Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 May 15.
Published in final edited form as: Brain Res. 2020 Feb 28;1735:146742. doi: 10.1016/j.brainres.2020.146742

Dynorphin and its role in alcohol use disorder

Anushree N Karkhanis 1, Ream Al-Hasani 2
PMCID: PMC7111141  NIHMSID: NIHMS1573415  PMID: 32114059

Abstract

The dynorphin / kappa opioid receptor (KOR) system has been implicated in many aspects that influence neuropsychiatric disorders. Namely, this system modulates neural circuits that primarily regulate reward seeking, motivation processing, stress responsivity, and pain sensitivity, thus affecting the development of substance and alcohol use disorder (AUD). The effects of this system are often bidirectional and depend on projection targets. To date, a majority of the studies focusing on this system have examined the KOR function using agonists and antagonists. Indeed, there are studies that have examined prodynorphin and dynorphin levels by measuring mRNA and tissue content levels; however, static levels of the neuropeptide and its precursor do not explain complete and online function of the peptide as would be explained by measuring dynorphin transmission in real time. New and exciting methods using optogenetics, chemogenetics, genetic sensors, fast scan cyclic voltammetry are now being developed to detect various neuropeptides with a focus on opioid peptides, including dynorphin. In this review we discuss studies that examine dynorphin projections in areas involved in AUD, its functional involvement in AUD and vulnerability to develop AUD at various ages. Moreover, we discuss dynorphin’s role in promoting AUD by dysregulation motivation circuits and how advancements in opioid peptide detection will further our understanding.

1. Introduction to dynorphin

Dynorphin, the endogenous peptide for the kappa opioid receptor, is a complex neuropeptide known to exist in many forms, yet we do not have a good grasp of the function of each of these forms or fragments. Dynorphin was first purified, sequenced and named Dynorphin A (1–17)(Cox et al., 1975; Goldstein et al., 1981, 1979). Interestingly, the amino-acid sequence was found to be the same as leu-enkephalin (YGGFL) but with a basic carboxy-terminal extension. This is one of reasons why detecting peptide changes in vivo has been so challenging. The discovery of Dynorphin B (13 amino acids) or rimorphin followed, in conjunction with the c-terminus extended form of dynorphin B, Big Dynorphin and leumorphin (Fischli et al., 1982; Kilpatrick et al., 1982; see Ferre et al., 2019 for sequences). This was preceded by discovery of the endogenous morphine-like substances and opioid peptides methionine and leucine enkephalin (Cox et al., 1976, 1975; Goldstein et al., 1971; Hughes et al., 1975; Li and Chung, 1976; Teschemacher et al., 1975; Tperenius and Wahlström, 1975). This era has been fondly described as the ‘golden era of neuropeptide discovery’ (Chavkin, 2013) a critical period during which our understanding of neuropeptide function was enhanced following new methods of neuropeptide detection and purification. In response to this and in the period of time that followed there was a drive to better understand the biochemistry and structural biology of these opioid peptides and as a result the cDNA was cloned and sequenced for prodynorphin, the precursor for dynorphin (Kakidani et al., 1982).

During the same period there was also a push to define and better understand the existence of specific opioid receptors mu, sigma and kappa (Martin et al., 1976), as well as mu-, delta-, kappa-type action using in vitro bioassays (Lord et al., 1977). There were a number of limitations with these assays at the time that prevented determination of receptor selectivity (Chavkin, 2013) until the development of selective-receptor protection strategy using the nitrogen mustard analog of naltrexone, β-chlornaltrexamine (β-CNA) (Portoghese et al., 1978), which ultimately resulted in confirmation that dynorphin A was an endogenous ligand for the ‘physically distinct, non-converting kappa opioid receptor’ (Chavkin, 2013; Chavkin et al., 1982). Following the discovery that different opioid receptors exist and have selective endogenous peptides, the focus of the field shifted primarily towards using this knowledge to ascertain the function of these opioid receptor systems in analgesia and addiction (discussed in detail further in the review).

Importantly the cloning of the opioid receptor genes (Chen et al., 1993; Evans et al., 1992; Kieffer et al., 1992; Mollereau et al., 1994; Yasuda et al., 1993), very much advanced our knowledge of the dynorphin/KOR signaling system as part of the G protein-coupled receptor (GPCR) superfamily, specifically Gi/Go-coupled (for a detailed review see (Al-Hasani and Bruchas, 2011). For example, we now know that a distinct characteristic of KOR is the high negative potential at extracellular loop 2, that has been shown experimentally to increase the surrounding concentration of dynorphin due to its highly positive charge (Björnerås et al., 2014; Wang et al., 1994).

To date we know that prodynorphin is primarily expressed in the central nervous system, where upon depolarization of prodynorphin-containing neurons, protein convertase 2 processing occurs to release dynorphin from within these presynaptic neurosecretory vesicles. Dynorphin in turn primarily activates KOR thereby modulating neurotransmitter release and postsynaptic activity (Chavkin et al., 1983; Civelli et al., 1985; Ferré et al., 2019; Molineaux and Cox, 1982; Pickel et al., 1993; Wagner et al., 1993, 1991). However, we do not have a clear grasp of the recycling and degradation kinetics of dynorphin following release. Furthermore, it is not clear which or when specific dynorphin fragments are released or necessary for signaling. Interestingly, it seems that the content/distribution of unprocessed prodynorphin and dynorphin peptides may differ in axon terminals depending on the brain region (Yakovleva et al., 2006), which suggests that regional differences in trafficking and processing of the protein exists, further complicating matters.

Over the years research efforts have flip flopped between focusing on better understanding the function and dynamics of dynorphin versus KOR receptor function and binding to allow us to better understand their mechanisms of action. In the last 10 or so years there has been increased effort to better visualize and solve the structure of not only the KOR receptor in their active and inactive state but GPCRs in general, using crystallography and electron microscopy (Che et al., 2018; Wu et al., 2012). Despite these impressive advances it is now more apparent than ever that in conjunction with a better understanding of KOR receptor function there are limitations in our ability to study the function of dynorphin in vivo, which the field is currently trying to address and is discussed below.

2. Dynorphin projections

Initial traces of dynorphin were observed as early as embryonic age 14 in mouse brain (Rius et al., 1991). In the adult brain, dynorphin is found in various regions. Relatively high levels of dynorphin were measured in the amygdala, entorhinal cortex, dentate gyrus, hypothalamus, midbrain, striatum, hippocampus, and medulla-pons (Fallon and Leslie, 1986; Goldstein and Ghazarossian, 1980; Schwarzer, 2009). In comparison, lower levels of dynorphin were observed in the cerebellum and cortex (Goldstein and Ghazarossian, 1980). These studies indicate that dynorphin is widely distributed in the brain, particularly in regions that are critical in the affective modulation of a number of neuropsychiatric diseases including addiction and mood disorders. In the following section we will explore dynorphin projections in the mesolimbic regions, including the extended amygdala.

2.1. Ventral Tegmental Area

The ventral tegmental area (VTA) provides dopaminergic innervation to many brain areas and receives afferents containing various neurotransmitters and neuromodulators from other regions. Neurons in the VTA have been known to be affected by ethanol. For example, acute ethanol directly excites dopamine neurons in the VTA (Brodie et al., 1999) and results in increased AMPA/NMDA ratio (Saal et al., 2003). Moreover, intra-VTA infusion of ethanol via self-administration is reinforcing and increases subsequent self-administration (Rodd et al., 2004). During withdrawal however, spontaneous activity dopamine neurons is reduced (Diana et al., 1992; Shen and Chiodo, 1993). Furthermore, acute alcohol exposure inhibits firing of GABA neurons in the VTA (Steffensen et al., 2009) followed by a large increase in GABA neuron activity during acute and chronic ethanol induced withdrawal (Gallegos et al., 1999; Steffensen et al., 2009). A recent study has shown that withdrawal following both acute and chronic ethanol modulate glutamate transmission in the VTA (Williams et al., 2018). Specifically, acute-ethanol induced withdrawal resulted in decreased spontaneous excitatory presynaptic currents (EPSCs) and amplitudes of miniature EPSCc; in contrast, withdrawal from chronic intermittent ethanol exposure resulted in greater frequency of presynaptic events (spontaneous and miniature EPSCs). Moreover, systemic administration of ethanol acutely results in elevation of dynorphin levels (Jarjour et al., 2009). Together, these studies show that the VTA is a crucial structure involve is ethanol-related processing and the overall responsivity of this brain region may influence the development of alcohol use disorder (AUD). Though not studied in detail, VTA receives dynorphinergic projections from the lateral hypothalamus and the striatum. In the following sub-sections, we will discuss the impact of that these afferents may have on reward seeking behaviors based on their synaptic targets within the VTA.

Lateral hypothalamic afferents

The VTA receives inputs from several regions containing prodynorphin expressing neurons. These neurons impinge upon VTA projection neurons and modulate reward processing and seeking behaviors. Neurons in the lateral hypothalamus, a region with a substantial projection to the VTA (Fallon et al., 1985; Iyer et al., 2018), have been recently implicated in reward seeking (España et al., 2011, 2010; Richardson and Aston-Jones, 2012). Within the VTA, these projection neurons make connections with dopaminergic cell bodies which in turn project to the medial and lateral nucleus accumbens shell (NAc shell) and the basolateral amygdala (BLA), and inhibit them via activation of its cognate receptor, the KOR (Baimel et al., 2017). Behaviorally, dynorphin release in the VTA via activation of these lateral hypothalamic neurons results in a reduction of reward seeking, possibly due to their inhibitory effects on mesolimbic dopamine projection neurons (Muschamp et al., 2014).

Striatal afferents

In addition to the lateral hypothalamus, striatal neurons also project to the VTA. The output neurons of dorsal and ventral striatum consist of medium spiny neurons (MSNs) expressing either dopamine-1 (D1) receptor or dopamine-2 (D2) receptor. Both, D1- and D2-MSNs are inhibitory and release GABA. D1-MSN neurons express prodynorphin and release dynorphin along with GABA (Al-Hasani et al., 2015). The striatum is divided in dorsal and ventral aspects consisting of the caudate-putamen and the nucleus accumbens (NAc), respectively. D1-MSN projections from dorsal striatum terminate in substantia nigra pars compacta and the VTA while the ones from ventral striatum only project to the VTA (Fallon et al., 1985). Ventral striatum, as referred to as the NAc, provides one of the most prominent feedback projections to the VTA (Watabe-Uchida et al., 2012). These striatal D1-MSN neurons preferentially impinge upon non-dopaminergic GABA neurons in the VTA resulting in disinhibition of dopamine neurons (Bocklisch et al., 2013; Chuhma et al., 2011; Xia et al., 2011). However, Yang and colleagues (Yang et al., 2018) recently showed a regionally specific projection of D1-MSNs from the NAc to the VTA. Specifically, they showed that D1-MSNs from medial NAc shell impinge upon VTA dopamine neurons predominantly projecting to medial NAc shell. In contrast, the same study showed that D1-MSNs from lateral NAc shell target VTA GABA neurons. Both projections were observed to exert inhibitory control over their respective targets, therefore resulting in opposing actions. Dopaminergic and non-dopaminergic neurons in the VTA express KORs; activation of these receptors results in a reduction of firing rate of these neurons (Ford et al., 2006; Margolis et al., 2006, 2005, 2003). Thus, even though D1-MSNs corelease GABA and dynorphin, it is possible that at least some of the inhibition is driven by dynorphin transmission. It is important to note however, that while VTA dopamine neurons projecting to NAc have been shown to not express KORs, KORs are present on VTA dopamine neurons projecting to prefrontal cortex (PFC) and the BLA (Margolis et al., 2006). Actions of dynorphin via KORs have been reported at both, pre- and post-synaptic locations (Ford et al., 2007).

In summary, dynorphin containing projections form the lateral hypothalamus and the NAc impinge upon both dopaminergic as well as GABAergic neurons in the VTA. Data from studies in the recent years have certainly added substantial understanding of how affective and reward-seeking behaviors are shaped by dynorphin and have helped identify targeted cell-types of these projection neurons. Nevertheless, more studies are necessary to fully identify targets of these projections, including topographical organization. Complete understanding at that level will enable us to clarify the role of dynorphin in addictive behaviors.

2.2. Nucleus Accumbens

Similar to VTA, the NAc is implicated in AUD based on findings of preclinical studies demonstrating changes in accumbal neuron activity following acute and chronic ethanol exposure. Acute systemic administration of ethanol (physiologically relevant doses, 0.5 – 2 g/kg) increases extracellular levels of dopamine in the NAc in ethanol naïve rodents (Imperato and Di Chiara, 1986; Karkhanis et al., 2016b, 2014; Yim et al., 2000) while very high doses of ethanol decrease dopamine levels (Imperato and Di Chiara, 1986; Yim and Gonzales, 2000). A similar effect, inhibition of dopamine at high concentrations of ethanol, is observed in ex vivo preparations (Yorgason et al., 2015, 2014); this effect is modulated by nicotinic acetylcholine receptors (Yorgason et al., 2015). Chronic intermittent ethanol exposure however, results in hypodopamine state in rodents and macaques possibly driving anhedonia, subsequently leading to excessive ethanol consumption (Karkhanis et al., 2016a; Rose et al., 2016; Siciliano et al., 2016a, 2016b). Furthermore, following chronic ethanol exposure low to moderate doses of ethanol administered systemically during acute withdrawal result in reduced extracellular levels of dopamine (Karkhanis et al., 2016a). Surprisingly, the opposite effect is observed in rats exposed to social isolation during adolescence, that is, systemic administration of ethanol following prolong adolescent social isolation, which also results in excessive ethanol consumption, augments ethanol-induced elevation in dopamine (Karkhanis et al., 2016b).

In addition to dopamine, glutamatergic transmission within the NAc also plays an important role in alcohol seeking and consumption. Activation of cortical glutamatergic afferents into the NAc is essential for cue-induced alcohol seeking (Keistler et al., 2017). In contrast, activation of glutamatergic projection from the basolateral amygdala to the NAc disrupted cue-induced alcohol seeking (Keistler et al., 2017). Moreover, glutamate receptors on medium spiny neurons in the NAc alter their morphology after chronic ethanol exposure, further affecting alcohol consumption and seeking (Laguesse et al., 2017). A recent study has shown increased c-Fos expression following chronic intermittent ethanol exposure during acute withdrawal in the NAc shell and core; this augmented c-Fos expression is present during protracted withdrawal (7 days post final ethanol exposure) in NAc shell (Smith et al., 2019).

In addition to the neurotransmitter changes highlighted above, neuropeptides, such as dynorphin and met-enkephalin are also altered by ethanol in the NAc. For example, systemic administration of ethanol elevates dynorphin and met-enkephalin levels (Marinelli et al., 2006, 2005). Interestingly, the dopamine inhibiting and augmented elevating effects of dopamine following chronic intermittent ethanol exposure and social isolation, respectively, are reversed following KOR blockade with norBNI (Karkhanis et al., 2016b, 2016a). These studies indicate that the dynorphin/KOR system interacts with neurotransmitters in the NAc and are involved in shaping the progression towards maladaptive reward and alcohol seeking behaviors as well as AUD. NAc consists of local and projecting dynorphin neurons; we will discuss these projections and the influence dynorphin may have in shaping synaptic transmission of various neurotransmitters within the NAc.

Lateral hypothalamic afferents

Like the VTA, NAc receives dynorphin projections from lateral hypothalamus. In fact, many of the dynorphin expressing neurons projecting from the lateral hypothalamus to the VTA reside in the same topographical location and are similar in response properties to the ones that project to NAc (Iyer et al., 2018). In fact, this study showed that about 21% of the neurons from the lateral hypothalamus project to both NAc and the VTA. Topographical survey of the terminal location of these lateral hypothalamic neurons indicates that these neurons terminate in the NAc core as well as the medial and lateral aspects of NAc shell (Baldo et al., 2003; Brown et al., 2013; Iyer et al., 2018), however information identifying the target neurons of these hypothalamic projections within the NAc is lacking.

Local collaterals on dopamine terminals

D1-MSNs residing in the NAc co-release dynorphin and GABA. As mentioned in section 2.1, these neurons project to the VTA and impinge upon dopamine projection neurons and local GABAergic interneurons within the VTA. Axon collaterals from these neurons also synapse onto local dopamine terminals arising from the VTA and other cell types within the NAc. Thus, dynorphin containing cell bodies and axon terminals are present in the NAc (Van Bockstaele et al., 1995). The KORs are located presynaptically on dopamine terminals in the NAc (Svingos et al., 2001; Werling et al., 1988) and their activation reduces dopamine release (Britt and McGehee, 2008; Karkhanis et al., 2016a, 2016b; Thompson et al., 2000) while increasing dopamine uptake, leading to an overall decrease in extracellular dopamine levels (Thompson et al., 2000). Likewise, dopamine transmission in the NAc regulates dynorphin; ablation of mesostriatal dopamine neurons with 6-hydroxydopamine resulted in a reduction of dynorphin mRNA levels and increasing dopamine levels by apomorphine administration increased dynorphin mRNA levels (Gerfen et al., 1991). Together, studies showing bidirectional relationship between dopamine and dynorphin imply that while dynorphin release-mediated excitation of KORs can inhibit dopamine release, dopamine transmission can control the excitability of these same dynorphin containing D1-MSNs.

Local collaterals on other neurotransmitter terminals

In addition to presynaptic location on dopamine terminals, KORs are also expressed presynaptically on glutamatergic and serotoninergic neurons, and pre- and post-synaptically on GABAergic neurons (Land et al., 2009; Svingos et al., 2001; Tao and Auerbach, 2002). KORs on serotonin terminals in the NAc promote serotonin transporter insertion following repeated exposure to stress, thus resulting in a reduction in serotonin transmission, driven by endogenous dynorphin release (Schindler et al., 2012). NAc receives most of its glutamatergic afferents from the PFC, ventral hippocampus, and the basolateral amydgala (BLA) (Britt et al., 2012). KORs inhibit glutamate presynaptically in a pathway specific manner; this inhibitory control results in attenuation of D1-MSN excitation specifically by activation of KORs on BLA glutamate projection but not on hippocampal glutamate projections (Tejeda et al., 2017). Effects of KOR activation on cortical glutamate afferents into the NAc are still unknown. Serotonin afferents into the NAc arise from the dorsal raphe and these terminals express KORs (Land et al., 2009; Tao and Auerbach, 2002). KORs are present on both D1- and D2-MSNs and their activation inhibit D2-MSNs collaterals; however, this inhibitory effect is minimal in comparison to KOR mediated inhibition of D1-MSNs (Tejeda et al., 2017). Therefore, the effect of dynorphin in the overall excitation of output D1-MSNs of NAc is the sum of inhibition and disinhibition of glutamate, serotonin, and dopamine transmission.

Regional differences

Previous literature indicates that there are regional differences in dynorphinergic and KOR control over hedonic and dysphoric behavioral responses. Studies have shown that local activation of both D1-MSNs and KORs is essential to modulate aggression and rejection of conspecifics in prairie voles (Aragona et al., 2006; Resendez et al., 2012) and local activation of dynorphin-expressing D1-MSNs in ventral NAc shell drives aversive behavioral response (Al-Hasani et al., 2015). Surprisingly, photostimulation of dynorphin-expressing D1-MSNs in dorsal NAc shell resulted in positively reinforced motivated behaviors (Al-Hasani et al., 2015). Both the aversive and reinforcing responses were inhibited by KOR blockade, therefore implying that these behavioral differences were produced by dynorphin (Al-Hasani et al., 2015). Differences in “liking” and “wanting” have also been observed along the rostro-caudal axis within NAc shell. Specifically, stimulation of KORs in the rostro-dorsal quadrant of NAc shell resulted in hedonic reactions, while stimulation of KORs in the caudal half of NAc shell resulted in a suppression of these hedonic reactions (Castro and Berridge, 2014). These regional differences and the collective effect of dynorphin on various neurotransmitter systems shape affective and motivated behaviors, which are grossly impacted by stress (Castro and Bruchas, 2019). The complexity of the nature of interaction of all these systems highlight the importance of using methods that enable us to selectively modulate specific circuits.

2.3. Bed Nucleus of Stria Terminalis

The bed nucleus of stria terminalis (BNST) is a part of the extended amygdala and plays a major role in stress-induced intake of ethanol. Inhibition of excitatory neural activity within the BNST reduced the magnitude of conditioned place preference for ethanol (Pina et al., 2015). Furthermore, ethanol-induced conditioned place preference, which involves contextual learning and repeated exposures to ethanol, is associated with overall augmented neuronal excitability within the ventral BNST; in contrast, ethanol-mediated conditioned place aversion results in overall attenuation of presynaptic glutamatergic transmission with no change in GABAergic transmission (Pati et al., 2019). Studies have also shown modulation in catecholamine transmission following ethanol administration. Specifically, acute systemic administration of ethanol elevates extracellular levels of dopamine (Carboni et al., 2000) and norepinephrine (Jadzic et al., 2019). Chronic intermittent exposure to ethanol results in increased c-Fos expression in the dorsal and ventral BNST during acute withdrawal; however, this effect is observed only after 26 hours into withdrawal in dorsal BNST (Smith et al., 2019).

In addition to changes in neurotransmitters, the dynorphin/KOR system within the BNST also regulates behaviors associated with ethanol self-administration. Dynorphin in the BNST has been shown to regulate anxiolytic phenotype; specifically, deletion of KORs on glutamate inputs in the BNST augments anxiety-like behaviors (Crowley et al., 2016). This alteration in anxiety-like behavior may, at least in part, further regulate ethanol-seeking behaviors. Activation of KORs by systemic administration of the agonist increases stress-aversive and anxiety-like behaviors and promotes ethanol intake and seeking as well as reinstatement of ethanol seeking; subsequent local inhibition of KORs within the BNST reverses these behaviors (Lê et al., 2018). Together these studies indicate that BNST is involved in AUD and dynorphin activity within the BNST may shape its contribution to the development and/or maintenance of AUD. In the following sub-sections, we will discuss neural circuits that are modulated by dynorphin within the BNST.

Dynorphin influence on GABA transmission

KORs are broadly expressed in the BNST (Poulin et al., 2009). It has been shown that high levels of expression of prodynorphin mRNA, and therefore dynorphin, were observed in fusiform, oval, rhomboid, and anterior lateral nuclei of the BNST, with moderate levels of expression in anteromedial and interfascicular nuclei (Crowley et al., 2016; Fallon and Leslie, 1986; Kash et al., 2014; Poulin et al., 2009). These dynorphin containing neurons also contain GABA (Crowley et al., 2016). High density of KORs were observed in the dorsomedial, principal, and ventral nuclei of the BNST (Marchant et al., 2007; Poulin et al., 2009). Interestingly, the expression patterns of dynorphin and the cognate KOR are complementary as labeling for prodynorphin and KORs was not observed in the same nuclei, except for rhomboid nucleus, where expression of both prodynorphin and KORs were observed (Poulin et al., 2009). The BNST receives a strong GABAergic projection, co-expressing dynorphin, from lateral central amygdala (CeA; Marchant et al., 2007) and the paraventricular hypothalamus (Crowley et al., 2016). Neurons in lateral CeA have been observed to express prodynorphin, and about 33% of these prodynorphin-expressing neurons, co-express corticotrophin releasing hormone (Marchant et al., 2007). An overall reduction in GABA transmission via presynaptic mechanisms was observed following KOR activation in the BNST; selective circuit manipulation using optogenetics revealed that this KOR-mediated reduction in GABA transmission was specific to the CeA→BNST projection (Li et al., 2012; Normandeau et al., 2018).

Dynorphin influence on glutamate transmission

In addition to GABAergic afferents, BNST receives glutamatergic projections from the BLA and the PFC (Kash, 2012). Surprisingly, KOR activation using an agonist as well as local dynorphin neuron stimulation within the BNST selectively inhibit BLA glutamate transmission but not PFC glutamatergic projection in the BNST (Crowley et al., 2016). The glutamate afferents from the BLA target both dynorphin positive and negative neurons; however, their control over dynorphin positive neurons is significantly greater than dynorphin negative neurons (Crowley et al., 2016).

In summary, of all the GABAergic and glutamatergic afferents in the BNST, the dynorphin/KOR system selectively inhibits GABA transmission in the CeA→BNST projections and glutamate transmission of the BLA→BNST projections. While we don’t know the exact net result of this mixed modulation, what we do know is that deletion of KORs from glutamatergic BLA terminals in the BNST results in augmentation of anxiolytic behavior (Crowley et al., 2016). These data imply that akin to the NAc, the effect of the dynorphin/KOR system is specific to targeted activation or inhibition.

In conclusion, dynorphin release and therefore KOR activation results in mixed behavioral effects even though the receptors couple to inhibitory G-protein therefore reducing neurotransmission. Specifically, activation of dynorphin containing neurons in dorsal NAc results in positively reinforced motivated behaviors while activation of dynorphin neurons in the ventral NAc promote aversive behaviors. Furthermore, opposing effects of KOR stimulation have been observed along the rostro-caudal axis in the NAc, with activation in rostral regions driving hedonic response and caudal regions suppressing these hedonic responses. In the BNST activation of KORs results in anxiolytic effects, which is the opposite of what is found in other areas. Collection of these studies highlights the importance of receptors, cell, and circuit specificity. While these are results that we observe in rodents naïve to experimental perturbations such as stress and drug exposure, the complexity of the system can only hint to differential responses that may be observed following chronic stress and drug exposure. The following sections will focus on these altered states.

3. Role of dynorphin in alcohol use disorder

The dynorphin/KOR system is modulated by both chronic and acute exposure to ethanol across various brain regions. A number of studies using the chronic intermittent ethanol (CIE) exposure paradigm using inhalation chambers have thoroughly examined the effect of CIE exposure on KOR function (Anderson et al., 2016a; Karkhanis et al., 2016a; Rose et al., 2016). These studies show that CIE exposure during adulthood results in increased KOR function in the NAc measured by recording KOR-activation mediated inhibition of dopamine release; these effects were observed immediately after and 72 hours after cessation of the last ethanol exposure (Karkhanis et al., 2016a; Rose et al., 2016). CIE-induced escalation in ethanol intake was inhibited by KOR blockade (Anderson et al., 2016a; Rose et al., 2016). Likewise, after six months of chronic ethanol consumption in monkeys, KORs were observed to be hyperfunctioning in the NAc (Siciliano et al., 2016a, 2018). An increase in the KOR gene, Oprk1, mRNA levels has been observed in the BNST following long-term ethanol self - administration in rats (Erikson et al., 2018). Various studies have shown that activation of KORs or augmented KOR function/signaling elevates ethanol consumption (Anderson et al., 2016b, 2016a; Kissler et al., 2014; Rose et al., 2016; Karkhanis et al., 2016b). Though much of the alcohol literature has focused on understanding the function and expression of KORs, it only explains ethanol’s effects partially. To fully understand the effects of ethanol on the dynorphin/KOR system, it is crucial to study the ligand. In fact, one study has shown that ethanol application to striatal neurons in vitro potentiates prodynorphin expression (Logrip et al., 2008). The following section focuses on studies that show changes in dynorphin or prodynorphin expression following acute and chronic ethanol exposure at various ontogenetic timepoints.

3.1. Ethanol exposure in adulthood

Acute ethanol exposure

Despite the difficulty involved in measuring dynorphin levels, a few studies have shown that acute and chronic ethanol exposure positively modulates dynorphin levels across various brain areas. A single ethanol exposure resulted in an upregulation of prodynorphin gene in the PFC and the amygdala (D’Addario et al., 2013). For example, single systemic administration of ethanol results in elevation in dynorphin levels in the NAc (Marinelli et al., 2006), VTA (Jarjour et al., 2009), CeA (Lam et al., 2008), and the paraventricular nucleus (Chang et al., 2007). The peak increase in dynorphin in the NAc was observed at 30 mins following ethanol administration, with a greater elevation at a high dose (3.2 g/kg) of ethanol (Marinelli et al., 2006). Importantly, complementary studies how that dopamine levels do not increase above baseline at high doses beyond 2.5 g/kg (Yim et al., 2000). Furthermore, at this timepoint, 30 mins post injection, dopamine levels have been observed to return to baseline (Karkhanis et al., 2014; Yim et al., 2000). Parametric studies of ethanol, however, have shown a dissociation between brain ethanol concentration and dopamine levels, where ethanol concentrations in the brain last for twice the amount of time after dopamine levels peak (Yim et al., 2000). Together, these studies suggest that the initial elevation of dopamine levels following acute systemic ethanol administration possibly activate D1-MSNs, thus increasing dynorphin levels, which then inhibits dopamine.

Chronic or repeated ethanol exposure

Repeated and chronic exposure to ethanol also increases dynorphin expression in various brain regions. For example, tissue content of dynorphin in the NAc and the periaqueductal gray increased by 30 mins and remained high for 21 days after the final ethanol dose following a 13-day regimen of systemic ethanol administration (Lindholm et al., 2000). This continuous elevation of tissue content opposes the transient elevation observed by Marinelli and colleagues (2006) following a single systemic administration. This difference could be due to multiple exposures verses a single exposure; however, the difference could also be driven by the fact that extracellular levels of dynorphin were measured in one study using microdialysis followed by radioimmunoassay (Marinelli et al., 2006), while total content – intra- and extra-cellular – was measured in the other (Lindholm et al., 2000). While we don’t fully understand the exact differences between the availability of usable intracellular versus extracellular levels of dynorphin, it is possible that total content levels will not necessarily match extracellular levels of dynorphin. This may also be dependent on the brain region in question.

In addition to systemic administration, voluntary consumption of ethanol has also been shown to increase expression of dynorphin or mRNA encoding dynorphin and prodynorphin. Specifically, expression of mRNA encoding dynorphin has been reported to be elevated in the CeA and the hypothalamus in alcohol preferring rats compared to non-preferring rats following voluntary consumption (Zhou et al., 2013). During acute withdrawal, expression of mRNA encoding dynorphin and pro-dynorphin was observed to be increased in the NAc after one-month of voluntary ethanol consumption; the expression levels returned to baseline after 96 hours (Przewłocka et al., 1997). In addition to changes in the NAc, manipulation of dynorphin transmission in the CeA also regulated ethanol intake. Chemogenetic inhibition of dynorphin-containing neurons in the CeA reduced binge-like ethanol intake and activation of KORs using the agonist in the presence of dynorphin inhibition increased binge-like ethanol intake (Anderson et al., 2019). In comparison, chronic ethanol exposure augmented dynorphin expression in the CeA of dependent rats during acute withdrawal (Kissler et al., 2014). Likewise, dependence-induced heightened consumption of ethanol was reversed by KOR blockade, potentially by inhibiting the actions of dynorphin (Berger et al., 2013). This effect was not observed in non-dependent rats. Together these studies show that inhibition of dynorphin transmission in the CeA rescues compulsive ethanol consumption and excessive ethanol exposure augments dynorphin production. Furthermore, blocking dynorphin effects by systemic administration of KOR antagonist also reduces compulsive ethanol intake. It is important to note that in another study, KOR blockade in adult rats with limited access to ethanol (30 min access to 10% ethanol every other day for a total of 3 exposures), resulted in elevation of ethanol intake in male rats and a reduction in female rats (Morales et al., 2014). This study confirms results from other studies showing a lack of KOR-blockade effect in non-dependent male rat. A majority of the studies have been primarily in male animals, which highlights the paucity in literature examining effects of ethanol on the dynorphin/KOR system in females. Clearly, there are differences that need to be studied.

Since rodents do not readily consume ethanol to the point of intoxication or dependence, ethanol vapor exposure is generally employed to induce dependence. Most of the studies examining changes in the dynorphin/KOR system following CIE exposure have focused on examining KOR function and manipulating it. Therefore, results from these studies only implicate the indirect impact of CIE on dynorphin function and transmission. CIE-induced escalation in ethanol intake is reversed by KOR blockade in mice (Anderson et al., 2016a; Rose et al., 2016). Moreover, dependence-induced elevation in operant responding for ethanol is also reduced by KOR blockade in rats (Walker and Koob, 2008). The effect of an acute ethanol challenge on extracellular levels of dopamine in the NAc in dependent mice (CIE exposed), as measured by microdialysis is inhibitory; which is reversed by inhibition of KORs (Karkhanis et al., 2016a). In non-dependent rats, an acute ethanol challenge increases dopamine levels and KOR blockade has no effects on this ethanol-mediated elevation in dopamine (Karkhanis et al., 2016a). The reversal of inhibitory effect of ethanol challenge on dopamine levels and reversal of escalated ethanol consumption in dependent mice and rats following KOR blockade suggests that dynorphin effects, which were blocked, were essential for producing those effects.

Relapse following protracted abstinence

The studies mentioned in the acute and chronic alcohol exposure subsections above have investigated dynorphin role following acute withdrawal – 0 to 72 hours post last ethanol exposure. However, it is important to note that the KOR/dynorphin system may in fact be involved in promoting relapse drinking and/or reinstatement of alcohol seeking after protracted abstinence in non-dependent and dependent animals. For example, KOR activation in the BNST promoted reinstatement of alcohol seeking following seven days of extinction in rats with history of alcohol self-administration; this effect was blocked by inhibition of KORs with norBNI (Lê et al., 2018). Furthermore, KOR activation in dependent rats promoted alcohol seeking at a greater propensity compared to non-dependent animals after 10 days of abstinence (Funk et al., 2019). In addition, reinstatement of ethanol seeking driven by ethanol conditioned stimulus was blocked by naltrexone, an opioid antagonist (Liu and Weiss, 2002), indicating a possible involvement of dynorphin in promoting context dependent reinstatement.

It is known that high anxiety levels, depression, comorbid affective disorders, and exposure to stress increase the propensity and risk of AUD in humans (McCaul et al., 2017; Strine et al., 2012). This anxiety and depression can be a result of preexisting neuropsychiatric disorders or it can be driven by excessive alcohol consumption. Collection of results from rodent studies suggest that stress influences the progression of AUD; stress enhances alcohol reward and alcohol alleviates stress; chronic alcohol compromises behavioral coping and alters stress responsivity; together these changes increase vulnerability to stress triggered relapse (see Becker, 2017 for review). Inhibition of KORs following protracted abstinence has been shown to reduce depressive-like behavior (Jarman et al., 2018), possibly also reducing occurrence of relapse and reinstatement (Lê et al., 2018). Together these studies indicate that in adults, the relationship between dynorphin/KOR system and alcohol is bidirectional. While acute and chronic alcohol exposure modulate dynorphin and KOR function, alcohol seeking, craving, and consumption are in fact also driven by changes in the dynorphin/KOR system.

3.2. Ethanol exposure during adolescence

While the above-mentioned studies show changes in dynorphin following acute or chronic ethanol exposure during adulthood, some studies have examined changes in dynorphin following adolescent ethanol exposure. Adolescence is an vulnerable period during which dynorphin mediated changes are heightened, for example, one study has shown that dynorphin-induced hyperpolarization in paraventricular nucleus neurons increases through puberty (approximately four weeks of age), followed by a return to levels observed at two weeks of age (Chen et al., 2015). Furthermore, ethanol exposure during adolescence promotes heightened ethanol consumption later in life (Spear, 2015). Adolescent ethanol exposure, in alcohol preferring rats, alters the mesolimbic dopamine system to be more sensitive and responsive to ethanol (Toalston et al., 2014), however mechanisms driving this high intake of ethanol are still unknown. A few studies have shown fluctuations in dynorphin levels following repeated ethanol exposure during adolescence. Orogastric exposure to ethanol (2 g/kg) for three consecutive days per week between four and nine weeks of age resulted in augmented dynorphin expression in the substantia nigra when examined three weeks after the last ethanol exposure, but not 2 hours post final exposure (Granholm et al., 2018). Surprisingly, this study did not show an effect of treatment (water vs. ethanol) on dynorphin expression in the NAc at either time points. Another study using voluntary ethanol intake in adolescent rats showed that KOR blockade did not reduce ethanol intake after a three-day repeated ethanol exposure regimen (10% ethanol in supersac solution, every other day); an effect that was observed in adult female rats (Morales et al., 2014). This may be due to housing effects as Morales and colleagues (2014) housed their rats singly (1 rat/cage) starting post-natal day 25. Interestingly, alcohol-induced effects on dynorphin in adolescent rats were dependent on housing condition: dynorphin expression levels were lower in single housed (each single housing session = 24 hours) adolescent rats in the amygdala when measured two hours after the final isolation session, but this effect was rescued by ethanol (20% ethanol, voluntary intake for 3 consecutive days of the week for six weeks; (Palm and Nylander, 2014). Surprisingly, no differences in dynorphin expression levels between group and single housed rats or between water and ethanol drinking single housed rats were found in the NAc (Palm and Nylander, 2014). Because many neural processes are changing during adolescence development, it is difficult to identify specific mechanisms that may drive the observed behavioral changes. The lack of changes in dynorphin observed in some studies may be due to, at least in part, continuous development of processes and pruning during this age.

3.3. Prenatal ethanol exposure

Similar to adolescent ethanol exposure, exposure to prenatal ethanol exposure promotes high ethanol consumption later in life. It is surprising however, that very few studies have actually examined these differences. One study has shown that repeated prenatal ethanol exposure (gestational day 17 – 20; 1 or 2 g/kg) results in significantly greater dynorphin mRNA levels following the 2 g/kg dose, and lower levels of protein content following 1 and 2 g/kg when examined on post-natal day 8 in the NAc (Bordner and Deak, 2015). In the same study, an increase in dynorphin mRNA was observed in the VTA following 2 g/kg when examined on post-natal day 4 (Brodner and Deak, 2015). This study indicates that prenatal ethanol exposure possibly results in an upregulation in dynorphin signaling. Similar results were observed in a separate study, which showed that prenatal (gestation day 17 – 20) ethanol exposure (2 g/kg) resulted in increased prodynorphin mRNA expression in the VTA and a reduction of DNA methylation at the prodynorphin gene of infant and adolescent rats (Wille-Bille et al., 2018). However, the opposite effect, a reduction in prodynorphin mRNA levels was observed in the NAc (Wille-Bille et al., 2018). This prodynorphin mRNA reduction observed in the NAc is likely due to a compensatory adaptation. Overall, rats in this study were observed to have an anxiety-like phenotype. Some adult studies have shown that repeated stress exposure during adolescence and adulthood result in reduced dynorphin mRNA (Donahue et al., 2015) and content (Karkhanis et al., 2016b) and increased anxiety-like behaviors along with facilitation of ethanol consumption (Karkhanis et al., 2016b). It is known that prenatal ethanol exposure promotes elevated ethanol consumption later in life. Thus, this augmented ethanol consumption may be driven by fluctuations and instability in the dynorphin/KOR system or by compensatory mechanisms yet to be identified. To date, no studies have examined the effect of ethanol on dynorphin transmission in animals exposed to ethanol prenatally.

In summary, ethanol and the dynorphin/KOR system have a circular relationship. Facilitation of dynorphin transmission promotes ethanol drinking behaviors and excessive ethanol consumption potentiates dynorphin expression and therefore dynorphin transmission. Chronic ethanol exposure, however, at least in adults downregulates dynorphin possibly promoting other compensatory mechanisms. Dynorphin is indeed regulated by other neuromodulatory systems, such as the corticotropin releasing factor (M. R. Bruchas et al., 2010; Koob, 2015; Land et al., 2008) and brain derived neurotropic factor (Logrip et al., 2008). These systems play a crucial role in shaping dynorphin responses to ethanol and stress and perhaps the modulatory effects are ontogeny specific. In the next section, we address the role of motivation and dynorphin in the modulation of reward seeking behaviors.

4. Role of dynorphin in motivation regulation

Drug addiction literature revolves around dopamine neurotransmission in the mesolimbic areas because numerous studies over decades have shown supraphysiological elevations of dopamine levels in the striatum following drug, including ethanol, exposure (Volkow et al., 2009). Increase in dopamine levels in the ventral striatum/NAc results in activation of D1-MSNs. These recurring supraphysiological surges of dopamine and activation and the subsequent activation of D1-MSNs promote various neuroadaptations. One such adaptation is change in upregulation of dynorphin expression via activation of the transcriptional factor CREB (cAMP response element-binding protein) within the neurons of NAc (Muschamp and Carlezon, 2013). This increase in dynorphin results in inhibition of dopamine release via activation of KORs, which are localized on dopamine terminals in most terminal regions and on cell bodies in the VTA. Due to this overall reduction in dopamine transmission, there is a shift in homeostatic point of dopamine, promoting hyperkatifeia, which is defined as hypersensitivity to emotional distress and pain as well as anhedonia, defined as the inability to experience pleasures (Koob, 2019; Shurman et al., 2010).

In order to fully understand the dysregulated motivation processes, it is important to examine behavior and neurobiology in animals that demonstrate vulnerability to excessive drug use. Mice lacking dynorphin have been shown to consume significantly more ethanol compared to their wildtype counterparts (Rácz et al., 2013). It is interesting that dynorphin is not necessary for ethanol preference, in fact, dynorphin deficiency may promote ethanol self-administration; however, it is crucial to reinstate stress-induced ethanol drinking (Rácz et al., 2013). In this study, mice lacking dynorphin did not demonstrate stress-induced facilitation of ethanol consumption. Moreover, rats genetically predisposed to consume alcohol appear to have reduced dynorphin tone. For example, dynorphin levels were lower in alcohol preferring rats compared to the alcohol non-preferring rats in the NAc at baseline (Nylander et al., 1994). Similarly, Lewis rats, an inbred strain, that readily consume ethanol express less dynorphin compared to the Fischer 344 inbred rat strain (Nylander et al., 1995a). Conversely, ethanol-avoiding DBA/2 mice were reported to have greater dynorphin expression in the striatum compared to the ethanol-preferring C57BL/6j mice (Jamensky and Gianoulakis, 1997; Ploj et al., 2000). Perhaps animals with genetic modifications that either reduce or eliminate dynorphin, are highly motivated to take drugs because the feedback signal necessary to reduce drug-induced dopamine release, is dysregulated or lacking.

Animals exposed to chronic stress exhibit elevated alcohol consumption; for example, Rats and mice exposed to adolescent social isolation stress and repeated forced swim stress consume more ethanol compared to their stress naïve counterparts; an effect reversed by KOR blockade (Anderson et al., 2016a; Karkhanis et al., 2016b). Acute stress exposure, such as single exposure to social defeat, three-hours of immobilization stress, and one force swim stress exposure, elevates dynorphin expression in NAc shell and core, as well as the hippocampus for up to two hours following the cessation of stress exposure (Shirayama et al., 2004). This increase in dynorphin is transient, as the levels were different from those in stress naïve rats when examined five hours after stress exposure terminated (Shirayama et al., 2004). However, repeated exposure to stress or adversity attenuates prodynorphin mRNA levels and tissue content in the NAc (Donahue et al., 2015; Karkhanis et al., 2016b). Together, these studies indicate that in fact, low levels of dynorphin may increase vulnerability to develop maladaptive behaviors, such as, compulsive drug taking behaviors. In fact, it has been shown that exposure to chronic stress results in increased cocaine seeking (Brodnik et al., 2019; Groblewski et al., 2015) and ethanol consumption (Anderson et al., 2016a; Karkhanis et al., 2016b). This may in fact be due to altered reinforcing efficacy of ethanol and reward valency experienced by the animal as dynorphin is no longer as effective in regulating ethanol-induced reinforcement. There needs to be more work in order to elucidate the detailed mechanism involved here.

It is important to note that chronic ethanol exposure, voluntary or involuntary, as well as chronic stress exposure, both result in increased consumption of ethanol and the mechanisms driving these differences may in fact be bidirectional. Most of the studies examining dynorphin content or expression have reported changes during acute withdrawal, but no changes during protracted withdrawal. Although an indirect measure, ethanol mediated inhibition of dopamine in the NAc, which then reverses with KOR blockade (Karkhanis et al., 2016a), indicates that it is possible that the observed inhibition of dopamine is promoted at least in part due to a surge of ethanol-induced dynorphin release. This dynorphin release may occur from lateral hypothalamic neurons innervating the NAc. The stress response system is critically involved in driving addictive behaviors. It is known that chronic or repeated stress exposure attenuates dopamine tone, potentially due to hyperactive KORs, causing a hypodopaminergic state which promotes anhedonic and dysphoric behaviors. Thus, it increases vulnerability to substance and alcohol use disorder and facilitates relapse and reinstatement of drug use mostly via dysregulation of motivation-related neural processes. Stress and addiction are bidirectionally related, while stress promotes addictive and maladaptive behaviors, excessive drug consumption and cycles of withdrawal are stressful. In both situations, motivation and reward-seeking processes are dysregulated. While animals exposed to ethanol chronically may exhibit compulsive drinking to abolish negative reinforcement, it is possible that animals exposed to chronic stress may drink compulsively for the positive reinforcing values of the drug.

5. Sex differences in dynorphin function

A majority of the studies in the literature have been conducted in male animal models; some studies, however, have reported sex differences in the expression and function of dynorphin and KORs (see Chartoff and Mavrikaki, 2015 for review). Expression levels of dynorphin are sex-dependent (Becker et al., 2012). These disparities are based on differential chromosomal makeup and other variables. For example, one study has shown that the expression of prodynorphin mRNA is greater in the NAc and dorsal striatum of female mice with intact XX sex chromosomes compared to male XY mice and XO mice (Chen et al., 2009). Furthermore, expression levels of dynorphin fluctuate during the course of the estrous; while the expression of dynorphin A and B is reported to be stable throughout the estrous cycle in the NAc, expression of dynorphin A and B is observed to be lower in substantia nigra and dynorphin A is observed to be lower in caudate putamen during the estrous phase (Roman et al., 2006). Together, these studies indicate that sex differences may be involved in differential effects of KOR-activation on motivation and drug seeking observed in studies discussed in the following paragraph.

The greater levels of prodynorphin and dynorphin observed in females in areas involved in reinforcement and motivation processing may explain the heightened vulnerability of females to drug dependence. The idea here is that greater availability of dynorphin would potentially augmenting KOR activation, subsequently leading to stress-aversive effects and drug seeking. Indeed, human studies have shown evidence that stress-related disorders, including anxiety and mood disorders are twice as frequently diagnosed in females as males, and are more prone to developing drug dependence (World Health Organization, 2014). Though it is intuitive that greater dynorphin levels would enhance KOR-activation therefore resulting in greater stress-aversive states, preclinical literature demonstrates a different picture. Specifically, female rats are less sensitive than males to the KOR-activation-mediated increases in intra-cranial self-stimulation threshold of the medial forebrain bundle suggesting that females are less sensitive to depression-like effects mediated by KOR-activation (Russell et al., 2014). Activation of c-Fos following KOR activation with its agonist is observed to be greater in BNST of female compared to male rats (Russell et al., 2014). The BNST consists of a heterogeneous population of neurons; therefore, this non-specific increase in c-Fos activation is inconclusive. This same study has shown that KOR-activation augments c-Fos activation in the corticotropin releasing factor containing neurons in the paraventricular nucleus, suggesting that the enhanced depressive effects observed in females is likely driven by dynorphin/KOR system interaction with the corticotropin releasing factor system.

With respect to drug addiction and self-administration, prior activation of KORs in male rodents enhances rewarding and hyperactivity effects of cocaine and reinstates cocaine seeking (Chartoff et al., 2016; McLaughlin et al., 2006; Redila and Chavkin, 2008; Sershen et al., 1998); these effects were not observed in female mice (Sershen et al., 1998). Continuous activation of KORs with U50,488 in male rhesus monkeys augments cocaine choice (Negus, 2004). Together, these studies suggest that a KOR-mediated enhancement in relative reinforcing effects of cocaine across species. KOR modulation affects consumption of other drug classes. For example, KOR-inhibition increases ethanol intake in adult females while reducing it in adult males (Morales et al., 2014). Conversely, another study has shown a reduction in ethanol intake and preference following a combination treatment of KOR agonist Mesyl Salvanorin B or Nalfurafine in conjunction with Naltrexone, a mu opioid receptor antagonist with a lower affinity to KORs in male and female mice (Zhou et al., 2017; Zhou and Kreek, 2019a). Combination of Nalfurafine and Naltrexone also reduces relapse-like drinking in male and female mice (Zhou and Kreek, 2019b). In rodents exposed to chronic stress or ethanol, KOR-inhibition results in a reduction in ethanol intake (Anderson et al., 2016a; Karkhanis et al., 2016b; Rose et al., 2016). In the human literature, a PET imaging study has shown greater KOR availability in males compared to females in the anterior cingulate cortex, insula and the ventral pallidum among other areas (Vijay et al., 2016). A recent human study examining mechanisms via which naltrexone reduced alcohol intake shows that naltrexone-induced reduction in alcohol consumption is negatively correlated with availability of KORs in the cingulate cortex, pallidum, and the striatum (de Laat et al., 2019). Although this study enlisted male and female participants, they do not address sex differences.

Taken together, this collection of studies examining drug taking and seeking behaviors, along with other affective behaviors is rather inconclusive. Unpublished evidence from our laboratories suggests that there are individual differences within dynorphin/KOR system. These individual differences of dynorphin and KORs (both numbers and affinity) likely drive disparities observed in studies mentioned in this section. More research directly comparing males and females is crucial to advance our knowledge regarding sex differences in order to develop refined pharmacotherapies to treat neuropsychiatric disorders, including addiction.

Most of the studies presented in this review thus far have examined the KORs, therefore indirectly studying the effects of dynorphin on behavior and neurotransmission. Although extremely informative, studies that have measured tissue content or expression levels of dynorphin, prodynorphin, and their respective mRNA are not able to explain the effects of and on dynorphin to the fullest as they are not measuring online transmission of dynorphin. Methods enabling direct measure of dynorphin release and transmission are currently being developed.

6. Where do we stand with dynorphin detection?

Neuropeptides have proven very difficult to detect as they can be rapidly cleaved by peptidases and undergo posttranslational modifications. They are also found at much lower concentrations compared to classical neurotransmitters. Most of our knowledge on neuropeptide changes following perturbation in rodent models has been through by proxy looking at mRNA or using immunoaffinity-based techniques due to their high sensitivity (Kendrick, 1990; Maidment et al., 1991, 1989; Marinelli et al., 2005; Nieto et al., 2002; Nylander et al., 1995b), however selectivity remains an issue. In addition to all dynorphin isoforms deriving from prodynorphin, α- and β-neoendorphin, which are leu-enkephalin-based opioid peptides are also derived from prodynorphin (Chavkin et al, 1983, Kakidani et al, 1982, Weber et al, 1982, Zamir et al, 1984). This in turn makes selectivity a continuous issue.

At a number of different time points through history, groups have tried to measure opioid peptide changes in vivo with somewhat limited success. A number of groups were able to measure enkephalins using a combination of microdialysis and mass spectrometry but this was primarily in rat models (Baseski et al., 2005; DiFeliceantonio et al., 2012; Gobaille et al., 1994; Haskins et al., 2001; Lisi and Sluka, 2006; Mabrouk et al., 2011; Shen et al., 1997). At the time one of the major concerns was the damage created by the implantation of the microdialysis probe. Since then microdialysis probe size has decreased and laboratories have been working to optimize detection methods to allow for detection of opioid peptides at lower concentrations from smaller volumes (Zhou et al., 2015). In general, the field has been more successful in detecting enkephalins as opposed to dynorphin in rats versus mice for the following reasons.

Dynorphin has been described to come online following a perturbation/insult to the system, such as in the event of a stressful stimuli and is thought to act to regulate mood states (Bruchas et al., 2010). Furthermore, we know dynorphin exists in many different forms and fragments, yet we do not have a clear idea of the dynamics/role of these fragments in the context of different behavioral states. Interestingly, early work in the 1980’s studying the structural analysis of the dynorphin sequence gave some insight into how specific amino-acid residues may be responsible for KOR selectivity. For example, the strongly basic residues in the C-terminal domain of dynorphin, arginine-7, lysine-11 and lysine-13, seem to be critical for KOR selectivity in Dynorphin-A (Chavkin, 2013; Chavkin and Goldstein, 1981). Furthermore, it was shown that Dynorphin-B also has strong basic residues in similar positions that influence KOR selectivity (James and Goldstein, 1984). Dynorphin fragments, described as natural fragments, such as Dynorphin a (1–8), do not have the C-terminal lysines and as a result have lower KOR potency and selectivity (Chavkin and Goldstein, 1981; James and Goldstein, 1984). These findings are important and likely overlooked but should certainly be taken into account as the field pushes towards developing methods to detect dynorphin in vivo.

With the rise of optogenetic and chemogenetic tools, there has been an increased effort to use these tools, in conjunction with microdialysis/LC-MS detection, in mouse models to improve existing detection methods (Al-Hasani et al., 2018; Patterson et al., 2015). Most recently, a study showed increases in dynorphin release in two regions of the NAc shell following optogenetic stimulation of dynorphin cells using pdyn-cre transgenic mouse line (Al-Hasani et al., 2018). In addition to the observed changes in dynorphin release in vivo, both leu- and met-enkephalin were also measured, as well as dopamine, GABA and glutamate (Al-Hasani et al., 2018). The ability to optogenetically stimulate release of peptides allows for method development and optimization that would be difficult when trying to measure more subtle changes in peptides, particularly dynorphin, during chronic behavioral manipulations for example. Interestingly, a group developed a complementary approach that does not use a microdilaysis/LC/MS approach but uses multiple-scan-rate waveform to enable real-time voltammetric detection of tyrosine containing neuropeptides, specifically met-enkephalin allowing for more spaciotemporal resolution (Calhoun et al., 2018), which holds promise in its applicability detecting dynorphin.

The field is beginning to see an increase in alternative and complementary methods to not only detect in vivo release of dynorphin and opioid peptides but to learn more about their release dynamics and function. For example, photoactivable opioids were developed to achieve rapid and spatially delimited delivery of neuropeptides in mammalian brain tissue (Banghart and Sabatini, 2012). Analogs of [Leu5]-enkephalin and the 8 amino acid form of Dynorphin A (Dyn-8) were developed, which when exposed to ultraviolet light causes release of leu-enkephalin and dynorphin (Banghart and Sabatini, 2012).

With the improvement of genetically encoded sensors, such as dLight (Patriarchi et al., 2018), there has been a push to developing opioid receptor-based sensors to allow for the same robust, spatiotemporal resolution and detection of physiologically and behaviorally relevant changes in peptides as has been shown with dopamine (Patriarchi et al., 2018). Along a similar vein, earlier this year Neuropeptide Release Reporters (NPRRs) were developed using the Drosophila larval neuromuscular junction (NMJ) as a model (Ding et al., 2019). These are genetically-encoded sensors with high temporal resolution and genetic specificity to better inform us of trafficking and packaging of native neuropeptides as real-time changes in fluorescence intensity (Ding et al., 2019). In the future, this technology could perhaps be applied to detecting opioid peptides. In summary, though we still have a long way to go before we can reliably and consistently measure and distinguish all fragments of in vivo dynorphin release during acute and chronic behavioral manipulations, the field has made impressive strides in developing new technological approaches and the future of dynorphin and opioid peptide in vivo detection looks very promising.

7. Summary

In this review we have highlighted the integral role of dynorphin, the endogenous peptide for KOR, in AUD. This system is evidently extremely complex in modulation of AUD and its role seems very much dependent on the model of AUD, exposure paradigm and age of exposure. It is also becoming very clear that part of the limitation to our ability to interrogate dynorphin’s role in AUD is due to a lack of accessible techniques to measure real-time in vivo changes following behavioral manipulations and alcohol exposure. We are amidst an exciting time where multiple groups are working on a number of complementary methods that will allow the field to more readily measure changes in both opioid peptides and neuropeptides in general.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Bibliography

  1. Al-Hasani R, Bruchas MR, 2011. Molecular mechanisms of opioid receptor-dependent signaling and behavior. Anesthesiology 115, 1363–1381. 10.1097/ALN.0b013e318238bba6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Al-Hasani R, McCall JG, Shin G, Gomez AM, Schmitz GP, Bernardi JM, Pyo C-O, Park SI, Marcinkiewcz CM, Crowley NA, Krashes MJ, Lowell BB, Kash TL, Rogers JA, Bruchas MR, 2015. Distinct Subpopulations of Nucleus Accumbens Dynorphin Neurons Drive Aversion and Reward. Neuron 87, 1063–1077. 10.1016/j.neuron.2015.08.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Al-Hasani R, Wong J-MT, Mabrouk OS, McCall JG, Schmitz GP, Porter-Stransky KA, Aragona BJ, Kennedy RT, Bruchas MR, 2018. In vivo detection of optically-evoked opioid peptide release. eLife 7, e36520 10.7554/eLife.36520 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Anderson RI, Lopez MF, Becker HC, 2016a. Stress-Induced Enhancement of Ethanol Intake in C57BL/6J Mice with a History of Chronic Ethanol Exposure: Involvement of Kappa Opioid Receptors. Front Cell Neurosci 10, 45 10.3389/fncel.2016.00045 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Anderson RI, Lopez MF, Becker HC, 2016b. Forced swim stress increases ethanol consumption in C57BL/6J mice with a history of chronic intermittent ethanol exposure. Psychopharmacology (Berl.) 233, 2035–2043. 10.1007/s00213-016-4257-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Anderson RI, Lopez MF, Griffin WC, Haun HL, Bloodgood DW, Pati D, Boyt KM, Kash TL, Becker HC, 2019. Dynorphin-kappa opioid receptor activity in the central amygdala modulates binge-like alcohol drinking in mice. Neuropsychopharmacology 44, 1084–1092. 10.1038/s41386-018-0294-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Aragona BJ, Liu Y, Yu YJ, Curtis JT, Detwiler JM, Insel TR, Wang Z, 2006. Nucleus accumbens dopamine differentially mediates the formation and maintenance of monogamous pair bonds. Nat Neurosci 9, 133–139. 10.1038/nn1613 [DOI] [PubMed] [Google Scholar]
  8. Baimel C, Lau BK, Qiao M, Borgland SL, 2017. Projection-Target-Defined Effects of Orexin and Dynorphin on VTA Dopamine Neurons. Cell Reports 18, 1346–1355. 10.1016/j.celrep.2017.01.030 [DOI] [PubMed] [Google Scholar]
  9. Baldo BA, Daniel RA, Berridge CW, Kelley AE, 2003. Overlapping distributions of orexin/hypocretin- and dopamine-beta-hydroxylase immunoreactive fibers in rat brain regions mediating arousal, motivation, and stress. J. Comp. Neurol 464, 220–237. 10.1002/cne.10783 [DOI] [PubMed] [Google Scholar]
  10. Banghart MR, Sabatini BL, 2012. Photoactivatable neuropeptides for spatiotemporally precise delivery of opioids in neural tissue. Neuron 73, 249–259. 10.1016/j.neuron.2011.11.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Baseski HM, Watson CJ, Cellar NA, Shackman JG, Kennedy RT, 2005. Capillary liquid chromatography with MS3 for the determination of enkephalins in microdialysis samples from the striatum of anesthetized and freely–moving rats. Journal of Mass Spectrometry 40, 146–153. 10.1002/jms.733 [DOI] [PubMed] [Google Scholar]
  12. Becker HC, 2017. Influence of stress associated with chronic alcohol exposure on drinking. Neuropharmacology 122, 115–126. 10.1016/j.neuropharm.2017.04.028 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Becker JB, Perry AN, Westenbroek C, 2012. Sex differences in the neural mechanisms mediating addiction: a new synthesis and hypothesis. Biol Sex Differ 3, 14 10.1186/2042-6410-3-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Berger AL, Williams AM, McGinnis MM, Walker BM, 2013. Affective cue-induced escalation of alcohol self-administration and increased 22-kHz ultrasonic vocalizations during alcohol withdrawal: role of kappa-opioid receptors. Neuropsychopharmacology 38, 647–654. 10.1038/npp.2012.229 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Björnerås J, Kurnik M, Oliveberg M, Gräslund A, Mäler L, Danielsson J, 2014. Direct detection of neuropeptide dynorphin A binding to the second extracellular loop of the κ opioid receptor using a soluble protein scaffold. FEBS J. 281, 814–824. [DOI] [PubMed] [Google Scholar]
  16. Bocklisch C, Pascoli V, Wong JCY, House DRC, Yvon C, Roo M. de, Tan KR, Lüscher C, 2013. Cocaine Disinhibits Dopamine Neurons by Potentiation of GABA Transmission in the Ventral Tegmental Area. Science 341, 1521–1525. 10.1126/science.1237059 [DOI] [PubMed] [Google Scholar]
  17. Bordner K, Deak T, 2015. Endogenous opioids as substrates for ethanol intake in the neonatal rat: The impact of prenatal ethanol exposure on the opioid family in the early postnatal period. Physiol. Behav 148, 100–110. 10.1016/j.physbeh.2015.02.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Britt JP, Benaliouad F, McDevitt RA, Stuber GD, Wise RA, Bonci A, 2012. Synaptic and behavioral profile of multiple glutamatergic inputs to the nucleus accumbens. Neuron 76, 790–803. 10.1016/j.neuron.2012.09.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Britt JP, McGehee DS, 2008. Presynaptic opioid and nicotinic receptor modulation of dopamine overflow in the nucleus accumbens. J. Neurosci 28, 1672–1681. 10.1523/JNEUROSCI.4275-07.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Brodie MS, Pesold C, Appel SB, 1999. Ethanol directly excites dopaminergic ventral tegmental area reward neurons. Alcohol. Clin. Exp. Res 23, 1848–1852. [PubMed] [Google Scholar]
  21. Brodnik ZD, Black EM, España RA, 2019. Accelerated development of cocaine-associated dopamine transients and cocaine use vulnerability following traumatic stress. Neuropsychopharmacology. 10.1038/s41386-019-0526-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Brown RM, Khoo SY-S, Lawrence AJ, 2013. Central orexin (hypocretin) 2 receptor antagonism reduces ethanol self-administration, but not cue-conditioned ethanol-seeking, in ethanol-preferring rats. Int. J. Neuropsychopharmacol 16, 2067–2079. 10.1017/S1461145713000333 [DOI] [PubMed] [Google Scholar]
  23. Bruchas MR, Land BB, Chavkin C, 2010. The dynorphin/kappa opioid system as a modulator of stress-induced and pro-addictive behaviors. Brain Res. 1314, 44–55. 10.1016/j.brainres.2009.08.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bruchas MR, Land BB, Chavkin C, 2010. The dynorphin/kappa opioid system as a modulator of stress-induced and pro-addictive behaviors. Brain Res. 1314, 44–55. 10.1016/j.brainres.2009.08.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Calhoun SE, Meunier CJ, Lee CA, McCarty GS, Sombers LA, 2018. Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin. ACS Chem Neurosci 10, 2022–2032. 10.1021/acschemneuro.8b00351 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Carboni E, Silvagni A, Rolando MT, Di Chiara G, 2000. Stimulation of in vivo dopamine transmission in the bed nucleus of stria terminalis by reinforcing drugs. J. Neurosci 20, RC102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Castro DC, Berridge KC, 2014. Opioid hedonic hotspot in nucleus accumbens shell: mu, delta, and kappa maps for enhancement of sweetness “liking” and “wanting.” J. Neurosci 34, 4239–4250. 10.1523/JNEUROSCI.4458-13.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Castro DC, Bruchas MR, 2019. A Motivational and Neuropeptidergic Hub: Anatomical and Functional Diversity within the Nucleus Accumbens Shell. Neuron 102, 529–552. 10.1016/j.neuron.2019.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chang G-Q, Karatayev O, Ahsan R, Gaysinskaya V, Marwil Z, Leibowitz SF, 2007. Dietary fat stimulates endogenous enkephalin and dynorphin in the paraventricular nucleus: role of circulating triglycerides. Am. J. Physiol. Endocrinol. Metab 292, E561–570. 10.1152/ajpendo.00087.2006 [DOI] [PubMed] [Google Scholar]
  30. Chartoff EH, Ebner SR, Sparrow A, Potter D, Baker PM, Ragozzino ME, Roitman MF, 2016. Relative Timing Between Kappa Opioid Receptor Activation and Cocaine Determines the Impact on Reward and Dopamine Release. Neuropsychopharmacology 41, 989–1002. 10.1038/npp.2015.226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Chartoff EH, Mavrikaki M, 2015. Sex Differences in Kappa Opioid Receptor Function and Their Potential Impact on Addiction. Front Neurosci 9, 466 10.3389/fnins.2015.00466 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Chavkin C, 2013. Dynorphin--still an extraordinarily potent opioid peptide. Mol. Pharmacol 83, 729–736. 10.1124/mol.112.083337 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Chavkin C, Bakhit C, Weber E, Bloom FE, 1983. Relative contents and concomitant release of prodynorphin/neoendorphin-derived peptides in rat hippocampus. PNAS 80, 7669–7673. 10.1073/pnas.80.24.7669 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chavkin C, Goldstein A, 1981. Specific receptor for the opioid peptide dynorphin: structure--activity relationships. Proc Natl Acad Sci U S A 78, 6543–6547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Chavkin C, James IF, Goldstein A, 1982. Dynorphin is a specific endogenous ligand of the kappa opioid receptor. Science 215, 413–415. [DOI] [PubMed] [Google Scholar]
  36. Che T, Majumdar S, Zaidi SA, Ondachi P, McCorvy JD, Wang S, Mosier PD, Uprety R, Vardy E, Krumm BE, Han GW, Lee M-Y, Pardon E, Steyaert J, Huang X-P, Strachan RT, Tribo AR, Pasternak GW, Carroll FI, Stevens RC, Cherezov V, Katritch V, Wacker D, Roth BL, 2018. Structure of the Nanobody-Stabilized Active State of the Kappa Opioid Receptor. Cell 172, 55–67.e15. 10.1016/j.cell.2017.12.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Chen X, Grisham W, Arnold AP, 2009. X chromosome number causes sex differences in gene expression in adult mouse striatum. Eur. J. Neurosci 29, 768–776. 10.1111/j.1460-9568.2009.06610.x [DOI] [PubMed] [Google Scholar]
  38. Chen Y, Mestek A, Liu J, Hurley JA, Yu L, 1993. Molecular cloning and functional expression of a mu-opioid receptor from rat brain. Mol. Pharmacol 44, 8–12. [PubMed] [Google Scholar]
  39. Chen Z, Tang Y, Tao H, Li C, Zhang X, Liu Y, 2015. Dynorphin activation of kappa opioid receptor reduces neuronal excitability in the paraventricular nucleus of mouse thalamus. Neuropharmacology 97, 259–269. 10.1016/j.neuropharm.2015.05.030 [DOI] [PubMed] [Google Scholar]
  40. Chuhma N, Tanaka KF, Hen R, Rayport S, 2011. Functional Connectome of the Striatal Medium Spiny Neuron. J. Neurosci 31, 1183–1192. 10.1523/JNEUROSCI.3833-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Civelli O, Douglass J, Goldstein A, Herbert E, 1985. Sequence and expression of the rat prodynorphin gene. PNAS 82, 4291–4295. 10.1073/pnas.82.12.4291 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Cox BM, Goldstein A, Hi CH, 1976. Opioid activity of a peptide, beta-lipotropin-(61–91), derived from beta-lipotropin. PNAS 73, 1821–1823. 10.1073/pnas.73.6.1821 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Cox BM, Opheim KE, Teschemacher H, Goldstein A, 1975. A peptide-like substance from pituitary that acts like morphine. 2. Purification and properties. Life Sci. 16, 1777–1782. 10.1016/0024-3205(75)90272-6 [DOI] [PubMed] [Google Scholar]
  44. Crowley NA, Bloodgood DW, Hardaway JA, Kendra AM, McCall JG, Al-Hasani R, McCall NM, Yu W, Schools ZL, Krashes MJ, Lowell BB, Whistler JL, Bruchas MR, Kash TL, 2016. Dynorphin Controls the Gain of an Amygdalar Anxiety Circuit. Cell Reports 14, 2774–2783. 10.1016/j.celrep.2016.02.069 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. D’Addario C, Caputi FF, Rimondini R, Gandolfi O, Del Borrello E, Candeletti S, Romualdi P, 2013. Different alcohol exposures induce selective alterations on the expression of dynorphin and nociceptin systems related genes in rat brain. Addict Biol 18, 425–433. 10.1111/j.1369-1600.2011.00326.x [DOI] [PubMed] [Google Scholar]
  46. de Laat B, Goldberg A, Shi J, Tetrault JM, Nabulsi N, Zheng M-Q, Najafzadeh S, Gao H, Kapinos M, Ropchan J, O’Malley SS, Huang Y, Morris ED, Krishnan-Sarin S, 2019. The Kappa Opioid Receptor Is Associated With Naltrexone-Induced Reduction of Drinking and Craving. Biol. Psychiatry 86, 864–871. 10.1016/j.biopsych.2019.05.021 [DOI] [PubMed] [Google Scholar]
  47. Diana M, Pistis M, Muntoni A, Rossetti ZL, Gessa G, 1992. Marked decrease of A10 dopamine neuronal firing during ethanol withdrawal syndrome in rats. Eur. J. Pharmacol 221, 403–404. 10.1016/0014-2999(92)90734-l [DOI] [PubMed] [Google Scholar]
  48. DiFeliceantonio AG, Mabrouk OS, Kennedy RT, Berridge KC, 2012. Enkephalin surges in dorsal neostriatum as a signal to eat. Curr. Biol 22, 1918–1924. 10.1016/j.cub.2012.08.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Ding K, Han Y, Seid TW, Buser C, Karigo T, Zhang S, Dickman DK, Anderson DJ, 2019. Imaging neuropeptide release at synapses with a genetically engineered reporter. eLife 8, e46421 10.7554/eLife.46421 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Donahue RJ, Landino SM, Golden SA, Carroll FI, Russo SJ, Carlezon WA, 2015. Effects of acute and chronic social defeat stress are differentially mediated by the dynorphin/kappa-opioid receptor system. Behav Pharmacol 26, 654–663. 10.1097/FBP.0000000000000155 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Erikson CM, Wei G, Walker BM, 2018. Maladaptive behavioral regulation in alcohol dependence: Role of kappa-opioid receptors in the bed nucleus of the stria terminalis. Neuropharmacology 140, 162–173. 10.1016/j.neuropharm.2018.07.034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. España RA, Melchior JR, Roberts DCS, Jones SR, 2011. Hypocretin 1/orexin A in the ventral tegmental area enhances dopamine responses to cocaine and promotes cocaine self-administration. Psychopharmacology (Berl.) 214, 415–426. 10.1007/s00213-010-2048-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. España RA, Oleson EB, Locke JL, Brookshire BR, Roberts DCS, Jones SR, 2010. The hypocretin-orexin system regulates cocaine self-administration via actions on the mesolimbic dopamine system. Eur. J. Neurosci 31, 336–348. 10.1111/j.1460-9568.2009.07065.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Evans CJ, Keith DE, Morrison H, Magendzo K, Edwards RH, 1992. Cloning of a delta opioid receptor by functional expression. Science 258, 1952–1955. 10.1126/science.1335167 [DOI] [PubMed] [Google Scholar]
  55. Fallon JH, Leslie FM, 1986. Distribution of dynorphin and enkephalin peptides in the rat brain. Journal of Comparative Neurology 249, 293–336. 10.1002/cne.902490302 [DOI] [PubMed] [Google Scholar]
  56. Fallon JH, Leslie FM, Cone RI, 1985. Dynorphin-containing pathways in the substantia nigra and ventral tegmentum: A double labeling study using combined immunofluorescence and retrograde tracing. Neuropeptides, Proceedings of the International Narcotic Research Conference 5, 457–460. 10.1016/0143-4179(85)90053-8 [DOI] [PubMed] [Google Scholar]
  57. Ferré G, Czaplicki G, Demange P, Milon A, 2019. Structure and dynamics of dynorphin peptide and its receptor. Vitam. Horm 111, 17–47. 10.1016/bs.vh.2019.05.006 [DOI] [PubMed] [Google Scholar]
  58. Fischli W, Goldstein A, Hunkapiller MW, Hood LE, 1982. Isolation and amino acid sequence analysis of a 4,000-dalton dynorphin from porcine pituitary. Proc. Natl. Acad. Sci. U.S.A 79, 5435–5437. 10.1073/pnas.79.17.5435 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Ford CP, Beckstead MJ, Williams JT, 2007. Kappa opioid inhibition of somatodendritic dopamine inhibitory postsynaptic currents. J. Neurophysiol 97, 883–891. 10.1152/jn.00963.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Ford CP, Mark GP, Williams JT, 2006. Properties and opioid inhibition of mesolimbic dopamine neurons vary according to target location. J. Neurosci 26, 2788–2797. 10.1523/JNEUROSCI.4331-05.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Funk D, Coen K, Tamadon S, Lê AD, 2019. Effect of chronic alcohol vapor exposure on reinstatement of alcohol seeking induced by U50,488. Neuropharmacology 148, 210–219. 10.1016/j.neuropharm.2019.01.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Gallegos RA, Lee RS, Criado JR, Henriksen SJ, Steffensen SC, 1999. Adaptive responses of gamma-aminobutyric acid neurons in the ventral tegmental area to chronic ethanol. J. Pharmacol. Exp. Ther 291, 1045–1053. [PubMed] [Google Scholar]
  63. Gerfen CR, McGinty JF, Young WS, 1991. Dopamine differentially regulates dynorphin, substance P, and enkephalin expression in striatal neurons: in situ hybridization histochemical analysis. J. Neurosci 11, 1016–1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Gobaille S, Schmidt C, Cupo A, Herbrecht F, Maitre M, 1994. Characterization of methionine-enkephalin release in the rat striatum by in vivo dialysis: Effects of gamma-hydroxybutyrate on cellular and extracellular methionine-enkephalin levels. Neuroscience 60, 637–648. 10.1016/0306-4522(94)90492-8 [DOI] [PubMed] [Google Scholar]
  65. Goldstein A, Fischli W, Lowney LI, Hunkapiller M, Hood L, 1981. Porcine pituitary dynorphin: complete amino acid sequence of the biologically active heptadecapeptide. PNAS 78, 7219–7223. 10.1073/pnas.78.11.7219 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Goldstein A, Ghazarossian VE, 1980. Immunoreactive dynorphin in pituitary and brain. Proc. Natl. Acad. Sci. U.S.A 77, 6207–6210. 10.1073/pnas.77.10.6207 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Goldstein A, Lowney LI, Pal BK, 1971. Stereospecific and Nonspecific Interactions of the Morphine Congener Levorphanol in Subcellular Fractions of Mouse Brain. PNAS 68, 1742–1747. 10.1073/pnas.68.8.1742 [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Goldstein A, Tachibana S, Lowney LI, Hunkapiller M, Hood L, 1979. Dynorphin-(1–13), an extraordinarily potent opioid peptide. PNAS 76, 6666–6670. 10.1073/pnas.76.12.6666 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Granholm L, Segerström L, Nylander I, 2018. Episodic Ethanol Exposure in Adolescent Rats Causes Residual Alterations in Endogenous Opioid Peptides. Front Psychiatry 9 10.3389/fpsyt.2018.00425 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Groblewski PA, Zietz C, Willuhn I, Phillips PEM, Chavkin C, 2015. Repeated stress exposure causes strain-dependent shifts in the behavioral economics of cocaine in rats. Addict Biol 20, 297–301. 10.1111/adb.12123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Haskins WE, Wang Z, Watson CJ, Rostand RR, Witowski SR, Powell DH, Kennedy RT, 2001. Capillary LC–MS2 at the Attomole Level for Monitoring and Discovering Endogenous Peptides in Microdialysis Samples Collected in Vivo. Anal. Chem 73, 5005–5014. 10.1021/ac010774d [DOI] [PubMed] [Google Scholar]
  72. Hughes J, Smith TW, Kosterlitz HW, Fothergill LA, Morgan BA, Morris HR, 1975. Identification of two related pentapeptides from the brain with potent opiate agonist activity. Nature 258, 577–579. 10.1038/258577a0 [DOI] [PubMed] [Google Scholar]
  73. Imperato A, Di Chiara G, 1986. Preferential stimulation of dopamine release in the nucleus accumbens of freely moving rats by ethanol. J. Pharmacol. Exp. Ther 239, 219–228. [PubMed] [Google Scholar]
  74. Iyer M, Essner RA, Klingenberg B, Carter ME, 2018. Identification of discrete, intermingled hypocretin neuronal populations. J. Comp. Neurol 526, 2937–2954. 10.1002/cne.24490 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Jadzic D, Bassareo V, Carta AR, Carboni E, 2019. Nicotine, cocaine, amphetamine, morphine, and ethanol increase norepinephrine output in the bed nucleus of stria terminalis of freely moving rats. Addict Biol e12864 10.1111/adb.12864 [DOI] [PubMed] [Google Scholar]
  76. Jamensky NT, Gianoulakis C, 1997. Content of dynorphins and kappa-opioid receptors in distinct brain regions of C57BL/6 and DBA/2 mice. Alcohol. Clin. Exp. Res 21, 1455–1464. [PubMed] [Google Scholar]
  77. James IF, Goldstein A, 1984. Site-directed alkylation of multiple opioid receptors. I. Binding selectivity. Mol Pharmacol 25, 337–342. [PubMed] [Google Scholar]
  78. Jarjour S, Bai L, Gianoulakis C, 2009. Effect of acute ethanol administration on the release of opioid peptides from the midbrain including the ventral tegmental area. Alcohol. Clin. Exp. Res 33, 1033–1043. 10.1111/j.1530-0277.2009.00924.x [DOI] [PubMed] [Google Scholar]
  79. Jarman SK, Haney AM, Valdez GR, 2018. Kappa opioid regulation of depressive-like behavior during acute withdrawal and protracted abstinence from ethanol. PLoS ONE 13, e0205016 10.1371/journal.pone.0205016 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Kakidani H, Furutani Y, Takahashi H, Noda M, Morimoto Y, Hirose T, Asai M, Inayama S, Nakanishi S, Numa S, 1982. Cloning and sequence analysis of cDNA for porcine beta-neo-endorphin/dynorphin precursor. Nature 298, 245–249. 10.1038/298245a0 [DOI] [PubMed] [Google Scholar]
  81. Karkhanis AN, Huggins KN, Rose JH, Jones SR, 2016a. Switch from excitatory to inhibitory actions of ethanol on dopamine levels after chronic exposure: Role of kappa opioid receptors. Neuropharmacology 110, 190–197. 10.1016/j.neuropharm.2016.07.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Karkhanis AN, Locke JL, McCool BA, Weiner JL, Jones SR, 2014. Social isolation rearing increases nucleus accumbens dopamine and norepinephrine responses to acute ethanol in adulthood. Alcohol. Clin. Exp. Res 38, 2770–2779. 10.1111/acer.12555 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Karkhanis AN, Rose JH, Weiner JL, Jones SR, 2016b. Early-Life Social Isolation Stress Increases Kappa Opioid Receptor Responsiveness and Downregulates the Dopamine System. Neuropsychopharmacology 41, 2263–2274. 10.1038/npp.2016.21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Kash TL, 2012. The role of biogenic amine signaling in the bed nucleus of the stria terminals in alcohol abuse. Alcohol 46, 303–308. 10.1016/j.alcohol.2011.12.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Kash TL, Pleil KE, Marcinkiewcz CA, Lowery-Gionta EG, Crowley N, Mazzone C, Sugam J, Hardaway JA, McElligott, and Z.A., 2014. Neuropeptide Regulation of Signaling and Behavior in the BNST. Molecules and Cells 38, 1–13. 10.14348/molcells.2015.2261 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Keistler CR, Hammarlund E, Barker JM, Bond CW, DiLeone RJ, Pittenger C, Taylor JR, 2017. Regulation of Alcohol Extinction and Cue-Induced Reinstatement by Specific Projections among Medial Prefrontal Cortex, Nucleus Accumbens, and Basolateral Amygdala. J Neurosci 37, 4462–4471. 10.1523/JNEUROSCI.3383-16.2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Kendrick KM, 1990. Microdialysis measurement of in vivo neuropeptide release. Journal of Neuroscience Methods 34, 35–46. 10.1016/0165-0270(90)90040-M [DOI] [PubMed] [Google Scholar]
  88. Kieffer BL, Befort K, Gaveriaux-Ruff C, Hirth CG, 1992. The delta-opioid receptor: isolation of a cDNA by expression cloning and pharmacological characterization. PNAS 89, 12048–12052. 10.1073/pnas.89.24.12048 [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kilpatrick DL, Wahlstrom A, Lahm HW, Blacher R, Udenfriend S, 1982. Rimorphin, a unique, naturally occurring [Leu]enkephalin-containing peptide found in association with dynorphin and alpha-neo-endorphin. PNAS 79, 6480–6483. 10.1073/pnas.79.21.6480 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kissler JL, Sirohi S, Reis DJ, Jansen HT, Quock RM, Smith DG, Walker BM, 2014. The One-Two Punch of Alcoholism: Role of Central Amygdala Dynorphins/Kappa-Opioid Receptors. Biological Psychiatry 75, 774–782. 10.1016/j.biopsych.2013.03.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Koob GF, 2019. Neurobiology of Opioid Addiction: Opponent Process, Hyperkatifeia, and Negative Reinforcement. Biol. Psychiatry 10.1016/j.biopsych.2019.05.023 [DOI] [PubMed] [Google Scholar]
  92. Koob GF, 2015. The dark side of emotion: the addiction perspective. Eur. J. Pharmacol 753, 73–87. 10.1016/j.ejphar.2014.11.044 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Laguesse S, Morisot N, Shin JH, Liu F, Adrover MF, Sakhai SA, Lopez MF, Phamluong K, Griffin WC, Becker HC, Bender KJ, Alvarez VA, Ron D, 2017. Prosapip1-Dependent Synaptic Adaptations in the Nucleus Accumbens Drive Alcohol Intake, Seeking, and Reward. Neuron 96, 145–159.e8. 10.1016/j.neuron.2017.08.037 [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Lam MP, Marinelli PW, Bai L, Gianoulakis C, 2008. Effects of acute ethanol on opioid peptide release in the central amygdala: an in vivo microdialysis study. Psychopharmacology (Berl.) 201, 261–271. 10.1007/s00213-008-1267-8 [DOI] [PubMed] [Google Scholar]
  95. Land BB, Bruchas MR, Lemos JC, Xu M, Melief EJ, Chavkin C, 2008. The dysphoric component of stress is encoded by activation of the dynorphin kappa-opioid system. J. Neurosci 28, 407–414. 10.1523/JNEUROSCI.4458-07.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Land BB, Bruchas MR, Schattauer S, Giardino WJ, Aita M, Messinger D, Hnasko TS, Palmiter RD, Chavkin C, 2009. Activation of the kappa opioid receptor in the dorsal raphe nucleus mediates the aversive effects of stress and reinstates drug seeking. Proc. Natl. Acad. Sci. U.S.A 106, 19168–19173. 10.1073/pnas.0910705106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Lê AD, Funk D, Coen K, Tamadon S, Shaham Y, 2018. Role of κ-Opioid Receptors in the Bed Nucleus of Stria Terminalis in Reinstatement of Alcohol Seeking. Neuropsychopharmacology 43, 838–850. 10.1038/npp.2017.120 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Li C, Pleil KE, Stamatakis AM, Busan S, Vong L, Lowell BB, Stuber GD, Kash TL, 2012. Presynaptic Inhibition of Gamma-Aminobutyric Acid Release in the Bed Nucleus of the Stria Terminalis by Kappa Opioid Receptor Signaling. Biological Psychiatry, Cannabis: Mechanisms of Consumption, Craving, and Cognition 71, 725–732. 10.1016/j.biopsych.2011.11.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Li CH, Chung D, 1976. Isolation and structure of an untriakontapeptide with opiate activity from camel pituitary glands. PNAS 73, 1145–1148. 10.1073/pnas.73.4.1145 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Lindholm S, Ploj K, Franck J, Nylander I, 2000. Repeated ethanol administration induces short- and long-term changes in enkephalin and dynorphin tissue concentrations in rat brain. Alcohol 22, 165–171. 10.1016/s0741-8329(00)00118-x [DOI] [PubMed] [Google Scholar]
  101. Lisi TL, Sluka KA, 2006. A new electrochemical HPLC method for analysis of enkephalins and endomorphins. Journal of Neuroscience Methods 150, 74–79. 10.1016/j.jneumeth.2005.06.001 [DOI] [PubMed] [Google Scholar]
  102. Liu X, Weiss F, 2002. Additive effect of stress and drug cues on reinstatement of ethanol seeking: exacerbation by history of dependence and role of concurrent activation of corticotropin-releasing factor and opioid mechanisms. J. Neurosci 22, 7856–7861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Logrip ML, Janak PH, Ron D, 2008. Dynorphin is a downstream effector of striatal BDNF regulation of ethanol intake. FASEB J. 22, 2393–2404. 10.1096/fj.07-099135 [DOI] [PubMed] [Google Scholar]
  104. Lord JAH, Waterfield AA, Hughes J, Kosterlitz HW, 1977. Endogenous opioid peptides: multiple agonists and receptors. Nature 267, 495–499. 10.1038/267495a0 [DOI] [PubMed] [Google Scholar]
  105. Mabrouk OS, Li Q, Song P, Kennedy RT, 2011. Microdialysis and mass spectrometric monitoring of dopamine and enkephalins in the globus pallidus reveal reciprocal interactions that regulate movement. Journal of Neurochemistry 118, 24–33. 10.1111/j.1471-4159.2011.07293.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Maidment NT, Brumbaugh DR, Rudolph VD, Erdelyi E, Evans CJ, 1989. Microdialysis of extracellular endogenous opioid peptides from rat brain in vivo. Neuroscience 33, 549–557. [DOI] [PubMed] [Google Scholar]
  107. Maidment NT, Siddall BJ, Rudolph VR, Erdelyi E, Evans CJ, 1991. Dual determination of extracellular cholecystokinin and neurotensin fragments in rat forebrain: microdialysis combined with a sequential multiple antigen radioimmunoassay. Neuroscience 45, 81–93. [DOI] [PubMed] [Google Scholar]
  108. Marchant NJ, Densmore VS, Osborne PB, 2007. Coexpression of prodynorphin and corticotrophin-releasing hormone in the rat central amygdala: Evidence of two distinct endogenous opioid systems in the lateral division. Journal of Comparative Neurology 504, 702–715. 10.1002/cne.21464 [DOI] [PubMed] [Google Scholar]
  109. Margolis EB, Hjelmstad GO, Bonci A, Fields HL, 2005. Both kappa and mu opioid agonists inhibit glutamatergic input to ventral tegmental area neurons. J. Neurophysiol 93, 3086–3093. 10.1152/jn.00855.2004 [DOI] [PubMed] [Google Scholar]
  110. Margolis EB, Hjelmstad GO, Bonci A, Fields HL, 2003. Kappa-opioid agonists directly inhibit midbrain dopaminergic neurons. J. Neurosci 23, 9981–9986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Margolis EB, Lock H, Chefer VI, Shippenberg TS, Hjelmstad GO, Fields HL, 2006. Kappa opioids selectively control dopaminergic neurons projecting to the prefrontal cortex. Proc. Natl. Acad. Sci. U.S.A 103, 2938–2942. 10.1073/pnas.0511159103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Marinelli PW, Bai L, Quirion R, Gianoulakis C, 2005. A microdialysis profile of Met-enkephalin release in the rat nucleus accumbens following alcohol administration. Alcohol. Clin. Exp. Res 29, 1821–1828. 10.1097/01.alc.0000183008.62955.2e [DOI] [PubMed] [Google Scholar]
  113. Marinelli PW, Lam M, Bai L, Quirion R, Gianoulakis C, 2006. A microdialysis profile of dynorphin A(1–8) release in the rat nucleus accumbens following alcohol administration. Alcohol. Clin. Exp. Res 30, 982–990. 10.1111/j.1530-0277.2006.00112.x [DOI] [PubMed] [Google Scholar]
  114. Martin WR, Eades CG, Thompson JA, Huppler RE, Gilbert PE, 1976. The effects of morphine- and nalorphine- like drugs in the nondependent and morphine-dependent chronic spinal dog. J. Pharmacol. Exp. Ther 197, 517–532. [PubMed] [Google Scholar]
  115. McCaul ME, Hutton HE, Stephens MAC, Xu X, Wand GS, 2017. Anxiety, Anxiety Sensitivity, and Perceived Stress as Predictors of Recent Drinking, Alcohol Craving, and Social Stress Response in Heavy Drinkers. Alcohol. Clin. Exp. Res 41, 836–845. 10.1111/acer.13350 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. McLaughlin JP, Land BB, Li S, Pintar JE, Chavkin C, 2006. Prior activation of kappa opioid receptors by U50,488 mimics repeated forced swim stress to potentiate cocaine place preference conditioning. Neuropsychopharmacology 31, 787–794. 10.1038/sj.npp.1300860 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Molineaux CJ, Cox BM, 1982. Subcellular localization of immunoreactive dynorphin and vasopressin in rat pituitary and hypothalmus. Life Sciences 31, 1765–1768. 10.1016/0024-3205(82)90205-3 [DOI] [PubMed] [Google Scholar]
  118. Mollereau C, Parmentier M, Mailleux P, Butour JL, Moisand C, Chalon P, Caput D, Vassart G, Meunier JC, 1994. ORL1, a novel member of the opioid receptor family. Cloning, functional expression and localization. FEBS Lett. 341, 33–38. 10.1016/0014-5793(94)80235-1 [DOI] [PubMed] [Google Scholar]
  119. Morales M, Anderson RI, Spear LP, Varlinskaya EI, 2014. Effects of the kappa opioid receptor antagonist, nor-binaltorphimine, on ethanol intake: impact of age and sex. Dev Psychobiol 56, 700–712. 10.1002/dev.21137 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Muschamp JW, Carlezon WA, 2013. Roles of nucleus accumbens CREB and dynorphin in dysregulation of motivation. Cold Spring Harb Perspect Med 3, a012005 10.1101/cshperspect.a012005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Muschamp JW, Hollander JA, Thompson JL, Voren G, Hassinger LC, Onvani S, Kamenecka TM, Borgland SL, Kenny PJ, Carlezon WA, 2014. Hypocretin (orexin) facilitates reward by attenuating the antireward effects of its cotransmitter dynorphin in ventral tegmental area. Proc. Natl. Acad. Sci. U.S.A 111, E1648–1655. 10.1073/pnas.1315542111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Negus SS, 2004. Effects of the kappa opioid agonist U50,488 and the kappa opioid antagonist norbinaltorphimine on choice between cocaine and food in rhesus monkeys. Psychopharmacology (Berl.) 176, 204–213. 10.1007/s00213-004-1878-7 [DOI] [PubMed] [Google Scholar]
  123. Nieto MM, Wilson J, Cupo A, Roques BP, Noble F, 2002. Chronic Morphine Treatment Modulates the Extracellular Levels of Endogenous Enkephalins in Rat Brain Structures Involved in Opiate Dependence: A Microdialysis Study. J. Neurosci 22, 1034–1041. 10.1523/JNEUROSCI.22-03-01034.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Normandeau CP, Torruella Suárez ML, Sarret P, McElligott ZA, Dumont EC, 2018. Neurotensin and dynorphin Bi-Directionally modulate CeA inhibition of oval BNST neurons in male mice. Neuropharmacology 143, 113–121. 10.1016/j.neuropharm.2018.09.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Nylander I, Hyytiä P, Forsander O, Terenius L, 1994. Differences between alcohol-preferring (AA) and alcohol-avoiding (ANA) rats in the prodynorphin and proenkephalin systems. Alcohol. Clin. Exp. Res 18, 1272–1279. 10.1111/j.1530-0277.1994.tb00118.x [DOI] [PubMed] [Google Scholar]
  126. Nylander I, Vlaskovska M, Terenius L, 1995a. Brain dynorphin and enkephalin systems in Fischer and Lewis rats: effects of morphine tolerance and withdrawal. Brain Res. 683, 25–35. 10.1016/0006-8993(95)00279-y [DOI] [PubMed] [Google Scholar]
  127. Nylander I, Vlaskovska M, Terenius L, 1995b. The effects of morphine treatment and morphine withdrawal on the dynorphin and enkephalin systems in Sprague-Dawley rats. Psychopharmacology (Berl.) 118, 391–400. [DOI] [PubMed] [Google Scholar]
  128. Palm S, Nylander I, 2014. Alcohol-Induced Changes in Opioid Peptide Levels in Adolescent Rats Are Dependent on Housing Conditions. Alcohol Clin Exp Res 38, 2978–2987. 10.1111/acer.12586 [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Pati D, Pina MM, Kash TL, 2019. Ethanol-induced conditioned place preference and aversion differentially alter plasticity in the bed nucleus of stria terminalis. Neuropsychopharmacology 44, 1843–1854. 10.1038/s41386-019-0349-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Patriarchi T, Cho JR, Merten K, Howe MW, Marley A, Xiong W-H, Folk RW, Broussard GJ, Liang R, Jang MJ, Zhong H, Dombeck D, Zastrow M. von, Nimmerjahn A, Gradinaru V, Williams JT, Tian L, 2018. Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors. Science 360, eaat4422 10.1126/science.aat4422 [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Patterson CM, Wong J-MT, Leinninger GM, Allison MB, Mabrouk OS, Kasper CL, Gonzalez IE, Mackenzie A, Jones JC, Kennedy RT, Myers MG, 2015. Ventral Tegmental Area Neurotensin Signaling Links the Lateral Hypothalamus to Locomotor Activity and Striatal Dopamine Efflux in Male Mice. Endocrinology 156, 1692–1700. 10.1210/en.2014-1986 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Pickel VM, Chan J, Sesack SR, 1993. Cellular substrates for interactions between dynorphin terminals and dopamine dendrites in rat ventral tegmental area and substantia nigra. Brain Research 602, 275–289. 10.1016/0006-8993(93)90693-H [DOI] [PubMed] [Google Scholar]
  133. Pina MM, Young EA, Ryabinin AE, Cunningham CL, 2015. The bed nucleus of the stria terminalis regulates ethanol-seeking behavior in mice. Neuropharmacology 99, 627–638. 10.1016/j.neuropharm.2015.08.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Ploj K, Roman E, Gustavsson L, Nylander I, 2000. Basal levels and alcohol-induced changes in nociceptin/orphanin FQ, dynorphin, and enkephalin levels in C57BL/6J mice. Brain Res. Bull 53, 219–226. 10.1016/s0361-9230(00)00328-2 [DOI] [PubMed] [Google Scholar]
  135. Portoghese PS, Larson DL, Jiang JB, Takemori AE, Caruso TP, 1978. 6.beta.-[N,N-Bis(2-chloroethyl)amino]-17-(cyclopropylmethyl)-4,5.alpha.-epoxy-3,14-dihydroxymorphinan (chloranaltrexamine), a potent opioid receptor alkylating agent with ultralong narcotic antagonist activity. J. Med. Chem 21, 598–599. 10.1021/jm00205a002 [DOI] [PubMed] [Google Scholar]
  136. Poulin J-F, Arbour D, Laforest S, Drolet G, 2009. Neuroanatomical characterization of endogenous opioids in the bed nucleus of the stria terminalis. Progress in Neuro-Psychopharmacology and Biological Psychiatry 33, 1356–1365. 10.1016/j.pnpbp.2009.06.021 [DOI] [PubMed] [Google Scholar]
  137. Przewłocka B, Turchan J, Lasoń W, Przewłocki R, 1997. Ethanol withdrawal enhances the prodynorphin system activity in the rat nucleus accumbens. Neurosci. Lett 238, 13–16. 10.1016/s0304-3940(97)00829-x [DOI] [PubMed] [Google Scholar]
  138. Rácz I, Markert A, Mauer D, Stoffel-Wagner B, Zimmer A, 2013. Long-term ethanol effects on acute stress responses: modulation by dynorphin. Addict Biol 18, 678–688. 10.1111/j.1369-1600.2012.00494.x [DOI] [PubMed] [Google Scholar]
  139. Redila VA, Chavkin C, 2008. Stress-induced reinstatement of cocaine seeking is mediated by the kappa opioid system. Psychopharmacology (Berl.) 200, 59–70. 10.1007/s00213-008-1122-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Resendez SL, Kuhnmuench M, Krzywosinski T, Aragona BJ, 2012. κ-Opioid receptors within the nucleus accumbens shell mediate pair bond maintenance. J. Neurosci 32, 6771–6784. 10.1523/JNEUROSCI.5779-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Richardson KA, Aston-Jones G, 2012. Lateral hypothalamic orexin/hypocretin neurons that project to ventral tegmental area are differentially activated with morphine preference. J. Neurosci 32, 3809–3817. 10.1523/JNEUROSCI.3917-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Rius RA, Barg J, Bem WT, Coscia CJ, Loh YP, 1991. The prenatal developmental profile of expression of opioid peptides and receptors in the mouse brain. Brain Res Dev Brain Res 58, 237–241. 10.1016/0165-3806(91)90010-G [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Rodd ZA, Melendez RI, Bell RL, Kuc KA, Zhang Y, Murphy JM, McBride WJ, 2004. Intracranial self-administration of ethanol within the ventral tegmental area of male Wistar rats: evidence for involvement of dopamine neurons. J. Neurosci 24, 1050–1057. 10.1523/JNEUROSCI.1319-03.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Roman E, Ploj K, Gustafsson L, Meyerson BJ, Nylander I, 2006. Variations in opioid peptide levels during the estrous cycle in Sprague-Dawley rats. Neuropeptides 40, 195–206. 10.1016/j.npep.2006.01.004 [DOI] [PubMed] [Google Scholar]
  145. Rose JH, Karkhanis AN, Chen R, Gioia D, Lopez MF, Becker HC, McCool BA, Jones SR, 2016. Supersensitive Kappa Opioid Receptors Promotes Ethanol Withdrawal-Related Behaviors and Reduce Dopamine Signaling in the Nucleus Accumbens. Int. J. Neuropsychopharmacol 19 10.1093/ijnp/pyv127 [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Russell SE, Rachlin AB, Smith KL, Muschamp J, Berry L, Zhao Z, Chartoff EH, 2014. Sex differences in sensitivity to the depressive-like effects of the kappa opioid receptor agonist U-50488 in rats. Biol. Psychiatry 76, 213–222. 10.1016/j.biopsych.2013.07.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Saal D, Dong Y, Bonci A, Malenka RC, 2003. Drugs of Abuse and Stress Trigger a Common Synaptic Adaptation in Dopamine Neurons. Neuron 37, 577–582. 10.1016/S0896-6273(03)00021-7 [DOI] [PubMed] [Google Scholar]
  148. Schindler AG, Messinger DI, Smith JS, Shankar H, Gustin RM, Schattauer SS, Lemos JC, Chavkin NW, Hagan CE, Neumaier JF, Chavkin C, 2012. Stress Produces Aversion and Potentiates Cocaine Reward by Releasing Endogenous Dynorphins in the Ventral Striatum to Locally Stimulate Serotonin Reuptake. J. Neurosci 32, 17582–17596. 10.1523/JNEUROSCI.3220-12.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Schwarzer C, 2009. 30 years of dynorphins--new insights on their functions in neuropsychiatric diseases. Pharmacol. Ther 123, 353–370. 10.1016/j.pharmthera.2009.05.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Sershen H, Hashim A, Lajtha A, 1998. Gender differences in kappa-opioid modulation of cocaine-induced behavior and NMDA-evoked dopamine release. Brain Res. 801, 67–71. 10.1016/s0006-8993(98)00546-0 [DOI] [PubMed] [Google Scholar]
  151. Shen H, Lada MW, Kennedy RT, 1997. Monitoring of met-enkephalin in vivo with 5-min temporal resolution using microdialysis sampling and capillary liquid chromatography with electrochemical detection. Journal of Chromatography B: Biomedical Sciences and Applications 704, 43–52. 10.1016/S0378-4347(97)00436-2 [DOI] [PubMed] [Google Scholar]
  152. Shen R-Y, Chiodo LA, 1993. Acute withdrawal after repeated ethanol treatment reduces the number of spontaneously active dopaminergic neurons in the ventral tegmental area. Brain Research 622, 289–293. 10.1016/0006-8993(93)90831-7 [DOI] [PubMed] [Google Scholar]
  153. Shirayama Y, Ishida H, Iwata M, Hazama G, Kawahara R, Duman RS, 2004. Stress increases dynorphin immunoreactivity in limbic brain regions and dynorphin antagonism produces antidepressant-like effects. Journal of Neurochemistry 90, 1258–1268. 10.1111/j.1471-4159.2004.02589.x [DOI] [PubMed] [Google Scholar]
  154. Shurman J, Koob GF, Gutstein HB, 2010. Opioids, pain, the brain, and hyperkatifeia: a framework for the rational use of opioids for pain. Pain Med 11, 1092–1098. 10.1111/j.1526-4637.2010.00881.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Siciliano CA, Calipari ES, Yorgason JT, Lovinger DM, Mateo Y, Jimenez VA, Helms CM, Grant KA, Jones SR, 2016a. Increased presynaptic regulation of dopamine neurotransmission in the nucleus accumbens core following chronic ethanol self-administration in female macaques. Psychopharmacology (Berl.) 233, 1435–1443. 10.1007/s00213-016-4239-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Siciliano CA, Calipari ES, Yorgason JT, Mateo Y, Helms CM, Lovinger DM, Grant KA, Jones SR, 2016b. Chronic ethanol self-administration in macaques shifts dopamine feedback inhibition to predominantly D2 receptors in nucleus accumbens core. Drug Alcohol Depend 158, 159–163. 10.1016/j.drugalcdep.2015.10.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Siciliano CA, Karkhanis AN, Holleran KM, Melchior JR, Jones SR, 2018. Cross-Species Alterations in Synaptic Dopamine Regulation After Chronic Alcohol Exposure. Handb Exp Pharmacol 248, 213–238. 10.1007/164_2018_106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Smith RJ, Anderson RI, Haun HL, Mulholland PJ, Griffin WC, Lopez MF, Becker HC, 2019. Dynamic c-Fos changes in mouse brain during acute and protracted withdrawal from chronic intermittent ethanol exposure and relapse drinking. Addict Biol e12804 10.1111/adb.12804 [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Spear LP, 2015. ADOLESCENT ALCOHOL EXPOSURE: ARE THERE SEPARABLE VULNERABLE PERIODS WITHIN ADOLESCENCE? Physiol Behav 148, 122–130. 10.1016/j.physbeh.2015.01.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Steffensen SC, Walton CH, Hansen DM, Yorgason JT, Gallegos RA, Criado JR, 2009. Contingent and non-contingent effects of low-dose ethanol on GABA neuron activity in the ventral tegmental area. Pharmacol. Biochem. Behav 92, 68–75. 10.1016/j.pbb.2008.10.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Strine TW, Dube SR, Edwards VJ, Prehn AW, Rasmussen S, Wagenfeld M, Dhingra S, Croft JB, 2012. Associations between adverse childhood experiences, psychological distress, and adult alcohol problems. Am J Health Behav 36, 408–423. 10.5993/AJHB.36.3.11 [DOI] [PubMed] [Google Scholar]
  162. Svingos AL, Chavkin C, Colago EE, Pickel VM, 2001. Major coexpression of kappa-opioid receptors and the dopamine transporter in nucleus accumbens axonal profiles. Synapse 42, 185–192. 10.1002/syn.10005 [DOI] [PubMed] [Google Scholar]
  163. Tao R, Auerbach SB, 2002. Opioid receptor subtypes differentially modulate serotonin efflux in the rat central nervous system. J. Pharmacol. Exp. Ther 303, 549–556. 10.1124/jpet.102.037861 [DOI] [PubMed] [Google Scholar]
  164. Tejeda HA, Wu J, Kornspun AR, Pignatelli M, Kashtelyan V, Krashes MJ, Lowell BB, Carlezon WA, Bonci A, 2017. Pathway- and Cell-Specific Kappa-Opioid Receptor Modulation of Excitation-Inhibition Balance Differentially Gates D1 and D2 Accumbens Neuron Activity. Neuron 93, 147–163. 10.1016/j.neuron.2016.12.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Teschemacher H, Opheim KE, Cox BM, Goldstein A, 1975. A peptide-like substance from pituitary that acts like morphine. I. Isolation. Life Sci 16, 1771–1775. 10.1016/0024-3205(75)90271-4 [DOI] [PubMed] [Google Scholar]
  166. Thompson AC, Zapata A, Justice JB, Vaughan RA, Sharpe LG, Shippenberg TS, 2000. Kappa-opioid receptor activation modifies dopamine uptake in the nucleus accumbens and opposes the effects of cocaine. J. Neurosci 20, 9333–9340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Toalston JE, Deehan GA, Hauser SR, Engleman EA, Bell RL, Murphy JM, Truitt WA, McBride WJ, Rodd ZA, 2014. Reinforcing Properties and Neurochemical Response of Ethanol within the Posterior Ventral Tegmental Area Are Enhanced in Adulthood by Periadolescent Ethanol Consumption. J Pharmacol Exp Ther 351, 317–326. 10.1124/jpet.114.218172 [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Tperenius L, Wahlström A, 1975. Morphine-like ligand for opiate receptors in human CSF. Life Sci. 16, 1759–1764. 10.1016/0024-3205(75)90269-6 [DOI] [PubMed] [Google Scholar]
  169. Van Bockstaele EJ, Gracy KN, Pickel VM, 1995. Dynorphin-immunoreactive neurons in the rat nucleus accumbens: ultrastructure and synaptic input from terminals containing substance P and/or dynorphin. J. Comp. Neurol 351, 117–133. 10.1002/cne.903510111 [DOI] [PubMed] [Google Scholar]
  170. Vijay A, Wang S, Worhunsky P, Zheng M-Q, Nabulsi N, Ropchan J, Krishnan-Sarin S, Huang Y, Morris ED, 2016. PET imaging reveals sex differences in kappa opioid receptor availability in humans, in vivo. Am J Nucl Med Mol Imaging 6, 205–214. [PMC free article] [PubMed] [Google Scholar]
  171. Volkow ND, Fowler JS, Wang GJ, Baler R, Telang F, 2009. Imaging dopamine’s role in drug abuse and addiction. Neuropharmacology 56, 3–8. 10.1016/j.neuropharm.2008.05.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Wagner JJ, Evans CJ, Chavkin C, 1991. Focal Stimulation of the Mossy Fibers Releases Endogenous Dynorphins That Bind K1-Opioid Receptors in Guinea Pig Hippocampus. Journal of Neurochemistry 57, 333–343. 10.1111/j.1471-4159.1991.tb02132.x [DOI] [PubMed] [Google Scholar]
  173. Wagner JJ, Terman GW, Chavkin C, 1993. Endogenous dynorphins inhibit excitatory neurotransmission and block LTP induction in the hippocampus. Nature 363, 451–454. 10.1038/363451a0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Walker BM, Koob GF, 2008. Pharmacological evidence for a motivational role of kappa-opioid systems in ethanol dependence. Neuropsychopharmacology 33, 643–652. 10.1038/sj.npp.1301438 [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Wang JB, Johnson PS, Wu JM, Wang WF, Uhl GR, 1994. Human kappa opiate receptor second extracellular loop elevates dynorphin’s affinity for human mu/kappa chimeras. J. Biol. Chem 269, 25966–25969. [PubMed] [Google Scholar]
  176. Watabe-Uchida M, Zhu L, Ogawa SK, Vamanrao A, Uchida N, 2012. Whole-Brain Mapping of Direct Inputs to Midbrain Dopamine Neurons. Neuron 74, 858–873. 10.1016/j.neuron.2012.03.017 [DOI] [PubMed] [Google Scholar]
  177. Werling LL, Frattali A, Portoghese PS, Takemori AE, Cox BM, 1988. Kappa receptor regulation of dopamine release from striatum and cortex of rats and guinea pigs. J. Pharmacol. Exp. Ther 246, 282–286. [PubMed] [Google Scholar]
  178. Wille-Bille A, Miranda-Morales RS, Pucci M, Bellia F, D’Addario C, Pautassi RM, 2018. Prenatal ethanol induces an anxiety phenotype and alters expression of dynorphin & nociceptin/orphanin FQ genes. Prog. Neuropsychopharmacol. Biol. Psychiatry 85, 77–88. 10.1016/j.pnpbp.2018.04.005 [DOI] [PubMed] [Google Scholar]
  179. Williams SB, Yorgason JT, Nelson AC, Lewis N, Nufer TM, Edwards JG, Steffensen SC, 2018. Glutamate Transmission to Ventral Tegmental Area GABA Neurons Is Altered by Acute and Chronic Ethanol. Alcohol Clin Exp Res 42, 2186–2195. 10.1111/acer.13883 [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. World Health Organization, 2014, Global status report on alcohol and health. WHO Press, Geneva, Switzerland. [Google Scholar]
  181. Wu H, Wacker D, Mileni M, Katritch V, Han GW, Vardy E, Liu W, Thompson AA, Huang X-P, Carroll FI, Mascarella SW, Westkaemper RB, Mosier PD, Roth BL, Cherezov V, Stevens RC, 2012. Structure of the human κ-opioid receptor in complex with JDTic. Nature 485, 327–332. 10.1038/nature10939 [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Xia Y, Driscoll JR, Wilbrecht L, Margolis EB, Fields HL, Hjelmstad GO, 2011. Nucleus Accumbens Medium Spiny Neurons Target Non-Dopaminergic Neurons in the Ventral Tegmental Area. J. Neurosci 31, 7811–7816. 10.1523/JNEUROSCI.1504-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Yakovleva T, Bazov I, Cebers G, Marinova Z, Hara Y, Ahmed A, Vlaskovska M, Johansson B, Hochgeschwender U, Singh IN, Bruce-Keller AJ, Hurd YL, Kaneko T, Terenius L, Ekström TJ, Hauser KF, Pickel VM, Bakalkin G, 2006. Prodynorphin storage and processing in axon terminals and dendrites. FASEB J. 20, 2124–2126. 10.1096/fj.06-6174fje [DOI] [PubMed] [Google Scholar]
  184. Yang H, de Jong JW, Tak Y, Peck J, Bateup HS, Lammel S, 2018. Nucleus Accumbens Subnuclei Regulate Motivated Behavior via Direct Inhibition and Disinhibition of VTA Dopamine Subpopulations. Neuron 97, 434–449.e4. 10.1016/j.neuron.2017.12.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Yasuda K, Raynor K, Kong H, Breder CD, Takeda J, Reisine T, Bell GI, 1993. Cloning and functional comparison of kappa and delta opioid receptors from mouse brain. PNAS 90, 6736–6740. 10.1073/pnas.90.14.6736 [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Yim HJ, Robinson DL, White ML, Jaworski JN, Randall PK, Lancaster FE, Gonzales RA, 2000. Dissociation Between the Time Course of Ethanol and Extracellular Dopamine Concentrations in the Nucleus Accumbens After a Single Intraperitoneal Injection. Alcoholism: Clinical and Experimental Research 24, 781–788. 10.1111/j.1530-0277.2000.tb02056.x [DOI] [PubMed] [Google Scholar]
  187. Yorgason JT, Ferris MJ, Steffensen SC, Jones SR, 2014. Frequency-dependent effects of ethanol on dopamine release in the nucleus accumbens. Alcohol. Clin. Exp. Res 38, 438–447. 10.1111/acer.12287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Yorgason JT, Rose JH, McIntosh JM, Ferris MJ, Jones SR, 2015. Greater ethanol inhibition of presynaptic dopamine release in C57BL/6J than DBA/2J mice: Role of nicotinic acetylcholine receptors. Neuroscience 284, 854–864. 10.1016/j.neuroscience.2014.10.052 [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Zhou Y, Colombo G, Gessa GL, Kreek MJ, 2013. Effects of voluntary alcohol drinking on corticotropin-releasing factor and preprodynorphin mRNA levels in the central amygdala of Sardinian alcohol-preferring rats. Neurosci. Lett 554, 110–114. 10.1016/j.neulet.2013.08.071 [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Zhou Y, Crowley RS, Ben K, Prisinzano TE, Kreek MJ, 2017. Synergistic blockade of alcohol escalation drinking in mice by a combination of novel kappa opioid receptor agonist Mesyl Salvinorin B and naltrexone. Brain Res. 1662, 75–86. 10.1016/j.brainres.2017.02.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Zhou Y, Kreek MJ, 2019a. Combination of Clinically Utilized Kappa-Opioid Receptor Agonist Nalfurafine With Low-Dose Naltrexone Reduces Excessive Alcohol Drinking in Male and Female Mice. Alcohol. Clin. Exp. Res 43, 1077–1090. 10.1111/acer.14033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Zhou Y, Kreek MJ, 2019b. Clinically utilized kappa-opioid receptor agonist nalfurafine combined with low-dose naltrexone prevents alcohol relapse-like drinking in male and female mice. Brain Res. 1724, 146410 10.1016/j.brainres.2019.146410 [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Zhou Y, Wong J-MT, Mabrouk OS, Kennedy RT, 2015. Reducing adsorption to improve recovery and in vivo detection of neuropeptides by microdialysis with LC-MS. Anal. Chem 87, 9802–9809. 10.1021/acs.analchem.5b02086 [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES