Skip to main content
Full access
Reviews and Overviews
Published Online: 1 March 2020

Translating Across Circuits and Genetics Toward Progress in Fear- and Anxiety-Related Disorders

Abstract

Anxiety and fear-related disorders are common and disabling, and they significantly increase risk for suicide and other causes of morbidity and mortality. However, there is tremendous potential for translational neuroscience to advance our understanding of these disorders, leading to novel and powerful interventions and even to preventing their initial development. This overview examines the general circuits and processes thought to underlie fear and anxiety, along with the promise of translational research. It then examines some of the data-driven “next-generation” approaches that are needed for discovery and understanding but that do not always fit neatly into established models. From one perspective, these disorders offer among the most tractable problems in psychiatry, with a great deal of accumulated understanding, across species, of neurocircuit, behavioral, and, increasingly, genetic mechanisms, of how dysregulation of fear and threat processes contributes to anxiety-related disorders. One example is the progressively sophisticated understanding of how extinction underlies the exposure therapy component of cognitive-behavioral therapy approaches, which are ubiquitously used across anxiety and fear-related disorders. However, it is also critical to examine gaps in our understanding between reasonably well-replicated examples of successful translation, areas of significant deficits in knowledge, and the role of large-scale data-driven approaches in future progress and discovery. Although a tremendous amount of progress is still needed, translational approaches to understanding, treating, and even preventing anxiety and fear-related disorders offer great opportunities for successfully bridging neuroscience discovery to clinical practice.
Anxiety, posttraumatic stress disorder (PTSD), and other trauma- and fear-related disorders are debilitating and common, are important causes of suicide, and cost society tens of billions of dollars. Despite their costs, morbidity, and mortality, these disorders may be among the most tractable of psychiatric illnesses from a neuroscience perspective, for several reasons: 1) the shared neural circuitry underlying threat processing across mammals, which allows translational approaches; 2) major advances in the neurobiology of threat and neural plasticity over recent decades; 3) the ability to examine shared intermediate phenotypes of fear processing across species; 4) the shared behavioral and neural mechanisms that underlie extinction and exposure-based therapies; and 5) in at least some cases, such as PTSD, a specific causal event for illness onset (e.g., exposure to an external trauma) that can be utilized for research and intervention. Nonetheless, established new treatments for these disorders have been slow to arise from this knowledge. Modern psychiatry seeks to expand our understanding of the biological mechanisms of fear-, stress-, and trauma-related disorders with the goal of identifying targeted, rationally designed interventions for these devastating disorders, based on advances in biology and neuroscience.
To fully understand fear- and anxiety-related disorders, we need to know what behaviors are affected, which brain regions are involved, and which genes, molecules, and cell types might be involved. The prefrontal cortex, the hippocampus, the insula, and the amygdala are often associated with fear and threat processing (14). In particular, different amygdala subregions and specific cell populations are being determined to have distinct roles in fear and extinction processing (512). Furthermore, large-scale international consortia and collaborations have allowed genome-wide association studies (GWASs) of hundreds of thousands of individuals with PTSD and anxiety disorders, confirming their heritability and shared polygenic risk with other severe psychiatric disorders, as well as pointing to novel genes and pathways while also confirming their association with well-known molecular pathways (1316).
We are at a unique time in history, having an embarrassment of riches with new tools for understanding dynamic human brain activity, with powerful combinatorial machine learning approaches, and with enormous biological data sets, from hundreds of thousands of subjects with genome-wide genomics data to single-cell sequencing on millions of cells across brain regions and with causal tools for precise neural circuit manipulation. It is hoped that such promising new approaches will allow integration of amygdala-based microcircuits of threat processing across animal model systems (1, 2) with postmortem human amygdala (and other related circuits) tissue (e.g., PsychENCODE) (17) and with large-scale consortium initiatives, such as the Psychiatric Genomics Consortium for PTSD (18), the Army Study to Assess Risk and Resilience in Servicemembers (Army STARRS) (19), the Advancing Understanding of Recovery After Trauma (AURORA) Study (20), and the Million Veteran Program (21), discussed below. Such progress would allow integration of cell-type-specific regulation of behavior with genomic and biomarker discovery samples from human cohorts.
What is not yet clear is how to integrate the avalanche of new data and findings with our current understanding of the neural circuits of fear and threat processing. Furthermore, how are these circuits functionally and even structurally altered by interventions, from exposure therapy to targeted neurotherapeutics to neurobiologically targeted pharmacology? Such future approaches may allow for rational identification and testing of novel target compounds and biologicals as potential interventions, which can then be efficiently brought from bench to bedside, leading to the next generation of more powerful treatments for PTSD, anxiety, and other fear- and trauma-related disorders.

Intersection Of Neural Circuitry Of Threat And Extinction In Fear- And Anxiety-Related Disorders

Much of the translational progress in understanding mechanisms of anxiety- and fear-related disorders comes from the intersection of decades of work in understanding the biology of mammalian fear or threat conditioning, based on more than a century of work since Pavlov initially worked out the basic rules underlying classical conditioning theory (22). In the past few decades, groundbreaking work, primarily in rodents, has informed our understanding of the role of the amygdala, modulated by the medial prefrontal cortex (mPFC), hippocampus, thalamus, and insula, among other structures, in underlying the neural mechanisms of behavioral Pavlovian threat conditioning (4, 5, 7, 9, 2327). Importantly, this work and the principal conserved brain regions involved have been well replicated across species in nonhuman primates (2831). Finally, many studies have demonstrated that Pavlovian fear conditioning and its correlated neural mechanisms are dysregulated in human anxiety- and fear-related disorders such as PTSD (3241).
The past decade of progress has focused on more precise dissection of the behavioral, neural, and genetic components underlying fear-related disorders. Behaviorally, it is important to note that fear memory itself can be broken down further into its component parts. During the minutes to hours (to possibly days) following trauma exposure, the fear/trauma memory remains in a labile state, called the consolidation period (4244). A number of exciting areas of inquiry suggest that new pharmacotherapeutic and psychotherapeutic approaches can be initiated that may inhibit the emotional component of fear memory consolidation without markedly affecting the explicit memory formation (4548).
After consolidation of the initial aversive memory, multiple behavioral and cognitive pathways (with component underlying neural circuitry) contribute to healthy recovery from the initial fear or trauma memory (even if the initial triggering events are unknown, as may be the case in panic, obsessive-compulsive disorder [OCD], and phobias) (2, 46, 47, 49, 50). Processes such as sensitization on reexposure to cues, generalization of context and cue triggers, avoidance of reminder triggers, reinstatement following new threats, and renewal of fear and anxiety behaviors in different contexts following treatment are all associated with increased anxiety- and fear-related symptoms and further comorbidity. In contrast, extinction of fear, discrimination of cues and contexts, and actively facing reminder cues and contexts (allowing natural extinction to occur) are all processes that support recovery. We hope that by understanding multiple components of fear memory modulation in humans and mice, a number of novel, powerful, and targeted treatment and intervention approaches may become possible (46, 47, 51).

Extinction Of Fear As An Example Of Translational Success

The symptoms of PTSD, panic, and other fear- and anxiety-related disorders can be explained, at least in part, as an inability to inhibit learned fear even during times and contexts that are safe. Notably, multiple studies have demonstrated that fear inhibition or extinction is impaired across PTSD and other anxiety-related disorders (23, 35, 3941, 5259). Some of the earliest reports of naturalistic observation following trauma exposure suggested that a failure to naturally recover or to learn to inhibit trauma-based fear reactions is central to these disorders (60). Other work has demonstrated that amygdala activation is high in individuals with PTSD and that there are delays in inhibition or extinction of fear. Thus, deficits in inhibiting or extinguishing fear are considered a crucial central intermediate phenotype in PTSD and related anxiety- and fear-related disorders.
The observation of failure to extinguish as an important construct in fear- and anxiety-related disorders is exciting for several reasons. From a clinical perspective, exposure-based therapy techniques—which are a central element of cognitive-behavioral therapy (CBT) and its many more specific versions, including cognitive processing therapy, eye movement desensitization and reprocessing, skill training in affective regulation, and of course prolonged exposure therapy—are among the best empirically supported treatments across the range of fear- and anxiety-related disorders. Increasingly, the elements that underscore the emotional learning process of extinction of threat behaviors are appreciated to be the same elements underlying the efficacy of exposure to feared cues and contexts. This provides for a relatively straightforward target for intervention, in which neural circuitry and behavioral psychology intersect in a predictable, reliable, and robust manner.

The Promise Of Fear- And Anxiety Disorders As Translational “Wins” For Psychiatry

Notably, the Research Domain Criteria project (RDoC) of the National Institute of Mental Health (61) chose “negative valence systems” and, specifically, the constructs of acute threat (fear), potential threat (anxiety), and sustained threat (chronic stress) as among the first RDoC templates. This was in part due to the progress in understanding the neurocircuitry of fear, threat, and aversive behaviors across levels of analyses and model systems. The translation of extinction of fear across species and interventions provides a great example.
Pavlov initially defined extinction of conditioned fear in his classic studies (22). Decades later, historical leaders in the field such as Rescorla and Bouton provided behavioral evidence that extinction of a prior fear memory did not lead to its forgetting or erasure, but rather to an apparent new memory trace representing fear memory inhibition (6265). In the early 1990s, Michael Davis’s lab at Yale University initially demonstrated the critical observation that extinction of fear required active new amygdala-dependent synaptic plasticity (66, 67). In the past two decades, a considerable amount of additional research has examined cellular and molecular mechanisms of fear inhibition and extinction. Substantial data now support a variety of mechanisms underlying synaptic plasticity as critical for new extinction learning. For example, brain-derived neurotrophic factor (BDNF) and its receptor TrkB have been shown to be required for memory formation related to fear and its extinction (6871). Furthermore, following the demonstration that extinction memories required N-methyl-d-aspartate (NMDA)–dependent plasticity, we demonstrated that a partial agonist at the NMDA receptor, d-cycloserine, enhances the consolidation of extinction of fear in rodents (72) and began to translate this finding to enhancement of extinction paradigms in humans (73). This finding has now been replicated repeatedly in rodents and humans (7481).
Notably, additional studies that failed to replicate the d-cycloserine effect, along with a number of more recent negative findings, have now begun to be better understood with the observation that NMDA-dependent plasticity enhances both memory reconsolidation and memory extinction (43, 82, 83). These two opposing processes of strengthening versus inhibiting preexisting memory traces always appear to be at odds on reactivation of a prior fearful memory. This is particularly interesting with clinical translation, given the long-standing observation that reexposure is not always successful, and a variety of tools have been implemented clinically to try to ensure and improve extinction processes over sensitization or reconsolidation processes during memory exposure. The use of cognitive enhancers to improve extinction is now an accepted translational paradigm from bench to bedside and is considered a paradigm-shifting approach. However, this approach has the limitation that both BDNF- and NMDA-dependent plasticity support both extinction and fear learning. Next-generation approaches, as discussed later, hope to identify cell populations or microcircuits, based on translational neuroscience, that are valence-specific and support fear inhibition or extinction and not fear acquisition, expression, or reconsolidation. Targeting cells and circuits, either pharmacologically or neurotherapeutically, that directly enhance extinction or block reconsolidation could provide a true intersection of targeted intervention, combining neurobiological insights with psychotherapeutic tools to drive specific memories.

New Findings And Potential Translational Relevance

Memory Representation, Priming, Generalization, And Anxiety Disorders

In addition to the amygdala, the anterior cingulate cortex (ACC) is one region involved in associative memory formation, conflict and aversive memory expression, and PTSD and anxiety disorders (34, 84, 85). A study reported in this issue of the Journal by Yang et al. (86) examines both the behavioral and semantic representation, as well as neural activation of the ACC, in panic-related representations and exposure therapy in panic disorder. From a translational perspective, this work is exciting in a number of ways. While the concept of priming is primarily a human psychology construct, behavioral studies in model systems have long demonstrated the role of “preparedness” for a threat and overactivation of a threat network, via generalization of cues and contexts that are more likely to contribute to a threat response, despite a relatively low actual probability of danger. Extinction of cues (which appears to depend on the hippocampus/amygdala and mPFC/ACC) in large part is context based, and successful extinction often leads to generalization of the extinction contexts as well (50, 87). Thus, it is exciting and promising that exposure-based CBT for panic disorder results in decreased priming and, possibly, increased generalization of extinguished cues (with decreased negative valence), along with decreased brain network activity corresponding to such processes.

Data-Driven Approaches For Brain Activity-Based Subtyping In PTSD

Recent work has also examined unbiased data-driven approaches to understanding brain activity patterns that may underlie different biologically distinct forms of anxiety and fear-related disorders. In another article in this issue, Maron-Katz et al. (88) examine resting-state functional MRI (fMRI) and dynamic brain activity to detect subtypes of PTSD. In parallel with neuroimaging, new analytic approaches with high-density EEG are allowing unsurpassed collection of real-time brain activity in a low-cost fashion that might be more feasible to disseminate. Toll et al. (89) further demonstrate progress in biotyping in PTSD using EEG in this issue of the Journal. In addition to the possibility of novel pathways and subgrouping tools, it is interesting that among the primary fMRI findings in the Maron-Katz et al. study (88) was the identification of visual and sensorimotor findings important in PTSD. As noted by the authors, although abnormalities in frontoparietal connectivity and activity in PTSD case subjects have been reported in several studies, abnormalities in connectivity and activity of sensorimotor and visual regions have been less often seen. Importantly, the translational literature would suggest that these areas are highly likely to be involved in the aftermath of trauma exposure. In fact, in mice, repeated auditory tone–paired foot shocks have been shown to increase voxel-based morphometry gray matter density, likely a result of increased spine and dendritic spine density, in the auditory cortex, sensorimotor regions, and amygdala (9093). These data, combined with a robust literature on structural representation of memory engrams across a large range of primary and secondary sensory areas, suggest that while the amygdala, dorsal ACC, and hippocampus may be “hubs” for memory formation, the actual encoding and sensitization of trauma is likely distributed across the full network encoding the memory (2, 9496). Thus, in addition to the important and hopefully soon-to-be-clinically-utilized tools of EEG- and MRI-based biotyping for anxiety disorders, understanding the networks that may be differentially regulated in these biotypes will expand our neuroscience understanding of risk and resilience.

Deep Brain Stimulation of Specific Limbic Circuits for OCD Treatment

The “holy grail” of a neural circuit model of anxiety disorders would be to utilize the mechanistic understanding of neural circuitry for direct intervention and treatment. Denys et al. (97), in a study reported in this issue, add to the growing literature suggesting the effectiveness and safety of deep brain stimulation (DBS) in treatment-refractory OCD. Of note, their target—the anterior limb of the internal capsule—is a critical white matter juncture involved in amygdala, ventromedial PFC, and nucleus accumbens (NAcc) connectivity that has been demonstrated across animal model systems to mediate aspects of anxiety and repetitive behaviors (98101). Furthermore, amygdala-striatal circuit activation has been shown to decrease long-term fear and threat processing in model systems, and the NAcc and broader striatum have long been associated with avoidance responses in behavioral learning models (101103)—and, notably, avoidance is often among the most refractory of clinical symptoms in treatment of OCD. This literature also brings together interesting findings associating threat/fear/amygdala networks with addiction/appetitive/NAcc networks. Such preclinical work has recently influenced novel approaches to behavioral exposure therapy as well, by actively bringing reward systems online at the culmination of threat exposures, which appears to enhance the efficacy of extinction and exposure for anxiety- and fear-related disorders (104, 105).

Large-Scale Genomics Discovery of Anxiety-Related Gene Pathways

In addition to interesting neural circuit approaches for discovery and intervention, ongoing work by large groups of investigators, including the Psychiatric Genomics Consortium Anxiety and PTSD work groups, the UK Biobank, and the Million Veteran Program, have worked to gather data on hundreds of thousands of individuals for large-scale discovery of the genetic architecture underlying anxiety- and fear-related disorders. The latest of these groundbreaking studies is published in this issue by Levey et al. (106), who examine the largest GWAS of anxiety to date. Notably, with fear- and anxiety-related disorders having approximately twice the incidence in females as males, the authors’ identification of the estrogen receptor gene (ESR1) as one of the genes robustly associated with anxiety symptoms was particularly exciting. Additionally, they identified the CRHR1 gene (Corticotropin Releasing Hormone Receptor 1) in gene-based association analyses. This is particularly exciting given the previous findings that the locus—including single-nucleotide polymorphisms associated with gene expression within the CRHR1 locus (among a number of other genes)—was a top hit meeting genome-wide significance for hyperarousal symptoms in PTSD and alcohol intake in previous recent publications from the same group (21, 107). This work provides new insights into genetic risk mechanisms underpinning anxiety and related psychiatric disorders. From a translational perspective, some of these findings are enormously exciting. For example, the role of estrogen in the regulation of fear and fear extinction has been gaining much traction recently, and the underlying mechanisms for ESR1 functioning in the neural circuits of threat are being identified (108112). Certainly, identification of gene pathways that were previously unknown will lead to a whole array of new biology to integrate within the known neural circuit pathways. Perhaps the most exciting development from these recent large-scale GWASs is that they may also provide some GWAS-level biological support for the role of the hypothalamic-pituitary-adrenal (HPA) axis in fear- and anxiety-related disorders. Since its discovery, corticotropin-releasing hormone (CRH) has been a leading translational target in understanding amygdala-hypothalamic-pituitary and other subcortical regulation of the threat reflex (113118). Preclinically, overexpression of CRH or genetic activation of CRH-expressing neurons leads to increased fear and deficits in fear extinction. Conversely, CRH knockout or silencing these cells has opposite effects (114, 119123).
Of note, given the tremendous amount of support for HPA dysregulation and CRH overactivity in depression, PTSD, and fear-related disorders, antagonists targeting CRHR1 were developed with tremendous excitement. Unfortunately, recent trials of these compounds were negative in depression, social anxiety, and PTSD (124127). The assumption is that the drugs, based on our preclinical model of stress disorders, are not effective in the human disorders and that this may represent a failure of translation, which may still be true. However, an important and yet untested possibility is that at the level of precision of a specific neuropeptide antagonist, there may be either sufficient redundancy in the system that monotherapy is inadequate or, perhaps more likely, biological heterogeneity is such that until the appropriate, perhaps transdiagnostic, biological subtypes of patients with specific dysregulation in their CRH systems are stratified for treatment and the ideal treatment conditions (e.g., baseline or with exposure therapy, etc.) are identified, translation of such biological findings will not be successful. Notably, secondary analysis of the PTSD–CRHR1 antagonist study suggests exactly this—that it may be helpful for targeted patients and targeted intermediate phenotypes (124, 128).
The unbiased identification of the CRHR1 locus in large-scale anxiety- and fear-related disorder GWASs surely should make the field reexamine the disconnect between the biology and targeted pharmacotherapy trials to date. Importantly, how best to integrate data from many gene pathways, each with small effect size, into appropriate treatment targets, perhaps by targeting hub genes or identifying the correct combination of therapies to target the correct combination of critical genes, remains a challenge across all genome-driven approaches to new therapies.

Gaps In Our Understanding

Discussed above are examples of relative translational successes, or at least areas of clear intersection between neuroscience progress in model systems and human clinical neuroscience and clinical disorders. In contrast, below are outlined some clear examples of where our current large-scale data-driven studies do not yet have clear translation with our current framework for understanding the neurobiology of fear- and anxiety-related disorders.
Among the most exciting challenges we currently face is how we can utilize the powerful, large-scale, unbiased, data-driven (e.g., model-free) approaches together with many decades of detailed progress in understanding the neural bases of threat processing and extinction (e.g., model-based).
We are at a crossroads where large-scale studies from single-cell RNA sequencing of millions of cells throughout the brain may intersect with GWASs of hundreds of thousands of patients and neural imaging of tens of thousands of patients.
What are the best next-generation functional genomic approaches to combine with large-scale GWAS discovery, e.g., genetic signatures to predict gene transcription, cell-type-specific expression from the PsychENCODE project, or other functional genomic tools?
What are the right model systems, computational approaches, and mechanistic neurobiological questions to address the enormous complexity?
Clearly, animal model systems are needed to understand some level of mammalian circuit organization, in particular at a causal level that cannot be done in humans. Which research questions are most appropriate for mouse, nonhuman primate, or other model systems (e.g., organoids)?
The anxiety and fear-related disorders may be more constrained in their complexity than some other areas of psychiatric disease, because of their evolutionary conservation, relatively well understood circuitry, and at least some translational understanding of our best current therapy approaches. Nonetheless, the problems remain very complex, and integrating all levels of analyses—from genetic to cellular to circuit to behavior—will require sophisticated translational approaches.

Conclusions And Future Directions

While some of the deficits in knowledge described above may seem insurmountable, the neuroscientific understanding of fear and threat processing and the inhibition and extinction of these behaviors has rapidly advanced in recent years. Some ideas on ways to leverage this progress in crossing these knowledge gaps could include the following:
Increased translational approaches to understanding circuit dynamics in model systems with constrained behaviors (e.g., EEG/fMRI in humans undergoing fear conditioning and extinction with the same specific behaviors in rodent and nonhuman primate model systems, using concomitant circuit-driven mechanistic tools).
Integration of circuit knowledge with genetics and molecular tools; this is currently do-able with single-cell RNA sequencing and cell-specific opto-/chemogenetics in rodents, and methods have now been adapted to nonhuman primates. Combining such a mechanistic cell-type-specific understanding of fear and threat regulation with single-cell RNA sequencing from postmortem human brain (from healthy and diseased subjects, for example) to best understand at what level our molecular and circuit translation will hold.
Integration of GWASs and other large-scale “omics” with the single-cell and circuit-level understanding to complete a model of how the genetic contributions to risk intersect within known circuits that underlie risk and recovery from fear- and anxiety-related disorders.
Experimental-medicine approaches with further targeting of causal circuits related to fear and threat processing in anxiety disorders with neurotherapeutic (e.g., DBS, targeted transcranial magnetic stimulation [TMS] [e.g., with neuronavigation], etc.), combined with quantitative human data collection, both at the clinical and the fMRI level, but also with EEG, other physiological measures, and more sophisticated quantitative behavioral outcome measures.
Use of rationally driven approaches to enhance neural plasticity, combined with well-controlled behaviorally targeted extinction of fear and threat (e.g., with either more directly targeted plasticity enhancers that preferentially drive extinction instead of reconsolidation, or with more precise mechanisms of behavior regulation [e.g., targeted psychotherapy and virtual reality approaches] driving extinction over reconsolidation).
Translate the great progress in neural understanding of trauma and fear memory consolidation to specific, adaptive intervention and prevention trials in the early aftermath of trauma, done in a thoughtful, well-powered way with a number of potential mechanistic compounds.
In conclusion, while anxiety-, fear-, and trauma-related disorders are common and costly and significantly raise risk for mortality and morbidity, advances in our understanding of genetics and human neuroscience offer tremendous promise. Data-driven tools will provide much greater unbiased insight into the underlying neural circuits and underlying genomic pathways that both predict outcome and allow for stratification of biological subtypes of these clinically heterogeneous syndromes. Next-generation approaches envision combining targeted psychotherapy—as a means to specifically manipulate the memories that are susceptible to modulation by targeting synaptic plasticity—and pharmacologically or targeted neural circuit–based manipulations such as TMS or DBS. While the road is long, tremendous progress has already been made toward translating across circuits and genetics for progress in fear- and anxiety-related disorders—and, fortunately, the best is yet to come.

References

1.
Fenster RJ, Lebois LAM, Ressler KJ, et al: Brain circuit dysfunction in post-traumatic stress disorder: from mouse to man. Nat Rev Neurosci 2018; 19:535–551
2.
Maddox SA, Hartmann J, Ross RA, et al: Deconstructing the Gestalt: mechanisms of fear, threat, and trauma memory encoding. Neuron 2019; 102:60–74
3.
Ball TM, Knapp SE, Paulus MP, et al: Brain activation during fear extinction predicts exposure success. Depress Anxiety 2017; 34:257–266
4.
Delgado MR, Nearing KI, Ledoux JE, et al: Neural circuitry underlying the regulation of conditioned fear and its relation to extinction. Neuron 2008; 59:829–838
5.
Ciocchi S, Herry C, Grenier F, et al: Encoding of conditioned fear in central amygdala inhibitory circuits. Nature 2010; 468:277–282
6.
Dejean C, Courtin J, Rozeske RR, et al: Neuronal circuits for fear expression and recovery: recent advances and potential therapeutic strategies. Biol Psychiatry 2015; 78:298–306
7.
Fadok JP, Krabbe S, Markovic M, et al: A competitive inhibitory circuit for selection of active and passive fear responses. Nature 2017; 542:96–100
8.
Gozzi A, Jain A, Giovannelli A, et al: A neural switch for active and passive fear. Neuron 2010; 67:656–666
9.
Haubensak W, Kunwar PS, Cai H, et al: Genetic dissection of an amygdala microcircuit that gates conditioned fear. Nature 2010; 468:270–276
10.
Li H, Penzo MA, Taniguchi H, et al: Experience-dependent modification of a central amygdala fear circuit. Nat Neurosci 2013; 16:332–339
11.
McCullough KM, Choi D, Guo J, et al: Molecular characterization of Thy1 expressing fear-inhibiting neurons within the basolateral amygdala. Nat Commun 2016; 7:13149
12.
McCullough KM, Morrison FG, Ressler KJ: Bridging the gap: towards a cell-type specific understanding of neural circuits underlying fear behaviors. Neurobiol Learn Mem 2016; 135:27–39
13.
Duncan LE, Ratanatharathorn A, Aiello AE, et al: Largest GWAS of PTSD (N=20 070) yields genetic overlap with schizophrenia and sex differences in heritability. Mol Psychiatry 2018; 23:666–673
14.
Dunn EC, Wiste A, Radmanesh F, et al: Genome-wide association study (GWAS) and genome-wide by environment interaction study (GWEIS) of depressive symptoms in African American and Hispanic/Latina women. Depress Anxiety 2016; 33:265–280
15.
Koenen KC, Duncan LE, Liberzon I, et al: From candidate genes to genome-wide association: the challenges and promise of posttraumatic stress disorder genetic studies. Biol Psychiatry 2013; 74:634–636
16.
Nievergelt CM, Maihofer AX, Klengel T, et al: International meta-analysis of PTSD genome-wide association studies identifies sex- and ancestry-specific genetic risk loci. Nat Commun 2019; 10:4558
17.
Akbarian S, Liu C, Knowles JA, et al: The PsychENCODE project. Nat Neurosci 2015; 18:1707–1712
18.
Sullivan PF, Agrawal A, Bulik CM, et al: Psychiatric genomics: an update and an agenda. Am J Psychiatry 2018; 175:15–27
19.
Stein MB, Chen CY, Ursano RJ, et al: Genome-wide association studies of posttraumatic stress disorder in 2 cohorts of US Army soldiers. JAMA Psychiatry 2016; 73:695–704
20.
McLean SA, Ressler K, Koenen KC, et al: The AURORA Study: a longitudinal, multimodal library of brain biology and function after traumatic stress exposure. Mol Psychiatry (Epub ahead of print, November 19, 2019)
21.
Gelernter J, Sun N, Polimanti R, et al: Genome-wide association study of post-traumatic stress disorder reexperiencing symptoms in >165,000 US veterans. Nat Neurosci 2019; 22:1394–1401
22.
Pavlov I: Conditioned Reflexes. Oxford, UK, Oxford University Press, 1927
23.
Mobbs D, Adolphs R, Fanselow MS, et al: Viewpoints: Approaches to defining and investigating fear. Nat Neurosci 2019; 22:1205–1216
24.
Johansen JP, Cain CK, Ostroff LE, et al: Molecular mechanisms of fear learning and memory. Cell 2011; 147:509–524
25.
Beyeler A, Namburi P, Glober GF, et al: Divergent routing of positive and negative information from the amygdala during memory retrieval. Neuron 2016; 90:348–361
26.
Burgos-Robles A, Kimchi EY, Izadmehr EM, et al: Amygdala inputs to prefrontal cortex guide behavior amid conflicting cues of reward and punishment. Nat Neurosci 2017; 20:824–835
27.
Janak PH, Tye KM: From circuits to behaviour in the amygdala. Nature 2015; 517:284–292
28.
Birn RM, Shackman AJ, Oler JA, et al: Extreme early-life anxiety is associated with an evolutionarily conserved reduction in the strength of intrinsic functional connectivity between the dorsolateral prefrontal cortex and the central nucleus of the amygdala. Mol Psychiatry 2014; 19:853
29.
Fox AS, Oler JA, Birn RM, et al: Functional connectivity within the primate extended amygdala is heritable and associated with early-life anxious temperament. J Neurosci 2018; 38:7611–7621
30.
Fox AS, Oler JA, Shelton SE, et al: Central amygdala nucleus (Ce) gene expression linked to increased trait-like Ce metabolism and anxious temperament in young primates. Proc Natl Acad Sci USA 2012; 109:18108–18113
31.
Kalin NH, Fox AS, Kovner R, et al: Overexpressing corticotropin-releasing factor in the primate amygdala increases anxious temperament and alters its neural circuit. Biol Psychiatry 2016; 80:345–355
32.
Kim YJ, van Rooij SJH, Ely TD, et al: Association between posttraumatic stress disorder severity and amygdala habituation to fearful stimuli. Depress Anxiety 2019; 36:647–658
33.
Stevens JS, Jovanovic T, Fani N, et al: Disrupted amygdala-prefrontal functional connectivity in civilian women with posttraumatic stress disorder. J Psychiatr Res 2013; 47:1469–1478
34.
Stevens JS, Kim YJ, Galatzer-Levy IR, et al: Amygdala reactivity and anterior cingulate habituation predict posttraumatic stress disorder symptom maintenance after acute civilian trauma. Biol Psychiatry 2017; 81:1023–1029
35.
Michopoulos V, Norrholm SD, Stevens JS, et al: Dexamethasone facilitates fear extinction and safety discrimination in PTSD: a placebo-controlled, double-blind study. Psychoneuroendocrinology 2017; 83:65–71
36.
Galatzer-Levy IR, Andero R, Sawamura T, et al: A cross species study of heterogeneity in fear extinction learning in relation to FKBP5 variation and expression: Implications for the acute treatment of posttraumatic stress disorder. Neuropharmacology 2017; 116:188–195
37.
Orcutt HK, Hannan SM, Seligowski AV, et al: Fear-potentiated startle and fear extinction in a sample of undergraduate women exposed to a campus mass shooting. Front Psychol 2017; 7:2031
38.
Norrholm SD, Jovanovic T, Gerardi M, et al: Baseline psychophysiological and cortisol reactivity as a predictor of PTSD treatment outcome in virtual reality exposure therapy. Behav Res Ther 2016; 82:28–37
39.
Norrholm SD, Glover EM, Stevens JS, et al: Fear load: the psychophysiological over-expression of fear as an intermediate phenotype associated with trauma reactions. Int J Psychophysiol 2015; 98:270–275
40.
Jovanovic T, Ely T, Fani N, et al: Reduced neural activation during an inhibition task is associated with impaired fear inhibition in a traumatized civilian sample. Cortex 2013; 49:1884–1891
41.
Norrholm SD, Jovanovic T, Olin IW, et al: Fear extinction in traumatized civilians with posttraumatic stress disorder: relation to symptom severity. Biol Psychiatry 2011; 69:556–563
42.
Kindt M, Soeter M, Vervliet B: Beyond extinction: erasing human fear responses and preventing the return of fear. Nat Neurosci 2009; 12:256–258
43.
Lee JL, Milton AL, Everitt BJ: Reconsolidation and extinction of conditioned fear: inhibition and potentiation. J Neurosci 2006; 26:10051–10056
44.
McGaugh JL: Memory: a century of consolidation. Science 2000; 287:248–251
45.
Sartori SB, Singewald N: Novel pharmacological targets in drug development for the treatment of anxiety and anxiety-related disorders. Pharmacol Ther 2019; 204:107402
46.
Singewald N, Schmuckermair C, Whittle N, et al: Pharmacology of cognitive enhancers for exposure-based therapy of fear, anxiety, and trauma-related disorders. Pharmacol Ther 2015; 149:150–190
47.
Bowers ME, Ressler KJ: An overview of translationally informed treatments for posttraumatic stress disorder: animal models of Pavlovian fear conditioning to human clinical trials. Biol Psychiatry 2015; 78:E15–E27
48.
Parsons RG, Ressler KJ: Implications of memory modulation for post-traumatic stress and fear disorders. Nat Neurosci 2013; 16:146–153
49.
Ross DA, Arbuckle MR, Travis MJ, et al: An integrated neuroscience perspective on formulation and treatment planning for posttraumatic stress disorder: an educational review. JAMA Psychiatry 2017; 74:407–415
50.
Myers KM, Davis M: Mechanisms of fear extinction. Mol Psychiatry 2007; 12:120–150
51.
Singewald N, Holmes A: Rodent models of impaired fear extinction. Psychopharmacology (Berl) 2019; 236:21–32
52.
Graham BM, Milad MR: The study of fear extinction: implications for anxiety disorders. Am J Psychiatry 2011; 168:1255–1265
53.
Marin MF, Milad MR: Extending the examination of the fear extinction network beyond anxiety and fear-based disorders: insight into autism spectrum disorder. Biol Psychiatry Cogn Neurosci Neuroimaging 2016; 1:302–304
54.
Milad MR, Rauch SL, Pitman RK, et al: Fear extinction in rats: implications for human brain imaging and anxiety disorders. Biol Psychol 2006; 73:61–71
55.
Rothbaum BO, Davis M: Applying learning principles to the treatment of post-trauma reactions. Ann N Y Acad Sci 2003; 1008:112–121
56.
Benito KG, Machan J, Freeman JB, et al: Measuring fear change within exposures: functionally defined habituation predicts outcome in three randomized controlled trials for pediatric OCD. J Consult Clin Psychol 2018; 86:615–630
57.
Foa EB, McLean CP: The efficacy of exposure therapy for anxiety-related disorders and its underlying mechanisms: the case of OCD and PTSD. Annu Rev Clin Psychol 2016; 12:1–28
58.
Lueken U, Straube B, Konrad C, et al: Neural substrates of treatment response to cognitive-behavioral therapy in panic disorder with agoraphobia. Am J Psychiatry 2013; 170:1345–1355
59.
Michael T, Blechert J, Vriends N, et al: Fear conditioning in panic disorder: enhanced resistance to extinction. J Abnorm Psychol 2007; 116:612–617
60.
Stein MB, Rothbaum BO: 175 Years of progress in PTSD therapeutics: learning from the past. Am J Psychiatry 2018; 175:508–516
61.
Stoyanov D, Telles-Correia D, Cuthbert BN: The Research Domain Criteria (RDoC) and the historical roots of psychopathology: a viewpoint. Eur Psychiatry 2019; 57:58–60
62.
Rescorla RA, Heth CD: Reinstatement of fear to an extinguished conditioned stimulus. J Exp Psychol Anim Behav Process 1975; 1:88–96
63.
Bouton ME, Moody EW: Memory processes in classical conditioning. Neurosci Biobehav Rev 2004; 28:663–674
64.
Bouton ME, Bolles RC: Role of conditioned contextual stimuli in reinstatement of extinguished fear. J Exp Psychol Anim Behav Process 1979; 5:368–378
65.
Bouton ME, King DA: Contextual control of the extinction of conditioned fear: tests for the associative value of the context. J Exp Psychol Anim Behav Process 1983; 9:248–265
66.
Falls WA, Miserendino MJ, Davis M: Extinction of fear-potentiated startle: blockade by infusion of an NMDA antagonist into the amygdala. J Neurosci 1992; 12:854–863
67.
Myers KM, Davis M: Behavioral and neural analysis of extinction. Neuron 2002; 36:567–584
68.
Andero R, Heldt SA, Ye K, et al: Effect of 7,8-dihydroxyflavone, a small-molecule TrkB agonist, on emotional learning. Am J Psychiatry 2011; 168:163–172
69.
Andero R, Ressler KJ: Fear extinction and BDNF: translating animal models of PTSD to the clinic. Genes Brain Behav 2012; 11:503–512
70.
Chhatwal JP, Stanek-Rattiner L, Davis M, et al: Amygdala BDNF signaling is required for consolidation but not encoding of extinction. Nat Neurosci 2006; 9:870–872
71.
Heldt SA, Stanek L, Chhatwal JP, et al: Hippocampus-specific deletion of BDNF in adult mice impairs spatial memory and extinction of aversive memories. Mol Psychiatry 2007; 12:656–670
72.
Walker DL, Ressler KJ, Lu KT, et al: Facilitation of conditioned fear extinction by systemic administration or intra-amygdala infusions of d-cycloserine as assessed with fear-potentiated startle in rats. J Neurosci 2002; 22:2343–2351
73.
Ressler KJ, Rothbaum BO, Tannenbaum L, et al: Cognitive enhancers as adjuncts to psychotherapy: use of d-cycloserine in phobic individuals to facilitate extinction of fear. Arch Gen Psychiatry 2004; 61:1136–1144
74.
Guastella AJ, Richardson R, Lovibond PF, et al: A randomized controlled trial of d-cycloserine enhancement of exposure therapy for social anxiety disorder. Biol Psychiatry 2008; 63:544–549
75.
Hofmann SG, Meuret AE, Smits JA, et al: Augmentation of exposure therapy with d-cycloserine for social anxiety disorder. Arch Gen Psychiatry 2006; 63:298–304
76.
Kushner MG, Kim SW, Donahue C, et al: d-Cycloserine augmented exposure therapy for obsessive-compulsive disorder. Biol Psychiatry 2007; 62:835–838
77.
Ledgerwood L, Richardson R, Cranney J: d-Cycloserine facilitates extinction of conditioned fear as assessed by freezing in rats, in Society for Neuroscience Annual Meeting, Orlando, Fla, November 3–7, 2002
78.
Ledgerwood L, Richardson R, Cranney J: Effects of d-cycloserine on extinction of conditioned freezing. Behav Neurosci 2003; 117:341–349
79.
Ledgerwood L, Richardson R, Cranney J: d-Cycloserine and the facilitation of extinction of conditioned fear: consequences for reinstatement. Behav Neurosci 2004; 118:505–513
80.
Ledgerwood L, Richardson R, Cranney J: d-Cycloserine facilitates extinction of learned fear: effects on reacquisition and generalized extinction. Biol Psychiatry 2005; 57:841–847
81.
Wilhelm S, Buhlmann U, Tolin DF, et al: Augmentation of behavior therapy with d-cycloserine for obsessive-compulsive disorder. Am J Psychiatry 2008; 165:335–341
82.
Mataix-Cols D, Fernández de la Cruz L, Monzani B, et al: d-Cycloserine augmentation of exposure-based cognitive behavior therapy for anxiety, obsessive-compulsive, and posttraumatic stress disorders: a systematic review and meta-analysis of individual participant data. JAMA Psychiatry 2017; 74:501–510
83.
Rosenfield D, Smits JAJ, Hofmann SG, et al: Changes in dosing and dose timing of d-cycloserine explain its apparent declining efficacy for augmenting exposure therapy for anxiety-related disorders: an individual participant-data meta-analysis. J Anxiety Disord 2019; 68:102149
84.
Liberzon I, Taylor SF, Amdur R, et al: Brain activation in PTSD in response to trauma-related stimuli. Biol Psychiatry 1999; 45:817–826
85.
Milad MR, Pitman RK, Ellis CB, et al: Neurobiological basis of failure to recall extinction memory in posttraumatic stress disorder. Biol Psychiatry 2009; 66:1075–1082
86.
Yang Y, Lueken U, Richter J, et al: Effect of CBT on biased semantic network in panic disorder: a multicenter fMRI study using semantic priming. Am J Psychiatry 2020; 177:254–264
87.
Bouton ME: Context and behavioral processes in extinction. Learn Mem 2004; 11:485–494
88.
Maron-Katz A, Zhang Y, Narayan M, et al: Individual patterns of abnormality in resting-state functional connectivity reveal two data-driven PTSD subgroups. Am J Psychiatry 2020; 177:244–253
89.
Toll RT, Wu W, Naparstek S, et al: An electroencephalography connectomic profile of posttraumatic stress disorder. Am J Psychiatry 2020; 177:233–243
90.
Keifer OP Jr, Hurt RC, Gutman DA, et al: Voxel-based morphometry predicts shifts in dendritic spine density and morphology with auditory fear conditioning. Nat Commun 2015; 6:7582
91.
Banerjee SB, Gutzeit VA, Baman J, et al: Perineuronal nets in the adult sensory cortex are necessary for fear learning. Neuron 2017; 95:169–179.e3, e163
92.
Ressler KJ, Paschall G, Zhou XL, et al: Regulation of synaptic plasticity genes during consolidation of fear conditioning. J Neurosci 2002; 22:7892–7902
93.
Lai CSW, Adler A, Gan WB: Fear extinction reverses dendritic spine formation induced by fear conditioning in the mouse auditory cortex. Proc Natl Acad Sci USA 2018; 115:9306–9311
94.
Dudai Y: The neurobiology of consolidations, or, how stable is the engram? Annu Rev Psychol 2004; 55:51–86
95.
Josselyn SA, Köhler S, Frankland PW: Finding the engram. Nat Rev Neurosci 2015; 16:521–534
96.
Park S, Kramer EE, Mercaldo V, et al: Neuronal allocation to a hippocampal engram. Neuropsychopharmacology 2016; 41:2987–2993
97.
Denys D, Graat I, Mocking R, et al: Efficacy of deep brain stimulation of the ventral anterior limb of the internal capsule for refractory obsessive-compulsive disorder: a clinical cohort of 70 patients. Am J Psychiatry 2020; 177:265–271
98.
Eden AS, Schreiber J, Anwander A, et al: Emotion regulation and trait anxiety are predicted by the microstructure of fibers between amygdala and prefrontal cortex. J Neurosci 2015; 35:6020–6027
99.
Correia SS, McGrath AG, Lee A, et al: Amygdala-ventral striatum circuit activation decreases long-term fear. eLife 2016; 5 ( https://doi.org/10.7554/eLife.12669)
100.
Abraham AD, Neve KA, Lattal KM: Dopamine and extinction: a convergence of theory with fear and reward circuitry. Neurobiol Learn Mem 2014; 108:65–77
101.
Wendler E, Gaspar JC, Ferreira TL, et al: The roles of the nucleus accumbens core, dorsomedial striatum, and dorsolateral striatum in learning: performance and extinction of Pavlovian fear-conditioned responses and instrumental avoidance responses. Neurobiol Learn Mem 2014; 109:27–36
102.
Ramirez F, Moscarello JM, LeDoux JE, et al: Active avoidance requires a serial basal amygdala to nucleus accumbens shell circuit. J Neurosci 2015; 35:3470–3477
103.
Wenzel JM, Oleson EB, Gove WN, et al: Phasic dopamine signals in the nucleus accumbens that cause active avoidance require endocannabinoid mobilization in the midbrain. Curr Biol 2018; 28:1392–1404.e5, e1395
104.
Craske MG, Treanor M, Conway CC, et al: Maximizing exposure therapy: an inhibitory learning approach. Behav Res Ther 2014; 58:10–23
105.
Zbozinek TD, Craske MG: Positive affect predicts less reacquisition of fear: relevance for long-term outcomes of exposure therapy. Cogn Emotion 2017; 31:712–725
106.
Levey DF, Gelernter J, Polimanti R, et al: Reproducible genetic risk loci for anxiety: results from ∼200,000 participants in the Million Veteran Program. Am J Psychiatry 2020; 177:223–232
107.
Gelernter J, Sun N, Polimanti R, et al: Genome-wide association study of maximum habitual alcohol intake in >140,000 US European and African American veterans yields novel risk loci. Biol Psychiatry 2019; 86:365–376
108.
Graham BM, Milad MR: Inhibition of estradiol synthesis impairs fear extinction in male rats. Learn Mem 2014; 21:347–350
109.
Zeidan MA, Igoe SA, Linnman C, et al: Estradiol modulates medial prefrontal cortex and amygdala activity during fear extinction in women and female rats. Biol Psychiatry 2011; 70:920–927
110.
Glover EM, Jovanovic T, Mercer KB, et al: Estrogen levels are associated with extinction deficits in women with posttraumatic stress disorder. Biol Psychiatry 2012; 72:19–24
111.
Glover EM, Mercer KB, Norrholm SD, et al: Inhibition of fear is differentially associated with cycling estrogen levels in women. J Psychiatry Neurosci 2013; 38:341–348
112.
Lynch JF 3rd, Winiecki P, Vanderhoof T, et al: Hippocampal cytosolic estrogen receptors regulate fear generalization in females. Neurobiol Learn Mem 2016; 130:83–92
113.
Bale TL, Picetti R, Contarino A, et al: Mice deficient for both corticotropin-releasing factor receptor 1 (CRFR1) and CRFR2 have an impaired stress response and display sexually dichotomous anxiety-like behavior. J Neurosci 2002; 22:193–199
114.
Dedic N, Kühne C, Gomes KS, et al: Deletion of CRH from GABAergic forebrain neurons promotes stress resilience and dampens stress-induced changes in neuronal activity. Front Neurosci 2019; 13:986
115.
Gafford G, Jasnow AM, Ressler KJ: Grin1 receptor deletion within CRF neurons enhances fear memory. PLoS One 2014; 9:e111009
116.
Heinrichs SC, Menzaghi F, Merlo Pich E, et al: The role of CRF in behavioral aspects of stress. Ann N Y Acad Sci 1995; 771:92–104
117.
Nemeroff CB: The corticotropin-releasing factor (CRF) hypothesis of depression: new findings and new directions. Mol Psychiatry 1996; 1:336–342
118.
Radulovic J, Rühmann A, Liepold T, et al: Modulation of learning and anxiety by corticotropin-releasing factor (CRF) and stress: differential roles of CRF receptors 1 and 2. J Neurosci 1999; 19:5016–5025
119.
Dedic N, Kühne C, Jakovcevski M, et al: Chronic CRH depletion from GABAergic long-range projection neurons in the extended amygdala reduces dopamine release and increases anxiety. Nat Neurosci 2018; 21:803–807
120.
Flandreau EI, Ressler KJ, Owens MJ, et al: Chronic overexpression of corticotropin-releasing factor from the central amygdala produces HPA axis hyperactivity and behavioral anxiety associated with gene-expression changes in the hippocampus and paraventricular nucleus of the hypothalamus. Psychoneuroendocrinology 2012; 37:27–38
121.
Gafford GM, Guo JD, Flandreau EI, et al: Cell-type specific deletion of GABA(A)α1 in corticotropin-releasing factor-containing neurons enhances anxiety and disrupts fear extinction. Proc Natl Acad Sci USA 2012; 109:16330–16335
122.
Keen-Rhinehart E, Michopoulos V, Toufexis DJ, et al: Continuous expression of corticotropin-releasing factor in the central nucleus of the amygdala emulates the dysregulation of the stress and reproductive axes. Mol Psychiatry 2009; 14:37–50
123.
Sink KS, Walker DL, Freeman SM, et al: Effects of continuously enhanced corticotropin releasing factor expression within the bed nucleus of the stria terminalis on conditioned and unconditioned anxiety. Mol Psychiatry 2013; 18:308–319
124.
Dunlop BW, Binder EB, Iosifescu D, et al: Corticotropin-releasing factor receptor 1 antagonism is ineffective for women with posttraumatic stress disorder. Biol Psychiatry 2017; 82:866–874
125.
Dunlop BW, Rothbaum BO, Binder EB, et al: Evaluation of a corticotropin releasing hormone type 1 receptor antagonist in women with posttraumatic stress disorder: study protocol for a randomized controlled trial. Trials 2014; 15:240
126.
Binneman B, Feltner D, Kolluri S, et al: A 6-week randomized, placebo-controlled trial of CP-316,311 (a selective CRH1 antagonist) in the treatment of major depression. Am J Psychiatry 2008; 165:617–620
127.
Coric V, Feldman HH, Oren DA, et al: Multicenter, randomized, double-blind, active comparator and placebo-controlled trial of a corticotropin-releasing factor receptor-1 antagonist in generalized anxiety disorder. Depress Anxiety 2010; 27:417–425
128.
Jovanovic T, Duncan EJ, Kaye J, et al: Psychophysiological treatment outcomes: Corticotropin-releasing factor type 1 receptor antagonist increases inhibition of fear-potentiated startle in PTSD patients. Psychophysiology 2020; 57:e13356

Information & Authors

Information

Published In

Go to American Journal of Psychiatry
Go to American Journal of Psychiatry
American Journal of Psychiatry
Pages: 214 - 222
PubMed: 32114783

History

Accepted: 15 January 2020
Published online: 1 March 2020
Published in print: March 01, 2020

Keywords

  1. Anxiety Disorders
  2. Posttraumatic Stress Disorder
  3. Amygdala
  4. Anterior Cingulate
  5. GWAS
  6. Fear Extinction

Authors

Affiliations

Kerry J. Ressler, M.D., Ph.D. [email protected]
McLean Hospital, Harvard Medical School, Belmont, Mass.

Notes

Send correspondence to Dr. Ressler ([email protected]).

Funding Information

National Institute of Mental Healthhttp://dx.doi.org/10.13039/100000025: P50-MH115874, R01-MH106595, R01-MH108665, R01-MH117292, U01-MH110925
Supported by funding from NIH (grants P50-MH115874, R01-MH108665, R01-MH117292, R01-MH106595, and U01-MH110925) and the Frazier Institute at McLean Hospital.Dr. Ressler has served as a consultant for Alkermes and serves on scientific advisory boards for Janssen, Resilience Therapeutics, and Verily; he has received sponsored research support from Brainsway and Takeda. Dr. Kalin has reviewed this editorial and found no evidence of influence from these relationships.

Metrics & Citations

Metrics

Citations

Export Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

For more information or tips please see 'Downloading to a citation manager' in the Help menu.

Format
Citation style
Style
Copy to clipboard

There are no citations for this item

View Options

View options

PDF/ePub

View PDF/ePub

Get Access

Login options

Already a subscriber? Access your subscription through your login credentials or your institution for full access to this article.

Personal login Institutional Login Open Athens login
Purchase Options

Purchase this article to access the full text.

PPV Articles - American Journal of Psychiatry

PPV Articles - American Journal of Psychiatry

Not a subscriber?

Subscribe Now / Learn More

PsychiatryOnline subscription options offer access to the DSM-5-TR® library, books, journals, CME, and patient resources. This all-in-one virtual library provides psychiatrists and mental health professionals with key resources for diagnosis, treatment, research, and professional development.

Need more help? PsychiatryOnline Customer Service may be reached by emailing [email protected] or by calling 800-368-5777 (in the U.S.) or 703-907-7322 (outside the U.S.).

Media

Figures

Other

Tables

Share

Share

Share article link

Share