Hostname: page-component-7c8c6479df-p566r Total loading time: 0 Render date: 2024-03-19T04:56:08.854Z Has data issue: false hasContentIssue false

Maternal separation in rodents: a journey from gut to brain and nutritional perspectives

Published online by Cambridge University Press:  28 June 2019

Marion Rincel*
Affiliation:
Univ. Bordeaux, INRA, Bordeaux INP, NutriNeuro, UMR 1286, F-33000, Bordeaux, France
Muriel Darnaudéry
Affiliation:
Univ. Bordeaux, INRA, Bordeaux INP, NutriNeuro, UMR 1286, F-33000, Bordeaux, France
*
*Corresponding author: Marion Rincel, email marion.rincel@hotmail.fr
Rights & Permissions [Opens in a new window]

Abstract

The developmental period constitutes a critical window of sensitivity to stress. Indeed, early-life adversity increases the risk to develop psychiatric diseases, but also gastrointestinal disorders such as the irritable bowel syndrome at adulthood. In the past decade, there has been huge interest in the gut–brain axis, especially as regards stress-related emotional behaviours. Animal models of early-life adversity, in particular, maternal separation (MS) in rodents, demonstrate lasting deleterious effects on both the gut and the brain. Here, we review the effects of MS on both systems with a focus on stress-related behaviours. In addition, we discuss more recent findings showing the impact of gut-directed interventions, including nutrition with pre- and probiotics, illustrating the role played by gut microbiota in mediating the long-term effects of MS. Overall, preclinical studies suggest that nutritional approaches with pro- and prebiotics may constitute safe and efficient strategies to attenuate the effects of early-life stress on the gut–brain axis. Further research is required to understand the complex mechanisms underlying gut–brain interaction dysfunctions after early-life stress as well as to determine the beneficial impact of gut-directed strategies in a context of early-life adversity in human subjects.

Type
Conference on ‘Optimal diet and lifestyle strategies for the management of cardio-metabolic risk’
Copyright
Copyright © The Authors 2019

Mounting evidence suggests a pivotal role of gut microbiota in the aetiology of psychiatric symptoms in stress-related diseases such as anxiety disorders and depression(Reference Rogers, Keating and Young1,Reference Valles-Colomer, Falony and Darzi2) . The mechanisms underlying this microbiota–gut–brain communication are beginning to be unravelled (see(Reference Cryan and Dinan3Reference Mayer5) for reviews). In particular, certain gut bacteria can have a beneficial effect on mood and emotional behaviour and, as such, have been proposed for potential therapeutic interventions in psychiatry (concept of psychobiotics)(Reference Dinan, Stanton and Cryan6,Reference Sherwin, Rea and Dinan7) . The bidirectional interplay between gut and brain is illustrated in population survey studies revealing a strong correlation between anxiety, depression and functional gastrointestinal (GI) disorders. Furthermore, psychological distress can predict later onset of a functional GI disorder and the converse is also true(Reference Koloski, Jones and Kalantar8).

Early postnatal life is a critical period during which both brain and gut undergo important maturation(Reference Borre, O'Keeffe and Clarke9,Reference Sharon, Sampson and Geschwind10) . Moreover, this maturation is greatly influenced by gut microbiota colonisation and diversification during the lactating period. Exposure to stressful events during childhood has been repeatedly associated with increased vulnerability to both psychiatric and GI disorders such as the irritable bowel syndrome (IBS)(Reference Chitkara, van Tilburg and Blois-Martin11Reference O'Mahony, Clarke and Dinan13). IBS is defined as a disorder of the gut–brain interaction. According to Rome IV classification, it is characterised by abdominal pain and altered bowel habits(Reference Drossman14), but also increased intestinal permeability and gut dysbiosis. Chronic disruption of the mother–infant relationship in rodents, best known as maternal separation (MS), is a useful preclinical tool since it models the co-morbidity between IBS and psychiatric disorders. Indeed, it induces a wide range of brain and gut alterations in offspring(Reference O'Mahony, Hyland and Dinan15). In the following, we concisely overview the adverse consequences of MS, which is the most used model of early adversity in gut–brain axis research. We then discuss the effects of gut-directed interventions on the microbiota–gut–brain axis, with a particular focus on stress-related behaviours.

The maternal separation model

Pioneering work from Harlow in non-human primates and Levine, Denenberg, Meaney and Plotsky in rodents has shown that the early environment, in particular the quality of maternal care, shapes emotional behaviour as well as stress responsivity in adult life(Reference Denenberg, Ottinger and Stephens16Reference Seay, Hansen and Harlow20). The work of Hofer also revealed the deleterious impact of early weaning on offspring physiology, including intestinal physiology(Reference Ackerman, Hofer and Weiner21). Since then, a vast body of literature has documented the effects of early mother–infant separations in rats during the first weeks of life (1–3 weeks). The most common MS paradigm consists in daily 3 h separations between postnatal days (PND) 2 and 14(Reference Lippmann, Bress and Nemeroff22). However, there are other models using different separation durations (3–8 h daily) or an acute 24 h separation(Reference Barna, Bálint and Baranyi23Reference Viveros, Llorente and López-Gallardo26). MS results in different degrees of perceived stress in dams and pups according to the protocol used (litter isolated in the homecage without the mother or litter isolated in a novel environment; pups individually separated or not; undisturbed control or ‘handling’ i.e. short separation episode (15 min)). The different models and their respective effects are reviewed in(Reference Korosi and Baram27Reference Vetulani29). In any case, pups are deprived of maternal care during the separation period. Importantly, the absence of the dam implies that the pups cannot benefit from dams' heat and milk. Temperature issues can be easily corrected by maintaining the room at 28–29°C during separation sessions. However, the lack of milk intake likely contributes to the short and long-term effects of 24 h MS(Reference Suchecki, Rosenfeld and Levine30,Reference Van Oers, de Kloet and Whelan31) . Mother–infant separation-based models have also been developed in other rodents (e.g. guinea pigs and mice) and in primates (rhesus macaques)(Reference Cirulli, Francia and Berry32), but the largest literature still involves rats, with mice being more and more used; we will focus on these rodent species in the present review. It appears that mice are less sensitive to early-life stress than rats(Reference Tan, Ho and Song33) (see(Reference Millstein and Holmes34) for review). This might be attributable to species specificities in neurodevelopment and maternal care patterns. Another possible reason is that mouse studies more often involve inbred strains (while outbred strains are used in rats) as well as transgenic strains that exhibit different levels of sensitivity to stress(Reference Millstein and Holmes34). To produce significant behavioural alterations in mice, MS is often combined with others stressors such as unpredictable stress in dams(Reference Franklin, Russig and Weiss35,Reference Gapp, Soldado-Magraner and Alvarez-Sánchez36) , early weaning(Reference Tchenio, Lecca and Valentinova37) or a combination of perinatal stressors(Reference Rincel, Aubert and Chevalier38).

Maternal separation and emotional vulnerability

Long-term psychoneuroendocrine alterations

Behaviour

The long-term consequences of MS on emotional behaviour have been extensively documented. Available tools to evaluate emotionality are mostly limited to tests with good predictive validity (i.e. sensitive to anxiolytics or antidepressants) such as the elevated plus maze, open-field or light–dark tests for anxiety and the forced swimming test or tail suspension test for depression. These tests have however a poor construct validity contrary to other tests such as sucrose preference or female urine sniffing tests used to assess reward deficiency as index of anhedonia (see(Reference Cryan and Holmes39Reference York, Blevins and Baynard41)).

Typically, MS leads to increased anxiety- and depressive-like behaviours. Indeed, animals exposed to MS during early-life display reduced exploration of the open areas in the elevated plus maze, light–dark box and open-field tests compared with non-separated controls(Reference Amini-Khoei, Haghani-Samani and Beigi42Reference Yang, Cheng and Tang72). Moreover, it has been shown that exposure to a novel stress at adulthood aggravates these anxiety-like behaviours(Reference Eiland and McEwen73,Reference Marais, van Rensburg and van Zyl74) . Numerous studies also report increased depressive-like behaviours in the forced swimming test or tail suspension test. Indeed, adult MS rodents show greater immobility time in these tests compared with controls(Reference Amini-Khoei, Haghani-Samani and Beigi42,Reference Gracia-Rubio, Moscoso-Castro and Pozo53,Reference Lambás-Señas, Mnie-Filali and Certin58,Reference Lee, Kim and Kim59,Reference Liu, Hao and Zhu61,Reference Maniam and Morris62,Reference Portero-Tresserra, Gracia-Rubio and Cantacorps66,Reference Amini-Khoei, Mohammadi-Asl and Amiri75Reference Yamawaki, Nishida and Harada91) . MS has been associated with decreased sucrose preference(Reference Shu, Xiao and Tang69,Reference Wang, Dong and Wang71,Reference Amini-Khoei, Mohammadi-Asl and Amiri75Reference Bai, Zhu and Zhang77,Reference Hui, Zhang and Liu79Reference Li, Yang and Yao81,Reference Masrour, Peeri and Azarbayjani83,Reference Sadeghi, Peeri and Hosseini87,Reference Dallé, Daniels and Mabandla92Reference Yang, Xu and Zhang96) and decreased social behaviour with a conspecific(Reference Rincel, Lépinay and Delage67,Reference Yang, Cheng and Tang72,Reference Kundakovic, Lim and Gudsnuk93,Reference Farrell, Holland and Shansky97Reference Zimmerberg and Sageser99) . The effects of MS are not limited to the above alterations of emotional behaviours; numerous studies also report that MS exacerbates motivation for alcohol and drugs of abuse (see(Reference Moffett, Vicentic and Kozel100) for review).

Finally, several studies have also shown deleterious effects of MS on cognition (see(Reference Kosten, Kim and Lee101) for review). Briefly, these effects include impaired hippocampal-dependent spatial learning and memory(Reference Dandi, Kalamari and Touloumi47,Reference Rincel, Lépinay and Delage67,Reference Hui, Zhang and Liu79,Reference Couto, Batalha and Valadas102Reference Wang, Jiao and Dulawa107) , altered non-spatial memory(Reference Banqueri, Méndez and Arias44,Reference Reshetnikov, Kovner and Lepeshko105,Reference Wang, Jiao and Dulawa107Reference Wang, Nie and Li115) and impairments in prefrontal cortex (PFC)-dependent tasks (working memory, extinction, cognitive flexibility)(Reference Feifel, Shair and Schmauss50,Reference Yang, Cheng and Tang72,Reference Wang, Jiao and Dulawa107,Reference Baudin, Blot and Verney116Reference Thomas, Caporale and Wu120) . In contrast, amygdala-dependent aversive memory (e.g. fear conditioning) seems to be enhanced by MS(Reference Wilber, Southwood and Wellman121Reference Xiong, Yang and Cao127).

Endocrine response and neurobiological correlates

MS exerts long-lasting effects on hypothalamic–pituitary–adrenal (HPA) axis function, leading in most of the studies to endocrine hyper-responsivity to a novel stress(Reference Plotsky and Meaney19,Reference Dandi, Kalamari and Touloumi47,Reference Aisa, Tordera and Lasheras108,Reference Biagini, Pich and Carani128Reference Slotten, Kalinichev and Hagan135) . Within the central nervous system, this HPA axis hyper-reactivity is associated with an up-regulation of corticotrophin-releasing hormone (CRF) expression in the paraventricular nucleus (PVN) of the hypothalamus and amygdala but also with high CRF concentration and increased CRF receptor density in the locus coeruleus and raphe nucleus(Reference Plotsky and Meaney19,Reference Ladd, Owens and Nemeroff130,Reference Ladd, Huot and Thrivikraman131,Reference de Almeida Magalhães, Correia and de Carvalho136) (see(Reference Rivarola and Renard137)) as well as altered oxytocin and vasopressin expression (either up- or down-regulated) in the PVN (see(Reference Veenema138) for review). MS also decreases glucocorticoid receptor (GR) expression in the hippocampus and PFC(Reference Wilber, Southwood and Wellman121,Reference Rivarola and Suárez139) , two main brain areas involved in HPA axis negative feedback. Numerous neurotransmission systems are affected by MS. MS decreases the number of type A γ-aminobutyric acid (GABA-A) receptors in noradrenergic neurons of the locus coeruleus and in the nucleus tractus solitarius(Reference Caldji, Francis and Sharma46) and hippocampus(Reference Yang, Xu and Zhang96). The gabaergic system plays a role in CRF synthesis inhibition in the central amygdala, allowing a buffering of the noradrenergic response to stress. In addition, MS impairs glutamatergic(Reference Chen, Chen and Tang140Reference Pickering, Gustafsson and Cebere142), serotonergic(Reference Daniels, Pietersen and Carstens48,Reference Liu, Hao and Zhu61,Reference Masrour, Peeri and Azarbayjani83,Reference Bravo, Dinan and Cryan143Reference Wu, Ziea and Lao147) , dopaminergic(Reference Moya-Pérez, Perez-Villalba and Benítez-Páez64,Reference Kawakami, Quadros and Machado145,Reference Arborelius and Eklund148Reference Romano-López, Méndez-Díaz and García153) , opioidergic(Reference Ploj, Roman and Nylander152,Reference Ploj, Roman and Nylander154) and endocannabinoidergic(Reference Portero-Tresserra, Gracia-Rubio and Cantacorps66) transmission. In the central nervous system, serotonin is involved in neuronal development(Reference Daubert and Condron155), emotionality and also pain modulation(Reference Kim and Camilleri156,Reference Sommer157) . Among other effects, MS reduces the expression of the serotonin transporter in the raphe nucleus(Reference Bravo, Dinan and Cryan143). Interestingly, selective serotonin reuptake inhibitor antidepressants such as paroxetine normalise HPA axis function as well as emotional behaviour in MS rats.

MS induces both functional and structural changes in several brain regions including the PFC, hippocampus, amygdala and nucleus accumbens(Reference Li, Xue and Shao150,Reference Chocyk, Bobula and Dudys158Reference Soztutar, Colak and Ulupinar163) . More specifically, impaired synaptic long-term potentiation, dendritic atrophy as well as reduced dendritic spine density have been reported in the medial PFC and hippocampus of adolescent and adult MS rats(Reference de Melo, de David Antoniazzi and Hossain63,Reference Shin, Han and Woo68,Reference Farrell, Holland and Shansky97,Reference Guo, Liang and Liang104,Reference Baudin, Blot and Verney116,Reference Chen, Chen and Tang140,Reference Romano-López, Méndez-Díaz and García153,Reference Chocyk, Bobula and Dudys158,Reference Bock, Gruss and Becker164Reference Sousa, Vital and Costenla171) . By contrast, MS induces dendritic hypertrophy in the amygdala(Reference Koe, Ashokan and Mitra57). A recent study reported that mice deficient for motopsin, a serine protease secreted from neuronal cells to induce filopodia, precursor structures of dendritic spines, are resistant to MS-induced increase in anxiety in the open field test(Reference Hidaka, Kashio and Uchigaki111). In addition, it has been shown that MS leads to hypomyelination in the medial PFC(Reference Yang, Cheng and Tang72).

MS is also accompanied by decreased expression of neurotrophins such as nerve growth factor and brain-derived neurotrophic factor (BDNF), that are known to play critical roles in dendrite growth and spinogenesis(Reference Lippmann, Bress and Nemeroff22,Reference Dandi, Kalamari and Touloumi47,Reference Marais, van Rensburg and van Zyl74,Reference Réus, Stringari and Ribeiro85,Reference Aisa, Elizalde and Tordera172,Reference De Lima, Presti-Torres and Vedana173) (see(Reference Park and Poo174) for review). In addition, MS leads to alterations of hippocampal neurogenesis (either decreased or increased) at adulthood(Reference Hulshof, Novati and Sgoifo112,Reference Mirescu, Peters and Gould175Reference Lajud, Roque and Cajero177) . Interestingly, decreased hippocampal BDNF and neurogenesis are consistent observations in post-mortem brains of depressed subjects and there is mounting evidence that BDNF is involved in emotional vulnerability (see(Reference Autry and Monteggia178) for review).

Peripheral and central inflammation

There is substantial evidence that MS activates inflammatory processes both systemically and within the central nervous system, although the underlying mechanisms remain to be explored. Indeed, increased circulating levels of IL-1β(Reference Wang, Dong and Wang71) and IL-6(Reference Wieck, Andersen and Brenhouse179) have been reported in MS animals. In addition, MS offspring display neuroinflammatory marks such as increased Tnfa, Il-1b and Tlr4 expression or increased reactive oxygen species levels and decreased Il-10 expression in the hippocampus(Reference Wang, Dong and Wang71,Reference Amini-Khoei, Mohammadi-Asl and Amiri75,Reference Sadeghi, Peeri and Hosseini87,Reference Pinheiro, de Lima and Portal114) , PFC(Reference Wang, Dong and Wang71) and PVN(Reference Tang, Zhang and Ji180). Recent studies have shown a decrease in the levels of the astrocytic marker GFAP (glial fibrillary acidic protein) in the PFC of MS animals(Reference Yamawaki, Nishida and Harada91) and the opposite effect in the locus coeruleus of MS females only(Reference Nakamoto, Aizawa and Kinoshita181).

Inconsistencies in the maternal separation literature

A number of studies did not replicate the abovementioned findings, reporting no alteration of certain emotional behaviours(Reference Tan, Ho and Song33,Reference Aya-Ramos, Contreras-Vargas and Rico43Reference Bondar, Lepeshko and Reshetnikov45,Reference Feifel, Shair and Schmauss50,Reference de Melo, de David Antoniazzi and Hossain63,Reference Rincel, Lépinay and Delage67,Reference Eiland and McEwen73,Reference Marais, van Rensburg and van Zyl74,Reference Sadeghi, Peeri and Hosseini87,Reference Uchida, Hara and Kobayashi89,Reference Øines, Murison and Mrdalj95,Reference Reshetnikov, Kovner and Lepeshko105,Reference Hulshof, Novati and Sgoifo112,Reference de Almeida Magalhães, Correia and de Carvalho136,Reference Ershov, Bondar and Lepeshko182Reference Zimmerberg and Kajunski192) , cognitive function(Reference Dandi, Kalamari and Touloumi47,Reference Wang, Jiao and Dulawa107,Reference Thomas, Caporale and Wu120,Reference Hill, Klug and Kiss Von Soly185,Reference Pryce, Bettschen and Nanz-Bahr193Reference Zhu, Wang and Yao197) or HPA axis signalling(Reference Daniels, Pietersen and Carstens48,Reference Hulshof, Novati and Sgoifo112) in male or in female MS animals. In addition, others studies reported opposite effects (e.g. lower anxiety or lower depressive-like behaviour)(Reference Aya-Ramos, Contreras-Vargas and Rico43,Reference Maniam and Morris94,Reference Yang, Xu and Zhang96,Reference Tsuda, Yamaguchi and Nakata98,Reference Reshetnikov, Kovner and Lepeshko105,Reference Furukawa, Tsukahara and Tomita110,Reference Slotten, Kalinichev and Hagan135,Reference Ershov, Bondar and Lepeshko182,Reference Chocyk, Majcher-Maślanka and Przyborowska198Reference Mourlon, Baudin and Blanc202) .

In some cases, these discrepancies may be attributed to the use of different MS protocols (number of separated pups, separation duration and control group), age of investigation, animal strain and sex, housing conditions (individual or collective cages, light–dark cycle, enrichment), but also other testing protocol issues (e.g. habituation prior testing, brightness, sucrose concentration for the sucrose preference test). Notably, the vast majority of the findings were obtained using males only. However, numerous recent studies report sex-specific behavioural alterations in MS animals.

Nevertheless, differential effects of MS have also been reported in studies using the same MS protocol, age, sex, strain or type of stressor. A recent study suggests that the effects of early adversity (maternal immune activation) depend upon the gut microbiota profile of the dams, in particular the presence of commensal segmented filamentous bacteria (which differs across animal suppliers, i.e. Jackson Laboratories and Taconic Biosciences)(Reference Kim, Kim and Yim203). Therefore, the gut microbiota profile may also influence the susceptibility to MS.

Possible early mechanisms at the origin of maternal separation programming

The mechanisms underlying the long-term effects of MS are not fully understood. Multiple, possibly synergistic effects in both dams and pups have been reported (see(Reference Korosi204) for review).

Mother–infant communication and maternal care

Maternal care is thought to play an important role in brain maturation and later vulnerability to stress. It has been established that rodent pups vocalise in response to isolation (30–90 Hz ultrasounds)(Reference Branchi, Santucci and Alleva205,Reference Hofer, Shair and Brunelli206) and MS has been shown to increase the number of these vocalisations compared with undisturbed pups in several mouse strains(Reference Feifel, Shair and Schmauss50). Because these isolation calls elicit retrieval behaviour in the mother, they are thought to serve mother–pup communication and stimulate maternal care towards their pups(Reference Brunelli, Curley and Gudsnuk207,Reference D'Amato, Scalera and Sarli208) . In the MS model, pups are deprived of maternal care during several consecutive hours, which may constitute a mechanism for the adverse effects of this early-life stress. Indeed, it has been demonstrated that the long-term behavioural effects of acute 24 h MS can be prevented by pup tactile stimulation(Reference Van Oers, de Kloet and Whelan31). Nevertheless, the role of maternal care in the long-term effects of MS remains controversial.

MS also constitutes a potent stressor for the dams. Indeed, it has been reported that this psychological stress induces anxiety and depressive-like behaviours in dams(Reference Aguggia, Suárez and Rivarola209Reference Maniam and Morris211). As a matter of fact, several studies suggest that dam's perceived stress plays an important role in the effects of separation in the offspring. Interestingly, MS-induced HPA hyper-response to stress in the offspring can be counteracted by providing a foster litter to the dam while its own litter is being separated(Reference Huot, Gonzalez and Ladd212). Furthermore, it has been reported that the offspring of dams with an experience of separation with a previous litter exhibit MS-like fear behaviour without direct exposure to the early stress(Reference Kan, Callaghan and Richardson213).

Endocrine, immune and neurobiological effects of maternal separation in developing pups

The HPA axis is almost silenced during a short window of early postnatal development (i.e. from PND4 to 14(Reference Boersma, Bale and Casanello214Reference Vázquez217)). This stress hypo-responsive period is characterised by extremely low basal corticosterone levels in the plasma as well as blunted adrenocorticotropic hormone and corticosterone response to stress. Nevertheless, this stress hypo-responsive period is not absolute, since a potent stressor such as MS is able to induce HPA axis activation(Reference Vázquez217Reference Xu, Qin and Shi220). It has been proposed that stress and immune activation result in a cross-sensitisation of both systems that possibly creates a self-perpetuating cycle contributing to the emergence of the alterations in animals subjected to early stress. Bacterial translocation into the liver and the spleen has been detected after MS in juvenile PND10 rats(Reference Moussaoui, Braniste and Ait-Belgnaoui221). In addition, altered circulating pro-inflammatory IL-1β, IL-6 and TNFα were observed in MS pups(Reference Réus, Fernandes and de Moura86,Reference Pinheiro, de Lima and Portal114,Reference Do Prado, Narahari and Holland118,Reference Wieck, Andersen and Brenhouse179,Reference Roque, Ochoa-Zarzosa and Torner222,Reference Roque, Mesquita and Palha223) . Furthermore, MS juveniles display increased activated microglia in the PFC and hippocampus(Reference Gracia-Rubio, Moscoso-Castro and Pozo53) and decreased number of astrocytes in the same areas(Reference Yamawaki, Nishida and Harada91,Reference Musholt, Cirillo and Cavaliere224,Reference Saavedra, Fenton Navarro and Torner225) along with increased Il-6, Il-1b and Tnfa expression compared with controls(Reference Park, Kim and Kang65,Reference Roque, Ochoa-Zarzosa and Torner222) . Increased microglia numbers and activation patterns have also been recently reported in the nucleus of the solitary tract of MS juveniles(Reference Yamawaki, Nishida and Harada91,Reference Baldy, Fournier and Boisjoly-Villeneuve226) . Interestingly, increased cytokine expression and microglial density have also been reported in the hippocampus of juvenile mice submitted to short MS (15 min) from PND1 to PND21, which led to increased anxiety similar to prolonged MS(Reference Delpech, Wei and Hao227).

Both altered HPA axis activity and neuroinflammation during development have been shown to be deleterious for the immature brain. MS disrupts the normal course of brain development and produces functional and structural alterations including delayed GABA excitatory-to-inhibitory functional switch(Reference Furukawa, Tsukahara and Tomita110), delayed synaptic maturity(Reference Andersen and Teicher228), decreased spine density(Reference Rincel, Lépinay and Janthakhin161) and increased neuronal and glial cell death(Reference Mirescu, Peters and Gould175,Reference Kuma, Miki and Matsumoto229,Reference Zhang, Levine and Dent230) . Altered expression of neurotrophins such as BDNF and nerve growth factor in separated pups could contribute to these effects(Reference Kuma, Miki and Matsumoto229,Reference Cirulli, Alleva and Antonelli231,Reference Roceri, Cirulli and Pessina232) . In addition, MS disturbs the serotonergic system during development. Indeed, reduced expression of the serotonin receptor 5HTr1A in the hippocampus and PFC has been reported in 7-d-old pups (Reference Ohta, Miki and Warita233). A recent study demonstrates that transient juvenile, but not adult, knockdown of orthodenticle homoeobox 2 in the ventral tegmental area mimics early-life stress by increasing stress susceptibility, whereas its overexpression reverses the effects of early-life stress(Reference Peña, Kronman and Walker234). Moreover, developmental decrease of the transcription repressor Rest4 (RE-1 silencing transcription factor 4) in the PFC of pups submitted to MS may play a causal role in the long-term effects of MS(Reference Rincel, Lépinay and Delage67,Reference Uchida, Hara and Kobayashi89) . We recently demonstrated that exposure to a high-fat diet (HFD) during the perinatal period can prevent the long-term MS-associated neurobehavioural alterations, possibly via a protective effect on gene expression in the PFC(Reference Rincel, Lépinay and Delage67). Indeed, perinatal HFD prevented the MS-induced alterations of Rest4, Bdnf and 5HTr1A expression in this brain area. A recent work demonstrated that chemogenetic inhibition of MS-induced neuronal hyperactivity in the lateral habenula of mice aged 35 d attenuates depressive-like behaviours(Reference Tchenio, Lecca and Valentinova37).

Epigenetic changes in maternal-separation offspring

Epigenetic marks are dynamic and highly sensitive to environmental factors; furthermore they can last in time and even be transferred across generations(Reference Bohacek and Mansuy235). As such, they represent a potential mechanism that could underlie the long-term effects of early-life stress(Reference Heim and Binder236Reference Silberman, Acosta and Zorrilla Zubilete239). Indeed, a number of studies have reported persistent epigenetic marks in the genome of animals submitted to MS (see(Reference Jawahar, Murgatroyd and Harrison240) for review). In particular, changes in DNA methylation of specific regulatory sites in key genes for stress processing such as Crf, Avp, GR or Bdnf in the PVN, hippocampus and PFC of maternally separated animals, have been documented(Reference Zhu, Wang and Yao197,Reference Meaney and Szyf241Reference Wang, Cattaneo and Ryno245) . It has been shown that administration of a DNA methyltransferase inhibitor prevents the decreased prefrontal Bdnf mRNA expression induced by MS(Reference Roth, Lubin and Funk244). Moreover, DNA methylation in the offspring has been shown to be associated with the level of maternal care(Reference Weaver, Cervoni and Champagne246). Nonetheless, the group of Mansuy provided evidence for epigenetically-mediated transmission of behavioural traits induced by early-life stress across generations irrespective of cross fostering(Reference Weiss, Franklin and Vizi247).

Another major epigenetic process is histone modification, especially acetylation by histone acetyltransferases or deacetylation by histone deacetylases. Histone acetylation patterns as well as histone acetyltransferase and histone deacetylase expressions in the brain are also altered by MS(Reference Pusalkar, Suri and Kelkar248). For instance, MS leads to decreased Bdnf and GR mRNA expressions in the hippocampus, and these effects were accompanied by decreased levels of histone acetylation at their respective promoters(Reference Park, Lee and Seo249,Reference Seo, Ly and Lee250) . Furthermore, a recent study suggests that there is a cross-talk between histone acetylation and DNA methylation(Reference Zhu, Wang and Yao197). Indeed, treatment with a histone deacetylase inhibitor reversed the MS-induced increased DNA methylation in the GR promoter region.

Finally, the possible role of brain miRNA in mediating the long-term effects of MS has been addressed in a few studies. Uchida and colleagues were the first to report changes in expression of several miRNA in the PFC of MS rats(Reference Uchida, Hara and Kobayashi89). Another MS study reported an increase in miR-16 in the hippocampus that was negatively correlated with Bdnf expression in the same brain area and also negatively correlated with sucrose preference(Reference Bai, Zhu and Zhang77).

Maternal separation as a model of irritable bowel syndrome: impact on the gastrointestinal tract

As mentioned earlier, MS is also widely used as a model of IBS (see(Reference O'Mahony, Hyland and Dinan15,Reference Barreau, Ferrier and Fioramonti251,Reference Moloney, Johnson and O'Mahony252) for reviews). In addition to its effects on stress vulnerability, it leads to several GI dysfunctions, in particular increased visceral sensitivity to painful stimuli, and increases the vulnerability to experimental colitis.

Effects of maternal separation on the enteric nervous system, visceral sensitivity and motility

MS induces dynamic structural and functional changes in the enteric nervous system(Reference Barreau, Salvador-Cartier and Houdeau253,Reference Tominaga, Fujikawa and Tanaka254) . For instance, MS increases nerve density and synaptogenesis in juveniles, but these effects are no longer present at adulthood(Reference Barreau, Salvador-Cartier and Houdeau253). In contrast, the levels of the neuronal marker PGP 9·5 (anti-protein gene product 9·5) in the colon are increased in adult MS animals but not in juveniles. Interestingly, early-life adversity has been shown to affect enteric nervous system development in a sex-dependent manner, with females being more sensitive than males(Reference Million and Larauche255). MS also produces increased intestinal motility in response to stress, as evidenced by reduced total transit time and increased number of faecal pellets(Reference Li, Yang and Yao81,Reference Hyland, O'Mahony and O'Malley256Reference Yi, Zhang and Sun260) . It has been extensively reported that MS rats display visceral hyperalgesia during colorectal distension(Reference Felice, Gibney and Gosselin51,Reference Rincel, Lépinay and Delage67,Reference Chen, Chen and Tang140,Reference Chen, Huang and Xu144,Reference O'Mahony, Chua and Quigley146,Reference Wu, Ziea and Lao147,Reference Tang, Zhang and Ji180,Reference Xu, Qin and Shi220,Reference Hyland, O'Mahony and O'Malley256,Reference Moloney, Stilling and Dinan257,Reference Schwetz, McRoberts and Coutinho259Reference Zhang, Li and Leung283) . A recent study demonstrated that MS-induced visceral hypersensitivity is dependent on Paneth cell defects and associated Escherichia coli expansion in the gut(Reference Riba, Olier and Lacroix-Lamandé284). MS-induced visceral hypersensitivity is lost in mice deficient for Toll-like receptor 4 (TLR4)(Reference Tang, Zhang and Ji180). This study suggests that TLR4 signalling in the PVN mediates increased CRF immunostaining and visceral hypersensitivity associated with MS. Interestingly, multiple MS-induced intestinal phenotypes, including visceral hyperalgesia and gut leakiness, can be prevented by CRF receptor antagonist administration(Reference Tang, Zhang and Ji180,Reference Schwetz, McRoberts and Coutinho259,Reference Barreau, Cartier and Leveque285Reference Million, Wang and Wang287) . GR antagonists or agonists of the metabotropic glutamate receptor type 7 (mGluR7) also prevent stress-induced visceral hyperalgesia(Reference Shao, Liu and Xiao278,Reference Myers and Greenwood-Van Meerveld288Reference Zhou, Sha and Huang290) .

The hyper-sensitivity to colorectal distension after MS is larger in females than in males and visceral hyperalgesia is greater when all pups are separated from the dam than when only half of littermates is removed, suggesting that sex and dam's perceived stress play a role in the long-term effects of MS on visceral sensitivity(Reference Rosztóczy, Fioramonti and Jármay277). Indeed, it has been demonstrated that MS-induced visceral hypersensitivity is transferred across generations and that this effect likely depends upon maternal care(Reference Van den Wijngaard, Stanisor and van Diest291).

Effects of maternal separation on gut microbiota composition

A growing number of studies have reported altered gut microbiota composition in MS animals. However, the use of different species, strains, sex, MS protocols, nature of the sample, microbiota analysis method and age of investigation renders between-studies comparisons difficult, and yet, there is no clear microbial pattern associated with MS.

The first study that has investigated the effects of MS on the gut microbiota was carried out by Bailey and Coe in rhesus monkeys(Reference Bailey and Coe292). The authors investigated the stability of gut microbiota 3 d after separation and found a significant decrease in faecal bacteria, in particular from the Lactobacillus genus. A few years later, O'Mahony and colleagues reported overall reduced bacterial diversity in MS rats v. controls(Reference O'Mahony, Marchesi and Scully293). This finding has been replicated in more recent studies(Reference Qian, Lu and Huang294,Reference Zhou, Li and Li295) . However, another recent study reports no change in diversity(Reference Moya-Pérez, Perez-Villalba and Benítez-Páez64). Qualitatively, MS was shown to increase the Firmicutes:Bacteroidetes ratio at the phylum level in some studies(Reference De Palma, Blennerhassett and Lu49,Reference Li, Lee and Filler286,Reference Zhou, Li and Li295,Reference El Aidy, Ramsteijn and Dini-Andreote296) , but again this finding is not consistent across studies as some report opposite(Reference Pusceddu, El Aidy and Crispie297) or no effects(Reference Zhou, Li and Li295). A consistent finding, however, is that the effects of MS on microbiota composition vary both qualitatively and quantitatively with respect to the age of investigation. Indeed, several studies comparing at least two time points show completely different patterns(Reference Moya-Pérez, Perez-Villalba and Benítez-Páez64,Reference Zhou, Li and Li295,Reference Barouei, Moussavi and Hodgson298,Reference García-Ródenas, Bergonzelli and Nutten299) . Overall, Bacteroides and Lachnospiraceae (including Clostridium XIVa) species seem to be consistently altered (either enriched or depleted) across several studies(Reference De Palma, Blennerhassett and Lu49,Reference Murakami, Kamada and Mizushima258,Reference Zhou, Li and Li295,Reference Ilchmann-Diounou, Olier and Lencina300) . Interestingly, it has been shown that changes in several bacterial taxa after MS are abrogated by adrenalectomy, suggesting that corticosterone signalling in response to stress is responsible for at least part of its effects on the microbiota(Reference Amini-Khoei, Haghani-Samani and Beigi42). More studies using global 16S-sequencing approaches are needed to better document the effects of MS on gut microbiota and potentially identify candidate species or genera associated with the behavioural effects of MS. Furthermore, considering the importance of sex differences in both stress effects and basal gut microbiota composition, more studies should be conducted in both males and females(Reference Rincel, Aubert and Chevalier38,Reference El Aidy, Ramsteijn and Dini-Andreote296) .

Effects of maternal separation on the gut mucosa

MS has been associated with alterations in the differentiation and distribution of enteroendocrine cells in the gut epithelium(Reference Estienne, Claustre and Clain-Gardechaux301) and a defect in Paneth cells(Reference Riba, Olier and Lacroix-Lamandé276,Reference Riba, Olier and Lacroix-Lamandé284) . Notably, the numbers of enterochromaffin cells in the colon are increased in MS animals compared with controls(Reference Bian, Zhang and Han264,Reference Bian, Qin and Tian265,Reference Ren, Wu and Yew275) . Accordingly, MS animals exhibit substantial increases in the levels of circulating and colonic serotonin (mainly produced by enterochromaffin cells)(Reference Chen, Huang and Xu144,Reference Wu, Ziea and Lao147,Reference Bian, Zhang and Han264,Reference Bian, Qin and Tian265,Reference Ren, Wu and Yew275,Reference Yang, Xian and Ip282) .

In addition, MS animals were shown to display colonic tissue damage including decreased crypt length and altered number of goblet cells and are more engaged in epithelial cell proliferation(Reference Barreau, Ferrier and Fioramonti262,Reference Li, Lee and Filler286,Reference Li, Zani and Lee302Reference O'Malley, Julio-Pieper and Gibney304) . Moreover, MS rats show more colonic damage after dextran sulphate sodium or 2,4,6-trinitrobenzenesulphonic acid-induced colitis than non-stressed animals and as a result, they also lose more weight, indicating that they are more sensitive to experimental colitis(Reference Ghia, Blennerhassett and Collins305Reference Veenema, Reber and Selch307). There is mounting evidence that MS produces long-term gut paracellular and transcellular hyper-permeability to ions and macromolecules(Reference Li, Yang and Yao81,Reference Barreau, Ferrier and Fioramonti262,Reference Miquel, Martín and Lashermes272,Reference Riba, Olier and Lacroix-Lamandé276,Reference García-Ródenas, Bergonzelli and Nutten299,Reference Li, Zani and Lee302,Reference Varghese, Verdú and Bercik306,Reference Barreau, Cartier and Ferrier308Reference Söderholm, Yates and Gareau311) . Remarkably, stress-induced intestinal hyperpermeability appears to be glucocorticoid-dependent, as it is evoked by the synthetic glucocorticoid dexamethasone and prevented by administration of a GR antagonist, similarly to an inhibitor of the myosin light chain kinase controlling epithelial cytoskeleton contraction(Reference Moussaoui, Braniste and Ait-Belgnaoui221). In addition, exposure to a novel stress at adulthood potentiates gut hyperpermeability in maternally separated rats(Reference Øines, Murison and Mrdalj95,Reference Söderholm, Yates and Gareau311) . Furthermore, it has been shown that acute MS induces immediate passage of macromolecules across the colonic mucosa and can lead to increased number of bacterial cells penetrating the gut epithelium(Reference Moussaoui, Braniste and Ait-Belgnaoui221,Reference Barreau, Ferrier and Fioramonti262,Reference Gareau, Jury and Yang312) .

MS also produces several immune alterations in the colon. Indeed, MS animals show an infiltration of immune cells (i.e. polymorphonuclear neutrophils)(Reference Barreau, Ferrier and Fioramonti262,Reference Ghia, Blennerhassett and Collins305) and an increase in mucosal mast cell density(Reference Li, Yang and Yao81,Reference Barreau, Salvador-Cartier and Houdeau253,Reference Barreau, Ferrier and Fioramonti262,Reference Hyland, Julio-Pieper and O'Mahony271,Reference Barreau, Cartier and Ferrier308) . MS also increases the expression of numerous cytokines including IL-6, IL-1β, TNFα, IFNγ, IL-4, IL-2 and IL-22 in the colonic mucosa(Reference Amini-Khoei, Haghani-Samani and Beigi42,Reference Moya-Pérez, Perez-Villalba and Benítez-Páez64,Reference Barreau, Ferrier and Fioramonti262,Reference Distrutti, Cipriani and Mencarelli268,Reference Riba, Olier and Lacroix-Lamandé276,Reference Shao, Liu and Xiao278,Reference Li, Lee and Filler286,Reference Li, Lee and Martin303,Reference Ghia, Blennerhassett and Collins305,Reference Barouei, Moussavi and Hodgson313) . Increased IFNγ and decreased IL-10 expression were prevented by mGluR7 agonist administration(Reference Shao, Liu and Xiao278) in MS animals.

It has been previously shown that MS increases IFNγ and TNF secretion by mesenteric lymph node cells(Reference Veenema, Reber and Selch307). In addition, increased mRNA expression of TLR3, 4 and 5 has been reported in the colonic mucosa of MS adult rats(Reference McKernan, Nolan and Brint314).

Impact of nutrition and microbiota-directed interventions in maternal separation offspring

An early study using the 24 h maternal deprivation paradigm suggested that feeding the pups during separation could prevent its effects on the HPA axis(Reference Van Oers, de Kloet and Whelan31). In the past decade, a growing number of studies have demonstrated that nutrition can modulate the long-term effects of early-life stress on brain and behaviour, although the underlying mechanisms remain unknown. Recent evidence suggests that the direct impact of nutrition on gut physiology and microbiota could counteract the stress-induced disruption of gut homoeostasis and promote a new state of equilibrium.

Nutritional strategies and maternal separation

Choline and vitamins

Several studies demonstrate a preventive effect of dietary choline and other vitamins in animals submitted to MS, and suggest that early nutritional interventions (before adulthood) have the strongest impact. In one study, the maternal diet was enriched with a mixture of essential C1 metabolism-associated micronutrients containing choline, betaine, methionine, folic acid, zinc, vitamins B6 and B12 during the course of MS. This treatment fully prevented the increased plasma corticosterone levels in MS pups at PND9 and further prevented later alterations of object recognition memory, but not spatial memory in adult MS mice(Reference Naninck, Oosterink and Yam315). In another study, dietary choline exposure from weaning to adulthood attenuated object recognition impairments in MS male rats(Reference Moreno Gudiño, Carías Picón and de Brugada Sauras113). In contrast, supplementation with a cocktail of methyl donors (choline, betaine, folic acid and vitamin B12) in adult maternally separated female rats failed to reverse the deleterious effect of MS on object recognition memory, but did prevent depressive-like behaviour in the forced swim test(Reference Paternain, Martisova and Campión84).

PUFA

Some evidence suggests that n-3 PUFA deficiency potentiates the effects of MS. For instance, dietary n-3 PUFA deficiency acts in synergy with MS to increase sucrose consumption in adulthood, an effect prevented by desipramine(Reference Ferreira, Bernardi and Krolow184,Reference Mathieu, Denis and Lavialle316) . It was further shown that the same dietary intervention also exacerbates MS-induced anxiety in the open-field test(Reference Mathieu, Oualian and Denis317).

Conversely, it has been reported that supplementation with either n-3, folic acid or n-acetylcysteine during peri-adolescence could prevent the MS-induced depressive-like behaviour in the forced swim test, likely through antioxidant effects within the brain(Reference Réus, Maciel and Abelaira318). Interestingly, supplementation with a mixture of EPA and DHA from adolescence onwards reverses MS-induced gut-microbiota dysbiosis in adult female rats(Reference Pusceddu, El Aidy and Crispie297). However, there was no major effect of the same treatment on anxiety and depressive-like behaviours or cognition in MS animals(Reference Pusceddu, Kelly and Ariffin319), yet no effect of MS per se was observed in this study. Nevertheless, in another study, dietary supplementation with PUFA-rich tuna oil failed to affect long-term visceral hypersensitivity in MS rats, but the diet was only administered after the induction of visceral hypersensitivity by acute stress(Reference van Diest, van den Elsen and Klok320).

High-fat diet

Previous studies have shown that palatable food consumption in adulthood can attenuate the deleterious effects of MS on anxiety and depressive-like behaviours and basal corticosterone levels(Reference Maniam and Morris62,Reference Maniam and Morris94) .

We reported that the long-term effects of MS on anxiety, social behaviour and stress endocrine response, but also visceral sensitivity, can be prevented by exposing the dams to HFD during gestation and lactation(Reference Rincel, Lépinay and Delage67). In addition to this protective effect of perinatal HFD in adult animals, we found similar beneficial effects on the developing brain(Reference Rincel, Lépinay and Janthakhin161). Indeed, maternal HFD exposure attenuated the stress-induced changes in mRNA expression of key genes involved in neuronal maturation and structural plasticity in the PFC of PND10 pups. The mechanisms underlying this protective effect of maternal HFD are elusive. We provided evidence that a comfort food effect of HFD in stressed mothers but also a modulation of the gut microbiota and/or gut barrier function by HFD in pups could contribute to its effects on brain and emotional behaviour(Reference Rincel, Lépinay and Delage67,Reference Rincel, Lépinay and Janthakhin161) .

Microbiota-directed interventions and maternal separation

The gut microbiota is highly sensitive to the environment and alterations of its composition (dysbiosis) have been described under conditions ranging from IBS and obesity to depression and autism(Reference Collins321Reference Zhao324). In particular, early-life environment, including diet and stressful experience, shapes the gut microbiota towards health and disease later in life(Reference Dong and Gupta325). However, the mechanisms underlying the ability of stress to modulate microbiota composition remain to be unravelled. Moreover, it is unclear whether dysbiosis is a causative factor in the aetiology of the abovementioned pathologies. Interestingly, studies using different, but complementary, gut microbiota-directed interventions (germ-free (GF) rodents, antibiotics, faecal microbiota transplantation, probiotics and prebiotics) have demonstrated that gut bacteria can have a beneficial effect on emotional behaviours and, as such, psychobiotics have been proposed for potential therapeutic interventions(Reference Dinan, Stanton and Cryan6,Reference Sherwin, Rea and Dinan7) .

Germ-free animals and microbiota transplantation experiments

Germ free

The study of GF (or axenic) animals served as a proof of concept for the role of gut microbiota in the regulation of brain function and behaviour. A large number of studies have explored GF-associated alterations both in the gut and the brain (see(Reference Luczynski, McVey Neufeld and Oriach326) for review).

Interestingly, many of the GF phenotypes are normalised by colonisation, although the effects largely depend upon the age of colonisation and the animal species and strain(Reference Braniste, Al-Asmakh and Kowal327Reference Sudo, Chida and Aiba332). Accordingly, Sudo et al. reported the first evidence that colonisation during early development, but not at a later age, could attenuate the increased HPA axis response to stress in GF mice(Reference Wang and Kasper323). In line with this study, further showed that locomotor hyperactivity in GF mice could be reversed by colonisation early in life, whereas colonisation at adulthood had no effect(Reference Heijtz, Wang and Anuar333).

A landmark study by De Palma and colleagues using GF mice exposed to MS revealed that the microbiota is necessary for the long-term effects of MS. Indeed, early-life stress fails to induce long-term endocrine and behavioural alterations in GF mice compared with SPF (specific-pathogen-free) controls(Reference De Palma, Blennerhassett and Lu49). Interestingly, colonisation with the gut microbiota of a conventional SPF control mouse unmasked the effects of early-life stress in GF mice. However, colonisation with the microbiota of an early-stressed animal did not transfer the stress-associated behavioural phenotype in naive GF mice, suggesting that gut bacteria are necessary but not sufficient to mediate the behavioural effects of early-life stress. Although the authors did not measure intestinal permeability, gut leakiness associated with MS could also contribute to the deleterious effects of MS on behaviour. An important limitation of the GF animal model is that the GF status is not specific at all to intestinal microbes. Previous studies suggest that maternal vaginal microbiota also impacts offspring neurodevelopment(Reference Jašarević, Howard and Morrison334). Furthermore, GF animals are housed in isolators with limited handling that constitutes a stressful environment and in most of the study the control groups are not housed in similar isolators and thus are not comparable. Together, these studies suggest that gut dysbiosis may be responsible for some, but not all of the MS-associated phenotypes later in life.

Faecal transplantation

The important role of gut microbiota in the regulation of behaviour was further confirmed by demonstrating the successful adoptive transfer of host behavioural phenotype between mice of different strains and with different behavioural profiles (see(Reference Collins, Kassam and Bercik335) for review). In animals, faecal transplantation can be achieved by oral gavage of fresh faecal content or by transient co-housing with the donor. The stability of the transplanted microbiota can vary depending upon several factors (strain, sex, age, housing conditions). The first evidence of gut–brain effects following faecal transplantation in animals showed a critical role of gut microbiota in host metabolism and energy balance(Reference Turnbaugh, Bäckhed and Fulton336,Reference Wang, Koonen and Hofker337) . Since then, accumulating data have demonstrated that faecal transplantation can affect brain and behaviour in rodents. For instance, social deficits in offspring from HFD-fed dams could be reversed by co-housing with offspring from dams fed a regular diet(Reference Buffington, Di Prisco and Auchtung338), an effect accompanied by restored synaptic plasticity in the brain following social interaction.

Conversely, it has recently been shown that naive rats receiving faecal microbiota from MS donors displayed MS-like intestinal hypermotility(Reference Zhao, Huang and Lu339). Interestingly, colonisation with the microbiota of IBS patients v. healthy controls recapitulated several features of IBS in GF mice, including faster GI transit, intestinal barrier dysfunction, innate immune activation, but also anxiety-like behaviour(Reference De Palma, Lynch and Lu340).

The faecal mycobiome of MS rats is altered relative to control animals(Reference Botschuijver, Roeselers and Levin341). Furthermore, fungicide treatment in adult MS rats prevented the visceral hypersensitivity induced by water avoidance stress. Strikingly, transplantation of the microbiota from MS rats could re-establish visceral hypersensitivity in the absence of water avoidance stress, an effect that was absent when the donor microbiota came from fungicide-treated rats. These findings highlight the need of considering exhaustively gut microbiota composition (i.e. bacteria but also viruses and fungi) and of better understanding the complex interactions between stress and gut microbes.

Overall, the potential clinical value of faecal transplantation for the treatment of disorders of the gut–brain axis is promising(Reference Borody and Khoruts342,Reference Brandt and Aroniadis343) and currently represents an active area of research. To date, the only indication for faecal transplantation in human subjects is the treatment of severe infections with Clostridium difficile, resulting in high success rates(Reference Khoruts344). In the recent years, two double-blind, placebo-controlled, randomised trials have investigated the impact of faecal microbiota transplantation in IBS patients(Reference Halkjær, Christensen and Lo345,Reference Johnsen, Hilpüsch and Cavanagh346) . However, evidence for clinical improvement of GI symptoms and psychiatric symptoms is unclear and has to be further established in larger studies.

Probiotics

The term probiotic, defined as ‘a live microbial feed supplement, which beneficially affects the host by improving its intestinal microbial balance’ was coined in 1953 by Werner Kollath to contrast with antibiotics(Reference Eberl347). The use of probiotics in animal studies has provided evidence that the gut microbiota posesses psychobiotic properties (i.e. antidepressant and/or anxiolytic-like activity) (see(Reference Dinan, Stanton and Cryan6,Reference Joseph and Law348,Reference Sarkar, Lehto and Harty349) for reviews).

Probiotic interventions are generally restricted to one or few bacterial species, thereby allowing the association between a given bug and a particular behavioural effect. The most used are members of the Bifidobacterium and Lactobacillus genera. Beneficial effects of probiotics have been reported in paradigms involving early-life stress. Several studies have shown anti-nociceptive effects of different probiotics (i.e. Faecalibacterium prausnitzii; Bifidobacterium breve or VSL#3)(Reference Barrett, Fitzgerald and Dinan263,Reference Distrutti, Cipriani and Mencarelli268,Reference Miquel, Martín and Lashermes272) . Moreover, the probiotic Bifidobacterium infantis chronically administered at adulthood (from PND50 to PND95) was reported to exert antidepressant-like effects in animals exposed to MS(Reference Desbonnet, Garrett and Clarke78). In addition, the increased peripheral levels of the proinflammatory cytokine IL-6 as well as the increased CRF mRNA levels in the amygdala in stressed animals were also normalised. Similarly, a lactobacillus strain, Lactobacillus plantarum PS128, has antidepressant-like effects in MS mice treated from weaning onwards in both the sucrose preference test and forced-swim test, but has no effect on MS anxiety(Reference Liu, Liu and Wu350). Moreover, serum increase in corticosterone (both at baseline and in response to stress), increase in IL-6 and decrease in IL-10 were all reversed by the probiotic. In addition to the beneficial effects of probiotics in adult animals, an increasing body of evidence shows that probiotics supplementation during early-life can have long-term preventive effects. Indeed, it has been shown that a mixture of Lactobacillus rhamnosus and Lactobacillus helveticus could prevent the elevation in basal plasma corticosterone observed in MS juvenile rats (PND20), in addition to mitigating the associated increased gut permeability(Reference Gareau, Jury and Perdue309). Similar findings have been reported in a mouse model of MS where mice received the probiotic Bifidobacterium pseudocatenulatum during the perinatal period(Reference Moya-Pérez, Perez-Villalba and Benítez-Páez64). Compared with their placebo-fed stressed counterparts, probiotic-fed mice exposed to early stress showed attenuated HPA axis reactivity and intestinal inflammation at weaning, as well as lower anxiety levels during adolescence. These findings were extended to other probiotic strains belonging to Bifidobacteria and Lactobacilli. Indeed, in juvenile rats, MS-induced hypercorticosteronaemia, intestinal hyper permeability and dysbiosis were all prevented by neonatal treatment with Bifidobacterium bifidum G9-1(Reference Fukui, Oshima and Tanaka351). Pretreatment with L. fermentum CECT 5716 was also able to attenuate the effects of a single 4 h-separation episode at PND10 (i.e. hypercorticosteronaemia and intestinal hyperpermeability)(Reference Vanhaecke, Aubert and Grohard352). In contrast, although maternal probiotics treatment with Bifidobacterium animalis subsp. lactis BB-12H and Propionibacterium jensenii 702 was shown to prevent the increase in plasma IFNγ in adult MS offspring(Reference Pusceddu, Kelly and Ariffin319), the same treatment increased plasma IL-6 in juveniles. In line with the latter, Barouei and collaborators showed that the maternal probiotic intervention induces MS-like dysbiosis along with increased levels of circulating corticosterone and adrenocorticotropic hormone in non-stressed developing offspring(Reference Barouei, Moussavi and Hodgson298). A recent study also reports preventive effects of maternal treatment with the probiotic Lacidofil® (L. rhamnosus R0011 and L. helveticus R0052), via the maternal drinking-water during the period of stress, on abnormal mPFC neural fear circuitry development in stressed pups(Reference Cowan, Stylianakis and Richardson353). Interestingly, the effects of neonatal probiotics in MS models are not restricted to stress response and depressive-like behaviours. It has been reported that MS disturbs puberty onset in a sex-dependent manner, but this effect is prevented by probiotic neonatal administration with Lacidofil®(Reference Cowan and Richardson354). In another study, MS rats transmitted their conditioned aversive memory to the next generation, but this effect was abolished if the F0 fathers or the F1 offspring was supplemented with Lacidofil®(Reference Callaghan, Cowan and Richardson355).

Notably, the majority of these findings were obtained using males only. Since a consistent gender effect has been reported in the prevalence of anxiety and depression, but also IBS, with higher rates in women than men(Reference Steel, Marnane and Iranpour356), additional preclinical studies using female animals are required. Moreover, the translational potential of these findings is currently limited by methodological and technical issues. It is not clear whether probiotic strains survive under aerobic conditions and are able to efficiently colonise the gut of the recipient. Future studies should systematically assess post-treatment colonisation to draw conclusions. In this regard, comparing heat-killed v. live probiotics can also be helpful to better understand their underlying mechanisms of action. Furthermore, there is a need to improve the dosage, treatment duration and route of administration. Indeed, only a few studies in animal models addressed the dose-dependency of the effects of probiotics. It is suggested that multi-strain probiotic combinations may provide greater health benefits, but this hypothesis has not been clearly tested. The systematic comparison of the effects of probiotics with that of a clinical drug such as anxiolytics or antidepressants appears critical to quantify the benefits.

It has been proposed that probiotics might represent an adjuvant therapy in psychiatric disorders including major depressive disorder, although well-designed clinical trials are needed to make clear conclusions(Reference Vlainić, Šuran and Vlainić357). A recent study reported that pregnant women supplemented with L. rhamnosus until 6 months postpartum had significantly lower depression and anxiety scores in the postpartum period(Reference Slykerman, Hood and Wickens358). To date, antidepressant effects of probiotics have been reported in three double-blind studies conducted in subjects diagnosed with significant anxiety or depression symptoms(Reference Majeed, Nagabhushanam and Arumugam359,Reference Romijn, Rucklidge and Kuijer360) and in major depressive disorder patients(Reference Akkasheh, Kashani-Poor and Tajabadi-Ebrahimi361). However, based on preclinical data, psychotropic-like effects of probiotics on mood and anxiety in subjects exposed to early-life adversity still need to be confirmed in human trials(Reference Rackers, Thomas and Williamson362). One study has explored the effects of probiotic strains L. rhamnosus HN001 or B. animalis subsp. lactis HN019 in 11-year-old children supplemented from fetal life to age 2 years on neurodevelopment, but found no major effect of probiotics(Reference Slykerman, Kang and Van Zyl363); yet the impact on emotional behaviours and especially in early-stressed patients remain unknown.

Prebiotics and symbiotics

Prebiotics are nutrients that can be fermented by microbes in the gut and thus favour the growth of certain microbial communities(Reference Gibson and Roberfroid364). In comparison with probiotics, a much smaller number of studies have examined the effects of prebiotics on behaviour (see(Reference Kao, Harty and Burnet365) for review). These include investigations of galacto-oligosaccharides (GOS) and fructo-oligosaccharides (FOS), which are sources of nutrition for Bifidobacteria and Lactobacilli. The effects of FOS and GOS have been tested in C57BL/6J male mice in basal and chronic stress situations(Reference Burokas, Arboleya and Moloney366). GOS of FOS alone showed some levels of protective effects but to a much lower extent compared with GOS and FOS. Conversely, human-milk oligosaccharide prebiotics have been reported to impact brain development and cognitive functions(Reference Wang and Brand-Miller367,Reference Wang, McVeagh and Petocz368) . Mice supplemented with human-milk oligosaccharides in their diet (2 weeks) were protected against stress-induced hyperanxiety(Reference Tarr, Galley and Fisher369). Apart from these effects on emotional behaviours, other studies have reported improved learning and memory performance in animals supplemented with different oligosaccharides including human-milk oligosaccharides(Reference Jia, Lu and Gao370Reference Yen, Wang and Wu373). Together, these findings suggest that combining several probiotics and/or prebiotics can improve the treatment outcome. For instance, increased intestinal permeability in adolescent MS rats was prevented by a symbiotic diet containing arachidonic acid and DHA, GOS and FOS and Lactobacillus paracasei NCC2461 (Reference García-Ródenas, Bergonzelli and Nutten299). In another study, MS rats were treated with either the prebiotics polydextrose and GOS, the probiotic L. rhamnosus GG or the symbiotic combination from weaning onwards(Reference McVey Neufeld, O'Mahony and Hoban374). Only the combination of pre- and probiotics was able to normalise anxiety in the open field test, although it impaired corticosterone negative feedback following acute restraint stress. In addition, expression of GABA receptor A2 (Gabra2) in the hippocampus was restored only by the combination of pre- and probiotics, whereas expression of GR (Nr3c1) was restored by L. rhamnosus GG alone.

Conclusion

MS induces a variety of long-term alterations similar to that observed in human subjects with a history of childhood adversity. In this review, we have outlined the specific effects of MS on both the brain and the gut, illustrating the validity of this model with respect to clinical data. In addition, the pivotal role played by gut microbiota in mediating the lasting imprinting by MS is highlighted in numerous studies using microbiota-directed interventions such as probiotics treatments. Preclinical studies suggest that nutritional approaches with pro- and prebiotics may constitute safe and efficient strategies to attenuate the effects of early-life stress on the gut–brain axis. However, it is still not clear whether gut dysbiosis, leakiness or inflammation precede each other and if they are the cause or consequence of stress-induced alterations within the brain. In this respect, studies are needed to understand how chronic neonatal stress disrupts gut–brain homoeostasis during development and which molecular mechanisms underlie the subsequent long-term imprinting. Moreover, despite widespread sex differences in both GI and neuropsychiatric vulnerability, there is still a gap to fill in the literature as regards the issue of sex. Meta-analyses on the impact of probiotics on anxiety and depressive-like symptoms exist, but the vast majority of the studies are conducted in healthy subjects and recent findings demonstrate that the effects of probiotics may differ between stressed and unstressed subjects(Reference Papalini, Michels and Kohn375). Future studies should develop nutritional strategies combining multiple prebiotics and probiotics, in addition to usual pharmacological strategies, to examine their impact at adult age on symptoms associated with early-life adversity using randomised placebo-control trials, with an effort to adapt these strategies according to sex(Reference Mayer and Hsiao376). Furthermore, prebiotics and probiotics effects should also be examined during development in populations exposed to stress. In human trials, it would be particularly valuable to study the potential preventive effects of prebiotics and probiotics after different stress experiences such as early-life traumas, but also parental depression, perinatal infections, premature birth or low parental socioeconomic status. Finally, early-life adversity is associated with poor diet quality at adulthood(Reference Gavrieli, Farr and Davis377). In this context, it would be crucial to improve health policies and to implement preventive interventions with nutritional advices in populations exposed to early-life adversity.

Financial Support

The authors were supported by the University of Bordeaux and INRA. M. R. was supported by the French ministry of research and education and Labex Brain Bordeaux Region Aquitaine Initiative for Neurosciences (PhD extension grant).

Conflict of Interest

None.

Authorship

M. R. drafted the first version of the manuscript then both authors revised and approved the manuscript.

Footnotes

Present address: Microenvironment & Immunity Unit, INSERM U1224, Department of Immunology, Institut Pasteur, 25, Rue du Dr. Roux 75724 Paris, France. Email: marion.rincel@pasteur.fr

References

1.Rogers, GB, Keating, DJ, Young, RL et al. (2016) From gut dysbiosis to altered brain function and mental illness: mechanisms and pathways. Mol Psychiatry 21, 738748.CrossRefGoogle ScholarPubMed
2.Valles-Colomer, M, Falony, G, Darzi, Y et al. (2019) The neuroactive potential of the human gut microbiota in quality of life and depression. Nat Microbiol 4, 623632.CrossRefGoogle ScholarPubMed
3.Cryan, JF & Dinan, TG (2012) Mind-altering microorganisms: the impact of the gut microbiota on brain and behaviour. Nat Rev Neurosci 13, 701712.CrossRefGoogle ScholarPubMed
4.Grenham, S, Clarke, G, Cryan, JF et al. (2011) Brain-Gut-Microbe Communication in Health and Disease. Front Physiol 2, 94.CrossRefGoogle ScholarPubMed
5.Mayer, EA (2011) Gut feelings: the emerging biology of gut–brain communication. Nat Rev Neurosci 12, 453466.CrossRefGoogle ScholarPubMed
6.Dinan, TG, Stanton, C & Cryan, JF (2013) Psychobiotics: a novel class of psychotropic. Biol Psychiatry 74, 720726.CrossRefGoogle ScholarPubMed
7.Sherwin, E, Rea, K, Dinan, TG et al. (2016) A gut (microbiome) feeling about the brain. Curr Opin Gastroenterol 32, 96102.CrossRefGoogle Scholar
8.Koloski, NA, Jones, M, Kalantar, J et al. (2012) The brain-gut pathway in functional gastrointestinal disorders is bidirectional: a 12-year prospective population-based study. Gut 61, 12841290.CrossRefGoogle ScholarPubMed
9.Borre, YE, O'Keeffe, GW, Clarke, G et al. (2014) Microbiota and neurodevelopmental windows: implications for brain disorders. Trends Mol Med 20, 509518.CrossRefGoogle ScholarPubMed
10.Sharon, G, Sampson, TR, Geschwind, DH et al. (2016) The central nervous system and the gut microbiome. Cell 167, 915932.CrossRefGoogle ScholarPubMed
11.Chitkara, DK, van Tilburg, MAL, Blois-Martin, N et al. (2008) Early life risk factors that contribute to irritable bowel syndrome in adults: a systematic review. Am J Gastroenterol 103, 765774; quiz 775.CrossRefGoogle ScholarPubMed
12.Nemeroff, CB (2016) Paradise lost: the neurobiological and clinical consequences of child abuse and neglect. Neuron 89, 892909.CrossRefGoogle ScholarPubMed
13.O'Mahony, SM, Clarke, G, Dinan, TG et al. (2017) Irritable bowel syndrome and stress-related psychiatric co-morbidities: focus on early life stress. Handb Exp Pharmacol 239, 219246.CrossRefGoogle ScholarPubMed
14.Drossman, DA (2016) Functional gastrointestinal disorders: history, pathophysiology, clinical features and Rome IV. Gastroenterology S0016-5085(16)00223-7.CrossRefGoogle ScholarPubMed
15.O'Mahony, SM, Hyland, NP, Dinan, TG et al. (2011) Maternal separation as a model of brain-gut axis dysfunction. Psychopharmacology 214, 7188.CrossRefGoogle ScholarPubMed
16.Denenberg, VH, Ottinger, DR & Stephens, MW (1962) Effects of maternal factors upon growth and behavior of the rat. Child Dev 33, 6571.Google ScholarPubMed
17.Levine, S (1957) Infantile experience and resistance to physiological stress. Science 126, 405.CrossRefGoogle ScholarPubMed
18.Meaney, MJ, Mitchell, JB, Aitken, DH et al. (1991) The effects of neonatal handling on the development of the adrenocortical response to stress: implications for neuropathology and cognitive deficits in later life. Psychoneuroendocrinology 16, 85103.CrossRefGoogle ScholarPubMed
19.Plotsky, PM & Meaney, MJ (1993) Early, postnatal experience alters hypothalamic corticotropin-releasing factor (CRF) mRNA, median eminence CRF content and stress-induced release in adult rats. Brain Res Mol Brain Res 18, 195200.CrossRefGoogle ScholarPubMed
20.Seay, B, Hansen, E & Harlow, HF (1962) Mother–infant separation in monkeys. J Child Psychol Psychiatry 3, 123132.CrossRefGoogle ScholarPubMed
21.Ackerman, SH, Hofer, MA & Weiner, H (1978) Early maternal separation increases gastric ulcer risk in rats by producing a latent thermoregulatory disturbance. Science 201, 373376.CrossRefGoogle ScholarPubMed
22.Lippmann, M, Bress, A, Nemeroff, CB et al. (2007) Long-term behavioural and molecular alterations associated with maternal separation in rats: Molecular adaptations after maternal separation. Eur J Neurosci 25, 30913098.CrossRefGoogle Scholar
23.Barna, I, Bálint, E, Baranyi, J et al. (2003) Gender-specific effect of maternal deprivation on anxiety and corticotropin-releasing hormone mRNA expression in rats. Brain Res Bull 62, 8591.CrossRefGoogle ScholarPubMed
24.Roman, E, Gustafsson, L, Berg, M et al. (2006) Behavioral profiles and stress-induced corticosteroid secretion in male Wistar rats subjected to short and prolonged periods of maternal separation. Horm Behav 50, 736747.CrossRefGoogle Scholar
25.Schmidt, M, Enthoven, L, van Woezik, JHG et al. (2004) The dynamics of the hypothalamic–pituitary–adrenal axis during maternal deprivation. J Neuroendocrinol 16, 5257.CrossRefGoogle ScholarPubMed
26.Viveros, MP, Llorente, R, López-Gallardo, M et al. (2009) Sex-dependent alterations in response to maternal deprivation in rats. Psychoneuroendocrinology 34(Suppl 1), S217S226.CrossRefGoogle ScholarPubMed
27.Korosi, A & Baram, TZ (2010) Plasticity of the stress response early in life: mechanisms and significance. Dev Psychobiol 52, 661670.CrossRefGoogle ScholarPubMed
28.Pryce, CR & Feldon, J (2003) Long-term neurobehavioural impact of the postnatal environment in rats: manipulations, effects and mediating mechanisms. Neurosci Biobehav Rev 27, 5771.CrossRefGoogle ScholarPubMed
29.Vetulani, J (2013) Early maternal separation: a rodent model of depression and a prevailing human condition. Pharmacol Rep 65, 14511461.CrossRefGoogle Scholar
30.Suchecki, D, Rosenfeld, P & Levine, S (1993) Maternal regulation of the hypothalamic–pituitary–adrenal axis in the infant rat: the roles of feeding and stroking. Brain Res Dev Brain Res 75, 185192.CrossRefGoogle ScholarPubMed
31.Van Oers, HJ, de Kloet, ER, Whelan, T et al. (1998) Maternal deprivation effect on the infant's neural stress markers is reversed by tactile stimulation and feeding but not by suppressing corticosterone. J Neurosci 18, 1017110179.CrossRefGoogle Scholar
32.Cirulli, F, Francia, N, Berry, A et al. (2009) Early life stress as a risk factor for mental health: role of neurotrophins from rodents to non-human primates. Neurosci Biobehav Rev 33, 573585.CrossRefGoogle ScholarPubMed
33.Tan, S, Ho, HS, Song, AY et al. (2017) Maternal separation does not produce a significant behavioral change in Mice. Exp Neurobiol 26, 390398.CrossRefGoogle Scholar
34.Millstein, RA & Holmes, A (2007) Effects of repeated maternal separation on anxiety- and depression-related phenotypes in different mouse strains. Neurosci Biobehav Rev 31, 317.CrossRefGoogle ScholarPubMed
35.Franklin, TB, Russig, H, Weiss, IC et al. (2010) Epigenetic transmission of the impact of early stress across generations. Biol Psychiatry 68, 408415.CrossRefGoogle Scholar
36.Gapp, K, Soldado-Magraner, S, Alvarez-Sánchez, M et al. (2014) Early life stress in fathers improves behavioural flexibility in their offspring. Nat Commun 5, 5466.CrossRefGoogle ScholarPubMed
37.Tchenio, A, Lecca, S, Valentinova, K et al. (2017) Limiting habenular hyperactivity ameliorates maternal separation-driven depressive-like symptoms. Nat Commun 8, 1135.CrossRefGoogle ScholarPubMed
38.Rincel, M, Aubert, P, Chevalier, J et al. (2019) Multi-hit early life adversity affects gut microbiota, brain and behavior in a sex-dependent manner. Brain Behav Immun S0889-1591(18)30570-1.CrossRefGoogle Scholar
39.Cryan, JF & Holmes, A (2005) The ascent of mouse: advances in modelling human depression and anxiety. Nat Rev Drug Discov 4, 775790.CrossRefGoogle ScholarPubMed
40.Nestler, EJ & Hyman, SE (2010) Animal models of neuropsychiatric disorders. Nat Neurosci 13, 11611169.CrossRefGoogle ScholarPubMed
41.York, JM, Blevins, NA, Baynard, T et al. (2012) Mouse testing methods in psychoneuroimmunology: an overview of how to measure sickness, depressive/anxietal, cognitive, and physical activity behaviors. Methods Mol Biol 934, 243276.CrossRefGoogle ScholarPubMed
42.Amini-Khoei, H, Haghani-Samani, E, Beigi, M et al. (2019) On the role of corticosterone in behavioral disorders, microbiota composition alteration and neuroimmune response in adult male mice subjected to maternal separation stress. Int Immunopharmacol 66, 242250.CrossRefGoogle ScholarPubMed
43.Aya-Ramos, L, Contreras-Vargas, C, Rico, JL et al. (2017) Early maternal separation induces preference for sucrose and aspartame associated with increased blood glucose and hyperactivity. Food Funct, 8, 25922600.CrossRefGoogle ScholarPubMed
44.Banqueri, M, Méndez, M & Arias, JL (2017) Behavioral effects in adolescence and early adulthood in two length models of maternal separation in male rats. Behav Brain Res 324, 7786.CrossRefGoogle ScholarPubMed
45.Bondar, NP, Lepeshko, AA & Reshetnikov, VV (2018) Effects of early-life stress on social and anxiety-like behaviors in adult mice: sex-specific effects. Behav Neurol 2018, 1538931.CrossRefGoogle ScholarPubMed
46.Caldji, C, Francis, D, Sharma, S et al. (2000) The effects of early rearing environment on the development of GABAA and central benzodiazepine receptor levels and novelty-induced fearfulness in the rat. Neuropsychopharmacology 22, 219229.CrossRefGoogle ScholarPubMed
47.Dandi, Ε, Kalamari, A, Touloumi, O et al. (2018) Beneficial effects of environmental enrichment on behavior, stress reactivity and synaptophysin/BDNF expression in hippocampus following early life stress. Int J Dev Neurosci 67, 1932.CrossRefGoogle Scholar
48.Daniels, WMU, Pietersen, CY, Carstens, ME et al. (2004) Maternal separation in rats leads to anxiety-like behavior and a blunted ACTH response and altered neurotransmitter levels in response to a subsequent stressor. Metab Brain Dis 19, 314.CrossRefGoogle Scholar
49.De Palma, G, Blennerhassett, P, Lu, J et al. (2015) Microbiota and host determinants of behavioural phenotype in maternally separated mice. Nat Commun 6, 7735.CrossRefGoogle ScholarPubMed
50.Feifel, AJ, Shair, HN & Schmauss, C (2017) Lasting effects of early life stress in mice: interaction of maternal environment and infant genes. Genes Brain Behav 16, 768780.Google ScholarPubMed
51.Felice, VD, Gibney, S M, Gosselin, RD et al. (2014) Differential activation of the prefrontal cortex and amygdala following psychological stress and colorectal distension in the maternally separated rat. Neuroscience 267, 252262.CrossRefGoogle ScholarPubMed
52.Francis, DD, Diorio, J, Plotsky, PM et al. (2002) Environmental enrichment reverses the effects of maternal separation on stress reactivity. J Neurosci 22, 78407843.CrossRefGoogle ScholarPubMed
53.Gracia-Rubio, I, Moscoso-Castro, M, Pozo, OJ et al. (2016) Maternal separation induces neuroinflammation and long-lasting emotional alterations in mice. Prog Neuropsychopharmacol Biol Psychiatry 65, 104117.CrossRefGoogle ScholarPubMed
54.Huot, RL, Thrivikraman, KV, Meaney, MJ et al. (2001) Development of adult ethanol preference and anxiety as a consequence of neonatal maternal separation in Long Evans rats and reversal with antidepressant treatment. Psychopharmacology 158, 366373.CrossRefGoogle ScholarPubMed
55.Kalinichev, M, Easterling, KW & Plotsky, PM (2002) Long-lasting changes in stress-induced corticosterone response and anxiety-like behaviors as a consequence of neonatal maternal separation in Long-Evans rats. Pharmacol Biochem Behav 73, 131140.CrossRefGoogle ScholarPubMed
56.Kiser, DP, Popp, S, Schmitt-Böhrer, AG et al. (2019) Early-life stress impairs developmental programming in cadherin 13 (CDH13)-deficient mice. Prog Neuropsychopharmacol Biol Psychiatry 89, 158168.CrossRefGoogle ScholarPubMed
57.Koe, AS, Ashokan, A & Mitra, R (2016) Short environmental enrichment in adulthood reverses anxiety and basolateral amygdala hypertrophy induced by maternal separation. Transl Psychiatry 6, e729.CrossRefGoogle ScholarPubMed
58.Lambás-Señas, L, Mnie-Filali, O, Certin, V et al. (2009) Functional correlates for 5-HT1A receptors in maternally deprived rats displaying anxiety and depression-like behaviors. Prog Neuropsychopharmacol Biol Psychiatry 33, 262268.CrossRefGoogle ScholarPubMed
59.Lee, JH, Kim, HJ, Kim, JG et al. (2007) Depressive behaviors and decreased expression of serotonin reuptake transporter in rats that experienced neonatal maternal separation. Neurosci Res 58, 3239.CrossRefGoogle ScholarPubMed
60.Lee, JH, Kim, JY & Jahng, JW (2014) Highly palatable food during adolescence improves anxiety-like behaviors and hypothalamic–pituitary–adrenal axis dysfunction in rats that experienced neonatal maternal separation. Endocrinol Metab 29, 169.CrossRefGoogle ScholarPubMed
61.Liu, C, Hao, S, Zhu, M et al. (2018) Maternal separation induces different autophagic responses in the hippocampus and prefrontal cortex of adult rats. Neuroscience 374, 287294.CrossRefGoogle ScholarPubMed
62.Maniam, J & Morris, MJ (2010) Voluntary exercise and palatable high-fat diet both improve behavioural profile and stress responses in male rats exposed to early life stress: role of hippocampus. Psychoneuroendocrinology 35, 15531564.CrossRefGoogle ScholarPubMed
63.de Melo, SR, de David Antoniazzi, CT, Hossain, S et al. (2018) Neonatal stress has a long-lasting sex-dependent effect on anxiety-like behavior and neuronal morphology in the prefrontal cortex and hippocampus. Dev Neurosci 40, 93103.CrossRefGoogle Scholar
64.Moya-Pérez, A, Perez-Villalba, A, Benítez-Páez, A et al. (2017) Bifidobacterium CECT 7765 modulates early stress-induced immune, neuroendocrine and behavioral alterations in mice. Brain Behav Immun 65, 4356.CrossRefGoogle ScholarPubMed
65.Park, HJ, Kim, SK, Kang, WS et al. (2014) Effects of essential oil from Chamaecyparis obtusa on cytokine genes in the hippocampus of maternal separation rats. Can J Physiol Pharmacol 92, 95101.CrossRefGoogle ScholarPubMed
66.Portero-Tresserra, M, Gracia-Rubio, I, Cantacorps, L et al. (2018) Maternal separation increases alcohol-drinking behaviour and reduces endocannabinoid levels in the mouse striatum and prefrontal cortex. Eur Neuropsychopharmacol 28, 499512.CrossRefGoogle ScholarPubMed
67.Rincel, M, Lépinay, AL, Delage, P et al. (2016) Maternal high-fat diet prevents developmental programming by early-life stress. Transl Psychiatry 6, e966.CrossRefGoogle ScholarPubMed
68.Shin, SY, Han, SH, Woo, RS et al. (2016) Adolescent mice show anxiety- and aggressive-like behavior and the reduction of long-term potentiation in mossy fiber-CA3 synapses after neonatal maternal separation. Neuroscience 316, 221231.CrossRefGoogle ScholarPubMed
69.Shu, C, Xiao, L, Tang, J et al. (2015) Blunted behavioral and molecular responses to chronic mild stress in adult rats with experience of infancy maternal separation. Tohoku J Exp Med 235, 8187.CrossRefGoogle ScholarPubMed
70.Troakes, C & Ingram, CD (2009) Anxiety behaviour of the male rat on the elevated plus maze: associated regional increase in c-fos mRNA expression and modulation by early maternal separation. Stress 12, 362369.CrossRefGoogle ScholarPubMed
71.Wang, Q, Dong, X, Wang, Y et al. (2017) Adolescent escitalopram prevents the effects of maternal separation on depression- and anxiety-like behaviours and regulates the levels of inflammatory cytokines in adult male mice. Int J Dev Neurosci 62, 3745.CrossRefGoogle ScholarPubMed
72.Yang, Y, Cheng, Z, Tang, H et al. (2016) Neonatal maternal separation impairs prefrontal cortical myelination and cognitive functions in rats through activation of Wnt signaling. Cereb Cortex 27, 28712884.Google Scholar
73.Eiland, L & McEwen, BS (2012) Early life stress followed by subsequent adult chronic stress potentiates anxiety and blunts hippocampal structural remodeling. Hippocampus 22(1), 8291.CrossRefGoogle ScholarPubMed
74.Marais, L, van Rensburg, SJ, van Zyl, JM et al. (2008) Maternal separation of rat pups increases the risk of developing depressive-like behavior after subsequent chronic stress by altering corticosterone and neurotrophin levels in the hippocampus. Neurosci Res 61, 106112.CrossRefGoogle ScholarPubMed
75.Amini-Khoei, H, Mohammadi-Asl, A, Amiri, S et al. (2017) Oxytocin mitigated the depressive-like behaviors of maternal separation stress through modulating mitochondrial function and neuroinflammation. Prog Neuropsychopharmacol Biol Psychiatry 76, 169178.CrossRefGoogle ScholarPubMed
76.Amiri, S, Amini-Khoei, H, Mohammadi-Asl, A et al. (2016) Involvement of D1 and D2 dopamine receptors in the antidepressant-like effects of selegiline in maternal separation model of mouse. Physiol Behav 163, 107114.CrossRefGoogle ScholarPubMed
77.Bai, M, Zhu, X, Zhang, Y et al. (2012) Abnormal hippocampal BDNF and miR-16 expression is associated with depression-like behaviors induced by stress during early life. PLoS ONE, 7, e46921.CrossRefGoogle ScholarPubMed
78.Desbonnet, L, Garrett, L, Clarke, G et al. (2010) Effects of the probiotic Bifidobacterium infantis in the maternal separation model of depression. Neuroscience 170, 11791188.CrossRefGoogle Scholar
79.Hui, JJ, Zhang, ZJ, Liu, SS et al. (2011) Hippocampal neurochemistry is involved in the behavioural effects of neonatal maternal separation and their reversal by post-weaning environmental enrichment: a magnetic resonance study. Behav Brain Res 217, 122127.CrossRefGoogle ScholarPubMed
80.Lesse, A, Rether, K, Gröger, N et al. (2017) Chronic postnatal stress induces depressive-like behavior in male mice and programs second-hit stress-induced gene expression patterns of OxtR and AvpR1a in adulthood. Mol Neurobiol 54, 48134819.CrossRefGoogle ScholarPubMed
81.Li, Y, Yang, T, Yao, Q et al. (2019) Metformin prevents colonic barrier dysfunction by inhibiting mast cell activation in maternal separation-induced IBS-like rats. Neurogastroenterol Motil 31, e13556.CrossRefGoogle ScholarPubMed
82.MacQueen, GM, Ramakrishnan, K, Ratnasingan, R et al. (2003) Desipramine treatment reduces the long-term behavioural and neurochemical sequelae of early-life maternal separation. Int J Neuropsychopharmacol 6, 391396.CrossRefGoogle ScholarPubMed
83.Masrour, FF, Peeri, M, Azarbayjani, MA et al. (2018) Voluntary exercise during adolescence mitigated negative the effects of maternal separation stress on the depressive-like behaviors of adult male rats: role of NMDA receptors. Neurochem Res 43, 10671074.CrossRefGoogle ScholarPubMed
84.Paternain, L, Martisova, E, Campión, J et al. (2016) Methyl donor supplementation in rats reverses the deleterious effect of maternal separation on depression-like behaviour. Behav Brain Res 299, 5158.CrossRefGoogle ScholarPubMed
85.Réus, GZ, Stringari, RB, Ribeiro, KF et al. (2011) Maternal deprivation induces depressive-like behaviour and alters neurotrophin levels in the rat brain. Neurochem Res 36, 460466.CrossRefGoogle ScholarPubMed
86.Réus, GZ, Fernandes, GC, de Moura, AB et al. (2017) Early life experience contributes to the developmental programming of depressive-like behaviour, neuroinflammation and oxidative stress. J Psychiatr Res 95, 196207.CrossRefGoogle ScholarPubMed
87.Sadeghi, M, Peeri, M & Hosseini, MJ (2016) Adolescent voluntary exercise attenuated hippocampal innate immunity responses and depressive-like behaviors following maternal separation stress in male rats. Physiol Behav 163, 177183.CrossRefGoogle ScholarPubMed
88.Sung, Y-H, Shin, M-S, Cho, S et al. (2010) Depression-like state in maternal rats induced by repeated separation of pups is accompanied by a decrease of cell proliferation and an increase of apoptosis in the hippocampus. Neurosci Lett 470, 8690.CrossRefGoogle Scholar
89.Uchida, S, Hara, K, Kobayashi, A et al. (2010) Early life stress enhances behavioral vulnerability to stress through the activation of REST4-mediated gene transcription in the medial prefrontal cortex of rodents. J Neurosci 30, 1500715018.CrossRefGoogle ScholarPubMed
90.Vargas, J, Junco, M, Gomez, C et al. (2016) Early life stress increases metabolic risk, HPA axis reactivity, and depressive-like behavior when combined with postweaning social isolation in rats. PLoS ONE 11, e0162665.CrossRefGoogle ScholarPubMed
91.Yamawaki, Y, Nishida, M, Harada, K et al. (2018) Data on the effect of maternal separation coupled with social isolation in a forced swim test and gene expression of glial fibrillary acid protein in the prefrontal cortex of rats. Data Brief 18, 496500.CrossRefGoogle Scholar
92.Dallé, E, Daniels, WMU & Mabandla, MV (2017) Fluvoxamine maleate normalizes striatal neuronal inflammatory cytokine activity in a Parkinsonian rat model associated with depression. Behav Brain Res 316, 189196.CrossRefGoogle Scholar
93.Kundakovic, M, Lim, S, Gudsnuk, K et al. (2013) Sex-specific and strain-dependent effects of early life adversity on behavioral and epigenetic outcomes. Front Psychiatry 4, 78.CrossRefGoogle ScholarPubMed
94.Maniam, J & Morris, MJ (2010) Palatable cafeteria diet ameliorates anxiety and depression-like symptoms following an adverse early environment. Psychoneuroendocrinology 35, 717728.CrossRefGoogle ScholarPubMed
95.Øines, E, Murison, R, Mrdalj, J et al. (2012) Neonatal maternal separation in male rats increases intestinal permeability and affects behavior after chronic social stress. Physiol Behav 105, 10581066.CrossRefGoogle ScholarPubMed
96.Yang, L, Xu, T, Zhang, K et al. (2016) The essential role of hippocampal alpha6 subunit-containing GABAA receptors in maternal separation stress-induced adolescent depressive behaviors. Behav Brain Res 313, 135143.CrossRefGoogle ScholarPubMed
97.Farrell, MR, Holland, FH, Shansky, RM et al. (2016) Sex-specific effects of early life stress on social interaction and prefrontal cortex dendritic morphology in young rats. Behav Brain Res 310, 119125.CrossRefGoogle ScholarPubMed
98.Tsuda, MC, Yamaguchi, N, Nakata, M et al. (2014) Modification of female and male social behaviors in estrogen receptor beta knockout mice by neonatal maternal separation. Front Neurosci 8, 274.CrossRefGoogle ScholarPubMed
99.Zimmerberg, B & Sageser, KA (2011) Comparison of two rodent models of maternal separation on juvenile social behavior. Front Psychiatry 2, 39.CrossRefGoogle ScholarPubMed
100.Moffett, M, Vicentic, A, Kozel, M et al. (2007) Maternal separation alters drug intake patterns in adulthood in rats. Biochem Pharmacol 73, 321330.CrossRefGoogle ScholarPubMed
101.Kosten, TA, Kim, JJ & Lee, HJ (2012) Early life manipulations alter learning and memory in rats. Neurosci Biobehav Rev 36, 19852006.CrossRefGoogle ScholarPubMed
102.Couto, FS, Batalha, VL, Valadas, JS et al. (2012) Escitalopram improves memory deficits induced by maternal separation in the rat. Eur J Pharmacol 695, 7175.CrossRefGoogle ScholarPubMed
103.Dalaveri, F, Nakhaee, N, Esmaeilpour, K et al. (2017) Effects of maternal separation on nicotine-induced conditioned place preference and subsequent learning and memory in adolescent female rats. Neurosci Lett 639, 151156.CrossRefGoogle ScholarPubMed
104.Guo, L, Liang, X, Liang, Z et al. (2018) Electroacupuncture ameliorates cognitive deficit and improves hippocampal synaptic plasticity in adult rat with neonatal maternal separation. Evid Based Complement Alternat Med 2018, 2468105.CrossRefGoogle ScholarPubMed
105.Reshetnikov, VV, Kovner, AV, Lepeshko, AA et al. (2018) Stress early in life leads to cognitive impairments, reduced numbers of CA3 neurons and altered maternal behavior in adult female mice. Genes Brain Behav e12541.CrossRefGoogle ScholarPubMed
106.Son, GH, Geum, D, Chung, S et al. (2006) Maternal stress produces learning deficits associated with impairment of NMDA receptor-mediated synaptic plasticity. J Neurosci 26, 33093318.CrossRefGoogle ScholarPubMed
107.Wang, L, Jiao, J & Dulawa, SC (2011) Infant maternal separation impairs adult cognitive performance in BALB/cJ mice. Psychopharmacology 216, 207218.CrossRefGoogle ScholarPubMed
108.Aisa, B, Tordera, R, Lasheras, B et al. (2007) Cognitive impairment associated to HPA axis hyperactivity after maternal separation in rats. Psychoneuroendocrinology 32, 256266.CrossRefGoogle ScholarPubMed
109.Benetti, F, Mello, PB, Bonini, JS et al. (2009) Early postnatal maternal deprivation in rats induces memory deficits in adult life that can be reversed by donepezil and galantamine. Int J Dev Neurosci 27, 5964.CrossRefGoogle ScholarPubMed
110.Furukawa, M, Tsukahara, T, Tomita, K et al. (2017) Neonatal maternal separation delays the GABA excitatory-to-inhibitory functional switch by inhibiting KCC2 expression. Biochem Biophys Res Commun 493, 12431249.CrossRefGoogle ScholarPubMed
111.Hidaka, C, Kashio, T, Uchigaki, D et al. (2018) Vulnerability or resilience of motopsin knockout mice to maternal separation stress depending on adulthood behaviors. Neuropsychiatr Dis Treat 14, 22552268.CrossRefGoogle ScholarPubMed
112.Hulshof, HJ, Novati, A, Sgoifo, A et al. (2011) Maternal separation decreases adult hippocampal cell proliferation and impairs cognitive performance but has little effect on stress sensitivity and anxiety in adult Wistar rats. Behav Brain Res 216, 552560.CrossRefGoogle ScholarPubMed
113.Moreno Gudiño, H, Carías Picón, D & de Brugada Sauras, I (2017) Dietary choline during periadolescence attenuates cognitive damage caused by neonatal maternal separation in male rats. Nutr Neurosci 20, 327335.CrossRefGoogle ScholarPubMed
114.Pinheiro, RMC, de Lima, MNM, Portal, BCD et al. (2014) Long-lasting recognition memory impairment and alterations in brain levels of cytokines and BDNF induced by maternal deprivation: effects of valproic acid and topiramate. J Neural Transm 122, 709719. doi:10.1007/s00702-014-1303-2.CrossRefGoogle ScholarPubMed
115.Wang, A, Nie, W, Li, H et al. (2014) Epigenetic upregulation of corticotrophin-releasing hormone mediates postnatal maternal separation-induced memory deficiency. PLoS ONE 9, e94394.CrossRefGoogle ScholarPubMed
116.Baudin, A, Blot, K, Verney, C et al. (2012) Maternal deprivation induces deficits in temporal memory and cognitive flexibility and exaggerates synaptic plasticity in the rat medial prefrontal cortex. Neurobiol Learn Mem 98, 207214.CrossRefGoogle ScholarPubMed
117.Boutros, N, Der-Avakian, A, Markou, A et al. (2017) Effects of early life stress and adolescent ethanol exposure on adult cognitive performance in the 5-choice serial reaction time task in Wistar male rats. Psychopharmacology 234, 15491556.CrossRefGoogle ScholarPubMed
118.Do Prado, CH, Narahari, T, Holland, FH et al. (2015) Effects of early adolescent environmental enrichment on cognitive dysfunction, prefrontal cortex development, and inflammatory cytokines after early life stress. Dev Psychobiol 58, 482491.CrossRefGoogle ScholarPubMed
119.Lejeune, S, Dourmap, N, Martres, MP et al. (2013) The dopamine D1 receptor agonist SKF 38393 improves temporal order memory performance in maternally deprived rats. Neurobiol Learn Mem 106, 268273.CrossRefGoogle ScholarPubMed
120.Thomas, AW, Caporale, N, Wu, C et al. (2016) Early maternal separation impacts cognitive flexibility at the age of first independence in mice. Dev Cogn Neurosci 18, 4956.CrossRefGoogle ScholarPubMed
121.Wilber, AA, Southwood, CJ & Wellman, CL (2009) Brief neonatal maternal separation alters extinction of conditioned fear and corticolimbic glucocorticoid and NMDA receptor expression in adult rats. Dev Neurobiol 69, 7387.CrossRefGoogle ScholarPubMed
122.Diehl, LA, Pereira, NSC, Laureano, DP et al. (2014) Contextual fear conditioning in maternal separated rats: the amygdala as a site for alterations. Neurochem Res 39, 384393.CrossRefGoogle ScholarPubMed
123.Elliott, ND & Richardson, R (2019) The effects of early life stress on context fear generalization in adult rats. Behav Neurosci 133, 5058.CrossRefGoogle ScholarPubMed
124.Mishra, PK, Kutty, BM & Laxmi, TR (2019) The impact of maternal separation and isolation stress during stress hyporesponsive period on fear retention and extinction recall memory from 5-week- to 1-year-old rats. Exp Brain Res 237, 181190.CrossRefGoogle ScholarPubMed
125.Sampath, D, Sabitha, KR, Hegde, P et al. (2014) A study on fear memory retrieval and REM sleep in maternal separation and isolation stressed rats. Behav Brain Res 273, 144154.CrossRefGoogle Scholar
126.Toda, H, Boku, S, Nakagawa, S et al. (2014) Maternal separation enhances conditioned fear and decreases the mRNA levels of the neurotensin receptor 1 gene with hypermethylation of this gene in the rat amygdala. PLoS ONE 9, e97421.CrossRefGoogle ScholarPubMed
127.Xiong, GJ, Yang, Y, Cao, J et al. (2015) Fluoxetine treatment reverses the intergenerational impact of maternal separation on fear and anxiety behaviors. Neuropharmacology 92, 17.CrossRefGoogle ScholarPubMed
128.Biagini, G, Pich, EM, Carani, C et al. (1998) Postnatal maternal separation during the stress hyporesponsive period enhances the adrenocortical response to novelty in adult rats by affecting feedback regulation in the CA1 hippocampal field. Int J Dev Neurosci 16, 187197.CrossRefGoogle ScholarPubMed
129.Cotella, EM, Mestres Lascano, I, Franchioni, L et al. (2013) Long-term effects of maternal separation on chronic stress response suppressed by amitriptyline treatment. Stress 16, 477481.CrossRefGoogle ScholarPubMed
130.Ladd, CO, Owens, MJ & Nemeroff, CB (1996) Persistent changes in corticotropin-releasing factor neuronal systems induced by maternal deprivation. Endocrinology 137, 12121218.CrossRefGoogle ScholarPubMed
131.Ladd, CO, Huot, RL, Thrivikraman, KV et al. (2000) Long-term behavioral and neuroendocrine adaptations to adverse early experience. Prog Brain Res 122, 81103.CrossRefGoogle ScholarPubMed
132.Lehmann, J, Russig, H, Feldon, J et al. (2002) Effect of a single maternal separation at different pup ages on the corticosterone stress response in adult and aged rats. Pharmacol Biochem Behav 73, 141145.CrossRefGoogle ScholarPubMed
133.Patchev, VK, Montkowski, A, Rouskova, D et al. (1997) Neonatal treatment of rats with the neuroactive steroid tetrahydrodeoxycorticosterone (THDOC) abolishes the behavioral and neuroendocrine consequences of adverse early life events. J Clin Invest 99, 962966.CrossRefGoogle ScholarPubMed
134.Rosenfeld, P, Wetmore, JB & Levine, S (1992) Effects of repeated maternal separations on the adrenocortical response to stress of preweanling rats. Physiol Behav 52, 787791.CrossRefGoogle ScholarPubMed
135.Slotten, HA, Kalinichev, M, Hagan, JJ et al. (2006) Long-lasting changes in behavioural and neuroendocrine indices in the rat following neonatal maternal separation: gender-dependent effects. Brain Res 1097, 123132.CrossRefGoogle ScholarPubMed
136.de Almeida Magalhães, T, Correia, D, de Carvalho, LM et al. (2018) Maternal separation affects expression of stress response genes and increases vulnerability to ethanol consumption. Brain Behav 8, e00841.CrossRefGoogle ScholarPubMed
137.Rivarola, MA & Renard, GM (2014) What we know about the long-term consequences of early maternal separation and neuroendocrine response to stress. Rev Farmacol Chile 7, 17.Google Scholar
138.Veenema, AH (2012) Toward understanding how early-life social experiences alter oxytocin- and vasopressin-regulated social behaviors. Horm Behav 61, 304312.CrossRefGoogle ScholarPubMed
139.Rivarola, MA & Suárez, MM (2009) Early maternal separation and chronic variable stress in adulthood changes the neural activity and the expression of glucocorticoid receptor in limbic structures. Int J Dev Neurosci 27, 567574.CrossRefGoogle ScholarPubMed
140.Chen, A, Chen, Y, Tang, Y et al. (2017) Hippocampal AMPARs involve the central sensitization of rats with irritable bowel syndrome. Brain Behav 7, e00650.CrossRefGoogle ScholarPubMed
141.Katsouli, S, Stamatakis, A, Giompres, P et al. (2014) Sexually dimorphic long-term effects of an early life experience on AMPA receptor subunit expression in rat brain. Neuroscience 257, 4964.CrossRefGoogle ScholarPubMed
142.Pickering, C, Gustafsson, L, Cebere, A et al. (2006) Repeated maternal separation of male Wistar rats alters glutamate receptor expression in the hippocampus but not the prefrontal cortex. Brain Res 1099, 101108.CrossRefGoogle Scholar
143.Bravo, JA, Dinan, TG & Cryan, JF (2014) Early-life stress induces persistent alterations in 5-HT1A receptor and serotonin transporter mRNA expression in the adult rat brain. Front Mol Neurosci 7, 24.CrossRefGoogle ScholarPubMed
144.Chen, YL, Huang, XQ, Xu, SJ et al. (2013) Relieving visceral hyperalgesia effect of Kangtai capsule and its potential mechanisms via modulating the 5-HT and NO level in vivo. Phytomedicine 20, 249257.CrossRefGoogle ScholarPubMed
145.Kawakami, SE, Quadros, IMH, Machado, RB et al. (2013) Sex-dependent effects of maternal separation on plasma corticosterone and brain monoamines in response to chronic ethanol administration. Neuroscience 253, 5566.CrossRefGoogle ScholarPubMed
146.O'Mahony, S, Chua, ASB, Quigley, EMM et al. (2008) Evidence of an enhanced central 5HT response in irritable bowel syndrome and in the rat maternal separation model. Neurogastroenterol Motil 20, 680688.CrossRefGoogle ScholarPubMed
147.Wu, JC, Ziea, ET, Lao, L et al. (2010) Effect of electroacupuncture on visceral hyperalgesia, serotonin and Fos expression in an animal model of irritable bowel syndrome. Neurogastroenterol Motil 16, 306314.CrossRefGoogle Scholar
148.Arborelius, L & Eklund, MB (2007) Both long and brief maternal separation produces persistent changes in tissue levels of brain monoamines in middle-aged female rats. Neuroscience 145, 738750.CrossRefGoogle ScholarPubMed
149.Brake, WG, Zhang, TY, Diorio, J et al. (2004) Influence of early postnatal rearing conditions on mesocorticolimbic dopamine and behavioural responses to psychostimulants and stressors in adult rats. Eur J Neurosci 19, 18631874.CrossRefGoogle ScholarPubMed
150.Li, M, Xue, X, Shao, S et al. (2013) Cognitive, emotional and neurochemical effects of repeated maternal separation in adolescent rats. Brain Res 1518, 8290.CrossRefGoogle ScholarPubMed
151.Matthews, K, Dalley, JW, Matthews, C et al. (2001) Periodic maternal separation of neonatal rats produces region- and gender-specific effects on biogenic amine content in postmortem adult brain. Synapse 40, 110.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
152.Ploj, K, Roman, E & Nylander, I (2003) Long-term effects of maternal separation on ethanol intake and brain opioid and dopamine receptors in male Wistar rats. Neuroscience 121, 787799.CrossRefGoogle ScholarPubMed
153.Romano-López, A, Méndez-Díaz, M, García, FG et al. (2016) Maternal separation and early stress cause long-lasting effects on dopaminergic and endocannabinergic systems and alters dendritic morphology in the nucleus accumbens and frontal cortex in rats. Dev Neurobiol 76, 819831.CrossRefGoogle ScholarPubMed
154.Ploj, K, Roman, E & Nylander, I (2003) Long-term effects of short and long periods of maternal separation on brain opioid peptide levels in male Wistar rats. Neuropeptides 37, 149156.CrossRefGoogle Scholar
155.Daubert, EA & Condron, BG (2010) Serotonin: a regulator of neuronal morphology and circuitry. Trends Neurosci 33, 424434.CrossRefGoogle ScholarPubMed
156.Kim, DY & Camilleri, M (2000) Serotonin: a mediator of the brain–gut connection. Am J Gastroenterol 95, 26982709.Google ScholarPubMed
157.Sommer, C (2004) Serotonin in pain and analgesia: actions in the periphery. Mol Neurobiol 30, 117125.CrossRefGoogle ScholarPubMed
158.Chocyk, A, Bobula, B, Dudys, D et al. (2013) Early-life stress affects the structural and functional plasticity of the medial prefrontal cortex in adolescent rats. Eur J Neurosci 38, 20892107.CrossRefGoogle ScholarPubMed
159.Danielewicz, J & Hess, G (2014) Early life stress alters synaptic modification range in the rat lateral amygdala. Behav Brain Res 265, 3237.CrossRefGoogle ScholarPubMed
160.Muhammad, A, Carroll, C & Kolb, B (2012) Stress during development alters dendritic morphology in the nucleus accumbens and prefrontal cortex. Neuroscience 216, 103109.CrossRefGoogle ScholarPubMed
161.Rincel, M, Lépinay, AL, Janthakhin, Y et al. (2017) Maternal high-fat diet and early life stress differentially modulate spine density and dendritic morphology in the medial prefrontal cortex of juvenile and adult rats. Brain Struct Funct 223, 883895.CrossRefGoogle ScholarPubMed
162.Sachs, BD, Tran, HL, Folse, E et al. (2018) Brain-region-specific molecular responses to maternal separation and social defeat stress in mice. Neuroscience 373, 122136.CrossRefGoogle ScholarPubMed
163.Soztutar, E, Colak, E & Ulupinar, E (2016) Gender- and anxiety level-dependent effects of perinatal stress exposure on medial prefrontal cortex. Exp Neurol 275, 274284.CrossRefGoogle ScholarPubMed
164.Bock, J, Gruss, M, Becker, S et al. (2005) Experience-induced changes of dendritic spine densities in the prefrontal and sensory cortex: correlation with developmental time windows. Cereb Cortex 15, 802808.CrossRefGoogle ScholarPubMed
165.Cao, X, Huang, S, Cao, J et al. (2014) The timing of maternal separation affects Morris water maze performance and long-term potentiation in male rats: timing of maternal separation on rats. Dev Psychobiol 56, 11021109.CrossRefGoogle ScholarPubMed
166.Chen, A, Bao, C, Tang, Y et al. (2015) Involvement of protein kinase ζ in the maintenance of hippocampal long-term potentiation in rats with chronic visceral hypersensitivity. J Neurophysiol 113, 30473055.CrossRefGoogle ScholarPubMed
167.Gos, T, Bock, J, Poeggel, G et al. (2008) Stress-induced synaptic changes in the rat anterior cingulate cortex are dependent on endocrine developmental time windows. Synapse 62, 229232.CrossRefGoogle ScholarPubMed
168.Gruss, M, Braun, K, Frey, JU et al. (2008) Maternal separation during a specific postnatal time window prevents reinforcement of hippocampal long-term potentiation in adolescent rats. Neuroscience 152, 17.CrossRefGoogle ScholarPubMed
169.Monroy, E, Hernández-Torres, E & Flores, G (2010) Maternal separation disrupts dendritic morphology of neurons in prefrontal cortex, hippocampus, and nucleus accumbens in male rat offspring. J Chem Neuroanat 40, 93101.CrossRefGoogle ScholarPubMed
170.Muhammad, A & Kolb, B (2011) Maternal separation altered behavior and neuronal spine density without influencing amphetamine sensitization. Behav Brain Res 223, 716.CrossRefGoogle ScholarPubMed
171.Sousa, VC, Vital, J, Costenla, AR et al. (2014) Maternal separation impairs long term-potentiation in CA1–CA3 synapses and hippocampal-dependent memory in old rats. Neurobiol Aging 35, 16801685.CrossRefGoogle ScholarPubMed
172.Aisa, B, Elizalde, N, Tordera, R et al. (2009) Effects of neonatal stress on markers of synaptic plasticity in the hippocampus: Implications for spatial memory. Hippocampus 19, 12221231.CrossRefGoogle ScholarPubMed
173.De Lima, MNM, Presti-Torres, J, Vedana, G et al. (2011) Early life stress decreases hippocampal BDNF content and exacerbates recognition memory deficits induced by repeated D-amphetamine exposure. Behav Brain Res 224, 100106.CrossRefGoogle ScholarPubMed
174.Park, H & Poo, M (2012) Neurotrophin regulation of neural circuit development and function. Nat Rev Neurosci 14, 723.CrossRefGoogle Scholar
175.Mirescu, C, Peters, JD & Gould, E (2004) Early life experience alters response of adult neurogenesis to stress. Nat Neurosci 7, 841846.CrossRefGoogle ScholarPubMed
176.Suri, D, Veenit, V, Sarkar, A et al. (2013) Early stress evokes age-dependent biphasic changes in hippocampal neurogenesis, Bdnf expression, and cognition. Biol Psychiatry 73, 658666.CrossRefGoogle ScholarPubMed
177.Lajud, N, Roque, A, Cajero, M et al. (2012) Periodic maternal separation decreases hippocampal neurogenesis without affecting basal corticosterone during the stress hyporesponsive period, but alters HPA axis and coping behavior in adulthood. Psychoneuroendocrinology 37, 410420.CrossRefGoogle ScholarPubMed
178.Autry, AE & Monteggia, LM (2012) Brain-derived neurotrophic factor and neuropsychiatric disorders. Pharmacol Rev 64, 238258.CrossRefGoogle ScholarPubMed
179.Wieck, A, Andersen, SL & Brenhouse, HC (2013) Evidence for a neuroinflammatory mechanism in delayed effects of early life adversity in rats: relationship to cortical NMDA receptor expression. Brain Behav Immun 28, 218226.CrossRefGoogle ScholarPubMed
180.Tang, HL, Zhang, G, Ji, NN et al. (2017) Toll-like receptor 4 in paraventricular nucleus mediates visceral hypersensitivity induced by maternal separation. Front Pharmacol 8, 309.CrossRefGoogle ScholarPubMed
181.Nakamoto, K, Aizawa, F, Kinoshita, M et al. (2017) Astrocyte activation in locus coeruleus is involved in neuropathic pain exacerbation mediated by maternal separation and social isolation stress. Front Pharmacol 8, 401.CrossRefGoogle ScholarPubMed
182.Ershov, NI, Bondar, NP, Lepeshko, AA et al. (2018) Consequences of early life stress on genomic landscape of H3K4me3 in prefrontal cortex of adult mice. BMC Genomics 19(Suppl 3), 93.CrossRefGoogle ScholarPubMed
183.Farkas, J, Reglodi, D, Gaszner, B et al. (2009) Effects of maternal separation on the neurobehavioral development of newborn Wistar rats. Brain Res Bull 79, 208214.CrossRefGoogle ScholarPubMed
184.Ferreira, CF, Bernardi, JR, Krolow, R et al. (2013) Vulnerability to dietary n-3 polyunsaturated fatty acid deficiency after exposure to early stress in rats. Pharmacol Biochem Behav 107, 1119.CrossRefGoogle ScholarPubMed
185.Hill, RA, Klug, M, Kiss Von Soly, S et al. (2014) Sex-specific disruptions in spatial memory and anhedonia in a ‘two hit’ rat model correspond with alterations in hippocampal brain-derived neurotrophic factor expression and signaling: sex-specific effects of stress on BDNF, cognition, and anhedonia. Hippocampus 24, 11971211.CrossRefGoogle Scholar
186.Klug, M & van den Buuse, M (2012) Chronic cannabinoid treatment during young adulthood induces sex-specific behavioural deficits in maternally separated rats. Behav Brain Res 233, 305313.CrossRefGoogle ScholarPubMed
187.Mourlon, V, Naudon, L, Giros, B et al. (2011) Early stress leads to effects on estrous cycle and differential responses to stress. Physiol Behav 102, 304310.CrossRefGoogle ScholarPubMed
188.Park, H, Yoo, D, Kwon, S et al. (2012) Acupuncture stimulation at HT7 alleviates depression-induced behavioral changes via regulation of the serotonin system in the prefrontal cortex of maternally-separated rat pups. J Physiol Sci 62, 351357.CrossRefGoogle ScholarPubMed
189.Rüedi-Bettschen, D, Pedersen, EM, Feldon, J et al. (2005) Early deprivation under specific conditions leads to reduced interest in reward in adulthood in Wistar rats. Behav Brain Res 156, 297310.CrossRefGoogle ScholarPubMed
190.Shalev, U & Kafkafi, N (2002) Repeated maternal separation does not alter sucrose-reinforced and open-field behaviors. Pharmacol Biochem Behav 73, 115122.CrossRefGoogle Scholar
191.Zhang, L, Hernández, VS, Liu, B et al. (2012) Hypothalamic vasopressin system regulation by maternal separation: Its impact on anxiety in rats. Neuroscience 215, 135148.CrossRefGoogle ScholarPubMed
192.Zimmerberg, B & Kajunski, EW (2004) Sexually dimorphic effects of postnatal allopregnanolone on the development of anxiety behavior after early deprivation. Pharmacol Biochem Behav 78, 465471.CrossRefGoogle ScholarPubMed
193.Pryce, CR, Bettschen, D, Nanz-Bahr, NI et al. (2003) Comparison of the effects of early handling and early deprivation on conditioned stimulus, context, and spatial learning and memory in adult rats. Behav Neurosci 117, 883893.CrossRefGoogle ScholarPubMed
194.Kosten, TA, Lee, HJ & Kim, JJ (2006) Early life stress impairs fear conditioning in adult male and female rats. Brain Res 1087, 142150.CrossRefGoogle ScholarPubMed
195.Lehmann, J, Pryce, CR, Bettschen, D et al. (1999) The maternal separation paradigm and adult emotionality and cognition in male and female Wistar rats. Pharmacol Biochem Behav 64, 705715.CrossRefGoogle ScholarPubMed
196.Stevenson, CW, Spicer, CH, Mason, R et al. (2009) Early life programming of fear conditioning and extinction in adult male rats. Behav Brain Res 205, 505510.CrossRefGoogle ScholarPubMed
197.Zhu, Y, Wang, Y, Yao, R et al. (2017) Enhanced neuroinflammation mediated by DNA methylation of the glucocorticoid receptor triggers cognitive dysfunction after sevoflurane anesthesia in adult rats subjected to maternal separation during the neonatal period. J Neuroinflammation 14, 6.CrossRefGoogle ScholarPubMed
198.Chocyk, A, Majcher-Maślanka, I & Przyborowska, A (2015) Early-life stress increases the survival of midbrain neurons during postnatal development and enhances reward-related and anxiolytic-like behaviors in a sex-dependent fashion. Int J Dev Neurosci 44, 3347.CrossRefGoogle Scholar
199.Eklund, MB & Arborelius, L (2006) Twice daily long maternal separations in Wistar rats decreases anxiety-like behaviour in females but does not affect males. Behav Brain Res 172, 278285.CrossRefGoogle Scholar
200.León Rodríguez, DA & Dueñas, Z (2013) Maternal separation during breastfeeding induces gender-dependent changes in anxiety and the GABA-A receptor α-subunit in adult Wistar rats. PLoS ONE, 8, e68010.CrossRefGoogle ScholarPubMed
201.Michaels, CC & Holtzman, SG (2007) Enhanced sensitivity to naltrexone-induced drinking suppression of fluid intake and sucrose consumption in maternally separated rats. Pharmacol Biochem Behav 86, 784796.CrossRefGoogle ScholarPubMed
202.Mourlon, V, Baudin, A, Blanc, O et al. (2010) Maternal deprivation induces depressive-like behaviours only in female rats. Behav Brain Res 213(2), 278287.CrossRefGoogle ScholarPubMed
203.Kim, S, Kim, H, Yim, YS et al. (2017) Maternal gut bacteria promote neurodevelopmental abnormalities in mouse offspring. Nature 549, 528532.CrossRefGoogle ScholarPubMed
204.Korosi, A (2009) The pathways from mother's love to baby's future. Front Behav Neurosci 3, 27.CrossRefGoogle ScholarPubMed
205.Branchi, I, Santucci, D & Alleva, E (2001) Ultrasonic vocalisation emitted by infant rodents: a tool for assessment of neurobehavioural development. Behav Brain Res 125, 4956.CrossRefGoogle ScholarPubMed
206.Hofer, MA, Shair, HN & Brunelli, SA (2002) Ultrasonic vocalizations in rat and mouse pups. Curr Protoc Neurosci 17, 8.14.18.14.16.Google Scholar
207.Brunelli, SA, Curley, JP, Gudsnuk, K et al. (2015) Variations in maternal behavior in rats selected for infant ultrasonic vocalization in isolation. Horm Behav 75, 7883.CrossRefGoogle Scholar
208.D'Amato, FR, Scalera, E, Sarli, C et al. (2005) Pups call, mothers rush: does maternal responsiveness affect the amount of ultrasonic vocalizations in mouse pups? Behav Genet 35, 103112.CrossRefGoogle ScholarPubMed
209.Aguggia, JP, Suárez, MM & Rivarola, MA (2013) Early maternal separation: neurobehavioral consequences in mother rats. Behav Brain Res 248, 2531.CrossRefGoogle ScholarPubMed
210.Boccia, ML, Razzoli, M, Prasad Vadlamudi, S et al. (2007) Repeated long separations from pups produce depression-like behavior in rat mothers. Psychoneuroendocrinology 32, 6571.CrossRefGoogle ScholarPubMed
211.Maniam, J & Morris, MJ (2010) Long-term postpartum anxiety and depression-like behavior in mother rats subjected to maternal separation are ameliorated by palatable high fat diet. Behav Brain Res 208, 7279.CrossRefGoogle ScholarPubMed
212.Huot, RL, Gonzalez, ME, Ladd, CO et al. (2004) Foster litters prevent hypothalamic–pituitary–adrenal axis sensitization mediated by neonatal maternal separation. Psychoneuroendocrinology 29, 279289.CrossRefGoogle ScholarPubMed
213.Kan, JM, Callaghan, BL & Richardson, R (2016) A mother's past can predict her offspring's future: previous maternal separation leads to the early emergence of adult-like fear behavior in subsequent male infant rat offspring. Behav Neurosci 130, 511520.CrossRefGoogle ScholarPubMed
214.Boersma, GJ, Bale, TL, Casanello, P et al. (2014) Long-term impact of early life events on physiology and behaviour. J Neuroendocrinol 26, 587602.CrossRefGoogle ScholarPubMed
215.Rosenfeld, P, Suchecki, D & Levine, S (1992) Multifactorial regulation of the hypothalamic–pituitary–adrenal axis during development. Neurosci Biobehav Rev 16, 553568.CrossRefGoogle ScholarPubMed
216.Sapolsky, RM & Meaney, MJ (1986) Maturation of the adrenocortical stress response: neuroendocrine control mechanisms and the stress hyporesponsive period. Brain Res 396, 6476.CrossRefGoogle ScholarPubMed
217.Vázquez, DM (1998) Stress and the developing limbic–hypothalamic–pituitary–adrenal axis. Psychoneuroendocrinology 23, 663700.CrossRefGoogle ScholarPubMed
218.Anisman, H, Zaharia, MD, Meaney, MJ et al. (1998) Do early-life events permanently alter behavioral and hormonal responses to stressors? Int J Dev Neurosci 16, 149164.CrossRefGoogle ScholarPubMed
219.Gutman, DA & Nemeroff, CB (2002) Neurobiology of early life stress: rodent studies. Semin Clin Neuropsychiatry 7, 8995.CrossRefGoogle ScholarPubMed
220.Xu, S, Qin, B, Shi, A et al. (2018) Oxytocin inhibited stress induced visceral hypersensitivity, enteric glial cells activation, and release of proinflammatory cytokines in maternal separated rats. Eur J Pharmacol 818, 578584.CrossRefGoogle ScholarPubMed
221.Moussaoui, N, Braniste, V, Ait-Belgnaoui, A et al. (2014) Changes in intestinal glucocorticoid sensitivity in early life shape the risk of epithelial barrier defect in maternal-deprived rats. PLoS ONE 9, e88382.CrossRefGoogle ScholarPubMed
222.Roque, A, Ochoa-Zarzosa, A & Torner, L (2016) Maternal separation activates microglial cells and induces an inflammatory response in the hippocampus of male rat pups, independently of hypothalamic and peripheral cytokine levels. Brain Behav Immun 55, 3948.CrossRefGoogle ScholarPubMed
223.Roque, S, Mesquita, AR, Palha, JA et al. (2014) The behavioral and immunological impact of maternal separation: a matter of timing. Front Behav Neurosci 8, 192.CrossRefGoogle ScholarPubMed
224.Musholt, K, Cirillo, G, Cavaliere, C et al. (2009) Neonatal separation stress reduces glial fibrillary acidic protein- and S100beta-immunoreactive astrocytes in the rat medial precentral cortex. Dev Neurobiol 69, 203211.CrossRefGoogle ScholarPubMed
225.Saavedra, LM, Fenton Navarro, B & Torner, L (2017) Early life stress activates glial cells in the hippocampus but attenuates cytokine secretion in response to an immune challenge in rat pups. Neuroimmunomodulation 24, 242255.CrossRefGoogle Scholar
226.Baldy, C, Fournier, S, Boisjoly-Villeneuve, S et al. (2018) The influence of sex and neonatal stress on medullary microglia in rat pups. Exp Physiol 103, 11921199.CrossRefGoogle ScholarPubMed
227.Delpech, J-C, Wei, L, Hao, J et al. (2016) Early life stress perturbs the maturation of microglia in the developing hippocampus. Brain Behav Immun 57, 7993.CrossRefGoogle ScholarPubMed
228.Andersen, SL & Teicher, MH (2004) Delayed effects of early stress on hippocampal development. Neuropsychopharmacology 29, 19881993.CrossRefGoogle ScholarPubMed
229.Kuma, H, Miki, T, Matsumoto, Y et al. (2004) Early maternal deprivation induces alterations in brain-derived neurotrophic factor expression in the developing rat hippocampus. Neurosci Lett 372, 6873.CrossRefGoogle ScholarPubMed
230.Zhang, LX, Levine, S, Dent, G et al. (2002) Maternal deprivation increases cell death in the infant rat brain. Brain Res Dev Brain Res 133, 111.CrossRefGoogle ScholarPubMed
231.Cirulli, F, Alleva, E, Antonelli, A et al. (2000) NGF expression in the developing rat brain: effects of maternal separation. Brain Res Dev Brain Res 123, 129134.CrossRefGoogle ScholarPubMed
232.Roceri, M, Cirulli, F, Pessina, C et al. (2004) Postnatal repeated maternal deprivation produces age-dependent changes of brain-derived neurotrophic factor expression in selected rat brain regions. Biol Psychiatry 55, 708714.CrossRefGoogle ScholarPubMed
233.Ohta, K, Miki, T, Warita, K et al. (2014) Prolonged maternal separation disturbs the serotonergic system during early brain development. Int J Dev Neurosci 33, 1521.CrossRefGoogle ScholarPubMed
234.Peña, CJ, Kronman, HG, Walker, DM et al. (2017) Early life stress confers lifelong stress susceptibility in mice via ventral tegmental area OTX2. Science 356, 11851188.CrossRefGoogle ScholarPubMed
235.Bohacek, J & Mansuy, IM (2013) Epigenetic inheritance of disease and disease risk. Neuropsychopharmacology 38, 220236.CrossRefGoogle ScholarPubMed
236.Heim, C & Binder, EB (2012) Current research trends in early life stress and depression: review of human studies on sensitive periods, gene–environment interactions, and epigenetics. Exp Neurol 233, 102111.CrossRefGoogle ScholarPubMed
237.Lutz, PE & Turecki, G (2014) DNA methylation and childhood maltreatment: from animal models to human studies. Neuroscience 264, 142156.CrossRefGoogle ScholarPubMed
238.Provençal, N & Binder, EB (2015) The effects of early life stress on the epigenome: from the womb to adulthood and even before. Exp Neurol 268, 1020.CrossRefGoogle Scholar
239.Silberman, DM, Acosta, GB & Zorrilla Zubilete, MA (2016) Long-term effects of early life stress exposure: Role of epigenetic mechanisms. Pharmacol Res 109, 6473.CrossRefGoogle ScholarPubMed
240.Jawahar, MC, Murgatroyd, C, Harrison, EL et al. (2015) Epigenetic alterations following early postnatal stress: a review on novel aetiological mechanisms of common psychiatric disorders. Clin Epigenetics 7, 122.CrossRefGoogle ScholarPubMed
241.Meaney, MJ & Szyf, M (2005) Environmental programming of stress responses through DNA methylation: life at the interface between a dynamic environment and a fixed genome. Dialogues Clin Neurosci 7, 103123.Google Scholar
242.Murgatroyd, C, Patchev, AV, Wu, Y et al. (2009) Dynamic DNA methylation programs persistent adverse effects of early-life stress. Nat Neurosci 12, 15591566.CrossRefGoogle ScholarPubMed
243.Roth, TL & Sweatt, JD (2011) Epigenetic marking of the BDNF gene by early-life adverse experiences. Horm Behav 59, 315320.CrossRefGoogle ScholarPubMed
244.Roth, TL, Lubin, FD, Funk, AJ et al. (2009) Lasting epigenetic influence of early-life adversity on the BDNF gene. Biol Psychiatry 65, 760769.CrossRefGoogle ScholarPubMed
245.Wang, X, Cattaneo, F, Ryno, L et al. (2014) The systemic amyloid precursor transthyretin (TTR) behaves as a neuronal stress protein regulated by HSF1 in SH-SY5Y human neuroblastoma cells and APP23 Alzheimer's disease model mice. J Neurosci 34, 72537265.CrossRefGoogle ScholarPubMed
246.Weaver, ICG, Cervoni, N, Champagne, FA et al. (2004) Epigenetic programming by maternal behavior. Nat Neurosci 7, 847854.CrossRefGoogle ScholarPubMed
247.Weiss, IC, Franklin, TB, Vizi, S et al. (2011) Inheritable effect of unpredictable maternal separation on behavioral responses in mice. Front Behav Neurosci 5, 3.CrossRefGoogle ScholarPubMed
248.Pusalkar, M, Suri, D, Kelkar, A et al. (2016) Early stress evokes dysregulation of histone modifiers in the medial prefrontal cortex across the life span. Dev Psychobiol 58, 198210.CrossRefGoogle ScholarPubMed
249.Park, SW, Lee, JG, Seo, MK et al. (2017) Epigenetic modification of glucocorticoid receptor promoter I7 in maternally separated and restraint-stressed rats. Neurosci Lett 650, 3844.CrossRefGoogle ScholarPubMed
250.Seo, MK, Ly, NN, Lee, CH et al. (2016) Early life stress increases stress vulnerability through BDNF gene epigenetic changes in the rat hippocampus. Neuropharmacology 105, 388397.CrossRefGoogle ScholarPubMed
251.Barreau, F, Ferrier, L, Fioramonti, J et al. (2007) New insights in the etiology and pathophysiology of irritable bowel syndrome: contribution of neonatal stress models. Pediatr Res 62, 240245.CrossRefGoogle ScholarPubMed
252.Moloney, RD, Johnson, AC, O'Mahony, SM et al. (2015) Stress and the microbiota–gut–brain axis in visceral pain: relevance to irritable bowel syndrome. CNS Neurosci Ther 22, 102117.CrossRefGoogle ScholarPubMed
253.Barreau, F, Salvador-Cartier, C, Houdeau, E et al. (2008) Long-term alterations of colonic nerve-mast cell interactions induced by neonatal maternal deprivation in rats. Gut 57, 582590.CrossRefGoogle ScholarPubMed
254.Tominaga, K, Fujikawa, Y, Tanaka, F et al. (2016) Structural changes in gastric glial cells and delayed gastric emptying as responses to early life stress and acute adulthood stress in rats. Life Sci 148, 254259.CrossRefGoogle ScholarPubMed
255.Million, M & Larauche, M (2016) Stress, sex, and the enteric nervous system. Neurogastroenterol Motil 28, 12831289.CrossRefGoogle ScholarPubMed
256.Hyland, NP, O'Mahony, SM, O'Malley, D et al. (2015) Early-life stress selectively affects gastrointestinal but not behavioral responses in a genetic model of brain–gut axis dysfunction. Neurogastroenterol Motil 27, 105113.CrossRefGoogle ScholarPubMed
257.Moloney, RD, Stilling, RM, Dinan, TG et al. (2015) Early-life stress-induced visceral hypersensitivity and anxiety behavior is reversed by histone deacetylase inhibition. Neurogastroenterol Motil 27, 18311836.CrossRefGoogle ScholarPubMed
258.Murakami, T, Kamada, K, Mizushima, K et al. (2017) Changes in intestinal motility and gut microbiota composition in a rat stress model. Digestion 95, 5560.CrossRefGoogle Scholar
259.Schwetz, I, McRoberts, JA, Coutinho, SV et al. (2005) Corticotropin-releasing factor receptor 1 mediates acute and delayed stress-induced visceral hyperalgesia in maternally separated long-Evans rats. Am J Physiol Gastrointest Liver Physiol 289, G704G712.CrossRefGoogle ScholarPubMed
260.Yi, L, Zhang, H, Sun, H et al. (2017) Maternal separation induced visceral hypersensitivity from childhood to adulthood. Neurogastroenterol Motil 23, 306315.CrossRefGoogle ScholarPubMed
261.Asano, T, Tanaka, KI, Tada, A et al. (2017) Aminophylline suppresses stress-induced visceral hypersensitivity and defecation in irritable bowel syndrome. Sci Rep 7, 40214.CrossRefGoogle ScholarPubMed
262.Barreau, F, Ferrier, L, Fioramonti, J et al. (2004) Neonatal maternal deprivation triggers long term alterations in colonic epithelial barrier and mucosal immunity in rats. Gut 53, 501506.CrossRefGoogle ScholarPubMed
263.Barrett, E, Fitzgerald, P, Dinan, TG et al. (2012) Bifidobacterium breve with α-Linolenic acid and linoleic acid alters fatty acid metabolism in the maternal separation model of irritable bowel syndrome. PLoS ONE 7, e48159.CrossRefGoogle ScholarPubMed
264.Bian, ZX, Zhang, M, Han, QB et al. (2010) Analgesic effects of JCM-16021 on neonatal maternal separation-induced visceral pain in rats. World J Gastroenterol 16, 837845.Google ScholarPubMed
265.Bian, ZX, Qin, HY, Tian, SL et al. (2011) Combined effect of early life stress and acute stress on colonic sensory and motor responses through serotonin pathways: differences between proximal and distal colon in rats. Stress 14, 448458.CrossRefGoogle ScholarPubMed
266.Chung, EKY, Zhang, X, Li, Z et al. (2007) Neonatal maternal separation enhances central sensitivity to noxious colorectal distention in rat. Brain Res 1153, 6877.CrossRefGoogle ScholarPubMed
267.Coutinho, SV, Plotsky, PM, Sablad, M et al. (2002) Neonatal maternal separation alters stress-induced responses to viscerosomatic nociceptive stimuli in rat. Am J Physiol Gastrointest Liver Physiol 282, G307G316.CrossRefGoogle ScholarPubMed
268.Distrutti, E, Cipriani, S, Mencarelli, A et al. (2013) Probiotics VSL#3 protect against development of visceral pain in murine model of irritable bowel syndrome. PLoS ONE 8, e63893.CrossRefGoogle ScholarPubMed
269.Gosselin, RD, O'Connor, RM, Tramullas, M et al. (2010) Riluzole normalizes early-life stress-induced visceral hypersensitivity in rats: role of spinal glutamate reuptake mechanisms. Gastroenterology 138, 24182425.CrossRefGoogle ScholarPubMed
270.Hu, XG, Xu, D, Zhao, Y et al. (2009) The alleviating pain effect of aqueous extract From Tong-Xie-Yao-Fang, on experimental visceral hypersensitivity and its mechanism. Biol Pharm Bull 32, 10751079.CrossRefGoogle ScholarPubMed
271.Hyland, NP, Julio-Pieper, M, O'Mahony, SM et al. (2009) A distinct subset of submucosal mast cells undergoes hyperplasia following neonatal maternal separation: a role in visceral hypersensitivity? Gut 58, 10291030; author reply 1030–1031.CrossRefGoogle ScholarPubMed
272.Miquel, S, Martín, R, Lashermes, A et al. (2016) Anti-nociceptive effect of Faecalibacterium prausnitzii in non-inflammatory IBS-like models. Sci Rep 6, 19399.CrossRefGoogle ScholarPubMed
273.Moloney, RD, Sajjad, J, Foley, T et al. (2016) Estrous cycle influences excitatory amino acid transport and visceral pain sensitivity in the rat: effects of early-life stress. Biol Sex Differ 7, 33.CrossRefGoogle ScholarPubMed
274.Prusator, DK & Greenwood-Van Meerveld, B (2016) Sex-related differences in pain behaviors following three early life stress paradigms. Biol Sex Differ 7, 29.CrossRefGoogle ScholarPubMed
275.Ren, TH, Wu, J, Yew, D et al. (2006) Effects of neonatal maternal separation on neurochemical and sensory response to colonic distension in a rat model of irritable bowel syndrome. Am J Physiol Gastrointest Liver Physiol 292, G849G856.CrossRefGoogle Scholar
276.Riba, A, Olier, M, Lacroix-Lamandé, S et al. (2018) Early life stress in mice is a suitable model for Irritable Bowel Syndrome but does not predispose to colitis nor increase susceptibility to enteric infections. Brain Behav Immun 73, 403415.CrossRefGoogle Scholar
277.Rosztóczy, A, Fioramonti, J, Jármay, K et al. (2003) Influence of sex and experimental protocol on the effect of maternal deprivation on rectal sensitivity to distension in the adult rat. Neurogastroenterol Motil 15, 679686.CrossRefGoogle ScholarPubMed
278.Shao, L, Liu, Y, Xiao, J et al. (2019) Activating metabotropic glutamate receptor-7 attenuates visceral hypersensitivity in neonatal maternally separated rats. Int J Mol Med 43, 761770.Google ScholarPubMed
279.Tjong, YW, Ip, SP, Lao, L et al. (2010) Neonatal maternal separation elevates thalamic corticotrophin releasing factor type 1 receptor expression response to colonic distension in rat. Neuro Endocrinol Lett 31, 215220.Google ScholarPubMed
280.Tjong, YW, Ip, SP, Lao, L et al. (2011) Role of neuronal nitric oxide synthase in colonic distension-induced hyperalgesia in distal colon of neonatal maternal separated male rats. Neurogastroenterol Motil 23, 666–e278.CrossRefGoogle ScholarPubMed
281.Tsang, SW, Zhao, M, Wu, J et al. (2012) Nerve growth factor-mediated neuronal plasticity in spinal cord contributes to neonatal maternal separation-induced visceral hypersensitivity in rats: nerve growth factor-mediated neuronal plasticity in spinal cord contributes to hypersensitivity in rats. Eur J Pain 16, 463472.CrossRefGoogle ScholarPubMed
282.Yang, JM, Xian, YF, Ip, PSP et al. (2012) Schisandra chinensis reverses visceral hypersensitivity in a neonatal-maternal separated rat model. Phytomedicine 19, 402408.CrossRefGoogle Scholar
283.Zhang, XJ, Li, Z, Leung, WM et al. (2008) The analgesic effect of paeoniflorin on neonatal maternal separation-induced visceral hyperalgesia in rats. J Pain 9(6), 497505.CrossRefGoogle ScholarPubMed
284.Riba, A, Olier, M, Lacroix-Lamandé, S et al. (2017) Paneth cell defects induce microbiota dysbiosis in mice and promote visceral hypersensitivity. Gastroenterology 153, 15941606.CrossRefGoogle ScholarPubMed
285.Barreau, F, Cartier, C, Leveque, M et al. (2007) Pathways involved in gut mucosal barrier dysfunction induced in adult rats by maternal deprivation: corticotrophin-releasing factor and nerve growth factor interplay. J Physiol 580, 347356.CrossRefGoogle ScholarPubMed
286.Li, B, Lee, C, Filler, T et al. (2017) Inhibition of corticotropin-releasing hormone receptor 1 and activation of receptor 2 protect against colonic injury and promote epithelium repair. Sci Rep 7, 46616.CrossRefGoogle ScholarPubMed
287.Million, M, Wang, L, Wang, Y et al. (2006) CRF2 receptor activation prevents colorectal distension induced visceral pain and spinal ERK1/2 phosphorylation in rats. Gut 55, 172181.CrossRefGoogle ScholarPubMed
288.Myers, B & Greenwood-Van Meerveld, B (2012) Differential involvement of amygdala corticosteroid receptors in visceral hyperalgesia following acute or repeated stress. Am J Physiol Gastrointest Liver Physiol 302, G260G266.CrossRefGoogle ScholarPubMed
289.Prusator, DK & Greenwood-Van Meerveld, B (2017) Amygdala-mediated mechanisms regulate visceral hypersensitivity in adult females following early life stress: importance of the glucocorticoid receptor and corticotropin-releasing factor. Pain 158, 296305.CrossRefGoogle ScholarPubMed
290.Zhou, XP, Sha, J, Huang, L et al. (2016) Nesfatin-1/NUCB2 in the amygdala influences visceral sensitivity via glucocorticoid and mineralocorticoid receptors in male maternal separation rats. Neurogastroenterol Motil 28, 15451553.CrossRefGoogle ScholarPubMed
291.Van den Wijngaard, RM, Stanisor, OI, van Diest, SA et al. (2013) Susceptibility to stress induced visceral hypersensitivity in maternally separated rats is transferred across generations. Neurogastroenterol Motil 25, e780e790.CrossRefGoogle Scholar
292.Bailey, MT & Coe, CL (1999) Maternal separation disrupts the integrity of the intestinal microflora in infant rhesus monkeys. Dev Psychobiol 35, 146155.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
293.O'Mahony, SM, Marchesi, JR, Scully, P et al. (2009) Early life stress alters behavior, immunity, and microbiota in rats: implications for irritable bowel syndrome and psychiatric illnesses. Biol Psychiatry 65, 263267.CrossRefGoogle ScholarPubMed
294.Qian, L, Lu, L, Huang, L et al. (2019) The effect of neonatal maternal separation on short-chain fatty acids and airway inflammation in adult asthma mice. Allergol Immunopathol 47, 211.CrossRefGoogle ScholarPubMed
295.Zhou, XY, Li, M, Li, X et al. (2016) Visceral hypersensitive rats share common dysbiosis features with irritable bowel syndrome patients. World J Gastroenterol 22, 52115227.CrossRefGoogle ScholarPubMed
296.El Aidy, S, Ramsteijn, AS, Dini-Andreote, F et al. (2017) Serotonin transporter genotype modulates the gut microbiota composition in young rats, an effect augmented by early life stress. Front Cell Neurosci 11, 222.CrossRefGoogle ScholarPubMed
297.Pusceddu, MM, El Aidy, S, Crispie, F et al. (2015) N-3 Polyunsaturated fatty acids (PUFAs) reverse the impact of early-life stress on the gut microbiota. PLoS ONE 10, e0139721.CrossRefGoogle ScholarPubMed
298.Barouei, J, Moussavi, M & Hodgson, DM (2012) Effect of maternal probiotic intervention on HPA axis, immunity and gut microbiota in a rat model of irritable bowel syndrome. PLoS ONE 7, e46051.CrossRefGoogle Scholar
299.García-Ródenas, CL, Bergonzelli, GE, Nutten, S et al. (2006) Nutritional approach to restore impaired intestinal barrier function and growth after neonatal stress in rats. J Pediatr Gastroenterol Nutr 43, 1624.CrossRefGoogle ScholarPubMed
300.Ilchmann-Diounou, H, Olier, M, Lencina, C et al. (2019). Early life stress induces type 2 diabetes-like features in ageing mice. Brain Behav Immun S0889-1591(18)30787-6.CrossRefGoogle ScholarPubMed
301.Estienne, M, Claustre, J, Clain-Gardechaux, G et al. (2010) Maternal deprivation alters epithelial secretory cell lineages in rat duodenum: role of CRF-related peptides. Gut 59, 744751.CrossRefGoogle ScholarPubMed
302.Li, B, Zani, A, Lee, C et al. (2016) Endoplasmic reticulum stress is involved in the colonic epithelium damage induced by maternal separation. J Pediatr Surg 51, 10011004.CrossRefGoogle ScholarPubMed
303.Li, B, Lee, C, Martin, Z et al. (2017) Intestinal epithelial injury induced by maternal separation is protected by hydrogen sulfide. J Pediatr Surg 52, 4044.CrossRefGoogle ScholarPubMed
304.O'Malley, D, Julio-Pieper, M, Gibney, SM et al. (2010) Distinct alterations in colonic morphology and physiology in two rat models of enhanced stress-induced anxiety and depression-like behaviour. Stress 13, 114122.CrossRefGoogle ScholarPubMed
305.Ghia, JE, Blennerhassett, P & Collins, SM (2008) Impaired parasympathetic function increases susceptibility to inflammatory bowel disease in a mouse model of depression. J Clin Invest 118, 22092218.Google Scholar
306.Varghese, AK, Verdú, EF, Bercik, P et al. (2006) Antidepressants attenuate increased susceptibility to colitis in a murine model of depression. Gastroenterology 130, 17431753.CrossRefGoogle Scholar
307.Veenema, AH, Reber, SO, Selch, S et al. (2008) Early life stress enhances the vulnerability to chronic psychosocial stress and experimental colitis in adult mice. Endocrinology 149, 27272736.CrossRefGoogle ScholarPubMed
308.Barreau, F, Cartier, C, Ferrier, L et al. (2004) Nerve growth factor mediates alterations of colonic sensitivity and mucosal barrier induced by neonatal stress in rats. Gastroenterology 127, 524534.CrossRefGoogle ScholarPubMed
309.Gareau, MG, Jury, J & Perdue, MH (2007) Neonatal maternal separation of rat pups results in abnormal cholinergic regulation of epithelial permeability. Am J Physiol Gastrointest Liver Physiol 293, G198G203.CrossRefGoogle ScholarPubMed
310.Gareau, MG, Jury, J, MacQueen, G et al. (2007) Probiotic treatment of rat pups normalises corticosterone release and ameliorates colonic dysfunction induced by maternal separation. Gut 56, 15221528.CrossRefGoogle ScholarPubMed
311.Söderholm, JD, Yates, DA, Gareau, MG et al. (2002) Neonatal maternal separation predisposes adult rats to colonic barrier dysfunction in response to mild stress. Am J Physiol Gastrointest Liver Physiol 283, G1257G1263.CrossRefGoogle ScholarPubMed
312.Gareau, MG, Jury, J, Yang, PC et al. (2006) Neonatal maternal separation causes colonic dysfunction in rat pups including impaired host resistance. Pediatr Res 59, 8388.CrossRefGoogle ScholarPubMed
313.Barouei, J, Moussavi, M & Hodgson, DM (2015) Perinatal maternal probiotic intervention impacts immune responses and ileal mucin gene expression in a rat model of irritable bowel syndrome. Benef Microbes 6, 8395.CrossRefGoogle Scholar
314.McKernan, DP, Nolan, A, Brint, EK et al. (2009) Toll-like receptor mRNA expression is selectively increased in the colonic mucosa of two animal models relevant to irritable bowel syndrome. PLoS ONE 4, e8226.CrossRefGoogle ScholarPubMed
315.Naninck, EFG, Oosterink, JE, Yam, KY et al. (2017). Early micronutrient supplementation protects against early stress-induced cognitive impairments. FASEB J 31, 505518.CrossRefGoogle ScholarPubMed
316.Mathieu, G, Denis, S, Lavialle, M et al. (2008). Synergistic effects of stress and omega-3 fatty acid deprivation on emotional response and brain lipid composition in adult rats. Prostaglandins Leukot Essent Fatty Acids 78, 391401.CrossRefGoogle ScholarPubMed
317.Mathieu, G, Oualian, C, Denis, I et al. (2011). Dietary n-3 polyunsaturated fatty acid deprivation together with early maternal separation increases anxiety and vulnerability to stress in adult rats. Prostaglandins Leukot Essent Fatty Acids 85, 129136.CrossRefGoogle ScholarPubMed
318.Réus, GZ, Maciel, AL, Abelaira, HM et al. (2018). ω-3 and folic acid act against depressive-like behavior and oxidative damage in the brain of rats subjected to early- or late-life stress. Nutrition 53, 120133.CrossRefGoogle ScholarPubMed
319.Pusceddu, MM, Kelly, P, Ariffin, N et al. (2015b). n-3 PUFAs have beneficial effects on anxiety and cognition in female rats: effects of early life stress. Psychoneuroendocrinology 58, 7990.CrossRefGoogle Scholar
320.van Diest, SA, van den Elsen, LWJ, Klok, AJ et al. (2015). Dietary marine n-3 PUFAs do not affect stress-induced visceral hypersensitivity in a rat maternal separation model. J Nutr, 145, 915922.CrossRefGoogle ScholarPubMed
321.Collins, SM (2014) A role for the gut microbiota in IBS. Nat Rev Gastroenterol Hepatol 11, 497505.CrossRefGoogle ScholarPubMed
322.Mangiola, F, Ianiro, G, Franceschi, F et al. (2016) Gut microbiota in autism and mood disorders. World J Gastroenterol 22, 361368.CrossRefGoogle ScholarPubMed
323.Wang, Y & Kasper, LH (2014) The role of microbiome in central nervous system disorders. Brain Behav Immun 38, 112.CrossRefGoogle ScholarPubMed
324.Zhao, L (2013) The gut microbiota and obesity: from correlation to causality. Nat Rev Microbiol 11, 639647.CrossRefGoogle ScholarPubMed
325.Dong, TS & Gupta, A (2019) Influence of early life, diet, and the environment on the microbiome. Clin Gastroenterol Hepatol 17, 231242.CrossRefGoogle ScholarPubMed
326Luczynski, P, McVey Neufeld, KA, Oriach, CS et al. (2016) Growing up in a bubble: using germ-free animals to assess the influence of the gut microbiota on brain and behavior. Int J Neuropsychopharmacol 19, 8.CrossRefGoogle Scholar
327.Braniste, V, Al-Asmakh, M, Kowal, C et al. (2014) The gut microbiota influences blood–brain barrier permeability in mice. Sci Transl Med 6, 263ra158.CrossRefGoogle ScholarPubMed
328.Clarke, G, Grenham, S, Scully, P et al. (2013) The microbiome–gut–brain axis during early life regulates the hippocampal serotonergic system in a sex-dependent manner. Mol Psychiatry 18, 666673.CrossRefGoogle Scholar
329.Desbonnet, L, Clarkz, G, Shanahan, F et al. (2014) Microbiota is essential for social development in the mouse. Mol Psychiatry 19, 146148.CrossRefGoogle ScholarPubMed
330.Luczynski, P, Tramullas, M, Viola, M et al. (2017). Microbiota regulates visceral pain in the mouse. ELife 6. https://elifesciences.org/articles/25887CrossRefGoogle ScholarPubMed
331.Neufeld, KAM, Kang, N, Bienenstock, J et al. (2011) Effects of intestinal microbiota on anxiety-like behavior. Commun Integr Biol 4, 492494.CrossRefGoogle ScholarPubMed
332.Sudo, N, Chida, Y, Aiba, Y et al. (2004) Postnatal microbial colonization programs the hypothalamic–pituitary–adrenal system for stress response in mice. J Physiol 558, 263275.CrossRefGoogle ScholarPubMed
333.Heijtz, RD, Wang, S, Anuar, F et al. (2011) Normal gut microbiota modulates brain development and behavior. Proc Natl Acad Sci U S A 108, 30473052.CrossRefGoogle Scholar
334.Jašarević, E, Howard, CD, Morrison, K et al. (2018) The maternal vaginal microbiome partially mediates the effects of prenatal stress on offspring gut and hypothalamus. Nat Neurosci 21, 10611071.CrossRefGoogle ScholarPubMed
335.Collins, SM, Kassam, Z & Bercik, P (2013) The adoptive transfer of behavioral phenotype via the intestinal microbiota: experimental evidence and clinical implications. Curr Opin Microbiol 16, 240245.CrossRefGoogle ScholarPubMed
336.Turnbaugh, PJ, Bäckhed, F, Fulton, L et al. (2008) Diet-induced obesity is linked to marked but reversible alterations in the mouse distal gut microbiome. Cell Host Microbe 3, 213223.CrossRefGoogle ScholarPubMed
337.Wang, Z, Koonen, D, Hofker, M et al. (2016) Gut microbiome and lipid metabolism: from associations to mechanisms. Curr Opin Lipidol 27, 216224.CrossRefGoogle ScholarPubMed
338.Buffington, SA, Di Prisco, GV, Auchtung, TA et al. (2016). Microbial reconstitution reverses maternal diet-induced social and synaptic deficits in offspring. Cell 165, 17621775.CrossRefGoogle ScholarPubMed
339.Zhao, L, Huang, Y, Lu, L et al. (2018) Saturated long-chain fatty acid-producing bacteria contribute to enhanced colonic motility in rats. Microbiome 6, 107.CrossRefGoogle ScholarPubMed
340.De Palma, G, Lynch, MDJ, Lu, J et al. (2017) Transplantation of fecal microbiota from patients with irritable bowel syndrome alters gut function and behavior in recipient mice. Sci Transl Med 9. https://stm.sciencemag.org/content/9/379/eaaf6397.shortCrossRefGoogle ScholarPubMed
341.Botschuijver, S, Roeselers, G, Levin, E et al. (2017) Intestinal fungal dysbiosis is associated with visceral hypersensitivity in patients with irritable bowel syndrome and rats. Gastroenterology 153, 10261039.CrossRefGoogle ScholarPubMed
342.Borody, TJ & Khoruts, A (2011) Fecal microbiota transplantation and emerging applications. Nat Rev Gastroenterol Hepatol 9, 8896.CrossRefGoogle ScholarPubMed
343.Brandt, LJ & Aroniadis, OC (2013) An overview of fecal microbiota transplantation: techniques, indications, and outcomes. Gastrointest Endosc 78(2), 240249.CrossRefGoogle ScholarPubMed
344.Khoruts, A (2014) Faecal microbiota transplantation in 2013: developing human gut microbiota as a class of therapeutics. Nat Rev Gastroenterol Hepatol 11, 7980.CrossRefGoogle ScholarPubMed
345.Halkjær, SI, Christensen, AH, Lo, BZS et al. (2018) Faecal microbiota transplantation alters gut microbiota in patients with irritable bowel syndrome: results from a randomised, double-blind placebo-controlled study. Gut 67, 21072115.CrossRefGoogle ScholarPubMed
346.Johnsen, PH, Hilpüsch, F & Cavanagh, JP (2018) Faecal microbiota transplantation versus placebo for moderate-to-severe irritable bowel syndrome: a double-blind, randomised, placebo-controlled, parallel-group, single-centre trial. Lancet Gastroenterol Hepatol 3, 1724.CrossRefGoogle ScholarPubMed
347.Eberl, G (2010) A new vision of immunity: homeostasis of the superorganism. Mucosal Immunol 3, 450460.CrossRefGoogle ScholarPubMed
348.Joseph, JM & Law, C (2018) Cross-species examination of single- and multi-strain probiotic treatment effects on neuropsychiatric outcomes. Neurosci Biobehav Rev 99, 160197.CrossRefGoogle ScholarPubMed
349.Sarkar, A, Lehto, SM, Harty, S et al. (2016) Psychobiotics and the manipulation of bacteria–gut–brain signals. Trends Neurosci 39, 763781.CrossRefGoogle ScholarPubMed
350.Liu, YW, Liu, WH, Wu, CC et al. (2016) Psychotropic effects of Lactobacillus plantarum PS128 in early life-stressed and naïve adult mice. Brain Res 1631, 112.CrossRefGoogle ScholarPubMed
351.Fukui, H, Oshima, T, Tanaka, Y et al. (2018) Effect of probiotic Bifidobacterium bifidum G9-1 on the relationship between gut microbiota profile and stress sensitivity in maternally separated rats. Sci Rep 8, 12384.CrossRefGoogle ScholarPubMed
352.Vanhaecke, T, Aubert, P, Grohard, PA et al. (2017) L. fermentum CECT 5716 prevents stress-induced intestinal barrier dysfunction in newborn rats. Neurogastroenterol Motil 29. https://onlinelibrary.wiley.com/doi/abs/10.1111/nmo.13069CrossRefGoogle ScholarPubMed
353.Cowan, CSM, Stylianakis, AA & Richardson, R (2019). Early-life stress, microbiota, and brain development: probiotics reverse the effects of maternal separation on neural circuits underpinning fear expression and extinction in infant rats. Dev Cogn Neurosci 37, 100627.CrossRefGoogle ScholarPubMed
354.Cowan, CSM & Richardson, R (2018) Early-life stress leads to sex-dependent changes in pubertal timing in rats that are reversed by a probiotic formulation. Dev Psychobiol [Epublication ahead of print version].Google ScholarPubMed
355.Callaghan, BL, Cowan, CSM & Richardson, R (2016) Treating generational stress: effect of paternal stress on development of memory and extinction in offspring is reversed by probiotic treatment. Psychol Sci 27, 11711180.CrossRefGoogle ScholarPubMed
356.Steel, Z, Marnane, C, Iranpour, C et al. (2014) The global prevalence of common mental disorders: a systematic review and meta-analysis 1980–2013. Int J Epidemiol 43, 476493.CrossRefGoogle ScholarPubMed
357.Vlainić, JV, Šuran, J, Vlainić, T et al. (2016) Probiotics as an adjuvant therapy in major depressive disorder. Curr Neuropharmacol 14, 952958.CrossRefGoogle ScholarPubMed
358.Slykerman, RF, Hood, F, Wickens, K et al. (2017) Effect of Lactobacillus rhamnosus HN001 in pregnancy on postpartum symptoms of depression and anxiety: a randomised double-blind placebo-controlled trial. EBioMedicine 24, 159165.CrossRefGoogle ScholarPubMed
359.Majeed, M, Nagabhushanam, K, Arumugam, S et al. (2018) Bacillus coagulans MTCC 5856 for the management of major depression with irritable bowel syndrome: a randomised, double-blind, placebo controlled, multi-centre, pilot clinical study. Food Nutr Res 62. https://foodandnutritionresearch.net/index.php/fnr/article/view/1218CrossRefGoogle ScholarPubMed
360.Romijn, AR, Rucklidge, JJ, Kuijer, RG et al. (2017) A double-blind, randomized, placebo-controlled trial of Lactobacillus helveticus and Bifidobacterium longum for the symptoms of depression. Aust N Z J Psychiatry 51, 810821.CrossRefGoogle ScholarPubMed
361.Akkasheh, G, Kashani-Poor, Z, Tajabadi-Ebrahimi, M et al. (2016) Clinical and metabolic response to probiotic administration in patients with major depressive disorder: a randomized, double-blind, placebo-controlled trial. Nutrition 32, 315320.CrossRefGoogle ScholarPubMed
362.Rackers, HS, Thomas, S, Williamson, K et al. (2018) Emerging literature in the microbiota–brain axis and perinatal mood and anxiety disorders. Psychoneuroendocrinology 95, 8696.CrossRefGoogle ScholarPubMed
363.Slykerman, RF, Kang, J, Van Zyl, N et al. (2018) Effect of early probiotic supplementation on childhood cognition, behaviour and mood a randomised, placebo-controlled trial. Acta Paediatr 107, 21722178.CrossRefGoogle Scholar
364.Gibson, GR & Roberfroid, MB (1995) Dietary modulation of the human colonic microbiota: introducing the concept of prebiotics. J Nutr 125, 14011412.CrossRefGoogle ScholarPubMed
365.Kao, ACC, Harty, S & Burnet, PWJ (2016) The influence of prebiotics on neurobiology and behavior. Int Rev Neurobiol 131, 2148.CrossRefGoogle ScholarPubMed
366.Burokas, A, Arboleya, S, Moloney, RD et al. (2017) Targeting the microbiota–gut–brain axis: prebiotics have anxiolytic and antidepressant-like effects and reverse the impact of chronic stress in mice. Biol Psychiatry 82, 472487.CrossRefGoogle ScholarPubMed
367.Wang, B & Brand-Miller, J (2003) The role and potential of sialic acid in human nutrition. Eur J Clin Nutr 57, 13511369.CrossRefGoogle ScholarPubMed
368.Wang, B, McVeagh, P, Petocz, P et al. (2003) Brain ganglioside and glycoprotein sialic acid in breastfed compared with formula-fed infants. Am J Clin Nutr 78, 10241029.CrossRefGoogle ScholarPubMed
369.Tarr, AJ, Galley, JD, Fisher, SE et al. (2015) The prebiotics 3′-sialyllactose and 6′-sialyllactose diminish stressor-induced anxiety-like behavior and colonic microbiota alterations: evidence for effects on the gut–brain axis. Brain Behav Immun 50, 166177.CrossRefGoogle ScholarPubMed
370.Jia, S, Lu, Z, Gao, Z et al. (2016) Chitosan oligosaccharides alleviate cognitive deficits in an amyloid-β1-42-induced rat model of Alzheimer's disease. Int J Biol Macromol 83, 416425.CrossRefGoogle Scholar
371.Oliveros, E, Ramirez, M, Vazquez, E et al. (2016) Oral supplementation of 2′-fucosyllactose during lactation improves memory and learning in rats. J Nutr Biochem 31, 2027.CrossRefGoogle ScholarPubMed
372.Vázquez, E, Barranco, A, Ramírez, M et al. (2015) Effects of a human milk oligosaccharide, 2′-fucosyllactose, on hippocampal long-term potentiation and learning capabilities in rodents. J Nutr Biochem 26, 455465.CrossRefGoogle ScholarPubMed
373.Yen, CH, Wang, CH, Wu, WT et al. (2017) Fructo-oligosaccharide improved brain β-amyloid, β-secretase, cognitive function, and plasma antioxidant levels in D-galactose-treated Balb/cJ mice. Nutr Neurosci 20, 228237.CrossRefGoogle ScholarPubMed
374.McVey Neufeld, KA, O'Mahony, SM, Hoban, AE et al. (2017) Neurobehavioural effects of Lactobacillus rhamnosus GG alone and in combination with prebiotics polydextrose and galactooligosaccharide in male rats exposed to early-life stress. Nutr Neurosci 22, 425434CrossRefGoogle ScholarPubMed
375.Papalini, S, Michels, F, Kohn, N et al. (2019). Stress matters: randomized controlled trial on the effect of probiotics on neurocognition. Neurobiol Stress 10, 100141.CrossRefGoogle ScholarPubMed
376.Mayer, EA & Hsiao, EY (2017) The gut and its microbiome as related to central nervous system functioning and psychological well-being: introduction to the special issue of psychosomatic medicine. Psychosom Med 79, 844846.CrossRefGoogle ScholarPubMed
377.Gavrieli, A, Farr, OM, Davis, CR et al. (2015). Early life adversity and/or posttraumatic stress disorder severity are associated with poor diet quality, including consumption of trans fatty acids, and fewer hours of resting or sleeping in a US middle-aged population: a cross-sectional and prospective study. Metabolism 64, 15971610.CrossRefGoogle ScholarPubMed