus of stressed rats when compared to con­
Accumulating evidence from the current literature on the link be­ trols in the resident-intruder paradigm. Kinase substrates of DUSP1,
tween DUSP genes and mental disorders such as depression, bipolar including phosphorylated MEK1/2 and ERK1/2, were decreased in the
disorder, autism spectrum disorder (ASDs), schizophrenia, post- hippocampus of stressed rats when compared to controls (Iio et al.,
traumatic stress disorder (PTSD), and substance abuse disorders are 2011). In another stress model – the chronic unpredictable stress (CUS)
described in the following section. The summary of the pathological model - the upregulated protein levels of DUSP1 in the hippocampus was
implication of DUSP family members in mental disorders is presented in rescued by the antidepressant fluoxetine, two weeks after inducing
Table 2. depressive-like behaviors in rodents (Duric et al., 2010). Similarly, the
upregulated protein level of DUSP1 in the ventrolateral orbital cortex of
3.1. Depression rats subjected to chronic unpredictable mild stress was attenuated by
enhancing miR-101 expression, and so was the depressive-like behavior
Depression is characterized by psychological and physiological (Zhao et al., 2017). miR-101 is a functional silencing small RNA tar­
symptoms, including negative thinking, anhedonia, fatigue, memory geting DUSP1 (Llorens et al., 2013; Zhu et al., 2010) and the amyloid
impairment, insomnia, extreme weight loss, or weight gain. The World precursor protein (APP) (Vilardo et al., 2010). Moreover, overexpression
Health Organization (WHO) reported that depression is a leading cause of DUSP1 in the hippocampus induced anhedonia-like behavior,
of disability worldwide. The neurobiological mechanisms underlying including a reduced preference for sucrose and increased frequency of
depression are complex, however, first-line antidepressants are effective failure-to-escape in the active avoidance test (Duric et al., 2010). DUSP1
in a subset of patients by reversing some of the symptoms of depression knockout mice were also resistant to CUS-induced depressive-like be­
(Thornicroft et al., 2017). However, not all individuals benefit from haviors (Duric et al., 2010).

9
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Antidepressant treatments have an impact on the expression of female patients diagnosed with depression. Females diagnosed with
DUSP1 in healthy animal subjects. For example, the administration of depression showed elevated levels of phospho-ERK1/2 in the prefrontal
fluoxetine reduced DUSP1 mRNA expression in the prefrontal cortex of cortex compared to healthy female controls. Elevated
healthy rats (Kodama et al., 2005). Moreover, electroconvulsive therapy phospho-ERK1/2-reactive cell density mainly localized in layers II/III
(ECT), a treatment for drug-resistant depression, induced upregulation and layers V/VI of the prefrontal cortex of female patients with
of DUSP1 mRNA levels in all hippocampal subregions, and the prefrontal depressive disorder (Labonte et al., 2017).
cortex of healthy rats (Kodama et al., 2005). Additionally, ECT and administration of antidepressants are shown
These results together strengthen the association between DUSP1 to have an impact on DUSP6 expression. For instance, ECT induces the
and depression. Future studies looking into treatments for depression upregulation of DUSP6 mRNA levels in the prefrontal cortex and DUSP6
should consider targeting DUSP1 as a therapeutic strategy by, for protein levels in the hippocampus and prefrontal cortex of healthy rats
example, making use of small RNAs, particularly miRNAs, to silence the (Kodama et al., 2005). The administration of fluoxetine, however,
expression of DUSP1 hence reversing signs and symptoms of depression. reduced the mRNA expression of DUSP6 in the prefrontal cortex of
healthy rats (Kodama et al., 2005). This discrepancy in results highlights
3.1.3. DUSP4 and depression the need to further investigate the involvement of DUSP6 in mood dis­
Analyses of postmortem brain tissue samples indicated increased orders such as depression.
protein levels of DUSP4 (aliases MKP-2, the ERK1/2 phosphatase) in
prefrontal cortical areas and the hippocampus of patients with major 3.1.5. DUSP2, DUSP12, DUSP19, DUSP23, DUSP24 and depression
depressive disorder following the death by suicide compared to non- The subregions of the hippocampus exhibit differential RNA
psychiatric control subjects. This increase was accompanied by expression of MKPs in post-mortem brain samples of patients diagnosed
decreased expression of mRNA and protein levels of ERK1 and ERK2, with depressive disorder. For instance, while DUSP2 and DUSP19 show
resulting in reduced MAPK activity (Dwivedi et al., 2001). Another study higher expression in the hippocampal DG than in the CA1, mRNA levels
showed sex-dependent differential expression of DUSP4 mRNA in the of DUSP12 and DUSP24 are increased in the CA1 region of postmortem
ventral subiculum of patients with depression compared to healthy brain tissue of patients diagnosed with depression (Duric et al., 2010) as
controls; with differences observed in male but not in female subjects compared to healthy controls. Another study showed differential
(Labonte et al., 2017). expression of DUSP19 mRNA in the ventral subiculum and of DUSP23
DUSP4 protein expression remained unchanged in the hippocampus mRNA in the nucleus accumbens and the Broadman area of male pa­
and the frontal cortex of rats subjected to prenatal stress (Budziszewska tients with depressive disorder, but not in female patients (Labonte
et al., 2010) and in a male rat model of depression, which was induced et al., 2017). These results indicate that distinct DUSP genes may be
by neonatal treatment with clomipramine (Feng et al., 2003). Treat­ linked with depression differently, including sex-specific effects. Despite
ments, including antidepressants and ECT, have led to changes in DUSP4 these observations, further experimental studies are needed to better
expression in healthy animals. For instance, ECT treatment in healthy understand the involved of MKPs in depression and eventually target
male rats induced increased expression of DUSP4 mRNA in the dentate their expression to reverse depressive symptoms.
gyrus (DG) of the hippocampus and the prefrontal cortex (Kodama et al.,
2005). 3.2. Bipolar disorder
Although these findings suggest DUPS4 changes in both human pa­
tients and animals subjected to antidepressant treatments ECS, the Bipolar disorder is characterized by mood instability, with episodes
available studies do not show DUSP4 change in the stress paradigm in of mania and depression. Bipolar disorder is a complex disorder with
animals, like the prenatal stress and depression model of neonatal high estimated heritability. Despite the accumulating evidence of the
treatment with clomipramine. These unchanged results might be due to etiology of bipolar disorder, the underlying biological mechanisms that
the limitations of the stress paradigm in animals. Besides, even though give rise to this mood disorder remain elusive (Harrison et al., 2018).
studies show DUSP4 level is sex-dependent in human depression studies, Besides underlying genetic factors, environmental risk factors have also
no studies show that the DUSP4 level is sex-dependent in animal models been identified as being partly responsible for the onset of bipolar dis­
of depression, which creates limitations in specifying the involvement of order (Rowland and Marwaha, 2018). Bipolar disorder is associated
DUSP4 in depression. with multiple dysregulations including disturbed brain development,
neuroplasticity, and chronobiology, specifically, neurotransmitter,
3.1.4. DUSP6 and depression neurotrophic factors, neuroinflammation, autoimmunity, cytokines,
In post-mortem brain tissue, DUSP6 mRNA was shown to be differ­ stress axis activity, oxidative stress, and mitochondrial dysfunctions
entially expressed in the ventromedial prefrontal cortex of patients with (Sigitova et al., 2017).
depressive disorder in a sex-specific manner (Labonte et al., 2017). Be­
sides, the DUSP6 substrates, phospho-ERK1/2, showed elevated protein 3.2.1. DUSP2 and bipolar disorder
expression in the prefrontal cortex of female patients with depressive It has been observed that patients diagnosed with bipolar disorder
disorder (Labonte et al., 2017). Downregulation of DUSP6 mediated by have increased proportions of monocytes in the blood or cerebrospinal
virus injection in the ventromedial prefrontal cortex of chronic stressed fluid (CSF) compared to controls (Barbosa et al., 2014; Jakobsson et al.,
induced depressive-like phenotype in female mice, but not male mice. 2015). It has also been found that the monocytes of patients with bipolar
The overexpression of DUSP6 mediated by a Herpes simplex virus vector disorder and the offspring of bipolar parents show aberrant levels in
rescued the depressive-like behavior in female mice (Labonte et al., DUSP2 mRNA expression. For instance, patients with bipolar disorder
2017). This sex-differential response of the stressed mice suggests that taking medication show elevated blood mRNA levels of DUSP2
DUSP6 exerts a sex-specific role in the stress response, possibly via an compared to healthy controls. DUSP2 carries a strong correlation to the
interaction with sex-specific hormones. mRNA expression of inflammatory cytokines (Drexhage et al., 2010).
Moreover, viral-mediated downregulation of DUSP6 was accompa­ DUSP2 mRNA expression has been shown to be significantly higher in
nied by increased phosphorylated ERK1/2 levels in the ventromedial monocytes of patients with mood disorder compared to healthy controls
prefrontal cortex of stressed female mice compared to control (Labonte (Padmos et al., 2008). Furthermore, lithium carbonate- and
et al., 2017). However, the total ERK1/2 protein levels in the antipsychotic-treated patients with bipolar disorder exhibited lower
DUSP6-downregulated female stressed mice remain unchanged in the levels of expression of DUSP2 mRNA in monocytes compared to
ventromedial prefrontal cortex compared to control. These findings non-lithium- and non-antipsychotic-treated patients with bipolar dis­
resemble findings observed in post-mortem brain tissue analyses of order (Padmos et al., 2008). Thus, DUSP2 may be an attractive target for

10
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

further analyses in patients with bipolar disorder. Interestingly, PTENm3m4 animals present an enlarged corpus callosum,
white matter abnormalities, and impaired learning and memory pro­
3.2.2. DUSP6 and bipolar disorder cesses (Frazier et al., 2015).
There is a positive association between the DUSP6 gene and patients
with bipolar disorder (Kim et al., 2012; Lee et al., 2006). A genetic study 3.3.2. DUSP15 and ASD
including 160 patients with a diagnosis of schizophrenia, 132 patients The analysis of peripheral blood from 255 children affected by ASD
with bipolar disorder, and 336 healthy controls, indicated that the G and 427 healthy controls revealed that DUSP15 could be a susceptibility
allele of the T/G polymorphism of the DUSP6 gene was significantly biomarker for ASD (Tian et al., 2017). Additionally, recurrent identical
more common in patients with bipolar disorder than controls. However, de novo mutations of DUSP15 were found via exome sequencing using
there was no difference between schizophrenia patients and controls. 175 samples from ASD cases and their parents (Neale et al., 2012).
This contrast suggests a specific association of the DUSP6 gene with Therefore, DUSP15 seems to carry a peculiar role in ASD risk and should
bipolar disorder but not with schizophrenia (Lee et al., 2006). However, be further investigated as a potential biomarker for ASD in children.
this association was not observed in male patients with bipolar disorder
(Kim et al., 2012), hence suggesting a sex-specific effect. 3.4. Schizophrenia
The lower expression of DUSP6 gene observed in postmortem brain
samples of patients with bipolar disorder has been found to show sex- Schizophrenia is a debilitating disease affecting various daily func­
specificity, with a reduced level of mRNA transcripts expression in fe­ tions, including self-care, social aspects, and occupational functions
male but not male patients with bipolar disorder (Struyf et al., 2008). (Mueser and McGurk, 2004; van Os et al., 2010). The symptoms of
Additionally, in vitro studies demonstrated functional Leu114Val and schizophrenia include hallucination, delusion, bizarre behavior, anhe­
Ser144Ala polymorphisms in DUSP6 blunted the effects of lithium on donia, and concentration problems (Andreasen and Olsen, 1982). Evi­
ERK1/2 activation by using SH-SY5Y human neuroblastoma cells dence shows that the factors contributing to schizophrenia include
infected with recombinant adenoviruses (Kim et al., 2012). Other evi­ genetic factors, early environmental influences, and social factors (e.g.
dence has suggested that DUSP6 may be linked with bipolar disorder, poverty) (Mueser and McGurk, 2004; van Os et al., 2010). Schizophrenia
possibly via the involvement of the ERK pathway and circadian rhythm is a complex disease affecting 1–3 % of the population. It is also
dysregulation (Lee et al., 2006). considered among the top ten causes of disability worldwide (Mueser
Also, DUSP6, a negative regulator of ERK1/2, has been linked to the and McGurk, 2004). Dysfunction of dopaminergic neurotransmission
disruption of the circadian rhythm in cell cultures of fibroblasts derived and synaptic function seems to contribute to psychotic symptoms and
from patients with bipolar disorder. The knock-down of DUSP6 in these abnormalities of neuronal connectivity, respectively (Owen et al.,
fibroblasts has been shown to reverse lithium-induced increases in 2016). Current first-line treatment mainly includes the administration of
amplitude of circadian rhythm. That being said, the inability of lithium antipsychotic drugs (such as chlorpromazine and haloperidol) combined
to regulate circadian rhythms in bipolar disorder may reflect reduced with psychotherapy, social support, and rehabilitation (Owen et al.,
ERK activity, which is partially regulated by DUSP6 (McCarthy et al., 2016).
2016). Thus, DUSP6 may be playing a crucial role in regulating circa­
dian rhythms and in the onset and course of bipolar disorder (Kim et al., 3.4.1. DUSP4 and schizophrenia
2012). A post-mortem brain study demonstrated decreased DUSP4 protein
levels in the cerebellar vermis of the patient with schizophrenia
3.3. Autism spectrum disorders (ASD) compared to control (Kyosseva et al., 1999). This is particularly inter­
esting given the increasing evidence showing the involvement of the
Autism spectrum disorder is a developmental disorder characterized cerebellum in psychiatric disorders, mainly schizophrenia. As substrates
primarily by a lack of social reciprocity accompanied by repetitive of DUSP4, ERK protein levels also seem to be disrupted in the post­
behavior such as stereotypical or repetitive motor movements. The mortem brains of schizophrenia patients. For instance, ERK1 protein
heritability of ASD is considerably high, and common genetic variants expression is reduced in the prefrontal cortex, while ERK2 is elevated in
have been shown to play a role in conferring risk to ASD (Grove et al., the thalamus (Hirayama-Kurogi et al., 2017).
2019). Besides, indirect evidence suggests a contribution of environ­
mental factors in interaction with genetic factors in the development of 3.4.2. DUSP22 and schizophrenia
ASD (Chaste and Leboyer, 2012). The hypermethylation of the DUSP22 promoter has been reported in
a recent study investigating genetic vulnerability to schizophrenia (Boks
3.3.1. PTEN and ASD et al., 2018). The blood and brain tissues of patients exhibited significant
PTEN gene mutations have been shown to be risk factors for ASDs hypermethylation at the DUSP22 gene promoter. Furthermore, the
associated with macrocephaly (Butler et al., 2005; Buxbaum et al., 2007; DUSP22 gene promoter showed higher DNA methylation levels in the
Herman et al., 2007). This specific type of ASD associated with macro­ famine-exposed schizophrenia patients compared to non-famine
cephaly is termed PTEN-ASD, typically characterized by reduced levels exposed groups. Thus, famine seemed to be a susceptible factor in the
of PTEN protein expression in conjunction with increased brain size and onset of schizophrenia. In an in vitro model of famine, nutritionally
cognition impairment (Frazier et al., 2015). In the last decades, PTEN deprived patient-derived fibroblasts showed hypermethylated DUSP22.
mutation frequencies in PTEN-ASD has been reported in ten human These results suggest an association between epigenetic changes on the
studies (Tilot et al., 2015). DUSP22 gene and increased susceptibility to the impact of an environ­
Moreover, PTEN loss in mice leads to alterations in synapses and mental risk factor on mental disorders (Boks et al., 2018). Although no
cytoarchitecture (Tilot et al., 2015). Ablation of PTEN in neural stem correlation was found to exist between DNA methylation and gene
cells in the subgranular zone of the hippocampus of mice leads to expression of the DUSP22, these results suggest changes in gene
increased proliferation and differentiation rate of the stem cells, which expression regulation of DUSP22, in response to extreme conditions like
later developed into hypertrophied neurons (Amiri et al., 2012). Several famine, may moderate or mediate risk for schizophrenia. Additionally,
mouse models characterized by PTEN deficiency or dysfunction show the hypermethylation of DUSP22 in the blood and brain of schizophrenia
autism-like behaviors, including social deficit and repetitive behavior patients that were not exposed to famine, also suggests that dysregula­
(Knafo and Esteban, 2017). One of these models, PTENm3m4, exhibits the tions in DUSP22 methylation are associated with underlying mecha­
same disrupted genes as those in human ASD (Tilot et al., 2016), nisms in the onset and development of schizophrenia. Further research
including genes related to myelination such as myelin basic protein. on the involvement of DUSP22 in neuronal development is required to

11
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

strengthen this association and better understand the changes that are medial prefrontal cortex of rats trained to self-administer heroin
brought by aberrant DUSP22 methylation. following 14 d of abstinence and the 90-minute extinction session
compared to the saline-treated controls. The importance of drug-seeking
3.4.3. MTMR2, MTMR9 and schizophrenia behavior and memory of previous drug-taking sessions suggest that such
MTMR9 mRNA level was found to be reduced approximately two- genes may be essential for relapse (Kuntz-Melcavage et al., 2009).
fold in peripheral blood lymphocytes from patients with schizophrenia DUSP5 mRNA expression level was significantly lower (0.67-fold) in the
compared to healthy controls (Bowden et al., 2006). Furthermore, nucleus accumbens following heroin treatment than after saline treat­
MTMR2 mRNA was shown to be lower in the superior temporal cortex of ment in mice (Seleman et al., 2014). This study indicated that DUSP5
patients diagnosed with schizophrenia as compared to controls without might be linked to both heroin exposure as well as to drug-seeking
a schizophrenia diagnosis (Schmitt et al., 2012). These results suggest behavior.
that further research into MTMR9 is required to investigate whether it
would eventually be qualified as a biomarker for the diagnosis of 3.6.2. DUSP1, DUSP6 and methamphetamine abuse
schizophrenia. DUSP1 and DUSP6 mRNAs were shown to be increased in the rat
striatum, thalamus, cortex, striatum, and hippocampus after single
3.5. Post-traumatic stress disorder (PTSD) methamphetamine administration (Ujike et al., 2002). Besides, another
study found that DUSP1 mRNA and protein levels were increased in a
PTSD symptoms include intrusions, avoidance/numbing, hyper­ range of rat brain regions, including the cortex, the striatum, and the
arousal, sensitization to stressors, and negative alterations in cognitions thalamus, after administration of methamphetamine (Takaki et al.,
and mood. The cause of the PTSD remains elusive, and various factors 2001). Elevated levels of DUSP6 mRNA have been observed in the cor­
have an impact on the pathology of PTSD, including genetic factors, tex, striatum, and hippocampus of rats after acute methamphetamine
trauma exposure, and interaction between gene and environment administration (Takaki et al., 2001). Together these results suggest the
(Mehta and Binder, 2012; Yehuda et al., 2015). The global burdens of involvement of DUSP genes not only in inferring vulnerability to sub­
PTSD in public health are substantial because the symptoms lead to stance abuse disorders but also in the development and exacerbation of
impaired functions in several aspects of one’s life, including health, this addictive disorder.
social, and professional life (Watson, 2019). Given the complexity of the
pathology of PTSD, the treatment of PTSD consists of various methods, 4. DUSP genes and neurological disorders
including pharmacologic approaches (Friedman and Bernardy, 2017;
Horn et al., 2016), psychotherapies (Steenkamp et al., 2015), and Accumulating evidence from the current literature on the link be­
mindfulness (Lang, 2017). tween DUSP genes and neurological disorders such as Alzheimer’s dis­
ease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), and
3.5.1. DUSP22 and PTSD epilepsy, is described in the following section. The summary of the
Differential DNA methylation at the DUSP22 gene is receiving pathological implication of DUSP family members in neurological dis­
increased attention to psychiatric disorders such as schizophrenia and orders is presented in Table 2.
PTSD. A recent longitudinal study in a Dutch military cohort identified
changes in DNA methylation at several differentially methylated posi­
tions, including DUSP22 (Rutten et al., 2018). Decreased DNA methyl­ 4.1. Alzheimer’s disease (AD)
ation around the DUSP22 gene was linked to increased PTSD symptoms
(Rutten et al., 2018). Although this observation was not replicated in an AD is characterized by progressive memory loss, impairments in
independent replication study, a thorough understanding of the role of cognition, and neuropsychiatric disturbances such as mood and per­
epigenetic changes around the DUSP22 gene in the face of extreme sonality changes, anxiety, and aggression (Scheltens et al., 2016). The
environmental conditions like traumatic stress and famine is required. prevalence of AD is 10–30 % in individuals over 65 years, with incidence
doubling every ten years after 60 (Eratne et al., 2018; Masters et al.,
3.6. Substance abuse disorders 2015). The pathophysiological hallmarks of AD are the accumulation of
extracellular amyloid-beta (Aβ) plaques and intracellular neurofibrillary
Substance abuse disorders represent complex behaviors or symptoms tangles of hyperphosphorylated-tau that lead to neuronal loss due to
involved in compulsive drug-seeking, including impaired control of neurotoxic effects. So far, there is no cure for AD, and the current
substance use and social interactions (Volkow et al., 2016). Low affect, treatment options have been shown to only slow down its progression
behavior self-regulation and interaction with social factors during child (Rafii, 2013).
development predispose to substance abuse in adolescence (Tarter,
2002). Besides, genetic factors and their interaction with the environ­ 4.1.1. DUSP1 and AD
ment contribute to the etiology of substance abuse and related behaviors The protein level of DUSP1 decreased in the hippocampus and
(McGue et al., 1996). The pathophysiology of substance abuse involves temporal cortex of patients diagnosed with AD compared to aged match
repeated perturbation of reward circuits, abnormalities in the prefrontal healthy controls (Du et al., 2019). Similarly, DUSP1 protein levels were
cortex, and molecular neuro-adaptations (Koob et al., 2008). The decreased in the hippocampus of APP/PS1 transgenic AD model mice at
treatment usually includes medication in combination with social and the age of 9 months compared to control mice (Du et al., 2019). DUSP1
psychotherapy (Tenegra and Leebold, 2016). inhibits the amyloidogenic process through the ERK/MAPK signaling
pathway, and DUSP1 reduces Aβ generation and plaque formation and
3.6.1. DUSP5, DUSP6, DUSP27 and heroin abuse alleviates synaptic and cognitive impairments in APP/PS1 mice (Du
Heroin impairs synaptic plasticity in neural projections extending et al., 2019). In PC12 cell culture, DUSP 1 mitigates Aβ-induced
from the prefrontal cortex to the nucleus accumbens in animals trained apoptosis, oxidative stress, and neuroinflammation by inhibiting the
to self-administer the drug. In human studies, the variant rs950302 of JNK signaling pathway, thereby playing a neuroprotective role. From
DUSP27 gene was significantly associated with heroin addiction the animal and cell culture studies, DUSP1 alleviates amyloid
vulnerability in African Americans (Nielsen et al., 2010). This gene beta-induced neurotoxicity (Gu et al., 2018). The results suggest that
encodes a recently identified member of the cytosolic dual-specificity DUSP1 impairment facilitates the pathogenesis of AD, whereas the
phosphatase family, which may be involved in energy metabolism. upregulation of DUSP1 plays a neuroprotective role to reduce Alzheimer
DUSP5 and DUSP6 mRNA expression increased by at least 1.4-fold in the related phenotypes (Du et al., 2019).

12
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

4.1.2. DUSP26 and AD postmortem brains and controls (Rickle et al., 2006). Instead, the ratio of
DUSP26 protein was observed to be elevated in the hippocampal of Ser380 p-PTEN / total PTEN protein reduced in temporal cortical ho­
postmortem brain tissue of patients with AD compared to controls (Jung mogenates in AD compared to control (Rickle et al., 2006), which
et al., 2016). In the APP-expressing HEK293/APP695 cell lines, over­ indicated reduced PTEN phosphorylation at residue Ser380 in AD.
expression of DUSP26 increased Aβ42 levels by twofold (Jung et al., Decreased PTEN and increased tau phosphorylation were evident in
2016). In contrast, the enzymatically inactive mutant of DUSP26 failed frontal cortex brain slices of AD (Zhang et al., 2006). This suggests that
to induce Aβ oligomers or APP processing. Thus, DUSP26 has been PTEN phosphorylation is involved in AD pathology and that PTEN dy­
linked to Aβ generation and APP processing in these cell lines (Jung namics in AD brain might be region dependent.
et al., 2016), thereby suggesting the involvement of DUSP26 in the PTEN contributes to AD pathology in animal and cell culture models
pathophysiology of AD. (Frere and Slutsky, 2016). Overexpression of PTEN induces synaptic
depression, similar to Aß-induced depression in transgenic mice (Knafo
4.1.3. SSH1, SSH3 and AD et al., 2016). Additionally, the overexpression of the PTEN protein in
SSH1 protein level showed a 45 % significant reduction in the frontal Chinese hamster ovary cells reduces tau phosphorylation (Kerr et al.,
cortex of patients diagnosed with AD compared to healthy controls. This 2006). Another study demonstrated that overexpression of PTEN de­
reduction was accompanied by unchanged pSSH1 levels, which led to creases the formation of tau aggregates in COS-7 cells.
the significant decrease of the pSSH1/SSH1 ratio indicative of the In contrast, the phosphatase-null or inactive PTEN increases tau
inactivation of SSH1 in human AD (Barone et al., 2014). Besides, SSH3 aggregation in rat cortical primary neurons transfected with a mutant
mRNA was shown to be differentially expressed in the hippocampus, the form of PTEN (Zhang et al., 2006). The loss of PTEN causes neuro­
temporal and frontal cortex, as well as in the whole blood of patients degeneration through the hyperphosphorylation of tau and neurofila­
diagnosed with AD in comparison with healthy controls (Wang and ments in mouse cerebellar neurons (Nayeem et al., 2007). However,
Wang, 2020). induction of PTEN is accompanied by okadaic acid-induced tau phos­
SSH phosphatase dephosphorylates cofilin, and the reduced protein phorylation, while the knockdown of PTEN reduced tau hyper­
expression and inactivation of SSH1 were further accompanied by an phosphorylation in SH-SY5Y neuroblastoma cells, and increased cell
increase in cofilin1 phosphorylation/inactivation in human and animal proliferation and survival. Inhibition of PTEN reduces tau phosphory­
studies (Barone et al., 2014). Hyperphosphorylation of cofilin can result lation in SH-SY5Y neuroblastoma cells (Chen et al., 2012). Furthermore,
in tau pathology, which can be induced by Aβ oligomers (Kang et al., inhibition of PTEN via intracerebroventricular delivery of a PTEN in­
2011). hibitor, VO-OHpic, rescued Aβ42-induced impairment in both basal
synaptic transmission and LTP, as well as spatial learning tasks, in
4.1.4. DUSP22 and AD APP/PS1 transgenic mouse model of AD (Knafo et al., 2016).
Analyses of blood/brain samples indicated that methylation of the The PDZ-binding domain of PTEN is central to Aβ-induced synaptic
region in DUSP22 correlated linearly and powerfully with the Braak toxicity and cognitive dysfunction (Knafo et al., 2016). Deletion of this
stages of neuropathology, an index of AD progression (Pearson corre­ PDZ-binding domain results in resistance to Aβ toxicity in postsynaptic
lation coefficient R = 0.95, P < 0.05) (Sanchez-Mut et al., 2014). In the neurons (Knafo et al., 2016). Overexpression of PTEN reduced tau
hippocampus of patients diagnosed with AD, hypermethylation of the phosphorylation in Chinese hamster ovary cells (Kerr et al., 2006). Thus,
DUSP22 promoter and decreased protein expression of DUSP22 as taken together, these findings indicate that PTEN may be centrally
compared to age-matched controls have been reported (Sanchez-Mut involved in moderating or mediating AD’s pathophysiology.
et al., 2014). Reductions in DUSP22 levels may lead to increased tau
phosphorylation due to weaker inhibitory control of protein kinase 4.2. Parkinson’s disease (PD)
A-mediated tau-phosphorylation, at least as suggested by findings in
neuronal cell lines (Sanchez-Mut et al., 2014). Besides, it has been found PD is a progressive disease with motor and non-motor symptoms,
in SK-N-BE(2) cell culture studies that are depleting DUSP22 through which consist of slow movements, tremors, rigidity, impaired balance
small hairpin RNA’s resulted in a higher survival capability than cells during walking, various disturbances in autonomic functions with
with control or healthy and overexpression of DUSP22 (Sanchez-Mut orthostatic hypotension, constipation, sleep disturbances, and a spec­
et al., 2014). It would be interesting for future studies on AD patho­ trum of neuropsychiatric symptoms (Sveinbjornsdottir, 2016). The
physiology to establish the regulatory role of DUSP22 in tau phos­ cause of PD is unknown but is believed to involve both genetic and
phorylation, Aß accumulation, and neuronal death AD. environmental factors (Kouli et al., 2018). Parkinson’s disease affects
1% of the population above 60 years old and is more frequently prev­
4.1.5. PTEN and AD alent in men than women (Tysnes and Storstein, 2017). The main hall­
AD is associated with excessive recruitment of PTEN into synapses, marks of PD are Lewy bodies and the degeneration of dopaminergic
leading to aberrant synaptic depression (Knafo and Esteban, 2017). It neurons in the substantia nigra. However, recent findings suggest that
has furthermore been observed that layer III of the temporal cortex in PD’s pathophysiology is heterogeneous. Other protein aggregates like
patients with AD showed a 15 % loss of PTEN immunoreactive neurons α-synuclein, DJ-1 (Waragai et al., 2010), tau, and β-amyloid also play a
compared to controls, while the majority of the layer III temporal cortex role (Parnetti et al., 2013) in the onset and progression of the disease.
were PTEN immunoreactive in control cases (Rickle et al., 2006). PTEN
protein levels decreased in the AD temporal cortex compared with 4.2.1. PTEN and PD
matched controls, and PTEN level has been negatively correlated with Protein level of nuclear PTEN has been shown to be significantly
the severity of neurofibrillary pathology or senile plaques (Griffin et al., increased by 5.6-fold in the substantia nigra of PD brain compared to the
2005). age-matched controls. PTEN downstream regulators, PI3K regulatory
It has been observed in AD, that PTEN delocalizes from the nucleus to subunit p85, PIP3, and Akt1/2/3 protein levels decreased by two-fold in
the cytoplasm and intracellular neurofibrillary tangles in postmortem nuclear of substantia nigra region of PD brain samples compared to the
brain tissues (Sonoda et al., 2010). The nuclear PTEN immunoreactivity age-matched controls (Sekar and Taghibiglou, 2018).
reduced in neurons of the CA1, subiculum, and entorhinal cortex of AD Downregulation of PTEN inhibits elevated levels of intracellular
cases, while the PTEN immunoreactivity increased in apical dendrites in reactive oxygen species and neuronal death in rat hippocampal and in
the CA1 and subiculum in AD cases compared with control (Griffin et al., human dopaminergic SH-SY5Y neurons caused by neurotoxin 1-methyl-
2005). However, in the temporal cortex, PTEN protein levels were not 4-phenylpyridinium iodide toxicity which mimic Parkinson’s disease
significantly different in either nucleus or membrane fractions in AD (Zhu et al., 2007). PTEN deletion in adult dopaminergic neurons protects

13
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

these neurons from 6-hydroxydopamine (6-OHDA) neurotoxicity and neuroprotective effects (Taylor et al., 2013). These findings suggest that
restores striatal dopamine levels in mouse models of PD (Domanskyi this DUSP1 regulated pathway may represent a novel candidate as a
et al., 2011). therapeutic target in HD.
PTEN has been involved in response to DNA damage repair in PD
(Ogino et al., 2016). Specific defects in DNA impact the dopaminergic 4.3.2. PTEN and HD
system and are associated with PD pathology in both cell and animal Elevated PTEN expression, together with amplification of BDNF
models (Sepe et al., 2016). Thus, PTEN does play a role in DNA damage signaling, seems to result in neuroplasticity abnormalities in the indirect
in PD. pathway of the spiny projection neurons from brain slice of the BACHD
mouse model and Q175 knock-in mouse model of HD (Plotkin and
4.2.2. DUSP1 and PD Surmeier, 2014). Furthermore, plasticity and LTP aberrations were
There are currently few studies on DUSP1 in patients with PD. rescued by inhibiting PTEN in indirect pathway spiny projection neu­
However, one study reported decreased DUSP1 mRNA expression in the rons of a transgenic HD mouse model (Plotkin et al., 2014).
dorsolateral prefrontal cortex of idiopathic PD patients (Renton, 2010). P53 has been shown to transiently upregulate PTEN protein and
In cell culture studies using neuronal PC12 cells, Serial Analysis of mRNA levels in medium spiny neurons in the striatum of TIF-IA knock-
Gene Expression-based study showed that acute (8 h) exposure to 6- out mice, a transgenic model of HD (Kreiner et al., 2013). TIF-IA inac­
OHDA, a dopaminergic neurotoxin commonly used to induce PD-like tivation leads to nucleolar disruption, which is also a common finding in
symptoms in experimental studies (Crotty et al., 2008; Walsh et al., HD pathology (Parlato and Kreiner, 2013). TIF-IA deletion, together
2011), induced a 35-fold increase of DUSP1 mRNA levels (Ryu et al., with PTEN or p53 deletion in mice, exhibited increased levels of
2005). Moreover, DUSP1 mRNA was transiently upregulated in the SN 4 apoptotic cells in the striatum (Kreiner et al., 2013). However, TIF-IA
days post-6-OHDA administration in the medial forebrain bundle lesion deletion combined with PTEN and p53 deletion did not show
model in rats (Collins et al., 2014). Besides, DUSP1 mRNA expression increased apoptotic cells in the striatum. Thus, PTEN and p53 prolong
was increased in rat striatum treated with 6- OHDA followed by injec­ neuronal survival upon nucleolar disruption (Kreiner et al., 2013). The
tion of SKF38393, a selective dopamine receptor D1 agonist (Berke upregulation of PTEN impairs kinase mammalian/mechanistic target of
et al., 1998). DUSP1 has been shown to promote the growth and elab­ rapamycin function in medium spiny neurons (Kreiner et al., 2013).
oration of dopaminergic neuronal processes, protecting them from the Future studies investigating PTEN regulation will contribute to un­
neurotoxic effects of 6-OHDA (Collins et al., 2013). This study indicates derstanding the etiology of HD and to the development of new thera­
that DUSP1 may have a neuroprotective effect, at least in PD rodent peutic strategies targeting PTEN.
models (Collins et al., 2013).
4.4. Epilepsy
4.3. Huntington’s disease (HD)
Epilepsy is characterized by recurrent unprovoked seizures and ac­
HD is a devastating heritable neurological disease characterized counts for the highest disability-adjusted life year rates among neuro­
primarily by progressive motor and cognitive impairments, as well as logical disorders (Beghi, 2016). Epilepsy is caused by various factors,
psychiatric symptoms, including affective disorder symptoms that often including genetic influence, head trauma, brain disease like brain tu­
precede other symptoms (Julien et al., 2007). HD is typically induced by mors or strokes, infectious diseases, prenatal injury, or developmental
a highly polymorphic CAG trinucleotide repeat expansion in exon-1 of disorders. Epilepsy affects more than 65 million people worldwide
the gene encoding the huntingtin protein (Bano et al., 2011). Huntingtin (Singh and Trevick, 2016). Gliosis, imbalance of ion and water ho­
protein is widely expressed during development and exhibits a complex meostasis, increased extracellular glutamate, altered neural circuits,
and dynamic distribution within cells (Schulte and Littleton, 2011). damaged blood-brain barrier are omnipresent in epilepsy in animal
Besides genetic factors, cerebral vitamin B5 deficiency is a potential models or patients (Patel et al., 2019). Antiepileptic drugs are the
cause of HD (Patassini et al., 2019). Most European populations show a first-line treatment, but alternative treatments, including surgical
relatively high prevalence (4–8 per 100,000), but HD is notably rare in resection of the seizure focus, ketogenic diets, vagus nerve stimulators,
Finland and in Japan (Harper, 1992). The underlying neuropathology is and implantable brain neurostimulators are available for patients with
characterized by neuronal loss, striatum microglial activation, and seizures that are not controlled with medication (Liu et al., 2017).
neuro-inflammation within the striatum (Novak and Tabrizi, 2011;
Reiner et al., 2011). Many genetically modified animal models of HD 4.4.1. Laforin and Lafora disease (LD)
recapitulate some of the pathophysiological features observed in LD is a progressive neurological disorder characterized by intrac­
humans, but many drug designs based on animal models of HD have table myoclonic seizures, emotional disturbance, and cognitive decline.
failed in clinical trials (Menalled and Brunner, 2014). Hallmarks of this disease are primarily attributed to the accumulation of
hyperphosphorylated insoluble poly-glucosan called Lafora bodies
4.3.1. DUSP1 and HD (LBs). LBs are caused by mutations in either the atypical DUSP Laforin,
Reduced DUSP1 levels were observed in animal models of HD EPM2A, or NHLRC1. Laforin genes encode phosphatases that dephos­
(Collins et al., 2015), and enhancing DUSP1 expression has been shown phorylate glycogen. Glycogen, a potent energy storage molecule in an­
to significantly reduce neuronal cell death in HD animal models induced imals, is degraded by glycogen phosphorylase and glycogen
by lentiviral infection and expression of a mutated huntingtin gene. In a debranching enzyme. In the brain, the accumulation of glycogen in
cell culture studies comprising primary striatal rat neurons exposed to a neurons can lead to neuronal loss, locomotive defects, and neuro­
pathological construct comprising of the N-terminal fragment of degeneration in mice and Drosophila. As such, a reduction of glycogen
polyglutamine-expanded huntingtin (Htt171–82Q), overexpression of synthesis may prevent LBs formation and subsequent neurodegeneration
DUSP1 inhibited apoptosis (Taylor et al., 2013). and seizure susceptibility, thus preventing LD progression (Duran and
This DUSP1-mediated neuroprotection has been suggested to be Guinovart, 2015; Kecmanovic et al., 2016; Roach, 2015).
dependent on the activity of phosphatases and occur through direct
regulation of JNKs and p38 s (Taylor et al., 2013). Mutant DUSP1 4.4.2. PTEN and epilepsy
selectively targeting JNK or p38, preserves significantly fewer PTEN mutations have been observed in patients with epilepsy and a
NeuN-positive cells in primary striatal neurons exposed to Htt171–82Q variety of comorbidities, including cancer (de Souza et al., 2015; Schick
fragments than in wild-type DUSP1 primary neuron HD models, indi­ et al., 2006). PTEN sequence analysis performed on a case where the
cating that dual targeting of JNK and p38 by DUSP1 may exert individual suffered from both epilepsy and Cowden syndrome, an

14
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

inherited disorder characterized by noncancerous growths, identified a King Faisal University, Employees Scholarship Program from the Saudi
heterozygous missense mutation in PTEN (Ueno et al., 2019). Resection Arabian Government (no. 1026374049). Work from Dr. Markus Roth­
of high-grade glioma tissue from patients with seizures exhibited ermel was supported by the DFG (RO4046/2-1 and /2-2, Emmy Noether
reduced PTEN expression compared to patients with glioma without Program; GRK2416). Work from B.P.F. Rutten was funded by a VIDI
seizures (Yang et al., 2016). award (no. 91718336) from the Netherlands Scientific Organization.
PTEN mutations have also been reported and used in several animal
models of cortical dysplasia, which is also a contributor to epilepsy in Appendix A. The Peer Review Overview and Supplementary
adults (Elia et al., 2012). For instance, the inhibition of PTEN rescued data
neuronal death in a mouse model of temporal lobe epilepsy, implying an
excitotoxic role for PTEN. This inhibition has also been shown to exert The Peer Review Overview and Supplementary data associated with
potent anti-inflammatory and neuroprotective effects (Grande et al., this article can be found in the online version, at doi:https://doi.org/10
2014). These results confer additional support to the role of PTEN in .1016/j.pneurobio.2020.101906.
neuronal dysfunction.
References
4.4.3. DUSP1 and epilepsy
A study on kainic acid-induced limbic seizures in rats reported that Akbarian, S., Davis, R.J., 2010. Keep the’ phospho’ on MAPK, be happy. Nat. Med. 16,
1187–1188.
DUSP1 protein expression was transiently induced in the dentate Amiri, A., Cho, W., Zhou, J., Birnbaum, S.G., Sinton, C.M., McKay, R.M., Parada, L.F.,
granule cells of the hippocampus, outer layers of the neocortex, and 2012. Pten deletion in adult hippocampal neural stem/progenitor cells causes
neurons of the lateral nucleus of the bed of the stria terminalis in rat with cellular abnormalities and alters neurogenesis. J. Neurosci. 32, 5880–5890.
Andreasen, N.C., Olsen, S., 1982. Negative v positive schizophrenia. Definition and
the seizures as compared to untreated controls (Gass et al., 1996). On a validation. Arch. Gen. Psychiatry 39, 789–794.
subcellular level, DUSP1 colocalizes with its substrate, MAP kinase, in Bano, D., Zanetti, F., Mende, Y., Nicotera, P., 2011. Neurodegenerative processes in
neuronal nuclei, which has been linked to inhibition of the seizure Huntington’s disease. Cell Death Dis. 2, e228.
Barbosa, I.G., Rocha, N.P., Assis, F., Vieira, E.L., Soares, J.C., Bauer, M.E., Teixeira, A.L.,
response in an animal model of temporal lobe epilepsy (Gass et al., 1996;
2014. Monocyte and lymphocyte activation in bipolar disorder: a new piece in the
Tai et al., 2017). Given these results, DUSP1 induction seems to have a puzzle of immune dysfunction in mood disorders. Int. J. Neuropsychopharmacol. 18.
partial role in the inhibition of MAP kinase activity following seizures. Barford, D., Das, A.K., Egloff, M.P., 1998. The structure and mechanism of protein
phosphatases: insights into catalysis and regulation. Annu. Rev. Biophys. Biomol.
Struct. 27, 133–164.
5. Conclusions Barone, E., Mosser, S., Fraering, P.C., 2014. Inactivation of brain Cofilin-1 by age,
Alzheimer’s disease and gamma-secretase. Biochim. Biophys. Acta 1842,
This review provides an extensive overview of research findings of 2500–2509.
Bayón, Y., Alonso, A., 2010. Atypical DUSPs: 19 phosphatases in search of a role.
the role of several DUSP genes and their involvement in the onset and Emerging Signaling Pathways in Tumor Biology, pp. 185–208.
development of certain mental and neurological disorders. The body of Beghi, E., 2016. Addressing the burden of epilepsy: many unmet needs. Pharmacol. Res.
literature on DUSP genes does show the great diversity of biological 107, 79–84.
Begley, M.J., Dixon, J.E., 2005. The structure and regulation of myotubularin
processes in which DUSP genes are involved, and it is therefore not phosphatases. Curr. Opin. Struct. Biol. 15, 614–620.
surprising that genomic variations in DUSP genes have been linked to Berke, J.D., Paletzki, R.F., Aronson, G.J., Hyman, S.E., Gerfen, C.R., 1998. A complex
several mental and somatic disorders. program of striatal gene expression induced by dopaminergic stimulation.
J. Neurosci. 18, 5301–5310.
It is furthermore noteworthy that the first wave of epigenetic studies Bermudez, O., Pages, G., Gimond, C., 2010. The dual-specificity MAP kinase
has identified epigenetic changes in particularly one DUSP gene, i.e., phosphatases: critical roles in development and cancer. Am. J. Physiol., Cell Physiol.
DUSP22, to be linked to a range of mental as well as neurological dis­ 299, C189–202.
Bermudez Brito, M., Goulielmaki, E., Papakonstanti, E.A., 2015. Focus on PTEN
orders, i.e., altered methylation in the DUSP22 gene has been observed regulation. Front. Oncol. 5, 166.
in PTSD patients, in patients diagnosed with schizophrenia and in AD Bessette, D.C., Qiu, D., Pallen, C.J., 2008. PRL PTPs: mediators and markers of cancer
patients as compared to controls. progression. Cancer Metastasis Rev. 27, 231–252.
Boks, M.P., Houtepen, L.C., Xu, Z., He, Y., Ursini, G., Maihofer, A.X., Rajarajan, P.,
The current state of the literature is nevertheless in a very early stage,
Yu, Q., Xu, H., Wu, Y., Wang, S., Shi, J.P., Hulshoff Pol, H.E., Strengman, E.,
and our review of the literature indicated a plethora of findings that are Rutten, B.P.F., Jaffe, A.E., Kleinman, J.E., Baker, D.G., Hol, E.M., Akbarian, S.,
most challenging to converge into a common pathway. We did find a Nievergelt, C.M., De Witte, L.D., Vinkers, C.H., Weinberger, D.R., Yu, J., Kahn, R.S.,
more comprehensive line of converging evidence from human obser­ 2018. Genetic vulnerability to DUSP22 promoter hypermethylation is involved in
the relation between in utero famine exposure and schizophrenia. NPJ Schizophr. 4,
vational studies and experimental studies using and model systems on 16.
the DUSP family genes SSH and PTEN suggesting intricate links to Bolino, A., Muglia, M., Conforti, F.L., LeGuern, E., Salih, M.A., Georgiou, D.M.,
neuroplasticity, cellular proliferation, survival, and cellular Christodoulou, K., Hausmanowa-Petrusewicz, I., Mandich, P., Schenone, A.,
Gambardella, A., Bono, F., Quattrone, A., Devoto, M., Monaco, A.P., 2000. Charcot-
architecture. Marie-Tooth type 4B is caused by mutations in the gene encoding myotubularin-
To reiterate, it is reasonable to suggest that DUSP genes are possibly related protein-2. Nat. Genet. 25, 17–19.
involved in biological processes underlying and/or mediating the onset Bolis, A., Coviello, S., Visigalli, I., Taveggia, C., Bachi, A., Chishti, A.H., Hanada, T.,
Quattrini, A., Previtali, S.C., Biffi, A., Bolino, A., 2009. Dlg1, Sec8, and Mtmr2
and course of mental health and neurological disorders. Because can­ regulate membrane homeostasis in Schwann cell myelination. J. Neurosci. 29,
didates from epidemiological cohort studies cannot be thoroughly tested 8858–8870.
for causality in observational studies, it would be interesting to consider Bowden, N.A., Weidenhofer, J., Scott, R.J., Schall, U., Todd, J., Michie, P.T., Tooney, P.
A., 2006. Preliminary investigation of gene expression profiles in peripheral blood
targeting the next phase of experimental cell and animal studies on lymphocytes in schizophrenia. Schizophr. Res. 82, 175–183.
manipulating the expression of distinct DUSP mentioned above family Bremmer, S.C., Hall, H., Martinez, J.S., Eissler, C.L., Hinrichsen, T.H., Rossie, S.,
genes, in a cell type- and/or circuit-specific as well as temporal- specific Parker, L.L., Hall, M.C., Charbonneau, H., 2012. Cdc14 phosphatases preferentially
dephosphorylate a subset of cyclin-dependent kinase (Cdk) sites containing
manner (e.g., by combining genetic editing with optogenetics) in order
phosphoserine. J. Biol. Chem. 287, 1662–1669.
to better understand their involvement in the etiopathogenesis of Budziszewska, B., Szymanska, M., Leskiewicz, M., Basta-Kaim, A., Jaworska-Feil, L.,
neurological and mental disorders, and as potential promising thera­ Kubera, M., Jantas, D., Lason, W., 2010. The decrease in JNK- and p38-MAP kinase
peutic targets. activity is accompanied by the enhancement of PP2A phosphate level in the brain of
prenatally stressed rats. J. Physiol. Pharmacol. 61, 207–215.
Butler, M.G., Dasouki, M.J., Zhou, X.P., Talebizadeh, Z., Brown, M., Takahashi, T.N.,
Acknowledgments Miles, J.H., Wang, C.H., Stratton, R., Pilarski, R., Eng, C., 2005. Subset of individuals
with autism spectrum disorders and extreme macrocephaly associated with germline
PTEN tumour suppressor gene mutations. J. Med. Genet. 42, 318–321.
Work from Ning An was supported by the China Scholarship Council Buxbaum, J.D., Cai, G., Chaste, P., Nygren, G., Goldsmith, J., Reichert, J.,
(File No.201506750020). Work from Ghazi I. Al Jowf was supported by Anckarsater, H., Rastam, M., Smith, C.J., Silverman, J.M., Hollander, E.,

15
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Leboyer, M., Gillberg, C., Verloes, A., Betancur, C., 2007. Mutation screening of the Endo, M., Ohashi, K., Sasaki, Y., Goshima, Y., Niwa, R., Uemura, T., Mizuno, K., 2003.
PTEN gene in patients with autism spectrum disorders and macrocephaly. Am. J. Control of growth cone motility and morphology by LIM kinase and Slingshot via
Med. Genet. B Neuropsychiatr. Genet. 144B, 484–491. phosphorylation and dephosphorylation of cofilin. J. Neurosci. 23, 2527–2537.
Cai, Q.Y., Chen, X.S., Zhong, S.C., Luo, X., Yao, Z.X., 2009. Differential expression of Endo, M., Ohashi, K., Mizuno, K., 2007. LIM kinase and slingshot are critical for neurite
PTEN in normal adult rat brain and upregulation of PTEN and p-Akt in the ischemic extension. J. Biol. Chem. 282, 13692–13702.
cerebral cortex. Anat. Rec. (Hoboken) 292, 498–512. Eratne, D., Loi, S.M., Farrand, S., Kelso, W., Velakoulis, D., Looi, J.C., 2018. Alzheimer’s
Cain, E.L., Beeser, A., 2013. Emerging roles of atypical dual specificity phosphatases in disease: clinical update on epidemiology, pathophysiology and diagnosis. Australas.
cancer. Oncogene and Cancer-From Bench to Clinic, p. 93. Psychiatry 26, 347–357.
Campbell, A.M., Zhang, Z.Y., 2014. Phosphatase of regenerating liver: a novel target for Feng, P., Guan, Z., Yang, X., Fang, J., 2003. Impairments of ERK signal transduction in
cancer therapy. Expert Opin. Ther. Targets 18, 555–569. the brain in a rat model of depression induced by neonatal exposure of
Carracedo, A., Pandolfi, P.P., 2008. The PTEN-PI3K pathway: of feedbacks and cross- clomipramine. Brain Res. 991, 195–205.
talks. Oncogene 27, 5527–5541. Finelli, M.J., Murphy, K.J., Chen, L., Zou, H., 2013. Differential phosphorylation of
Caunt, C.J., Keyse, S.M., 2013. Dual-specificity MAP kinase phosphatases (MKPs): Smad1 integrates BMP and neurotrophin pathways through Erk/Dusp in axon
shaping the outcome of MAP kinase signalling. FEBS J. 280, 489–504. development. Cell Rep. 3, 1592–1606.
Chaste, P., Leboyer, M., 2012. Autism risk factors: genes, environment, and gene- Frazier, T.W., Embacher, R., Tilot, A.K., Koenig, K., Mester, J., Eng, C., 2015. Molecular
environment interactions. Dialogues Clin. Neurosci. 14, 281–292. and phenotypic abnormalities in individuals with germline heterozygous PTEN
Chen, Z., Chen, B., Xu, W.F., Liu, R.F., Yang, J., Yu, C.X., 2012. Effects of PTEN inhibition mutations and autism. Mol. Psychiatry 20, 1132–1138.
on regulation of tau phosphorylation in an okadaic acid-induced neurodegeneration Frere, S., Slutsky, I., 2016. Targeting PTEN interactions for Alzheimer’s disease. Nat.
model. Int. J. Dev. Neurosci. 30, 411–419. Neurosci. 19, 416–418.
Chesney, E., Goodwin, G.M., Fazel, S., 2014. Risks of all-cause and suicide mortality in Friedman, M.J., Bernardy, N.C., 2017. Considering future pharmacotherapy for PTSD.
mental disorders: a meta-review. World Psychiatry 13, 153–160. Neurosci. Lett. 649, 181–185.
Clifford, D.M., Chen, C.T., Roberts, R.H., Feoktistova, A., Wolfe, B.A., Chen, J.S., Gass, P., Eckhardt, A., Schroder, H., Bravo, R., Herdegen, T., 1996. Transient expression
McCollum, D., Gould, K.L., 2008. The role of Cdc14 phosphatases in the control of of the mitogen-activated protein kinase phosphatase MKP-1 (3CH134/ERP1) in the
cell division. Biochem. Soc. Trans. 36, 436–438. rat brain after limbic epilepsy. Brain Res. Mol. Brain Res. 41, 74–80.
Collins, L.M., O’Keeffe, G.W., Long-Smith, C.M., Wyatt, S.L., Sullivan, A.M., Toulouse, A., Govender, D., Chetty, R., 2012. Gene of the month: PTEN. J. Clin. Pathol. 65, 601–603.
Nolan, Y.M., 2013. Mitogen-activated protein kinase phosphatase (MKP)-1 as a Grande, V., Manassero, G., Vercelli, A., 2014. Neuroprotective and anti-inflammatory
neuroprotective agent: promotion of the morphological development of midbrain roles of the phosphatase and tensin homolog deleted on chromosome ten (PTEN)
dopaminergic neurons. Neuromolecular Med. 15, 435–446. Inhibition in a Mouse Model of Temporal Lobe Epilepsy. PLoS One 9, e114554.
Collins, L.M., Gavin, A.M., Walsh, S., Sullivan, A.M., Wyatt, S.L., O’Keeffe, G.W., Gray, C.H., Good, V.M., Tonks, N.K., Barford, D., 2003. The structure of the cell cycle
Nolan, Y.M., Toulouse, A., 2014. Expression of endogenous Mkp1 in 6-OHDA rat protein Cdc14 reveals a proline-directed protein phosphatase. EMBO J. 22,
models of Parkinson’s disease. Springerplus 3, 205. 3524–3535.
Collins, L.M., Downer, E.J., Toulouse, A., Nolan, Y.M., 2015. Mitogen-activated protein Griffin, R.J., Moloney, A., Kelliher, M., Johnston, J.A., Ravid, R., Dockery, P.,
kinase phosphatase (MKP)-1 in nervous system development and disease. Mol. O’Connor, R., O’Neill, C., 2005. Activation of Akt/PKB, increased phosphorylation of
Neurobiol. 51, 1158–1167. Akt substrates and loss and altered distribution of Akt and PTEN are features of
Crotty, S., Fitzgerald, P., Tuohy, E., Harris, D.M., Fisher, A., Mandel, A., Bolton, A.E., Alzheimer’s disease pathology. J. Neurochem. 93, 105–117.
Sullivan, A.M., Nolan, Y., 2008. Neuroprotective effects of novel Grove, J., Ripke, S., Als, T.D., Mattheisen, M., Walters, R.K., Won, H., Pallesen, J.,
phosphatidylglycerol-based phospholipids in the 6-hydroxydopamine model of Agerbo, E., Andreassen, O.A., Anney, R., 2019. Identification of common genetic risk
Parkinson’s disease. Eur. J. Neurosci. 27, 294–300. variants for autism spectrum disorder. Nat. Genet. 51, 431–444.
Cunha, M.P., Budni, J., Ludka, F.K., Pazini, F.L., Rosa, J.M., Oliveira, A., Lopes, M.W., Gu, Y., Ma, L.-J., Bai, X.-X., Jie, J., Zhang, X.-F., Chen, D., Li, X.-P., 2018. Mitogen-
Tasca, C.I., Leal, R.B., Rodrigues, A.L.S., 2016. Involvement of PI3K/Akt signaling activated protein kinase phosphatase 1 protects PC12 cells from amyloid beta-
pathway and its downstream intracellular targets in the antidepressant-like effect of induced neurotoxicity. Neural Regen. Res. 13, 1842.
creatine. Mol. Neurobiol. 53, 2954–2968. Harper, P.S., 1992. The epidemiology of Huntington’s disease. Hum. Genet. 89, 365–376.
de Souza, P.V., de Rezende Pinto, W.B., Santos, A.J., 2015. Metastatic breast cancer in a Harrison, P.J., Geddes, J.R., Tunbridge, E.M., 2018. The emerging neurobiology of
man with nonprogressive ataxia and epilepsy. Neurology 85, 1183–1184. bipolar disorder. Trends Neurosci. 41, 18–30.
Deng, Z., Yuan, C., Yang, J., Peng, Y., Wang, W., Wang, Y., Gao, W., 2019. Behavioral Herman, G.E., Butter, E., Enrile, B., Pastore, M., Prior, T.W., Sommer, A., 2007.
defects induced by chronic social defeat stress are protected by Momordica charantia Increasing knowledge of PTEN germline mutations: two additional patients with
polysaccharides via attenuation of JNK3/PI3K/AKT neuroinflammatory pathway. autism and macrocephaly. Am. J. Med. Genet. A 143A, 589–593.
Ann. Transl. Med. 7. Hirayama-Kurogi, M., Takizawa, Y., Kunii, Y., Matsumoto, J., Wada, A., Hino, M.,
Domanskyi, A., Geissler, C., Vinnikov, I.A., Alter, H., Schober, A., Vogt, M.A., Gass, P., Akatsu, H., Hashizume, Y., Yamamoto, S., Kondo, T., Ito, S., Tachikawa, M., Niwa, S.
Parlato, R., Schutz, G., 2011. Pten ablation in adult dopaminergic neurons is I., Yabe, H., Terasaki, T., Setou, M., Ohtsuki, S., 2017. Downregulation of GNA13-
neuroprotective in Parkinson’s disease models. FASEB J. 25, 2898–2910. ERK network in prefrontal cortex of schizophrenia brain identified by combined
Drexhage, R.C., van der Heul-Nieuwenhuijsen, L., Padmos, R.C., van Beveren, N., focused and targeted quantitative proteomics. J. Proteomics.
Cohen, D., Versnel, M.A., Nolen, W.A., Drexhage, H.A., 2010. Inflammatory gene Hnia, K., Vaccari, I., Bolino, A., Laporte, J., 2012. Myotubularin phosphoinositide
expression in monocytes of patients with schizophrenia: overlap and difference with phosphatases: cellular functions and disease pathophysiology. Trends Mol. Med. 18,
bipolar disorder. A study in naturalistically treated patients. Int. J. 317–327.
Neuropsychopharmacol. 13, 1369–1381. Hopkins, B.D., Hodakoski, C., Barrows, D., Mense, S.M., Parsons, R.E., 2014. PTEN
Du, Y., Du, Y., Zhang, Y., Huang, Z., Fu, M., Li, J., Pang, Y., Lei, P., Wang, Y.T., Song, W., function: the long and the short of it. Trends Biochem. Sci. 39, 183–190.
2019. MKP-1 reduces Aβ generation and alleviates cognitive impairments in Horn, S.R., Charney, D.S., Feder, A., 2016. Understanding resilience: new approaches for
Alzheimer’s disease models. Signal Transduct. Target. Ther. 4, 1–12. preventing and treating PTSD. Exp. Neurol. 284, 119–132.
Dumaual, C.M., Sandusky, G.E., Crowell, P.L., Randall, S.K., 2006. Cellular localization Hsieh, S.H., Ferraro, G.B., Fournier, A.E., 2006. Myelin-associated inhibitors regulate
of PRL-1 and PRL-2 gene expression in normal adult human tissues. J. Histochem. cofilin phosphorylation and neuronal inhibition through LIM kinase and Slingshot
Cytochem. 54, 1401–1412. phosphatase. J. Neurosci. 26, 1006–1015.
Duran, J., Guinovart, J.J., 2015. Brain glycogen in health and disease. Mol. Aspects Med. Huang, C.Y., Tan, T.H., 2012. DUSPs, to MAP kinases and beyond. Cell Biosci. 2, 24.
46, 70–77. Iio, W., Matsukawa, N., Tsukahara, T., Kohari, D., Toyoda, A., 2011. Effects of chronic
Duric, V., Banasr, M., Licznerski, P., Schmidt, H.D., Stockmeier, C.A., Simen, A.A., social defeat stress on MAP kinase cascade. Neurosci. Lett. 504, 281–284.
Newton, S.S., Duman, R.S., 2010. A negative regulator of MAP kinase causes Jakobsson, J., Bjerke, M., Sahebi, S., Isgren, A., Ekman, C.J., Sellgren, C., Olsson, B.,
depressive behavior. Nat. Med. 16, 1328–1332. Zetterberg, H., Blennow, K., Palsson, E., Landen, M., 2015. Monocyte and microglial
Dwivedi, Y., Rizavi, H.S., Roberts, R.C., Conley, R.C., Tamminga, C.A., Pandey, G.N., activation in patients with mood-stabilized bipolar disorder. J. Psychiatry Neurosci.
2001. Reduced activation and expression of ERK1/2 MAP kinase in the post-mortem 40, 250–258.
brain of depressed suicide subjects. J. Neurochem. 77, 916–928. Julien, C.L., Thompson, J.C., Wild, S., Yardumian, P., Snowden, J.S., Turner, G.,
Dwivedi, Y., Rizavi, H.S., Conley, R.R., Roberts, R.C., Tamminga, C.A., Pandey, G.N., Craufurd, D., 2007. Psychiatric disorders in preclinical Huntington’s disease.
2003. Altered gene expression of brain-derived neurotrophic factor and receptor J. Neurol. Neurosurg. Psychiatry 78, 939–943.
tyrosine kinase B in postmortem brain of suicide subjects. Arch. Gen. Psychiatry 60, Jung, S.K., Jeong, D.G., Yoon, T.S., Kim, J.H., Ryu, S.E., Kim, S.J., 2007. Crystal structure
804–815. of human slingshot phosphatase 2. Proteins 68, 408–412.
Dwivedi, Y., Rizavi, H.S., Conley, R.R., Pandey, G.N., 2006. ERK MAP kinase signaling in Jung, S., Nah, J., Han, J., Choi, S.G., Kim, H., Park, J., Pyo, H.K., Jung, Y.K., 2016. Dual-
post-mortem brain of suicide subjects: differential regulation of upstream Raf kinases specificity phosphatase 26 (DUSP26) stimulates Abeta42 generation by promoting
Raf-1 and B-Raf. Mol. Psychiatry 11, 86–98. amyloid precursor protein axonal transport during hypoxia. J. Neurochem. 137,
Ebert, M., Fei, G., Herrera, P., Friess, H., Buechler, M., Malfertheiner, P., 2000. 770–781.
Overexpression of transforming growth factor beta-1 (TGF-B1) is associated with Kang, D.E., Roh, S.E., Woo, J.A., Liu, T., Bu, J.H., Jung, A.R., Lim, Y., 2011. The interface
reduced pten expression in pancreatic fibrosis and cancer. Gastroenterology 118, between cytoskeletal aberrations and mitochondrial dysfunction in alzheimer’s
A650. disease and related disorders. Exp. Neurobiol. 20, 67–80.
Elia, M., Amato, C., Bottitta, M., Grillo, L., Calabrese, G., Esposito, M., Carotenuto, M., Karege, F., Perroud, N., Burkhardt, S., Fernandez, R., Ballmann, E., La Harpe, R.,
2012. An atypical patient with Cowden syndrome and PTEN gene mutation Malafosse, A., 2011. Alterations in phosphatidylinositol 3-kinase activity and PTEN
presenting with cortical malformation and focal epilepsy. Brain Dev. 34, 873–876. phosphatase in the prefrontal cortex of depressed suicide victims.
Neuropsychobiology 63, 224–231.

16
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Kecmanovic, M., Keckarevic-Markovic, M., Keckarevic, D., Stevanovic, G., Jovic, N., Low, H.B., Zhang, Y., 2016. Regulatory roles of MAPK phosphatases in cancer. Immune
Romac, S., 2016. Genetics of Lafora progressive myoclonic epilepsy: current Netw. 16, 85–98.
perspectives. Appl. Clin. Genet. 9, 49–53. Ludka, F.K., Constantino, L.C., Dal-Cim, T., Binder, L.B., Zomkowski, A., Rodrigues, A.L.,
Kerr, F., Rickle, A., Nayeem, N., Brandner, S., Cowburn, R.F., Lovestone, S., 2006. PTEN, Tasca, C.I., 2016. Involvement of PI3K/Akt/GSK-3beta and mTOR in the
a negative regulator of PI3 kinase signalling, alters tau phosphorylation in cells by antidepressant-like effect of atorvastatin in mice. J. Psychiatr. Res. 82, 50–57.
mechanisms independent of GSK-3. FEBS Lett. 580, 3121–3128. Luo, Y., Liang, F., Zhang, Z.Y., 2009. PRL1 promotes cell migration and invasion by
Kim, E.K., Choi, E.J., 2015. Compromised MAPK signaling in human diseases: an update. increasing MMP2 and MMP9 expression through Src and ERK1/2 pathways.
Arch. Toxicol. 89, 867–882. Biochemistry 48, 1838–1846.
Kim, J.S., Huang, T.Y., Bokoch, G.M., 2009. Reactive oxygen species regulate a slingshot- Mailand, N., Lukas, C., Kaiser, B.K., Jackson, P.K., Bartek, J., Lukas, J., 2002.
cofilin activation pathway. Mol. Biol. Cell 20, 2650–2660. Deregulated human Cdc14A phosphatase disrupts centrosome separation and
Kim, S.H., Shin, S.Y., Lee, K.Y., Joo, E.J., Song, J.Y., Ahn, Y.M., Lee, Y.H., Kim, Y.S., chromosome segregation. Nat. Cell Biol. 4, 317–322.
2012. The genetic association of DUSP6 with bipolar disorder and its effect on ERK Masters, C.L., Bateman, R., Blennow, K., Rowe, C.C., Sperling, R.A., Cummings, J.L.,
activity. Prog. Neuropsychopharmacol. Biol. Psychiatry 37, 41–49. 2015. Alzheimer’s disease. Nat. Rev. Dis. Primers 1, 15056.
Klengel, T., Binder, E.B., 2015. Epigenetics of stress-related psychiatric disorders and Matter, W.F., Estridge, T., Zhang, C., Belagaje, R., Stancato, L., Dixon, J., Johnson, B.,
gene x environment interactions. Neuron 86, 1343–1357. Bloem, L., Pickard, T., Donaghue, M., Acton, S., Jeyaseelan, R., Kadambi, V.,
Kligys, K., Claiborne, J.N., DeBiase, P.J., Hopkinson, S.B., Wu, Y., Mizuno, K., Jones, J.C., Vlahos, C.J., 2001. Role of PRL-3, a human muscle-specific tyrosine phosphatase, in
2007. The slingshot family of phosphatases mediates Rac1 regulation of cofilin angiotensin-II signaling. Biochem. Biophys. Res. Commun. 283, 1061–1068.
phosphorylation, laminin-332 organization, and motility behavior of keratinocytes. McCarthy, M.J., Wei, H., Landgraf, D., Le Roux, M.J., Welsh, D.K., 2016. Disinhibition of
J. Biol. Chem. 282, 32520–32528. the extracellular-signal-regulated kinase restores the amplification of circadian
Knafo, S., Esteban, J.A., 2017. PTEN: local and global modulation of neuronal function in rhythms by lithium in cells from bipolar disorder patients. Eur.
health and disease. Trends Neurosci. 40, 83–91. Neuropsychopharmacol. 26, 1310–1319.
Knafo, S., Sanchez-Puelles, C., Palomer, E., Delgado, I., Draffin, J.E., Mingo, J., Wahle, T., McGue, M., Lykken, D.T., Iacono, W.G., 1996. Genotype-environment correlations and
Kaleka, K., Mou, L., Pereda-Perez, I., Klosi, E., Faber, E.B., Chapman, H.M., Lozano- interactions in the etiology of substance abuse and related behaviors. NIDA Res.
Montes, L., Ortega-Molina, A., Ordonez-Gutierrez, L., Wandosell, F., Vina, J., Monogr. 159, 49–72 discussion 73-80.
Dotti, C.G., Hall, R.A., Pulido, R., Gerges, N.Z., Chan, A.M., Spaller, M.R., Meberg, P.J., Ono, S., Minamide, L.S., Takahashi, M., Bamburg, J.R., 1998. Actin
Serrano, M., Venero, C., Esteban, J.A., 2016. PTEN recruitment controls synaptic and depolymerizing factor and cofilin phosphorylation dynamics: response to signals that
cognitive function in Alzheimer’s models. Nat. Neurosci. 19, 443–453. regulate neurite extension. Cell Motil. Cytoskeleton 39, 172–190.
Kodama, M., Russell, D.S., Duman, R.S., 2005. Electroconvulsive seizures increase the Mehta, D., Binder, E.B., 2012. Gene x environment vulnerability factors for PTSD: the
expression of MAP kinase phosphatases in limbic regions of rat brain. HPA-axis. Neuropharmacology 62, 654–662.
Neuropsychopharmacology 30, 360–371. Menalled, L., Brunner, D., 2014. Animal models of Huntington’s disease for translation to
Kondoh, K., Nishida, E., 2007. Regulation of MAP kinases by MAP kinase phosphatases. the clinic: best practices. Mov. Disord. 29, 1375–1390.
Biochim. Biophys. Acta 1773, 1227–1237. Minamide, L.S., Striegl, A.M., Boyle, J.A., Meberg, P.J., Bamburg, J.R., 2000.
Koob, G.F., Kandel, D., Volkow, N.D., 2008. Pathophysiology of addiction. Psychiatry Neurodegenerative stimuli induce persistent ADF/cofilin-actin rods that disrupt
354–378. distal neurite function. Nat. Cell Biol. 2, 628–636.
Kouli, A., Torsney, K.M., Kuan, W.-L., 2018. Parkinson’s disease: etiology, Mocciaro, A., Schiebel, E., 2010. Cdc14: a highly conserved family of phosphatases with
neuropathology, and pathogenesis. Exon Publications 3–26. non-conserved functions? J. Cell. Sci. 123, 2867–2876.
Kreiner, G., Bierhoff, H., Armentano, M., Rodriguez-Parkitna, J., Sowodniok, K., Mueser, K.T., McGurk, S.R., 2004. Schizophrenia. Lancet 363, 2063–2072.
Naranjo, J.R., Bonfanti, L., Liss, B., Schutz, G., Grummt, I., Parlato, R., 2013. Nayeem, N., Kerr, F., Naumann, H., Linehan, J., Lovestone, S., Brandner, S., 2007.
A neuroprotective phase precedes striatal degeneration upon nucleolar stress. Cell Hyperphosphorylation of tau and neurofilaments and activation of CDK5 and ERK1/
Death Differ. 20, 1455–1464. 2 in PTEN-deficient cerebella. Mol. Cell. Neurosci. 34, 400–408.
Kreis, P., Leondaritis, G., Lieberam, I., Eickholt, B.J., 2014. Subcellular targeting and Neale, B.M., Kou, Y., Liu, L., Ma’Ayan, A., Samocha, K.E., Sabo, A., Lin, C.-F., Stevens, C.,
dynamic regulation of PTEN: implications for neuronal cells and neurological Wang, L.-S., Makarov, V., 2012. Patterns and rates of exonic de novo mutations in
disorders. Front. Mol. Neurosci. 7, 23. autism spectrum disorders. Nature 485, 242–245.
Kuntz-Melcavage, K.L., Brucklacher, R.M., Grigson, P.S., Freeman, W.M., Vrana, K.E., Nestler, E.J., Pena, C.J., Kundakovic, M., Mitchell, A., Akbarian, S., 2016. Epigenetic
2009. Gene expression changes following extinction testing in a heroin behavioral basis of mental illness. Neuroscientist 22, 447–463.
incubation model. BMC Neurosci. 10, 95. Nielsen, D.A., Ji, F., Yuferov, V., Ho, A., He, C., Ott, J., Kreek, M.J., 2010. Genome-wide
Kyosseva, S.V., Elbein, A.D., Griffin, W.S., Mrak, R.E., Lyon, M., Karson, C.N., 1999. association study identifies genes that may contribute to risk for developing heroin
Mitogen-activated protein kinases in schizophrenia. Biol. Psychiatry 46, 689–696. addiction. Psychiatr. Genet. 20, 207–214.
Labonte, B., Engmann, O., Purushothaman, I., Menard, C., Wang, J., Tan, C., Scarpa, J.R., Niwa, R., Nagata-Ohashi, K., Takeichi, M., Mizuno, K., Uemura, T., 2002. Control of actin
Moy, G., Loh, Y.E., Cahill, M., Lorsch, Z.S., Hamilton, P.J., Calipari, E.S., Hodes, G.E., reorganization by Slingshot, a family of phosphatases that dephosphorylate ADF/
Issler, O., Kronman, H., Pfau, M., Obradovic, A.L.J., Dong, Y., Neve, R.L., Russo, S., cofilin. Cell 108, 233–246.
Kazarskis, A., Tamminga, C., Mechawar, N., Turecki, G., Zhang, B., Shen, L., Novak, M.J., Tabrizi, S.J., 2011. Huntington’s disease: clinical presentation and
Nestler, E.J., 2017. Sex-specific transcriptional signatures in human depression. Nat. treatment. Int. Rev. Neurobiol. 98, 297–323.
Med. 23, 1102–1111. Ogino, M., Ichimura, M., Nakano, N., Minami, A., Kitagishi, Y., Matsuda, S., 2016. Roles
Lang, A.J., 2017. Mindfulness in PTSD treatment. Curr. Opin. Psychol. 14, 40–43. of PTEN with DNA repair in Parkinson’s disease. Int. J. Mol. Sci. 17.
Lang, R., Hammer, M., Mages, J., 2006. DUSP meet immunology: dual specificity MAPK Ohashi, K., 2015. Roles of cofilin in development and its mechanisms of regulation. Dev.
phosphatases in control of the inflammatory response. J. Immunol. 177, 7497–7504. Growth Differ. 57, 275–290.
Laporte, J., Blondeau, F., Gansmuller, A., Lutz, Y., Vonesch, J.L., Mandel, J.L., 2002. The Ohta, Y., Kousaka, K., Nagata-Ohashi, K., Ohashi, K., Muramoto, A., Shima, Y., Niwa, R.,
PtdIns3P phosphatase myotubularin is a cytoplasmic protein that also localizes to Uemura, T., Mizuno, K., 2003. Differential activities, subcellular distribution and
Rac1-inducible plasma membrane ruffles. J. Cell. Sci. 115, 3105–3117. tissue expression patterns of three members of Slingshot family phosphatases that
LaSarge, C.L., Pun, R.Y., Muntifering, M.B., Danzer, S.C., 2016. Disrupted hippocampal dephosphorylate cofilin. Genes Cells 8, 811–824.
network physiology following PTEN deletion from newborn dentate granule cells. Ohtake, Y., Hayat, U., Li, S., 2015. PTEN inhibition and axon regeneration and neural
Neurobiol. Dis. 96, 105–114. repair. Neural Regen. Res. 10, 1363–1368.
Lee, J.O., Yang, H., Georgescu, M.M., Di Cristofano, A., Maehama, T., Shi, Y., Dixon, J.E., Owen, M.J., Sawa, A., Mortensen, P.B., 2016. Schizophrenia. Lancet 388, 86–97.
Pandolfi, P., Pavletich, N.P., 1999. Crystal structure of the PTEN tumor suppressor: Owens, D.M., Keyse, S.M., 2007. Differential regulation of MAP kinase signalling by dual-
implications for its phosphoinositide phosphatase activity and membrane specificity protein phosphatases. Oncogene 26, 3203–3213.
association. Cell 99, 323–334. Padmos, R.C., Hillegers, M.H., Knijff, E.M., Vonk, R., Bouvy, A., Staal, F.J., de Ridder, D.,
Lee, K.Y., Ahn, Y.M., Joo, E.J., Chang, J.S., Kim, Y.S., 2006. The association of DUSP6 Kupka, R.W., Nolen, W.A., Drexhage, H.A., 2008. A discriminating messenger RNA
gene with schizophrenia and bipolar disorder: its possible role in the development of signature for bipolar disorder formed by an aberrant expression of inflammatory
bipolar disorder. Mol. Psychiatry 11, 425–426. genes in monocytes. Arch. Gen. Psychiatry 65, 395–407.
Leslie, N.R., Downes, C.P., 2002. PTEN: the down side of PI 3-kinase signalling. Cell. Parlato, R., Kreiner, G., 2013. Nucleolar activity in neurodegenerative diseases: a missing
Signal. 14, 285–295. piece of the puzzle? J. Mol. Med. 91, 541–547.
Liang, H., He, S., Yang, J., Jia, X., Wang, P., Chen, X., Zhang, Z., Zou, X., McNutt, M.A., Parnetti, L., Castrioto, A., Chiasserini, D., Persichetti, E., Tambasco, N., El-Agnaf, O.,
Shen, W.H., Yin, Y., 2014. PTENα is a PTEN isoform translated through alternative Calabresi, P., 2013. Cerebrospinal fluid biomarkers in Parkinson disease. Nat. Rev.
initiation and regulates mitochondrial function. Cell Metab. 19, 836–848. Neurol. 9, 131–140.
Liang, H., Chen, X., Yin, Q., Ruan, D., Zhao, X., Zhang, C., McNutt, M.A., Yin, Y., 2017. Patassini, S., Begley, P., Xu, J., Church, S.J., Kureishy, N., Reid, S.J., Waldvogel, H.J.,
PTENbeta is an alternatively translated isoform of PTEN that regulates rDNA Faull, R.L., Snell, R.G., Unwin, R.D., 2019. Cerebral vitamin B5 (D-Pantothenic acid)
transcription. Nat. Commun. 8, 14771. deficiency as a potential cause of metabolic perturbation and neurodegeneration in
Liu, L.J., Zhu, C., Tian, H.J., Zheng, T.S., Ye, M.J., Li, H., 2016. Correlations of PTEN Huntington’s disease. Metabolites 9, 113.
genetic polymorphisms with the risk of depression and depressive symptoms in a Patel, D.C., Tewari, B.P., Chaunsali, L., Sontheimer, H., 2019. Neuron–glia interactions in
Chinese population. Gene 595, 77–82. the pathophysiology of epilepsy. Nat. Rev. Neurosci. 20, 282–297.
Liu, G., Slater, N., Perkins, A., 2017. Epilepsy: treatment options. Am. Fam. Phys. 96, Patterson, K.I., Brummer, T., O’Brien, P.M., Daly, R.J., 2009. Dual-specificity
87–96. phosphatases: critical regulators with diverse cellular targets. Biochem. J. 418,
Llorens, F., Banez-Coronel, M., Pantano, L., del Rio, J.A., Ferrer, I., Estivill, X., Marti, E., 475–489.
2013. A highly expressed miR-101 isomiR is a functional silencing small RNA. BMC
Genomics 14, 104.

17
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Peng, L., Xing, X., Li, W., Qu, L., Meng, L., Lian, S., Jiang, B., Wu, J., Shou, C., 2009. PRL- Sigitova, E., Fišar, Z., Hroudová, J., Cikánková, T., Raboch, J., 2017. Biological
3 promotes the motility, invasion, and metastasis of LoVo colon cancer cells through hypotheses and biomarkers of bipolar disorder. Psychiatry Clin. Neurosci. 71,
PRL-3-integrin beta1-ERK1/2 and-MMP2 signaling. Mol. Cancer 8, 110. 77–103.
Perez-Sen, R., Queipo, M.J., Gil-Redondo, J.C., Ortega, F., Gomez-Villafuertes, R., Miras- Singh, A., Trevick, S., 2016. The epidemiology of global epilepsy. Neurol. Clin. 34,
Portugal, M.T., Delicado, E.G., 2019. Dual-specificity phosphatase regulation in 837–847.
neurons and glial cells. Int. J. Mol. Sci. 20. Sonoda, Y., Mukai, H., Matsuo, K., Takahashi, M., Ono, Y., Maeda, K., Akiyama, H.,
Pittenger, C., Duman, R.S., 2008. Stress, depression, and neuroplasticity: a convergence Kawamata, T., 2010. Accumulation of tumor-suppressor PTEN in Alzheimer
of mechanisms. Neuropsychopharmacology 33, 88–109. neurofibrillary tangles. Neurosci. Lett. 471, 20–24.
Plotkin, J.L., Surmeier, D.J., 2014. Impaired striatal function in Huntington’s disease is Soosairajah, J., Maiti, S., Wiggan, O., Sarmiere, P., Moussi, N., Sarcevic, B., Sampath, R.,
due to aberrant p75NTR signaling. Rare Dis. 2, e968482. Bamburg, J.R., Bernard, O., 2005. Interplay between components of a novel LIM
Plotkin, J.L., Day, M., Peterson, J.D., Xie, Z., Kress, G.J., Rafalovich, I., Kondapalli, J., kinase-slingshot phosphatase complex regulates cofilin. EMBO J. 24, 473–486.
Gertler, T.S., Flajolet, M., Greengard, P., Stavarache, M., Kaplitt, M.G., Rosinski, J., Sparrow, N., Manetti, M.E., Bott, M., Fabianac, T., Petrilli, A., Bates, M.L., Bunge, M.B.,
Chan, C.S., Surmeier, D.J., 2014. Impaired TrkB receptor signaling underlies Lambert, S., Fernandez-Valle, C., 2012. The actin-severing protein cofilin is
corticostriatal dysfunction in Huntington’s disease. Neuron 83, 178–188. downstream of neuregulin signaling and is essential for Schwann cell myelination.
Previtali, S.C., Zerega, B., Sherman, D.L., Brophy, P.J., Dina, G., King, R.H., Salih, M.M., J. Neurosci. 32, 5284–5297.
Feltri, L., Quattrini, A., Ravazzolo, R., Wrabetz, L., Monaco, A.P., Bolino, A., 2003. Steenkamp, M.M., Litz, B.T., Hoge, C.W., Marmar, C.R., 2015. Psychotherapy for
Myotubularin-related 2 protein phosphatase and neurofilament light chain protein, military-related PTSD: a review of randomized clinical trials. JAMA 314, 489–500.
both mutated in CMT neuropathies, interact in peripheral nerve. Hum. Mol. Genet. Stegmeier, F., Amon, A., 2004. Closing mitosis: the functions of the Cdc14 phosphatase
12, 1713–1723. and its regulation. Annu. Rev. Genet. 38, 203–232.
Raess, M.A., Friant, S., Cowling, B.S., Laporte, J., 2017. WANTED - Dead or alive: Struyf, J., Dobrin, S., Page, D., 2008. Combining gene expression, demographic and
myotubularins, a large disease-associated protein family. Adv. Biol. Regul. 63, clinical data in modeling disease: a case study of bipolar disorder and schizophrenia.
49–58. BMC Genomics 9, 531.
Rafii, M.S., 2013. Update on Alzheimer’s disease therapeutics. Rev. Recent Clin. Trials 8, Sveinbjornsdottir, S., 2016. The clinical symptoms of Parkinson’s disease. J. Neurochem.
110–118. 139 (Suppl 1), 318–324.
Reiner, A., Dragatsis, I., Dietrich, P., 2011. Genetics and neuropathology of Huntington’s Tai, T.Y., Warner, L.N., Jones, T.D., Jung, S., Concepcion, F.A., Skyrud, D.W., Fender, J.,
disease. Int. Rev. Neurobiol. 98, 325–372. Liu, Y., Williams, A.D., Neumaier, J.F., 2017. Antiepileptic action of c-Jun N-
Renton, A.E.M., 2010. Gene Expression and Genetic Analyses in Parkinson’s Disease With terminal kinase (JNK) inhibition in an animal model of temporal lobe epilepsy.
and Without Dementia. UCL (University College London). Neuroscience 349, 35–47.
Rickle, A., Bogdanovic, N., Volkmann, I., Zhou, X., Pei, J.J., Winblad, B., Cowburn, R.F., Takaki, M., Ujike, H., Kodama, M., Takehisa, Y., Nakata, K., Kuroda, S., 2001. Two kinds
2006. PTEN levels in Alzheimer’s disease medial temporal cortex. Neurochem. Int. of mitogen-activated protein kinase phosphatases, MKP-1 and MKP-3, are
48, 114–123. differentially activated by acute and chronic methamphetamine treatment in the rat
Rios, P., Li, X., Kohn, M., 2013. Molecular mechanisms of the PRL phosphatases. FEBS J. brain. J. Neurochem. 79, 679–688.
280, 505–524. Tarter, R.E., 2002. Etiology of adolescent substance abuse: a developmental perspective.
Rios, P., Nunes-Xavier, C.E., Tabernero, L., Kohn, M., Pulido, R., 2014. Dual-specificity Am. J. Addict. 11, 171–191.
phosphatases as molecular targets for inhibition in human disease. Antioxid. Redox Taylor, D.M., Moser, R., Regulier, E., Breuillaud, L., Dixon, M., Beesen, A.A., Elliston, L.,
Signal. 20, 2251–2273. Silva Santos Mde, F., Kim, J., Jones, L., Goldstein, D.R., Ferrante, R.J., Luthi-
Roach, P.J., 2015. Glycogen phosphorylation and Lafora disease. Mol. Aspects Med. 46, Carter, R., 2013. MAP kinase phosphatase 1 (MKP-1/DUSP1) is neuroprotective in
78–84. Huntington’s disease via additive effects of JNK and p38 inhibition. J. Neurosci. 33,
Rosso, L., Marques, A.C., Weier, M., Lambert, N., Lambot, M.A., Vanderhaeghen, P., 2313–2325.
Kaessmann, H., 2008. Birth and rapid subcellular adaptation of a hominoid-specific Tenegra, J.C., Leebold, B., 2016. Substance abuse screening and treatment. Prim. Care
CDC14 protein. PLoS Biol. 6, e140. Clin. Off. Pract. 43, 217–227.
Rowland, T.A., Marwaha, S., 2018. Epidemiology and risk factors for bipolar disorder. Thornicroft, G., Chatterji, S., Evans-Lacko, S., Gruber, M., Sampson, N., Aguilar-
Ther. Adv. Psychopharmacol. 8, 251–269. Gaxiola, S., Al-Hamzawi, A., Alonso, J., Andrade, L., Borges, G., 2017.
Rutten, B.P.F., Vermetten, E., Vinkers, C.H., Ursini, G., Daskalakis, N.P., Pishva, E., de Undertreatment of people with major depressive disorder in 21 countries. Br. J.
Nijs, L., Houtepen, L.C., Eijssen, L., Jaffe, A.E., Kenis, G., Viechtbauer, W., van den Psychiatry 210, 119–124.
Hove, D., Schraut, K.G., Lesch, K.P., Kleinman, J.E., Hyde, T.M., Weinberger, D.R., Tian, Y., Wang, L., Jia, M., Lu, T., Ruan, Y., Wu, Z., Wang, L., Liu, J., Zhang, D., 2017.
Schalkwyk, L., Lunnon, K., Mill, J., Cohen, H., Yehuda, R., Baker, D.G., Maihofer, A. Association of oligodendrocytes differentiation regulator gene DUSP15 with autism.
X., Nievergelt, C.M., Geuze, E., Boks, M.P.M., 2018. Longitudinal analyses of the World J. Biol. Psychiatry 18, 143–150.
DNA methylome in deployed military servicemen identify susceptibility loci for post- Tielens, S., Huysseune, S., Godin, J.D., Chariot, A., Malgrange, B., Nguyen, L., 2016.
traumatic stress disorder. Mol. Psychiatry 23, 1145–1156. Elongator controls cortical interneuron migration by regulating actomyosin
Ryu, E.J., Angelastro, J.M., Greene, L.A., 2005. Analysis of gene expression changes in a dynamics. Cell Res. 26, 1131–1148.
cellular model of Parkinson disease. Neurobiol. Dis. 18, 54–74. Tilot, A.K., Frazier 2nd, T.W., Eng, C., 2015. Balancing proliferation and connectivity in
Sanchez-Mut, J.V., Aso, E., Heyn, H., Matsuda, T., Bock, C., Ferrer, I., Esteller, M., 2014. PTEN-associated autism spectrum disorder. Neurotherapeutics 12, 609–619.
Promoter hypermethylation of the phosphatase DUSP22 mediates PKA-dependent Tilot, A.K., Bebek, G., Niazi, F., Altemus, J.B., Romigh, T., Frazier, T.W., Eng, C., 2016.
TAU phosphorylation and CREB activation in Alzheimer’s disease. Hippocampus 24, Neural transcriptome of constitutional Pten dysfunction in mice and its relevance to
363–368. human idiopathic autism spectrum disorder. Mol. Psychiatry 21, 118–125.
Scheltens, P., Blennow, K., Breteler, M.M., de Strooper, B., Frisoni, G.B., Salloway, S., Trautmann, S., McCollum, D., 2002. Cell cycle: new functions for Cdc14 family
Van der Flier, W.M., 2016. Alzheimer’s disease. Lancet 388, 505–517. phosphatases. Curr. Biol. 12, R733–735.
Schick, V., Majores, M., Engels, G., Spitoni, S., Koch, A., Elger, C.E., Simon, M., Tysnes, O.B., Storstein, A., 2017. Epidemiology of Parkinson’s disease. J. Neural Transm.
Knobbe, C., Blumcke, I., Becker, A.J., 2006. Activation of Akt independent of PTEN (Vienna) 124, 901–905.
and CTMP tumor-suppressor gene mutations in epilepsy-associated Taylor-type focal Ueno, Y., Enokizono, T., Fukushima, H., Ohto, T., Imagawa, K., Tanaka, M., Sakai, A.,
cortical dysplasias. Acta Neuropathol. 112, 715–725. Suzuki, H., Uehara, T., Takenouchi, T., 2019. A novel missense PTEN mutation
Schmitt, A., Leonardi-Essmann, F., Durrenberger, P.F., Wichert, S.P., Spanagel, R., identified in a patient with macrocephaly and developmental delay. Hum. Genome
Arzberger, T., Kretzschmar, H., Zink, M., Herrera-Marschitz, M., Reynolds, R., Var. 6, 1–4.
Rossner, M.J., Falkai, P., Gebicke-Haerter, P.J., 2012. Structural synaptic elements Ujike, H., Takaki, M., Kodama, M., Kuroda, S., 2002. Gene expression related to
are differentially regulated in superior temporal cortex of schizophrenia patients. synaptogenesis, neuritogenesis, and MAP kinase in behavioral sensitization to
Eur. Arch. Psychiatry Clin. Neurosci. 262, 565–577. psychostimulants. Ann. N. Y. Acad. Sci. 965, 55–67.
Schulte, J., Littleton, J.T., 2011. The biological function of the Huntingtin protein and its van Os, J., Kenis, G., Rutten, B.P., 2010. The environment and schizophrenia. Nature
relevance to Huntington’s Disease pathology. Curr. Trends Neurol. 5, 65–78. 468, 203–212.
Sekar, S., Taghibiglou, C., 2018. Elevated nuclear phosphatase and tensin homolog Vilardo, E., Barbato, C., Ciotti, M., Cogoni, C., Ruberti, F., 2010. MicroRNA-101 regulates
(PTEN) and altered insulin signaling in substantia nigral region of patients with amyloid precursor protein expression in hippocampal neurons. J. Biol. Chem. 285,
Parkinson’s disease. Neurosci. Lett. 666, 139–143. 18344–18351.
Seleman, M., Chapy, H., Cisternino, S., Courtin, C., Smirnova, M., Schlatter, J., Volkow, N.D., Koob, G.F., McLellan, A.T., 2016. Neurobiologic advances from the brain
Chiadmi, F., Scherrmann, J.M., Noble, F., Marie-Claire, C., 2014. Impact of P- disease model of addiction. N. Engl. J. Med. 374, 363–371.
glycoprotein at the blood-brain barrier on the uptake of heroin and its main Walsh, S., Finn, D.P., Dowd, E., 2011. Time-course of nigrostriatal neurodegeneration
metabolites: behavioral effects and consequences on the transcriptional responses and neuroinflammation in the 6-hydroxydopamine-induced axonal and terminal
and reinforcing properties. Psychopharmacology (Berl.) 231, 3139–3149. lesion models of Parkinson’s disease in the rat. Neuroscience 175, 251–261.
Sepe, S., Milanese, C., Gabriels, S., Derks, K.W., Payan-Gomez, C., van, I.W.F., Rijksen, Y. Wang, Y., Wang, Z., 2020. Identification of dysregulated genes and pathways of different
M., Nigg, A.L., Moreno, S., Cerri, S., Blandini, F., Hoeijmakers, J.H., brain regions in Alzheimer’s disease. Int. J. Neurosci. 1–13.
Mastroberardino, P.G., 2016. Inefficient DNA repair is an aging-related modifier of Wang, Y., Shibasaki, F., Mizuno, K., 2005. Calcium signal-induced cofilin
parkinson’s disease. Cell Rep. 15, 1866–1875. dephosphorylation is mediated by Slingshot via calcineurin. J. Biol. Chem. 280,
Shen, W.H., Balajee, A.S., Wang, J., Wu, H., Eng, C., Pandolfi, P.P., Yin, Y., 2007. 12683–12689.
Essential role for nuclear PTEN in maintaining chromosomal integrity. Cell 128, Waragai, M., Sekiyama, K., Sekigawa, A., Takamatsu, Y., Fujita, M., Hashimoto, M.,
157–170. 2010. Alpha-synuclein and DJ-1 as potential biological fluid biomarkers for
Parkinson’s disease. Int. J. Mol. Sci. 11, 4257–4266.
Watson, P., 2019. PTSD as a public mental health priority. Curr. Psychiatry Rep. 21, 61.

18
N. An et al. Progress in Neurobiology xxx (xxxx) xxx

Wei, M., Korotkov, K.V., Blackburn, J.S., 2018. Targeting phosphatases of regenerating Yehuda, R., Hoge, C.W., McFarlane, A.C., Vermetten, E., Lanius, R.A., Nievergelt, C.M.,
liver (PRLs) in cancer. Pharmacol. Ther. 190, 128–138. Hobfoll, S.E., Koenen, K.C., Neylan, T.C., Hyman, S.E., 2015. Post-traumatic stress
Wen, Z., Han, L., Bamburg, J.R., Shim, S., Ming, G.L., Zheng, J.Q., 2007. BMP gradients disorder. Nat. Rev. Dis. Primers 1, 15057.
steer nerve growth cones by a balancing act of LIM kinase and Slingshot phosphatase Yuan, P., Zhou, R., Wang, Y., Li, X., Li, J., Chen, G., Guitart, X., Manji, H.K., 2010.
on ADF/cofilin. J. Cell Biol. 178, 107–119. Altered levels of extracellular signal-regulated kinase signaling proteins in
Willner, P., Scheel-Kruger, J., Belzung, C., 2013. The neurobiology of depression and postmortem frontal cortex of individuals with mood disorders and schizophrenia.
antidepressant action. Neurosci. Biobehav. Rev. 37, 2331–2371. J. Affect. Disord. 124, 164–169.
Wittchen, H.U., Jacobi, F., Rehm, J., Gustavsson, A., Svensson, M., Jonsson, B., Yuen, E.Y., Liu, W., Kafri, T., van Praag, H., Yan, Z., 2010. Regulation of AMPA receptor
Olesen, J., Allgulander, C., Alonso, J., Faravelli, C., Fratiglioni, L., Jennum, P., channels and synaptic plasticity by cofilin phosphatase Slingshot in cortical neurons.
Lieb, R., Maercker, A., van Os, J., Preisig, M., Salvador-Carulla, L., Simon, R., J. Physiol. 588, 2361–2371.
Steinhausen, H.C., 2011. The size and burden of mental disorders and other Zafar, S., Younas, N., Sheikh, N., Tahir, W., Shafiq, M., Schmitz, M., Ferrer, I.,
disorders of the brain in Europe 2010. Eur. Neuropsychopharmacol. 21, 655–679. Andreoletti, O., Zerr, I., 2017. Cytoskeleton-associated risk modifiers involved in
Worby, C.A., Dixon, J.E., 2014. Pten. Annual Review of Biochemistry, 83, pp. 641–669. early and rapid progression of sporadic creutzfeldt-jakob disease. Mol. Neurobiol.
Wu, Z., Wang, G., Wei, Y., Xiao, L., Wang, H., 2018. PI3K/AKT/GSK3β/CRMP-2- Zeng, Q., Hong, W., Tan, Y.H., 1998. Mouse PRL-2 and PRL-3, two potentially prenylated
mediated neuroplasticity in depression induced by stress. Neuroreport 29, protein tyrosine phosphatases homologous to PRL-1. Biochem. Biophys. Res.
1256–1263. Commun. 244, 421–427.
Wykes, T., Haro, J.M., Belli, S.R., Obradors-Tarrago, C., Arango, C., Ayuso-Mateos, J.L., Zhang, X., Li, F., Bulloj, A., Zhang, Y.W., Tong, G., Zhang, Z., Liao, F.F., Xu, H., 2006.
Bitter, I., Brunn, M., Chevreul, K., Demotes-Mainard, J., Elfeddali, I., Evans-Lacko, S., Tumor-suppressor PTEN affects tau phosphorylation, aggregation, and binding to
Fiorillo, A., Forsman, A.K., Hazo, J.B., Kuepper, R., Knappe, S., Leboyer, M., Lewis, S. microtubules. FASEB J. 20, 1272–1274.
W., Linszen, D., Luciano, M., Maj, M., McDaid, D., Miret, M., Papp, S., Park, A.L., Zhang, Y., Li, Z., Fan, X., Xiong, J., Zhang, G., Luo, X., Li, K., Jie, Z., Cao, Y., Huang, Z.,
Schumann, G., Thornicroft, G., van der Feltz-Cornelis, C., van Os, J., Wahlbeck, K., Wu, F., Xiao, L., Duan, G., Chen, H., 2018. PRL-3 promotes gastric cancer peritoneal
Walker-Tilley, T., Wittchen, H.U., consortium, R, 2015. Mental health research metastasis via the PI3K/AKT signaling pathway in vitro and in vivo. Oncol. Lett. 15,
priorities for Europe. Lancet Psychiatry 2, 1036–1042. 9069–9074.
Yang, P., Liang, T., Zhang, C., Cai, J., Zhang, W., Chen, B., Qiu, X., Yao, K., Li, G., Zhao, Y., Wang, S., Chu, Z., Dang, Y., Zhu, J., Su, X., 2017. MicroRNA-101 in the
Wang, H., Jiang, C., You, G., Jiang, T., 2016. Clinicopathological factors predictive ventrolateral orbital cortex (VLO) modulates depressive-like behaviors in rats and
of postoperative seizures in patients with gliomas. Seizure 35, 93–99. targets dual-specificity phosphatase 1 (DUSP1). Brain Res. 1669, 55–62.
Ye, B., Jiang, L.L., Xu, H.T., Zhou, D.W., Li, Z.S., 2012. Expression of PI3K/AKT pathway Zhu, Y., Hoell, P., Ahlemeyer, B., Sure, U., Bertalanffy, H., Krieglstein, J., 2007.
in gastric cancer and its blockade suppresses tumor growth and metastasis. Int. J. Implication of PTEN in production of reactive oxygen species and neuronal death in
Immunopathol. Pharmacol. 25, 627–636. in vitro models of stroke and Parkinson’s disease. Neurochem. Int. 50, 507–516.
Ye, Z., Al-Aidaroos, A.Q., Park, J.E., Yuen, H.F., Zhang, S.D., Gupta, A., Lin, Y., Shen, H. Zhu, Q.Y., Liu, Q., Chen, J.X., Lan, K., Ge, B.X., 2010. MicroRNA-101 targets MAPK
M., Zeng, Q., 2015. PRL-3 activates mTORC1 in cancer progression. Sci. Rep. 5, phosphatase-1 to regulate the activation of MAPKs in macrophages. J. Immunol.
17046. 185, 7435–7442.

19

You might also like