Next Article in Journal
What Can We Learn from FGF-2 Isoform-Specific Mouse Mutants? Differential Insights into FGF-2 Physiology In Vivo
Next Article in Special Issue
Immunological Dysfunction in Tourette Syndrome and Related Disorders
Previous Article in Journal
How the AHR Became Important in Cancer: The Role of Chronically Active AHR in Cancer Aggression
Previous Article in Special Issue
Sensitivity of Rodent Microglia to Kynurenines in Models of Epilepsy and Inflammation In Vivo and In Vitro: Microglia Activation Is Inhibited by Kynurenic Acid and the Synthetic Analogue SZR104
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuropsychiatric Disorders Due to Limbic Encephalitis: Immunologic Aspect

1
Department of Pediatrics, E-Da Hospital, Kaohsiung 82445, Taiwan
2
Department of Pediatrics, Shin Kong Wu Ho-Su Memorial Hospital, Taipei 11101, Taiwan
3
Department of Pediatrics, National Taiwan University Hospital, Taipei 100226, Taiwan
4
Department of Pediatrics, National Taiwan University College of Medicine, Taipei 100233, Taiwan
5
Graduate Institute of Brain and Mind Sciences, National Taiwan University College of Medicine, Taipei 100233, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(1), 389; https://doi.org/10.3390/ijms22010389
Submission received: 10 December 2020 / Revised: 26 December 2020 / Accepted: 28 December 2020 / Published: 31 December 2020
(This article belongs to the Special Issue Immunology of Neuropsychiatric Disorders)

Abstract

:
Limbic encephalitis (LE) is a rare cause of encephalitis presenting as an acute and subacute onset of neuropsychiatric manifestations, particularly with memory deficits and confusion as core features, along with seizure occurrence, movement disorders, or autonomic dysfunctions. LE is caused by neuronal antibodies targeting the cellular surface, synaptic, and intracellular antigens, which alter the synaptic transmission, especially in the limbic area. Immunologic mechanisms involve antibodies, complements, or T-cell-mediated immune responses in different degree according to different autoantibodies. Sensitive cerebrospinal fluid markers of LE are unavailable, and radiographic findings may not reveal a typical mesiotemporal involvement at neurologic presentations; therefore, a high clinical index of suspicions is pivotal, and a neuronal antibody testing is necessary to make early diagnosis. Some patients have concomitant tumors, causing paraneoplastic LE; therefore, tumor survey and treatment are required in addition to immunotherapy. In this study, a review on the molecular and immunologic aspects of LE was conducted to gain awareness of its peculiarity, which we found quite different from our knowledge on traditional psychiatric illness.

1. Introduction

Limbic encephalitis (LE) is an inflammatory encephalitis involving the limbic system, which encompasses the medial temporal lobe, hippocampus, and frontobasal and cingulate cortex [1]. LE is frequently associated with antibodies against the neuronal cell surface, synaptic vesicles, and intracellular proteins; therefore, it belongs to the autoimmune encephalitis (AE) category. Triggering factors include tumors, viral infections, or immune check point inhibitors (ICI’s) [2,3]. LE most often occurs in adults older than 45 years but can affect people of all ages. In addition, gender predominance varies with the type of antibody.
The clinical hallmark of LE includes acute and subacute onset of memory and cognitive deficits. Other symptoms include confusion, psychiatric symptoms (such as anxiety, depression, or psychosis), behavioral changes, seizures, movement disorders (such as ataxia, dystonia, or myoclonus), autonomic disturbances, and sleep disturbances [1,4]. The amygdala is a core region of the limbic system involved in the control of positive and negative affect, modulation of memory, and social and cognitive functions as well as behavioral adaptation to stress [5], which explains the core neuropsychiatric manifestations of LE. The basolateral complex of the amygdala is the main input site for sensory information from the thalamus and cortical regions, which plays a central role in seizure generation, particularly in the temporal lobe epilepsy [6].
According to Graus et al., the diagnosis of LE requires four criteria: (1) subacute onset (<3 months) of cognitive deficits, seizures, or psychiatric symptoms; (2) brain abnormalities in the medial temporal lobes on T2-weighted and fluid-attenuated inversion recovery magnetic resonance imaging (FLAIR MRI) images; (3) cerebrospinal fluid (CSF) pleocytosis (>5 cells per mm3) or an electroencephalography (EEG) showing epileptic discharges or slow-wave activity involving the temporal lobes; and (4) exclusion of alternative causes [1]. With time, the radiographic features of swelling and hyperintensity in the limbic structures may progress to mesiotemporal atrophy [7]. Some cases showed abnormalities in the basal ganglia, particularly those with leucine-rich glioma inactivated 1 (LGI1) antibody presenting with faciobrachial dystonic seizures [8,9]. In some cases, MRI may show only unilateral limbic or extralimbic involvement, and even normal finding, making the diagnosis a challenge. Fluorodeoxyglucose (FDG)-positron emission tomography (PET) is more sensitive than EEG and MRI in detecting abnormalities in LE, particularly in LGI1 and contactin-associated protein-like 2 (CASPR2) encephalitis [9,10]. A positive neuronal antibody is frequent but not mandatory for diagnosis as around 7–26% of LE cases had no detectable antibodies and are referred to as seronegative [11]. Positive antibody results are useful in diagnosing patients with atypical features or those who do not fulfill the Graus’s criteria [1].

2. Antibodies

Antibodies associated with LE are thought to be of peripheral origin that penetrates a leaky blood–brain barrier (BBB) or to be synthesized intrathecally. Most AE has higher serum than CSF antibody levels, implicating that antibodies are likely to be initiated by a peripheral immune response [12]. The autoantibodies are predominantly of IgG1 subclass. However, LGI1 and CASPR2 antibodies are mainly IgG4 subclass [13,14]. IgG4 antibody is hetero-bispecific (continuously undergoing half antibody exchange), hence less effective than IgG1 in crosslinking and internalizing the target antigen and does not fix complement [15]. In general, IgG1-associated AE (e.g., N-methyl-D-asparate receptor [NMDAR], γ-aminobutyric acid-B receptor [GABA-BR], and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate receptor [AMPAR]) have more inflammatory signs than IgG4-associated AE (e.g., LGI1 and CASPR2) [16]. Nevertheless, complement-mediated neuronal loss was still observed in LGI1 and CASPR2-associated neurologic syndromes [17,18]. These patients had higher complement-fixing IgG1 subclass with more prolonged clinical course, more severe cognitive impairment, and frequent hippocampal atrophy in advanced stages of the disease [13,19].
Antibodies associated with LE target: (1) cell-surface receptors (GABA, AMPA, and glycine receptors); (2) ion channels (voltage-gated potassium channels [VGKC]); (3) neighboring proteins that stabilize channel complex into the membrane (LGI1 and CASPR2); (4) enzymes that catalyze the formation of neurotransmitters glutamic acid decarboxylase (GAD); and (5) intracellular proteins (Hu, Ma2, collapsin response-mediator protein-5 [CRMP5], and amphiphysin) [20]. These antibodies are classified as neuronal cell surface, synaptic vesicle, and intracellular antibodies according to the location of their targeting antigens (Table 1). Antibodies against intracellular antigens are also known as onconeural antibodies because of their large association with tumors, and cause paraneoplastic syndromes (PNS) [21]. Pathologic effects include direct blockage of receptors and ion channels, indirect disruption with neighboring molecules interaction, and crosslinking and internalization of receptors to deplete them from the cell surface (Figure 1).
Antibodies may activate complements and natural-killer (NK) cells to induce complement-dependent cytolysis or antibody-dependent cell-mediated cytotoxicity, resulting in neuronal death. The cell-surface and synaptic antibodies alter neurotransmission leading to neuronal dysfunction, whereas intracellular antibodies are epiphenomenon of T-cell-mediated immune response rather than direct pathogenic mediators of neurological diseases. Intracellular (onconeural) antibodies are considered biomarkers for the presence of tumors [22,23]; therefore, aggressive tumor survey is necessary. Removal of tumors is mandatory to ameliorate PNS and enhance treatment response. Nevertheless, intracellular antibodies can be detected in patients without identifiable tumors or with cancer without PNS [24].
LE with cell-surface and synaptic antibodies can also be triggered by cancer, although this association has lower frequency. The most common antibodies affecting the limbic system are antibodies against LGI1, GAD, GABA-BR, AMPAR, CASPR2, Ma2, and Hu [11,25]. A latest systemic review comprising 24 studies with 263 Asian patients with LE showed that 53% had LGI1, 43% had GABA-BR, and 4% had CASPR2 antibodies [26]. About 7% of patients with LE were seronegative [11]. Rarely, ≥2 autoantibodies coexist in the same patient. Patients with LE having double or triple antibodies have worse outcomes than those with single antibody because of a higher chance of concomitant tumors [27,28,29].
In addition, patients with LE may have other comorbid autoimmune diseases, which should also be evaluated and treated. For example, about half of patients with Hashimoto encephalopathy had anti-NH2-terminal of α-enolase antibodies, and some manifested as LE [30].

3. Immunopathology

LE with antibodies against cell-surface antigens act mainly through antibody and/or complement-mediated mechanisms. However, LE with antibodies against intracellular antigens is predominantly mediated by cluster of differentiation (CD) 8+ cytotoxic T cells with frequent neuronal loss [17,25]. T cells are the major immune cells in the CSF, which play a major immunosurveillance role in the central nervous system (CNS) [31]. In all AE cases, CD3+ lymphocytes comprised the majority of inflammatory cells within the brain. A significant difference in the proportion of parenchymal CD8+ T cells (CD8/CD3 ratio) between intracellular and surface antibodies (mean 75% vs. 43%, respectively) was noted. The percentage of CD8+ T cells in the GAD antibody-mediated disease is intermediate (54%) [17]. The CD8+ T cells cause impairment of neuronal excitability and integrity with neuronal degeneration via two independent pathways: (1) granule cytotoxicity by perforin and granzyme-B and (2) ligation of death receptors [32,33]. Perforin is a Ca2+-dependent protein, which can form transmembranous pores, alter electrical excitability and signaling, and cause neuronal necrosis with swelling and rupture of cell membranes. Perforin mediates the delivery of granzymes into the target cells to promote apoptosis [34]. Apposition of multiple granzyme-B lymphocytes to single neurons, which is consistent with a specific cytotoxic T-cell attack, was observed in several onconeural antibody-mediated diseases. Therefore, higher CD8/CD3 ratio and more frequent appositions of granzyme-B+ cytotoxic T cells to neurons were associated with greater neuronal loss [17]. Cytotoxic T cells can also liberate neurotoxic cytokines, such as interferon-γ (IFN- γ) and tumor necrosis factor-α as well as excitotoxic glutamate [35]. CD8+ T cells also destroy the myelin sheath or glial cells in both white and gray matter [32].
In addition to antibodies and lymphocytes, microglia and astrocytes also play roles in mediating neuronal damages in the hippocampus and other brain regions. Microglia and astrocytes are required for synaptic integrity structurally and synaptic transmission functionally [36]. Activated microglia have dual functions, either anti-inflammatory or pro-inflammatory [37]. Previous studies showed that activated microglia in hippocampus may be associated with chronic epilepsy and was also found in VGKC-associated LE [38,39,40]. Immunotherapy may decrease the activation of microglia leading to improvement in seizure. Astrocytes and their autoimmune glial fibrillary acidic protein (GFAP) have also been shown to be associated with AE including NMDAR encephalitis and other encephalitis mimicking LE [41,42], and play important roles in neuroinflammation. The activation of astrocytes may trigger an astrocyte-neuron signaling cascade leading to persistent functional change in hippocampal excitatory synapses [43], which may be associated with cognitive impairment.

4. Inflammatory Mediators

Conventional CSF inflammatory markers such as white blood cells or protein are neither sensitive nor specific for LE, although the detection of oligoclonal band may increase diagnostic sensitivity [44]. A lower CD4/CD8+ T-cell ratio is detected in the CSF of all patients with LE [45].
The serum and CSF levels of C-X-C motif chemokine ligand 13 (CXCL13) were significantly higher in patients with LGI1 encephalitis [46]. The serum levels of CXCL10 were elevated in CASPR2 encephalitis. CXCL10 is a cytokine that recruits C-X-C motif chemokine receptor 3 cells such as activated T cells [47]. CASPR2 encephalitis seems to elicit a higher immune response than LGI1 encephalitis, with more intrathecal IgG and higher cytokine levels, indicated by higher CXCL13 and soluble intercellular adhesion molecule-1 (sICAM1) in the CSF. CXCL13 points to a B-cell mediated, whereas sICAM1 to a T-cell mediated neuroinflammation [48]. Patients with PNS also had high levels of chemokine CXCL10 in the CSF with the presence of IFN-γ receptors on their T cells [47].
CSF mediators are different in infectious and immune-mediated encephalitis [49]. CSF cytokines IL-21 and IFN-γ-induced protein 10 kDa (IP10) are promising in differentiating between viral encephalitis and AE [50]. IL-21 causes autoantibody production, which downregulates regulatory T cells leading to enhanced autoimmunity and increased CD8+ T cells and NK cells in AE [51], whereas IP10/CXCL10 is secreted in response to IFN-γ, which is produced as part of the Th1 in response to viral infection [52].

5. Genetic Factors

Human leukocyte antigen (HLA) is the main genetic factor related to autoimmunity. Anti-LGI1 encephalitis is strongly associated with HLA class II allele DRB1*07:01, which was carried by approximately 90% of patients, as well as B*44:03 and C*07:06 in the HLA class I allele [53,54]. DRB1*11:01 was detected in approximately 50% of the patients with CASPR2-antibody syndrome [54]. Anti-GAD neurological syndromes have frequent association with HLA class II haplotype DQA1*05:01-DQB1*02:01-DRB1*03:01 [55].

6. Cancers

Cancers occur in 20–60% of patients with LE [11,56]. LE can be classified as paraneoplastic or nonparaneoplastic, based on the presence or absence of an underlying malignancy [57].
Genetic alterations in tumor cells may trigger immune tolerance breakdown with initiation of PNS [58]. In PNS, the target antigens such as Hu, Ma, Yo, Ri, CRMP5, amphiphysin, and SOX-1 are proteins normally expressed by neurons and expressed ectopically by tumor cells. These antigens are recognized by onconeuronal antibodies or presented to T cells, which result in both cellular and humoral immune responses against neural structures [3,59]. The most common PNSs are LE and cerebellar degeneration, each accounting for approximately one-third of all PNSs [60]. Lung cancer, particularly small cell lung cancer (SCLC), testicular germ-cell tumor, thymoma, and lymphoma are the most frequent associated tumors [60,61,62]. Almost all patients with LE caused by Hu antibodies have SCLC [22]. Approximately 60% of patients with neurological syndromes associated with GABA-BR antibodies have underlying SCLC, and 60% of those with AMPAR antibodies have non-SCLC, breast cancer, or thymic tumors [63]. The presence of tumor is rare in patients with VGKC antibodies, especially LGI1 antibody, but is more common in patients with peripheral nervous system involvement such as Morvan syndrome or neuromyotonia caused by CASPR2 antibody. In the latter condition, a thymoma is common [64,65]. Therefore, screening for malignancy is usually required in patients with LE because the coexistence of tumors and their treatment influence the clinical outcome [61]. If initial tumor screening is negative, total-body FDG-PET is recommended [16], or screening can be repeated after 3–6 months, followed by screenings every 6 months for 4 years [66].

7. Infection

Occurrence of AE during the convalescence phase of viral CNS infections, particular NMDAR encephalitis after herpes simplex encephalitis (HSE), has been well-known [67]. Patients with HSE may also develop antibodies to GABA-BR, AMPAR, dopamine-2 receptors, or concurrently to NMDAR and LGI1, and manifest relapses of neurologic symptoms [2,68,69]. Proposed mechanisms include (1) induction of self-immunization due to a viral-induced inflammatory response leading to the release of neural tissue epitopes such as NMDAR and (2) molecular mimicry based on sequence homology or structural similarities between viral protein and neural tissue epitopes [20,70]. Human herpesvirus 6 was reported as a cause of LE in immunocompromised hosts after hematopoietic stem cell transplant [71] or triggering factor of GABA-BR encephalitis [2].

8. Immune Checkpoint Inhibitors

In addition to developing PNS, patients with cancers receiving biological cancer therapies with ICI’s are at risk for developing peripheral and CNS complications [72,73]. The main ICI’s are ipilimumab targeting the cytotoxic T-lymphocyte antigen 4 and pembrolizumab and nivolumab inhibiting the programmed cell death 1 pathway [74]. LE is the most frequent CNS complication of ICI’s [73]. Antibody-positive (CASPR2, Hu, Ma2, and GAD65 antibodies) or antibody-negative LE may develop weeks to 1 year after starting ICI’s therapy [75,76,77,78]. Anti-Ma2 antibody is the most common antibody found in ICI-induced LE, which carries a poor outcome [73].

9. Antigens Associated with LE

9.1. Voltage-Gated Potassium Channel Complex (VGKC)

VGKCs are located on the neuronal membranes of both the central and peripheral nervous system. The VGKC complex is composed of Kv1 subunits, which function with other extracellular proteins including LGI1, CASPR2, and contactin 2, membranous proteins a disintegrin and metalloproteinase 22 and 23 (ADAM22 and 23), and intracellular scaffolding proteins including postsynaptic density protein 95 (PSD-95) [79]. VGKC antibody is categorized into four subgroups: LGI1 positive, CASPR2 positive, VGKC-positive group lacking both LGI1 and CASPR2 antibodies (double-negative VGKC) [19], and rarely double positive group [8]. The extracellular proteins LGI1 and CASPR2 are direct antigenic epitopes of VGKC antibodies, whereas in double-negative VGKC, antibody may target another extracellular protein, the contactin-2 or intracellular epitopes [80]. VGKC antibodies cause VGKC hyperexcitability spectrum, including LE, Morvan syndrome, neuromyotonia (Isaacs’ syndrome), seizures, sleep disturbance, and autonomic dysfunction [81,82]. Patients with VGKC complex LE were more likely to have an autoimmune inflammation than patients with other syndromes [83], and may show perivascular lymphocytic infiltration and neuronal loss predominantly in the hippocampus and amygdala [38]. A complement-mediated neuronal death may play a prominent role [17]. A positive VGKC antibody, especially if tested by radioimmunoprecipitation assay, has uncertain pathogenicity and limited clinical specificity [8,84]. Testing LGI1 and CASPR2 antibodies separately is advocated instead of testing VGKC antibody alone [85].

9.2. Leucine-Rich, Glioma Inactivated 1 Protein (LGI1)

LGI1 is a secreted synaptic protein, which interacts with presynaptic and postsynaptic ADAM to form a trans-synaptic linkage between presynaptic ADAM23-Kv1 channel and postsynaptic ADAM22-PSD-95. LGI1 is important for AMPAR-mediated synaptic transmission [86]. LGI1 is located mainly in the hippocampus and temporal cortex, and the LGI1-ADAM22-AMPAR interaction influences both long-term depression (LTD) and long-term potentiation (LTP) [87]. LGI1 secreted from the excitatory pyramidal neurons contributes to LGI1-related epileptogenesis [88]. LGI1 antibodies inhibit binding of LGI1 to ADAM22/23, downregulate Kv1.1, reduce AMPAR clustering, and alter synaptic excitability, plasticity, and memory [89,90]. Faciobrachial dystonic seizures are very specific for LGI1-encephalitis, but only present in about 50% of patients [91]. One-third of patients had peripheral nervous system manifestations, commonly with pain and dysautonomia [8]. Comorbid tumors (mainly thymoma and SCLC) [92] occurred in <10% of patients with anti-LGI1 LE.

9.3. Contactin-Associated Protein-Like 2 (CASPR2)

CASPR2 is a transmembrane cell adhesion protein of the neurexin family that is located at the juxtaparanodal region of the nodes of Ranvier in myelinated axons of the central and peripheral nervous system. CASPR2 interacts with dimerized contactin-2 (also known as transient axonal glycoprotein-1) [93,94] and colocalizes with VGKC Kv1.1 and Kv1.2 [95]. CASPR2 antibodies do not reduce CASPR2 expression but inhibit the interaction of CASPR2 with its binding partner contactin 2 [96] and reduce the clustering and surface expression of Kv1, thereby interfering with axonal excitability and action potential conduction [93,97]. CASPR2 antibodies did not cause evident neuronal loss, except for a modest Purkinje cell loss in the cerebellum [12].
CASPR2 antibody-associated diseases almost exclusively affect older males. 80% of patients have neuropsychiatric features such as cognitive deficits, and 50% developed seizures. Many others presented with neuropathic pain associated with peripheral neuronal hyperexcitability such as myokymia, fasciculations, and muscle cramps [19,98]. Other syndromes include autonomic dysfunction, cerebellar symptoms, insomnia, or weight loss [19]. Cancer is more frequent in CASPR2 compared to LGI1-antibody positive patients, more commonly with thymoma [8].

9.4. AMPA Receptor

AMPARs are glutamatergic ionotropic transmembrane receptors composed of four subunits (GluA1–4) [99]. AMPARs are concentrated in the hippocampal CA3–CA1 region, subiculum, basal ganglia, cerebellum, and the cerebral cortex. They mediate most of the fast excitatory synaptic transmission in the brain and are critical for synaptic plasticity, learning, and memory [100]. LE patients’ antibodies usually target GluA1 or GluA2 subunits of AMPARs, which decrease AMPAR clustering at synapses through increased internalization of receptors but do not alter the synaptic density or cell viability [101,102]. In accordance with the decrease in AMPA-mediated excitatory postsynaptic currants, affected neurons also exhibit homeostatic decrease in inhibitory postsynaptic currants and reduced inhibitory synapse density, whereas the intrinsic neuronal excitability increases [102]. AMPAR antibodies do not cause significant neuronal apoptosis. Synaptic AMPAR cluster density was restored within a few days after removal of patients’ antibodies, and patients usually recovered to baseline after treatment [101].
In AMPAR encephalitis, confusion is the most frequent manifestation. Seizure occurs in approximately one-third of patients. Associated cancers were found in approximately 50–60% of patients, most commonly lung carcinoma and thymoma [27,103].

9.5. Metabotropic Glutamate Receptor 5 (mGluR5)

Glutamate receptors (GluRs) are the main excitatory synaptic transmission mediators in the brain. GluRs can be divided into ligand-gated ion channels (ionotropic) and G protein-coupled (metabotropic) GluRs (mGluRs). In the three groups of mGluRs, mGluR5 belongs to group I. mGluR5 are expressed primarily in the hippocampus and amygdala, which modulates LTD and LTP [104]. mGluR5 antibodies cause a decrease of mGluR5 cluster density at both synaptic and extrasynaptic locations [105]
All patients with anti-mGluR5 encephalitis had CSF pleocytosis and 75% had oligoclonal bands. MRI abnormalities revealed not only the limbic system involvement, but also the involvement of extralimbic regions such as thalamus, pons, frontal or parieto-occipital cortex, and cerebellum [105]. mGluR5 antibody is associated with LE and Hodgkin lymphoma, presented as Ophelia syndrome in nearly 50% of patients [105,106].

9.6. Dipeptidyl Peptidase-Like Protein 6 (DPP6)/DPPX

DPP6 (DPPX) is a membrane glycoprotein and an auxiliary subunit of the voltage-gated A-type (rapidly inactivating) potassium channel Kv4.2 [107]. Kv4.2 is involved in somatodendritic signal integration and attenuation of back-propagation of action potentials [107], which is responsible for transient, inhibitory currents in the peripheral and CNS [108]. DPPX is predominantly expressed in the hippocampus, cerebellum, and striatum [109]. DPPX antibodies are predominantly of IgG4 subtype. Loss of DPPX modulation of potassium currents leads to neuronal hyperexcitability [108].
Most patients with DPPX encephalitis had a combination of LE, brainstem dysfunction, severe diarrhea, and weight loss [107,110,111]. Many patients have characteristic gastrointestinal symptoms even before neuropsychiatric symptoms [111] because of the enriched expression of DPPX in myenteric plexus [112]. The course of neuropsychiatric symptoms in anti-DPPX encephalitis was prolonged, and chronic immunotherapy is often required [107].

9.7. GABA-B Receptor

GABA-BRs are transmembrane G-protein-coupled receptors composed of two subunits, the GABA-B1 and GABA-B2, and mediate slow and prolonged inhibitory neurotransmission [113]. GABA-BRs are ubiquitously distributed in the brain and spinal cord, but the highest levels are found in the hippocampus, thalamus, and cerebellum [114]. Most GABA-BR antibodies recognized the GABA-B1 subunit. GABA-BR antibodies do not decrease the synaptic levels of receptors but alter their synaptic function.
Co-expression of GABA-BR with other autoantibodies is relatively common [63,115,116,117,118]. Reported coexisting antibodies in GABA-BR encephalitis included potassium channel tetramerization domain containing 16 (KCTD16), GAD65, Hu, Ri, amphiphysin, SOX1, CV2/CRMP, 5N-type or P/Q-type voltage-gated calcium channels, or thyroid peroxidase [115,116,117,119,120]. Most of these cases were paraneoplastic.
Anti-GABA-BR encephalitis have highest rates for pleocytosis, protein concentration, and positive oligoclonal bands compared to other LE [121]. Increased CSF glutamate levels were found in GABA-BR associated encephalitis [122]. Increased CD138+ and CD19+ plasma cells and activated cytotoxic T cells in CSF corresponded to higher overall neuropsychological and memory deficits in patients with GABA-BR LE. A hippocampal specimen also revealed perivascular infiltrates of CD138+ plasma cells and CD4+ T cells, whereas cytotoxic CD8+ T cells were detected within the brain parenchyma in close contact to neurons [123].
Patients with GABA-BR LE had seizure, confusion, and memory deficits. 80–100% of patients presented an early or prominent seizure, and a large proportion had convulsive or nonconvulsive status epilepticus. Seizure was often of temporal lobe onset, usually refractory to medications but was well responsive to immunotherapy [63,115,116,120]. The seizure phase was followed by an encephalitic phase compatible with LE such as confusion and memory deficits. More than 50% of patients had tumors, mostly SCLC [115,116,124].

9.8. Glutamic Acid Decarboxylase (GAD)

GAD is a rate-limiting enzyme in presynaptic inhibitory neurons for synthesis of the inhibitory neurotransmitter GABA from glutamate, using pyridoxal-5′-phosphate as cofactors. GAD is widely distributed within the CNS and pancreas [125]. There are two isoforms: cytoplasmatic, a constitutively active isoform of 67 kDa (GAD67) for maintaining steady basal production of GABA and synaptic membrane-associated form of 65 kDa (GAD65), which undergoes auto-inactivation during enzyme activity and provides pulse production when rapid surge of GABA is required [126]. GAD65 is localized predominantly in nerve terminals, anchored to the cytoplasm-facing side of synaptic vesicles [127]. Compared with GAD67, GAD65 has greater antigenicity partly due to its surface electrostatic charge and greater mobility in the C-terminal and catalytic domains [127]. GAD is mostly found intracellularly but transiently exposed extracellularly during the dynamic process of neurotransmission and exocytosis [128].
GAD antibodies disrupt GABAergic signaling. GAD65 antibodies are detected in about 8% of general population, but always at low level [129]. GAD antibodies are frequently found in patients with type 1 diabetes mellitus (DM1) (80–90%), and around 40% of patients had been diagnosed with DM1 before the onset of neurologic symptoms [125,130]. Patients with neurologic syndromes associated with GAD usually have serum titers several hundred-fold higher than those with DM1 [131]. GAD antibodies from patients with DM1 did not react with brain tissue [132]. In rats, in vivo injection of GAD65 antibodies from patients with stiff-person syndrome (SPS) induced electrophysiologic changes in myelinated neurons, whereas GAD65 antibodies from patients with DM1 did not [133]. Antibodies seem to recognize different GAD65 epitopes in patients with DM1 and neurologic syndromes [134]. Pathological specimens showed T-cell infiltrates but no IgG deposition as well as acute necrosis and neuronal loss in the hippocampus, evolving to hippocampal atrophy [17,135]. Loss of GABAergic neurons in the brainstem has been observed [136].
Most patients harboring GAD65 antibodies are women, with a mean age of 23 years, which is younger compared to LE caused by other antibodies. According to one small pediatric case series of LE, 50% of patients had GAD antibodies [137]. Neurological syndromes associated with GAD antibodies include SPS, cerebellar ataxia, limbic and extralimbic encephalitis, nystagmus/oculomotor dysfunction, drug-resistant epilepsy, paraneoplastic SPS, and progressive encephalopathy with rigidity and myoclonus. The latter two are mainly related to amphiphysin and glycine receptor antibodies, respectively [138]. Seizures occurred in almost all patients with GAD-related LE, may present as status epilepticus, and were difficult to control even with immunotherapy [139,140,141].
GAD antibodies were detected in 7–17% of adult patients with LE [135]. Previous study reported a probable <15% cancer association in GAD antibody-associated LE [135]. GAD antibody-associated LE is ten times more likely to be paraneoplastic than other neurologic syndromes such as SPS or cerebellar ataxia. The most frequent tumors are lung and thymic neoplasms. Patients with PNS associated with GAD were older, male predominance, having coexisting antibodies than non-neoplastic GAD cases [116,119,142]. Cancer risk is seven times higher in patients with GAD and coexisting antibodies against neuronal cell-surface antigens, especially GABA-BR [119]. The prognosis varies and is considered intermediate between that of neurologic syndromes associated with antibodies against cell-surface and intracellular antibodies [138]. Approximately 70% of patients with high GAD65 antibody concentration improved partially after immunotherapy; however, none reached a complete recovery [130].

10. Intracellular Antibodies

Intracellular antigens such as Ma, Hu, Ri, Yo, CRMP5, amphiphysin, or SOX1 are considered inaccessible by their antibodies [143]. Intracellular antibodies are usually considered onconeural antibodies due to their strong association with tumors. Anti-Hu antibodies were the most frequently detected antibodies in PNS associated with onconeural antibodies. Cerebellar degeneration was the most frequent syndrome in PNS associated with onconeural antibodies, whereas brainstem encephalitis and subacute sensorimotor neuronopathy were the most frequent nonclassical syndromes [144]. LE is a rare neurologic presentation in this entity. Onconeural antibodies associated with paraneoplastic LE are Ma, Hu, Ri, CRMP5, and SOX1 [66,145], with Ma2 being the major intracellular antibody causing paraneoplastic LE [146]. In paraneoplastic LE associated with lung cancer, Hu and cell surface GABA-BR antibodies were the most frequently encountered antibodies, whereas in paraneoplastic LE associated with testicular tumor, Ma2 antibody is the most frequent antibody [147,148].

11. Treatment

11.1. First-Line Treatment

First-line treatment (Table 2) includes pulse therapy with intravenous methylprednisolone, intravenous immunoglobulin (IVIg), or plasmapheresis. Methylprednisolone is usually given first, followed by either IVIg or plasmapheresis within 3–7 days depending on clinical features and disease severity [16].
Although IVIg and plasmapheresis neutralize or eliminate systemic antibodies, they do not target plasma cells or intrathecal antibody synthesis. Both plasma exchange (PE) and immunoadsorption (IA) resulted in clinical improvement in 60–67% in patients with AE associated with NMDAR, LGI1, CASPR2, mGluR5, GAD, and Hu antibodies [149,150,151]. Plasmapheresis was more effective for encephalitis caused by cell-surface antigens (approximately 85%) than by intracellular or synaptic antigens (66.7%) [149,150,151]. Response rate was higher with early initiation of therapy [152].
In patients with encephalitis after ICI’s treatment, early immunotherapy with steroids or IVIg is essential [153].

11.2. Second-Line Treatment

Second-line treatment includes rituximab, cyclophosphamide, azathioprine, mycophenolate mofetil (MMF), or others. It is usually started within approximately 2 weeks when the first-line therapies fail to achieve improvement [19,154], or when the disease is severe or relapsing.
Rituximab and cyclophosphamide are the preferred second-line treatment. Rituximab is a monoclonal antibody targeting the CD20 expressed by B cells and plasmablasts. The lifetime of antibody-producing plasma cells is typically days to weeks [155], which accounts for the delayed response of rituximab [156]. In LE with common relapses such as AMPAR encephalitis, administration of second-line therapy resulted in less relapse [157]. Encephalitis caused by DPPX usually require rituximab [111].
Cyclophosphamide, a deoxyribonucleic acid alkylating agent, targets hematopoietic cells [158], which can cross the BBB and affects both B and T cells as well as the intrathecal antibody synthesis. In GAD-associated refractory status epilepticus, cyclophosphamide resulted in seizure control and decreased intrathecal production of antibody [159]. Both rituximab and cyclophosphamide target the antibody source, causing fewer relapses, and can be combined in most severe cases. However, combination therapy carries risks of progressive multifocal leukoencephalopathy or reactivation of infections such as hepatitis B [160].
Tocilizumab is an anti-IL-6 receptor monoclonal antibody. IL-6 is a key mediator for the survival of plasmablasts and plasma cells. IL-6 not only induces B-cell differentiation and proliferation, but also promotes the differentiation of IL-17-producing helper T cells and cytotoxic T cells [161,162]. For patients with AE having inadequate clinical response to the first-line therapy and rituximab, tocilizumab had been used with favorable outcome. Nearly 90% of the tocilizumab-treated patients maintained a long-term favorable clinical response [163]. Tocilizumab also resulted in improvement in LE cases caused by CASPR2 antibodies [164] and in pediatric case of AE caused by GAD antibodies refractory to first-line treatment [165].
Daratumumab, an anti-CD38 antibody that depletes plasma cells and modifies various T-cell functions, was used in a refractory case of CASPR2 encephalitis. Daratumumab resulted in antibody titers reduction, T-cell activation markers normalization, and neurological improvement. However, the reported patient died of septicemia [166].
Bortezomib, a proteasome inhibitor that interferes with NF-κB and ubiquitin-proteasome pathway to deplete plasma cells, is currently under clinical trial for patients with severe AE (NCT03993262).

11.3. Maintenance Therapy

Very limited evidence on the efficacy of long-term therapy were reported. In cases unresponsive to first or second-line immunotherapy, or in cases with relapses requiring maintenance therapy, drugs such as azathioprine, mycophenolate mofetil (MMF), or methotrexate can be utilized [167].
Monthly intravenous or oral steroid pulses, oral steroids with tapering dose, monthly IVIg, or rituximab redosing could also be tried [168]. In patients with LE having high relapse rates such as those with LGI1 and CASPR2 antibodies, a gradual corticosteroid tapering over around 18 months can decrease relapses [169].
In general, LE associated with cell-surface antibodies are more likely to respond to immunotherapy, resulting in good recovery in up to 70–80% of cases [56]. In PNS caused by onconeural antibodies, the tissue damage is thought to be primarily cell-mediated; therefore, immunotherapy often has only little effect. The most important treatment is prompt and complete removal of tumor in order to withdraw the auto-antigen expressed by the tumor that triggers the production of autoantibodies [61]. In treatment-resistant cases, gradual neuronal loss and brain atrophy were noted [25]. For these cases, if initial tumor screening is negative, tumor surveillance must be continued at follow-up, particularly depending on the antibody [16]. In addition to intracellular antibodies, certain cell-surface antibodies (e.g., AMPAR, GABA-BR, and mGluR5) are more often associated with tumors. If antibodies are not found and the clinical presentation fulfills the criteria of LE, screening with computed tomography of the thorax is recommended, followed by a total-body FDG-PET [11]. In LGI1 patients, seizure including FBDS usually improved after immunotherapy. However, a large proportion (60%) of GAD-LE patients was not seizure free.

12. Prognosis

The general mortality rate of LE was 9.7%, largely due to pulmonary infection or status epilepticus. GABA-BR antibody was attributed to the highest mortality rate (23.2%), largely due to tumor progression [26]. Cognitive dysfunction is a major complication in patients with long-term LE, and some patients are left disabled [56]. In LE caused by GAD, LGI1, or CASRP2 antibodies, the memory outcomes were poor even after clinical recovery with immunotherapy. Pathologic tearfulness was reported in 50% of patients, which correlated with changes in certain emotional brain networks, although not part of the neuropsychiatric comorbidities [170].

13. Conclusions

The majority of patients with LE have prominent neuropsychiatric symptoms at disease onset, which is quite common in clinical practice. Therefore, patients with acute or subacute neuropsychiatric symptoms with the core symptom of memory deficits, particularly if associated with seizures or abnormal signal changes on brain MRI, particularly in the mesiotemporal region, should be highly suspected for LE (Figure 2 and Table 3). A diagnosis of LE is extremely important because timely immunotherapy is even more important than symptomatic therapies with antipsychotic or antiepileptic drugs, because partial or complete remission or even cure can be achieved.

Author Contributions

Writing—original draft, Y.-C.K.; writing—review and editing, M.-IL., W.-C.W. and W.-T.L.; concept and design—W.-T.L.; funding acquisition, Y.-C.K. and W.-T.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the following research projects: 100-EDN03, 101-EDN01 and NCKUEDA10211.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The study was supported by grants from E-Da Hospital National Taiwan University Hospital Joint Research Program (100-EDN03, 101-EDN01) and National Cheng-Kung University E-Da Hospital Joint Program (NCKUEDA10211).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AEautoimmune encephalitis
AK5adenylate kinase 5
AMPARα-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor
BBBblood-brain barrier
BGbasal ganglia
CASPR2contactin-associated protein-like 2
CJDCreutzfeldt-Jakob disease
CNScentral nervous system
CRMP5collapsin response-mediator protein-5
CSFcerebrospinal fluid
DPPX/DPP6dipeptidyl peptidase-like protein 6
EEGelectroencephalography
FDG-PETfluorodeoxyglucose-positron emission tomography
FLAIRfluid-attenuated inversion recovery
GABA-BRγ-aminobutyric acid B receptor
GADglutamic acid decarboxylase
GlyRglycine receptor
HIVhuman immunodeficiency virus
HSVherpes simplex virus
HHV6human herpesvirus 6
ICIimmune checkpoint inhibitor
ILinterleukin
INF-γinterferon-γ
LElimbic encephalitis
LGI1leucine-rich, glioma inactivated 1
mGluR5metabotropic glutamate receptor subtype 5
MRImagnetic resonance imaging
NKnatural killer
NMDARN-methyl-D-aspartate receptor
OCBoligoclonal band
PNSparaneoplastic syndrome
SLEsystemic lupus erythematosus
SOX1sex-determining region Y (SRY)-box 1
SPSstiff-person syndrome
VZVvaricella zoster virus

References

  1. Graus, F.; Titulaer, M.J.; Balu, R.; Benseler, S.; Bien, C.G.; Cellucci, T.; Cortese, I.; Dale, R.C.; Gelfand, J.M.; Geschwind, M.; et al. A clinical approach to diagnosis of autoimmune encephalitis. Lancet Neurol. 2016, 15, 391–404. [Google Scholar] [CrossRef] [Green Version]
  2. Linnoila, J.J.; Binnicker, M.J.; Majed, M.; Klein, C.J.; McKeon, A. CSF herpes virus and autoantibody profiles in the evaluation of encephalitis. Neurol. Neuroimmunol. Neuroinflamm. 2016, 3, e245. [Google Scholar] [CrossRef] [Green Version]
  3. Graus, F.; Dalmau, J. Paraneoplastic neurological syndromes in the era of immune-checkpoint inhibitors. Nat. Rev. Clin. Oncol. 2019, 16, 535–548. [Google Scholar] [CrossRef] [PubMed]
  4. Budhram, A.; Leung, A.; Nicolle, M.W.; Burneo, J.G. Diagnosing autoimmune limbic encephalitis. CMAJ 2019, 191, E529–E534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Pape, H.C.; Pare, D. Plastic synaptic networks of the amygdala for the acquisition, expression, and extinction of conditioned fear. Physiol. Rev. 2010, 90, 419–463. [Google Scholar] [CrossRef] [Green Version]
  6. Aroniadou-Anderjaska, V.; Fritsch, B.; Qashu, F.; Braga, M.F. Pathology and pathophysiology of the amygdala in epileptogenesis and epilepsy. Epilepsy Res. 2008, 78, 102–116. [Google Scholar] [CrossRef] [Green Version]
  7. Urbach, H.; Soeder, B.M.; Jeub, M.; Klockgether, T.; Meyer, B.; Bien, C.G. Serial MRI of limbic encephalitis. Neuroradiology 2006, 48, 380–386. [Google Scholar] [CrossRef]
  8. Gadoth, A.; Pittock, S.J.; Dubey, D.; McKeon, A.; Britton, J.W.; Schmeling, J.E.; Smith, A.; Kotsenas, A.L.; Watson, R.E.; Lachance, D.H.; et al. Expanded phenotypes and outcomes among 256 LGI1/CASPR2-IgG-positive patients. Ann. Neurol. 2017, 82, 79–92. [Google Scholar] [CrossRef]
  9. Liu, X.; Shan, W.; Zhao, X.; Ren, J.; Ren, G.; Chen, C.; Shi, W.; Lv, R.; Li, Z.; Liu, Y.; et al. The Clinical Value of (18) F-FDG-PET in Autoimmune Encephalitis Associated With LGI1 Antibody. Front. Neurol. 2020, 11, 418. [Google Scholar] [CrossRef]
  10. Boyko, M.; Au, K.L.K.; Casault, C.; de Robles, P.; Pfeffer, G. Systematic review of the clinical spectrum of CASPR2 antibody syndrome. J. Neurol. 2020, 267, 1137–1146. [Google Scholar] [CrossRef]
  11. Graus, F.; Escudero, D.; Oleaga, L.; Bruna, J.; Villarejo-Galende, A.; Ballabriga, J.; Barcelo, M.I.; Gilo, F.; Popkirov, S.; Stourac, P.; et al. Syndrome and outcome of antibody-negative limbic encephalitis. Eur. J. Neurol. 2018, 25, 1011–1016. [Google Scholar] [CrossRef] [PubMed]
  12. Giannoccaro, M.P.; Menassa, D.A.; Jacobson, L.; Coutinho, E.; Prota, G.; Lang, B.; Leite, M.I.; Cerundolo, V.; Liguori, R.; Vincent, A. Behaviour and neuropathology in mice injected with human contactin-associated protein 2 antibodies. Brain 2019, 142, 2000–2012. [Google Scholar] [CrossRef] [PubMed]
  13. Thompson, J.; Bi, M.; Murchison, A.G.; Makuch, M.; Bien, C.G.; Chu, K.; Farooque, P.; Gelfand, J.M.; Geschwind, M.D.; Hirsch, L.J.; et al. The importance of early immunotherapy in patients with faciobrachial dystonic seizures. Brain 2018, 141, 348–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Joubert, B.; Saint-Martin, M.; Noraz, N.; Picard, G.; Rogemond, V.; Ducray, F.; Desestret, V.; Psimaras, D.; Delattre, J.Y.; Antoine, J.C.; et al. Characterization of a Subtype of Autoimmune Encephalitis with Anti-Contactin-Associated Protein-like 2 Antibodies in the Cerebrospinal Fluid, Prominent Limbic Symptoms, and Seizures. JAMA Neurol. 2016, 73, 1115–1124. [Google Scholar] [CrossRef]
  15. Vidarsson, G.; Dekkers, G.; Rispens, T. IgG subclasses and allotypes: From structure to effector functions. Front. Immunol. 2014, 5, 520. [Google Scholar] [CrossRef] [Green Version]
  16. Zuliani, L.; Nosadini, M.; Gastaldi, M.; Spatola, M.; Iorio, R.; Zoccarato, M.; Mariotto, S.; De Gaspari, P.; Perini, F.; Ferrari, S.; et al. Management of antibody-mediated autoimmune encephalitis in adults and children: Literature review and consensus-based practical recommendations. Neurol. Sci. 2019, 40, 2017–2030. [Google Scholar] [CrossRef]
  17. Bien, C.G.; Vincent, A.; Barnett, M.H.; Becker, A.J.; Blumcke, I.; Graus, F.; Jellinger, K.A.; Reuss, D.E.; Ribalta, T.; Schlegel, J.; et al. Immunopathology of autoantibody-associated encephalitides: Clues for pathogenesis. Brain 2012, 135, 1622–1638. [Google Scholar] [CrossRef] [Green Version]
  18. Bauer, J.; Bien, C.G. Neuropathology of autoimmune encephalitides. Handb. Clin. Neurol. 2016, 133, 107–120. [Google Scholar] [CrossRef]
  19. van Sonderen, A.; Petit-Pedrol, M.; Dalmau, J.; Titulaer, M.J. The value of LGI1, Caspr2 and voltage-gated potassium channel antibodies in encephalitis. Nat. Rev. Neurol. 2017, 13, 290–301. [Google Scholar] [CrossRef]
  20. Alexopoulos, H.; Dalakas, M.C. The immunobiology of autoimmune encephalitides. J. Autoimmun. 2019, 104, 102339. [Google Scholar] [CrossRef]
  21. Lancaster, E.; Dalmau, J. Neuronal autoantigens—Pathogenesis, associated disorders and antibody testing. Nat. Rev. Neurol. 2012, 8, 380–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. van Coevorden-Hameete, M.H.; de Graaff, E.; Titulaer, M.J.; Hoogenraad, C.C.; Sillevis Smitt, P.A. Molecular and cellular mechanisms underlying anti-neuronal antibody mediated disorders of the central nervous system. Autoimmun. Rev. 2014, 13, 299–312. [Google Scholar] [CrossRef] [PubMed]
  23. Irani, S.R.; Gelfand, J.M.; Al-Diwani, A.; Vincent, A. Cell-surface central nervous system autoantibodies: Clinical relevance and emerging paradigms. Ann. Neurol. 2014, 76, 168–184. [Google Scholar] [CrossRef] [PubMed]
  24. Pittock, S.J.; Kryzer, T.J.; Lennon, V.A. Paraneoplastic antibodies coexist and predict cancer, not neurological syndrome. Ann. Neurol. 2004, 56, 715–719. [Google Scholar] [CrossRef] [PubMed]
  25. Dalmau, J.; Geis, C.; Graus, F. Autoantibodies to Synaptic Receptors and Neuronal Cell Surface Proteins in Autoimmune Diseases of the Central Nervous System. Physiol. Rev. 2017, 97, 839–887. [Google Scholar] [CrossRef] [PubMed]
  26. Ghimire, P.K.U.; Gajurel, B.P.; Karn, R.; Rajbhandari, R.; Paudel, S.; Gautam, N.; Ojha, R. Anti-LGI1, anti-GABABR, and Anti-CASPR2 encephalitides in Asia: A systematic review. Brain Behav. 2020, 10, e01793. [Google Scholar] [CrossRef]
  27. Joubert, B.; Kerschen, P.; Zekeridou, A.; Desestret, V.; Rogemond, V.; Chaffois, M.O.; Ducray, F.; Larrue, V.; Daubail, B.; Idbaih, A.; et al. Clinical Spectrum of Encephalitis Associated with Antibodies Against the alpha-Amino-3-Hydroxy-5-Methyl-4-Isoxazolepropionic Acid Receptor: Case Series and Review of the Literature. JAMA Neurol. 2015, 72, 1163–1169. [Google Scholar] [CrossRef]
  28. Fukuda, T.G.; do Rosario, M.S.; Branco, R.C.C.; Fukuda, J.S.; de Souza, E.S.R.A.; Oliveira-Filho, J.; de Jesus, P.A.P. Multiple paraneoplastic antibodies (anti-SOX1, anti-Hu, and anti-Amphiphysin) detected in a patient with limbic encephalitis and small cell lung cancer. Neurol. India 2017, 65, 1127–1128. [Google Scholar] [CrossRef]
  29. Jia, Y.; Wang, J.; Xue, L.; Hou, Y. Limbic encephalitis associated with AMPA receptor and CRMP5 antibodies: A case report and literature review. Brain Behav. 2020, 10, e01528. [Google Scholar] [CrossRef]
  30. Yoneda, M. Hashimoto’s Encephalopathy and Autoantibodies. Brain Nerve 2018, 70, 305–314. [Google Scholar] [CrossRef]
  31. Pilli, D.; Zou, A.; Tea, F.; Dale, R.C.; Brilot, F. Expanding Role of T Cells in Human Autoimmune Diseases of the Central Nervous System. Front. Immunol. 2017, 8, 652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Melzer, N.; Meuth, S.G.; Wiendl, H. CD8+ T cells and neuronal damage: Direct and collateral mechanisms of cytotoxicity and impaired electrical excitability. FASEB J. 2009, 23, 3659–3673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ehling, P.; Melzer, N.; Budde, T.; Meuth, S.G. CD8(+) T Cell-Mediated Neuronal Dysfunction and Degeneration in Limbic Encephalitis. Front. Neurol. 2015, 6, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Waterhouse, N.J.; Sutton, V.R.; Sedelies, K.A.; Ciccone, A.; Jenkins, M.; Turner, S.J.; Bird, P.I.; Trapani, J.A. Cytotoxic T lymphocyte-induced killing in the absence of granzymes A and B is unique and distinct from both apoptosis and perforin-dependent lysis. J. Cell Biol. 2006, 173, 133–144. [Google Scholar] [CrossRef] [Green Version]
  35. Kreutzfeldt, M.; Bergthaler, A.; Fernandez, M.; Bruck, W.; Steinbach, K.; Vorm, M.; Coras, R.; Blumcke, I.; Bonilla, W.V.; Fleige, A.; et al. Neuroprotective intervention by interferon-gamma blockade prevents CD8+ T cell-mediated dendrite and synapse loss. J. Exp. Med. 2013, 210, 2087–2103. [Google Scholar] [CrossRef] [Green Version]
  36. Chung, W.S.; Welsh, C.A.; Barres, B.A.; Stevens, B. Do glia drive synaptic and cognitive impairment in disease? Nat. Neurosci. 2015, 18, 1539–1545. [Google Scholar] [CrossRef]
  37. Wesselingh, R.; Butzkueven, H.; Buzzard, K.; Tarlinton, D.; O’Brien, T.J.; Monif, M. Innate Immunity in the Central Nervous System: A Missing Piece of the Autoimmune Encephalitis Puzzle? Front. Immunol. 2019, 10, 2066. [Google Scholar] [CrossRef]
  38. Khan, N.L.; Jeffree, M.A.; Good, C.; Macleod, W.; Al-Sarraj, S. Histopathology of VGKC antibody-associated limbic encephalitis. Neurology 2009, 72, 1703–1705. [Google Scholar] [CrossRef]
  39. Amhaoul, H.; Hamaide, J.; Bertoglio, D.; Reichel, S.N.; Verhaeghe, J.; Geerts, E.; Van Dam, D.; De Deyn, P.P.; Kumar-Singh, S.; Katsifis, A.; et al. Brain inflammation in a chronic epilepsy model: Evolving pattern of the translocator protein during epileptogenesis. Neurobiol. Dis. 2015, 82, 526–539. [Google Scholar] [CrossRef] [Green Version]
  40. Yang, M.T.; Lin, Y.C.; Ho, W.H.; Liu, C.L.; Lee, W.T. Everolimus is better than rapamycin in attenuating neuroinflammation in kainic acid-induced seizures. J. Neuroinflamm. 2017, 14, 15. [Google Scholar] [CrossRef] [Green Version]
  41. Tomczak, A.; Su, E.; Tugizova, M.; Carlson, A.M.; Kipp, L.B.; Feng, H.; Han, M.H. A case of GFAP-astroglial autoimmunity presenting with reversible parkinsonism. Mult. Scler. Relat. Disord. 2019, 39, 101900. [Google Scholar] [CrossRef] [PubMed]
  42. Ismail, F.S.; Faustmann, P.M. Astrocytes and their potential role in anti-NMDA receptor encephalitis. Med. Hypotheses 2020, 139, 109612. [Google Scholar] [CrossRef] [PubMed]
  43. Habbas, S.; Santello, M.; Becker, D.; Stubbe, H.; Zappia, G.; Liaudet, N.; Klaus, F.R.; Kollias, G.; Fontana, A.; Pryce, C.R.; et al. Neuroinflammatory TNFalpha Impairs Memory via Astrocyte Signaling. Cell 2015, 163, 1730–1741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Hebert, J.; Gros, P.; Lapointe, S.; Amtashar, F.S.; Steriade, C.; Maurice, C.; Wennberg, R.A.; Day, G.S.; Tang-Wai, D.F. Searching for autoimmune encephalitis: Beware of normal CSF. J. Neuroimmunol. 2020, 345, 577285. [Google Scholar] [CrossRef]
  45. Hansen, N.; Schwing, K.; Onder, D.; Widman, G.; Leelaarporn, P.; Prusseit, I.; Surges, R.; Melzer, N.; Gross, C.; Becker, A.J.; et al. Low CSF CD4/CD8+ T-cell proportions are associated with blood-CSF barrier dysfunction in limbic encephalitis. Epilepsy Behav. 2020, 102, 106682. [Google Scholar] [CrossRef]
  46. Lin, Y.T.; Yang, X.; Lv, J.W.; Liu, X.W.; Wang, S.J. CXCL13 Is A Biomarker of Anti-Leucine-Rich Glioma-Inactivated Protein 1 Encephalitis Patients. Neuropsychiatr. Dis. Treat. 2019, 15, 2909–2915. [Google Scholar] [CrossRef] [Green Version]
  47. Roberts, W.K.; Blachere, N.E.; Frank, M.O.; Dousmanis, A.; Ransohoff, R.M.; Darnell, R.B. A destructive feedback loop mediated by CXCL10 in central nervous system inflammatory disease. Ann. Neurol. 2015, 78, 619–629. [Google Scholar] [CrossRef]
  48. Kortvelyessy, P.; Goihl, A.; Guttek, K.; Schraven, B.; Pruss, H.; Reinhold, D. Serum and CSF cytokine levels mirror different neuroimmunological mechanisms in patients with LGI1 and Caspr2 encephalitis. Cytokine 2020, 135, 155226. [Google Scholar] [CrossRef]
  49. Michael, B.D.; Griffiths, M.J.; Granerod, J.; Brown, D.; Davies, N.W.; Borrow, R.; Solomon, T. Characteristic Cytokine and Chemokine Profiles in Encephalitis of Infectious, Immune-Mediated, and Unknown Aetiology. PLoS ONE 2016, 11, e0146288. [Google Scholar] [CrossRef]
  50. Jiang, J.X.; Fewings, N.; Dervish, S.; Fois, A.F.; Duma, S.R.; Silsby, M.; Bandodkar, S.; Ramanathan, S.; Bleasel, A.; John, B.; et al. Novel Surrogate Markers of CNS Inflammation in CSF in the Diagnosis of Autoimmune Encephalitis. Front. Neurol. 2019, 10, 1390. [Google Scholar] [CrossRef]
  51. Spolski, R.; Leonard, W.J. Interleukin-21: A double-edged sword with therapeutic potential. Nat. Rev. Drug Discov. 2014, 13, 379–395. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, M.; Guo, S.; Hibbert, J.M.; Jain, V.; Singh, N.; Wilson, N.O.; Stiles, J.K. CXCL10/IP-10 in infectious diseases pathogenesis and potential therapeutic implications. Cytokine Growth Factor Rev. 2011, 22, 121–130. [Google Scholar] [CrossRef] [PubMed]
  53. Kim, T.J.; Lee, S.T.; Moon, J.; Sunwoo, J.S.; Byun, J.I.; Lim, J.A.; Shin, Y.W.; Jun, J.S.; Lee, H.S.; Lee, W.J.; et al. Anti-LGI1 encephalitis is associated with unique HLA subtypes. Ann. Neurol. 2017, 81, 183–192. [Google Scholar] [CrossRef] [PubMed]
  54. Binks, S.; Varley, J.; Lee, W.; Makuch, M.; Elliott, K.; Gelfand, J.M.; Jacob, S.; Leite, M.I.; Maddison, P.; Chen, M.; et al. Distinct HLA associations of LGI1 and CASPR2-antibody diseases. Brain 2018, 141, 2263–2271. [Google Scholar] [CrossRef] [PubMed]
  55. Muniz-Castrillo, S.; Ambati, A.; Dubois, V.; Vogrig, A.; Joubert, B.; Rogemond, V.; Picard, G.; Lin, L.; Fabien, N.; Mignot, E.; et al. Primary DQ effect in the association between HLA and neurological syndromes with anti-GAD65 antibodies. J. Neurol. 2020, 267, 1906–1911. [Google Scholar] [CrossRef] [PubMed]
  56. Shojima, Y.; Nishioka, K.; Watanabe, M.; Jo, T.; Tanaka, K.; Takashima, H.; Noda, K.; Okuma, Y.; Urabe, T.; Yokoyama, K.; et al. Clinical Characterization of Definite Autoimmune Limbic Encephalitis: A 30-case Series. Intern. Med. 2019, 58, 3369–3378. [Google Scholar] [CrossRef] [Green Version]
  57. Dalmau, J.; Lancaster, E.; Martinez-Hernandez, E.; Rosenfeld, M.R.; Balice-Gordon, R. Clinical experience and laboratory investigations in patients with anti-NMDAR encephalitis. Lancet Neurol. 2011, 10, 63–74. [Google Scholar] [CrossRef] [Green Version]
  58. Small, M.; Treilleux, I.; Couillault, C.; Pissaloux, D.; Picard, G.; Paindavoine, S.; Attignon, V.; Wang, Q.; Rogemond, V.; Lay, S.; et al. Genetic alterations and tumor immune attack in Yo paraneoplastic cerebellar degeneration. Acta Neuropathol. 2018, 135, 569–579. [Google Scholar] [CrossRef]
  59. Melzer, N.; Meuth, S.G.; Wiendl, H. Paraneoplastic and non-paraneoplastic autoimmunity to neurons in the central nervous system. J. Neurol. 2013, 260, 1215–1233. [Google Scholar] [CrossRef] [Green Version]
  60. Vogrig, A.; Gigli, G.L.; Segatti, S.; Corazza, E.; Marini, A.; Bernardini, A.; Valent, F.; Fabris, M.; Curcio, F.; Brigo, F.; et al. Epidemiology of paraneoplastic neurological syndromes: A population-based study. J. Neurol. 2020, 267, 26–35. [Google Scholar] [CrossRef]
  61. Gultekin, S.H.; Rosenfeld, M.R.; Voltz, R.; Eichen, J.; Posner, J.B.; Dalmau, J. Paraneoplastic limbic encephalitis: Neurological symptoms, immunological findings and tumour association in 50 patients. Brain 2000, 123 Pt 7, 1481–1494. [Google Scholar] [CrossRef]
  62. Erkmen, C.P.; Fadul, C.E.; Dalmau, J.; Erkmen, K. Thymoma-associated paraneoplastic encephalitis (TAPE): Diagnosis and treatment of a potentially fatal condition. J. Thorac. Cardiovasc. Surg. 2011, 141, e17–e20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. van Coevorden-Hameete, M.H.; de Bruijn, M.; de Graaff, E.; Bastiaansen, D.; Schreurs, M.W.J.; Demmers, J.A.A.; Ramberger, M.; Hulsenboom, E.S.P.; Nagtzaam, M.M.P.; Boukhrissi, S.; et al. The expanded clinical spectrum of anti-GABABR encephalitis and added value of KCTD16 autoantibodies. Brain 2019, 142, 1631–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Irani, S.R.; Alexander, S.; Waters, P.; Kleopa, K.A.; Pettingill, P.; Zuliani, L.; Peles, E.; Buckley, C.; Lang, B.; Vincent, A. Antibodies to Kv1 potassium channel-complex proteins leucine-rich, glioma inactivated 1 protein and contactin-associated protein-2 in limbic encephalitis, Morvan’s syndrome and acquired neuromyotonia. Brain 2010, 133, 2734–2748. [Google Scholar] [CrossRef] [PubMed]
  65. Irani, S.R.; Pettingill, P.; Kleopa, K.A.; Schiza, N.; Waters, P.; Mazia, C.; Zuliani, L.; Watanabe, O.; Lang, B.; Buckley, C.; et al. Morvan syndrome: Clinical and serological observations in 29 cases. Ann. Neurol. 2012, 72, 241–255. [Google Scholar] [CrossRef]
  66. Dalmau, J.; Rosenfeld, M.R. Paraneoplastic syndromes of the CNS. Lancet Neurol. 2008, 7, 327–340. [Google Scholar] [CrossRef] [Green Version]
  67. Armangue, T.; Spatola, M.; Vlagea, A.; Mattozzi, S.; Carceles-Cordon, M.; Martinez-Heras, E.; Llufriu, S.; Muchart, J.; Erro, M.E.; Abraira, L.; et al. Frequency, symptoms, risk factors, and outcomes of autoimmune encephalitis after herpes simplex encephalitis: A prospective observational study and retrospective analysis. Lancet Neurol. 2018, 17, 760–772. [Google Scholar] [CrossRef] [Green Version]
  68. Mohammad, S.S.; Sinclair, K.; Pillai, S.; Merheb, V.; Aumann, T.D.; Gill, D.; Dale, R.C.; Brilot, F. Herpes simplex encephalitis relapse with chorea is associated with autoantibodies to N-Methyl-D-aspartate receptor or dopamine-2 receptor. Mov. Disord. 2014, 29, 117–122. [Google Scholar] [CrossRef]
  69. Peters, J.; Wesley, S.F. Case of concurrent herpes simplex and autoimmune encephalitis. Neurol. Neuroimmunol. Neuroinflamm. 2020, 7. [Google Scholar] [CrossRef]
  70. Hoftberger, R.; Armangue, T.; Leypoldt, F.; Graus, F.; Dalmau, J. Clinical Neuropathology practice guide 4-2013: Post-herpes simplex encephalitis: N-methyl-Daspartate receptor antibodies are part of the problem. Clin. Neuropathol. 2013, 32, 251–254. [Google Scholar] [CrossRef]
  71. Seeley, W.W.; Marty, F.M.; Holmes, T.M.; Upchurch, K.; Soiffer, R.J.; Antin, J.H.; Baden, L.R.; Bromfield, E.B. Post-transplant acute limbic encephalitis: Clinical features and relationship to HHV6. Neurology 2007, 69, 156–165. [Google Scholar] [CrossRef] [PubMed]
  72. Dubey, D.; David, W.S.; Amato, A.A.; Reynolds, K.L.; Clement, N.F.; Chute, D.F.; Cohen, J.V.; Lawrence, D.P.; Mooradian, M.J.; Sullivan, R.J.; et al. Varied phenotypes and management of immune checkpoint inhibitor-associated neuropathies. Neurology 2019, 93, e1093–e1103. [Google Scholar] [CrossRef] [PubMed]
  73. Vogrig, A.; Muniz-Castrillo, S.; Joubert, B.; Picard, G.; Rogemond, V.; Marchal, C.; Chiappa, A.M.; Chanson, E.; Skowron, F.; Leblanc, A.; et al. Central nervous system complications associated with immune checkpoint inhibitors. J. Neurol. Neurosurg. Psychiatry 2020, 91, 772–778. [Google Scholar] [CrossRef] [PubMed]
  74. Ribas, A.; Wolchok, J.D. Cancer immunotherapy using checkpoint blockade. Science 2018, 359, 1350–1355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Papadopoulos, K.P.; Romero, R.S.; Gonzalez, G.; Dix, J.E.; Lowy, I.; Fury, M. Anti-Hu-Associated Autoimmune Limbic Encephalitis in a Patient with PD-1 Inhibitor-Responsive Myxoid Chondrosarcoma. Oncologist 2018, 23, 118–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Chung, M.; Jaffer, M.; Verma, N.; Mokhtari, S.; Ramsakal, A.; Peguero, E. Immune checkpoint inhibitor induced anti-glutamic acid decarboxylase 65 (Anti-GAD 65) limbic encephalitis responsive to intravenous immunoglobulin and plasma exchange. J. Neurol. 2020, 267, 1023–1025. [Google Scholar] [CrossRef]
  77. Vogrig, A.; Fouret, M.; Joubert, B.; Picard, G.; Rogemond, V.; Pinto, A.L.; Muniz-Castrillo, S.; Roger, M.; Raimbourg, J.; Dayen, C.; et al. Increased frequency of anti-Ma2 encephalitis associated with immune checkpoint inhibitors. Neurol. Neuroimmunol. Neuroinflamm. 2019, 6. [Google Scholar] [CrossRef] [Green Version]
  78. Brown, M.P.; Hissaria, P.; Hsieh, A.H.; Kneebone, C.; Vallat, W. Autoimmune limbic encephalitis with anti-contactin-associated protein-like 2 antibody secondary to pembrolizumab therapy. J. Neuroimmunol. 2017, 305, 16–18. [Google Scholar] [CrossRef]
  79. Hacohen, Y.; Singh, R.; Rossi, M.; Lang, B.; Hemingway, C.; Lim, M.; Vincent, A. Clinical relevance of voltage-gated potassium channel-complex antibodies in children. Neurology 2015, 85, 967–975. [Google Scholar] [CrossRef] [Green Version]
  80. Lang, B.; Makuch, M.; Moloney, T.; Dettmann, I.; Mindorf, S.; Probst, C.; Stoecker, W.; Buckley, C.; Newton, C.R.; Leite, M.I.; et al. Intracellular and non-neuronal targets of voltage-gated potassium channel complex antibodies. J. Neurol. Neurosurg. Psychiatry 2017, 88, 353–361. [Google Scholar] [CrossRef] [Green Version]
  81. Lilleker, J.B.; Jones, M.S.; Mohanraj, R. VGKC complex antibodies in epilepsy: Diagnostic yield and therapeutic implications. Seizure 2013, 22, 776–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. van Sonderen, A.; Schreurs, M.W.; Wirtz, P.W.; Sillevis Smitt, P.A.; Titulaer, M.J. From VGKC to LGI1 and Caspr2 encephalitis: The evolution of a disease entity over time. Autoimmun. Rev. 2016, 15, 970–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. van Sonderen, A.; Schreurs, M.W.; de Bruijn, M.A.; Boukhrissi, S.; Nagtzaam, M.M.; Hulsenboom, E.S.; Enting, R.H.; Thijs, R.D.; Wirtz, P.W.; Sillevis Smitt, P.A.; et al. The relevance of VGKC positivity in the absence of LGI1 and Caspr2 antibodies. Neurology 2016, 86, 1692–1699. [Google Scholar] [CrossRef] [PubMed]
  84. Klein, C.J.; Lennon, V.A.; Aston, P.A.; McKeon, A.; O’Toole, O.; Quek, A.; Pittock, S.J. Insights from LGI1 and CASPR2 potassium channel complex autoantibody subtyping. JAMA Neurol. 2013, 70, 229–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Michael, S.; Waters, P.; Irani, S.R. Stop testing for autoantibodies to the VGKC-complex: Only request LGI1 and CASPR2. Pr. Neurol. 2020, 20, 377–384. [Google Scholar] [CrossRef] [PubMed]
  86. Fukata, Y.; Yokoi, N.; Miyazaki, Y.; Fukata, M. The LGI1-ADAM22 protein complex in synaptic transmission and synaptic disorders. Neurosci. Res. 2017, 116, 39–45. [Google Scholar] [CrossRef] [PubMed]
  87. Ohkawa, T.; Satake, S.; Yokoi, N.; Miyazaki, Y.; Ohshita, T.; Sobue, G.; Takashima, H.; Watanabe, O.; Fukata, Y.; Fukata, M. Identification and characterization of GABA(A) receptor autoantibodies in autoimmune encephalitis. J. Neurosci. 2014, 34, 8151–8163. [Google Scholar] [CrossRef] [Green Version]
  88. Boillot, M.; Huneau, C.; Marsan, E.; Lehongre, K.; Navarro, V.; Ishida, S.; Dufresnois, B.; Ozkaynak, E.; Garrigue, J.; Miles, R.; et al. Glutamatergic neuron-targeted loss of LGI1 epilepsy gene results in seizures. Brain 2014, 137, 2984–2996. [Google Scholar] [CrossRef] [Green Version]
  89. Ohkawa, T.; Fukata, Y.; Yamasaki, M.; Miyazaki, T.; Yokoi, N.; Takashima, H.; Watanabe, M.; Watanabe, O.; Fukata, M. Autoantibodies to epilepsy-related LGI1 in limbic encephalitis neutralize LGI1-ADAM22 interaction and reduce synaptic AMPA receptors. J. Neurosci. 2013, 33, 18161–18174. [Google Scholar] [CrossRef]
  90. Petit-Pedrol, M.; Sell, J.; Planaguma, J.; Mannara, F.; Radosevic, M.; Haselmann, H.; Ceanga, M.; Sabater, L.; Spatola, M.; Soto, D.; et al. LGI1 antibodies alter Kv1.1 and AMPA receptors changing synaptic excitability, plasticity and memory. Brain 2018, 141, 3144–3159. [Google Scholar] [CrossRef]
  91. Irani, S.R.; Michell, A.W.; Lang, B.; Pettingill, P.; Waters, P.; Johnson, M.R.; Schott, J.M.; Armstrong, R.J.; Zagami, A.S.; Bleasel, A.; et al. Faciobrachial dystonic seizures precede Lgi1 antibody limbic encephalitis. Ann. Neurol. 2011, 69, 892–900. [Google Scholar] [CrossRef] [PubMed]
  92. Malter, M.P.; Frisch, C.; Schoene-Bake, J.C.; Helmstaedter, C.; Wandinger, K.P.; Stoecker, W.; Urbach, H.; Surges, R.; Elger, C.E.; Vincent, A.V.; et al. Outcome of limbic encephalitis with VGKC-complex antibodies: Relation to antigenic specificity. J. Neurol. 2014, 261, 1695–1705. [Google Scholar] [CrossRef] [PubMed]
  93. Bastiaansen, A.E.M.; van Sonderen, A.; Titulaer, M.J. Autoimmune encephalitis with anti-leucine-rich glioma-inactivated 1 or anti-contactin-associated protein-like 2 antibodies (formerly called voltage-gated potassium channel-complex antibodies). Curr. Opin. Neurol. 2017, 30, 302–309. [Google Scholar] [CrossRef] [PubMed]
  94. Saint-Martin, M.; Joubert, B.; Pellier-Monnin, V.; Pascual, O.; Noraz, N.; Honnorat, J. Contactin-associated protein-like 2, a protein of the neurexin family involved in several human diseases. Eur. J. Neurosci. 2018, 48, 1906–1923. [Google Scholar] [CrossRef]
  95. Ogawa, Y.; Oses-Prieto, J.; Kim, M.Y.; Horresh, I.; Peles, E.; Burlingame, A.L.; Trimmer, J.S.; Meijer, D.; Rasband, M.N. ADAM22, a Kv1 channel-interacting protein, recruits membrane-associated guanylate kinases to juxtaparanodes of myelinated axons. J. Neurosci. 2010, 30, 1038–1048. [Google Scholar] [CrossRef]
  96. Patterson, K.R.; Dalmau, J.; Lancaster, E. Mechanisms of Caspr2 antibodies in autoimmune encephalitis and neuromyotonia. Ann. Neurol. 2018, 83, 40–51. [Google Scholar] [CrossRef]
  97. Dawes, J.M.; Weir, G.A.; Middleton, S.J.; Patel, R.; Chisholm, K.I.; Pettingill, P.; Peck, L.J.; Sheridan, J.; Shakir, A.; Jacobson, L.; et al. Immune or Genetic-Mediated Disruption of CASPR2 Causes Pain Hypersensitivity Due to Enhanced Primary Afferent Excitability. Neuron 2018, 97, 806–822.e10. [Google Scholar] [CrossRef] [Green Version]
  98. Klein, C.J.; Lennon, V.A.; Aston, P.A.; McKeon, A.; Pittock, S.J. Chronic pain as a manifestation of potassium channel-complex autoimmunity. Neurology 2012, 79, 1136–1144. [Google Scholar] [CrossRef] [Green Version]
  99. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [Green Version]
  100. Sprengel, R. Role of AMPA receptors in synaptic plasticity. Cell Tissue Res. 2006, 326, 447–455. [Google Scholar] [CrossRef]
  101. Lai, M.; Hughes, E.G.; Peng, X.; Zhou, L.; Gleichman, A.J.; Shu, H.; Mata, S.; Kremens, D.; Vitaliani, R.; Geschwind, M.D.; et al. AMPA receptor antibodies in limbic encephalitis alter synaptic receptor location. Ann. Neurol. 2009, 65, 424–434. [Google Scholar] [CrossRef] [PubMed]
  102. Peng, X.; Hughes, E.G.; Moscato, E.H.; Parsons, T.D.; Dalmau, J.; Balice-Gordon, R.J. Cellular plasticity induced by anti-alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor encephalitis antibodies. Ann. Neurol. 2015, 77, 381–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Laurido-Soto, O.; Brier, M.R.; Simon, L.E.; McCullough, A.; Bucelli, R.C.; Day, G.S. Patient characteristics and outcome associations in AMPA receptor encephalitis. J. Neurol. 2019, 266, 450–460. [Google Scholar] [CrossRef] [PubMed]
  104. Nicoletti, F.; Bockaert, J.; Collingridge, G.L.; Conn, P.J.; Ferraguti, F.; Schoepp, D.D.; Wroblewski, J.T.; Pin, J.P. Metabotropic glutamate receptors: From the workbench to the bedside. Neuropharmacology 2011, 60, 1017–1041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Spatola, M.; Sabater, L.; Planaguma, J.; Martinez-Hernandez, E.; Armangue, T.; Pruss, H.; Iizuka, T.; Caparo Oblitas, R.L.; Antoine, J.C.; Li, R.; et al. Encephalitis with mGluR5 antibodies: Symptoms and antibody effects. Neurology 2018, 90, e1964–e1972. [Google Scholar] [CrossRef] [PubMed]
  106. Lancaster, E.; Martinez-Hernandez, E.; Titulaer, M.J.; Boulos, M.; Weaver, S.; Antoine, J.C.; Liebers, E.; Kornblum, C.; Bien, C.G.; Honnorat, J.; et al. Antibodies to metabotropic glutamate receptor 5 in the Ophelia syndrome. Neurology 2011, 77, 1698–1701. [Google Scholar] [CrossRef] [Green Version]
  107. Boronat, A.; Gelfand, J.M.; Gresa-Arribas, N.; Jeong, H.Y.; Walsh, M.; Roberts, K.; Martinez-Hernandez, E.; Rosenfeld, M.R.; Balice-Gordon, R.; Graus, F.; et al. Encephalitis and antibodies to dipeptidyl-peptidase-like protein-6, a subunit of Kv4.2 potassium channels. Ann. Neurol. 2013, 73, 120–128. [Google Scholar] [CrossRef]
  108. Nadal, M.S.; Ozaita, A.; Amarillo, Y.; Vega-Saenz de Miera, E.; Ma, Y.; Mo, W.; Goldberg, E.M.; Misumi, Y.; Ikehara, Y.; Neubert, T.A.; et al. The CD26-related dipeptidyl aminopeptidase-like protein DPPX is a critical component of neuronal A-type K+ channels. Neuron 2003, 37, 449–461. [Google Scholar] [CrossRef] [Green Version]
  109. Clark, B.D.; Kwon, E.; Maffie, J.; Jeong, H.Y.; Nadal, M.; Strop, P.; Rudy, B. DPP6 Localization in Brain Supports Function as a Kv4 Channel Associated Protein. Front. Mol. Neurosci. 2008, 1, 8. [Google Scholar] [CrossRef] [Green Version]
  110. Tobin, W.O.; Lennon, V.A.; Komorowski, L.; Probst, C.; Clardy, S.L.; Aksamit, A.J.; Appendino, J.P.; Lucchinetti, C.F.; Matsumoto, J.Y.; Pittock, S.J.; et al. DPPX potassium channel antibody: Frequency, clinical accompaniments, and outcomes in 20 patients. Neurology 2014, 83, 1797–1803. [Google Scholar] [CrossRef] [Green Version]
  111. Hara, M.; Arino, H.; Petit-Pedrol, M.; Sabater, L.; Titulaer, M.J.; Martinez-Hernandez, E.; Schreurs, M.W.; Rosenfeld, M.R.; Graus, F.; Dalmau, J. DPPX antibody-associated encephalitis: Main syndrome and antibody effects. Neurology 2017, 88, 1340–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Linnoila, J.J.; Rosenfeld, M.R.; Dalmau, J. Neuronal surface antibody-mediated autoimmune encephalitis. Semin. Neurol. 2014, 34, 458–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Emson, P.C. GABA(B) receptors: Structure and function. Prog. Brain Res. 2007, 160, 43–57. [Google Scholar] [CrossRef]
  114. Bettler, B.; Kaupmann, K.; Mosbacher, J.; Gassmann, M. Molecular structure and physiological functions of GABA(B) receptors. Physiol. Rev. 2004, 84, 835–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Lancaster, E.; Lai, M.; Peng, X.; Hughes, E.; Constantinescu, R.; Raizer, J.; Friedman, D.; Skeen, M.B.; Grisold, W.; Kimura, A.; et al. Antibodies to the GABA(B) receptor in limbic encephalitis with seizures: Case series and characterisation of the antigen. Lancet Neurol. 2010, 9, 67–76. [Google Scholar] [CrossRef] [Green Version]
  116. Hoftberger, R.; Titulaer, M.J.; Sabater, L.; Dome, B.; Rozsas, A.; Hegedus, B.; Hoda, M.A.; Laszlo, V.; Ankersmit, H.J.; Harms, L.; et al. Encephalitis and GABAB receptor antibodies: Novel findings in a new case series of 20 patients. Neurology 2013, 81, 1500–1506. [Google Scholar] [CrossRef] [Green Version]
  117. Jeffery, O.J.; Lennon, V.A.; Pittock, S.J.; Gregory, J.K.; Britton, J.W.; McKeon, A. GABAB receptor autoantibody frequency in service serologic evaluation. Neurology 2013, 81, 882–887. [Google Scholar] [CrossRef] [Green Version]
  118. Dogan Onugoren, M.; Deuretzbacher, D.; Haensch, C.A.; Hagedorn, H.J.; Halve, S.; Isenmann, S.; Kramme, C.; Lohner, H.; Melzer, N.; Monotti, R.; et al. Limbic encephalitis due to GABAB and AMPA receptor antibodies: A case series. J. Neurol. Neurosurg. Psychiatry 2015, 86, 965–972. [Google Scholar] [CrossRef]
  119. Arino, H.; Hoftberger, R.; Gresa-Arribas, N.; Martinez-Hernandez, E.; Armangue, T.; Kruer, M.C.; Arpa, J.; Domingo, J.; Rojc, B.; Bataller, L.; et al. Paraneoplastic Neurological Syndromes and Glutamic Acid Decarboxylase Antibodies. JAMA Neurol. 2015, 72, 874–881. [Google Scholar] [CrossRef]
  120. Zhang, X.; Lang, Y.; Sun, L.; Zhang, W.; Lin, W.; Cui, L. Clinical characteristics and prognostic analysis of anti-gamma-aminobutyric acid-B (GABA-B) receptor encephalitis in Northeast China. BMC Neurol. 2020, 20, 1. [Google Scholar] [CrossRef] [Green Version]
  121. Blinder, T.; Lewerenz, J. Cerebrospinal Fluid Findings in Patients With Autoimmune Encephalitis—A Systematic Analysis. Front. Neurol. 2019, 10, 804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Moser, A.; Hanssen, H.; Wandinger, K.P. Excessively increased CSF glutamate levels in GABAB-receptor antibody associated encephalitis: A case report. J. Neurol. Sci. 2018, 388, 10–11. [Google Scholar] [CrossRef] [PubMed]
  123. Golombeck, K.S.; Bonte, K.; Monig, C.; van Loo, K.M.; Hartwig, M.; Schwindt, W.; Widman, G.; Lindenau, M.; Becker, A.J.; Glatzel, M.; et al. Evidence of a pathogenic role for CD8(+) T cells in anti-GABAB receptor limbic encephalitis. Neurol. Neuroimmunol. Neuroinflamm. 2016, 3, e232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Cui, J.; Bu, H.; He, J.; Zhao, Z.; Han, W.; Gao, R.; Li, X.; Li, Q.; Guo, X.; Zou, Y. The gamma-aminobutyric acid-B receptor (GABAB) encephalitis: Clinical manifestations and response to immunotherapy. Int. J. Neurosci. 2018, 128, 627–633. [Google Scholar] [CrossRef]
  125. Solimena, M.; Folli, F.; Aparisi, R.; Pozza, G.; De Camilli, P. Autoantibodies to GABA-ergic neurons and pancreatic beta cells in stiff-man syndrome. N. Engl. J. Med. 1990, 322, 1555–1560. [Google Scholar] [CrossRef] [PubMed]
  126. Ali, F.; Rowley, M.; Jayakrishnan, B.; Teuber, S.; Gershwin, M.E.; Mackay, I.R. Stiff-person syndrome (SPS) and anti-GAD-related CNS degenerations: Protean additions to the autoimmune central neuropathies. J. Autoimmun. 2011, 37, 79–87. [Google Scholar] [CrossRef] [PubMed]
  127. Fenalti, G.; Buckle, A.M. Structural biology of the GAD autoantigen. Autoimmun. Rev. 2010, 9, 148–152. [Google Scholar] [CrossRef]
  128. Dalakas, M.C. Progress and stiff challenges in understanding the role of GAD-antibodies in stiff-person syndrome. Exp. Neurol. 2013, 247, 303–307. [Google Scholar] [CrossRef]
  129. Walikonis, J.E.; Lennon, V.A. Radioimmunoassay for glutamic acid decarboxylase (GAD65) autoantibodies as a diagnostic aid for stiff-man syndrome and a correlate of susceptibility to type 1 diabetes mellitus. Mayo Clin. Proc. 1998, 73, 1161–1166. [Google Scholar] [CrossRef]
  130. Munoz-Lopetegi, A.; de Bruijn, M.; Boukhrissi, S.; Bastiaansen, A.E.M.; Nagtzaam, M.M.P.; Hulsenboom, E.S.P.; Boon, A.J.W.; Neuteboom, R.F.; de Vries, J.M.; Sillevis Smitt, P.A.E.; et al. Neurologic syndromes related to anti-GAD65: Clinical and serologic response to treatment. Neurol. Neuroimmunol. Neuroinflamm. 2020, 7. [Google Scholar] [CrossRef] [Green Version]
  131. Raju, R.; Foote, J.; Banga, J.P.; Hall, T.R.; Padoa, C.J.; Dalakas, M.C.; Ortqvist, E.; Hampe, C.S. Analysis of GAD65 autoantibodies in Stiff-Person syndrome patients. J. Immunol. 2005, 175, 7755–7762. [Google Scholar] [CrossRef] [Green Version]
  132. Vianello, M.; Giometto, B.; Vassanelli, S.; Canato, M.; Betterle, C.; Mucignat, C. Peculiar labeling of cultured hippocampal neurons by different sera harboring anti-glutamic acid decarboxylase autoantibodies (GAD-Ab). Exp. Neurol. 2006, 202, 514–518. [Google Scholar] [CrossRef] [PubMed]
  133. Burton, A.R.; Baquet, Z.; Eisenbarth, G.S.; Tisch, R.; Smeyne, R.; Workman, C.J.; Vignali, D.A. Central nervous system destruction mediated by glutamic acid decarboxylase-specific CD4+ T cells. J. Immunol. 2010, 184, 4863–4870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Cheramy, M.; Hampe, C.S.; Ludvigsson, J.; Casas, R. Characteristics of in-vitro phenotypes of glutamic acid decarboxylase 65 autoantibodies in high-titre individuals. Clin. Exp. Immunol. 2013, 171, 247–254. [Google Scholar] [CrossRef]
  135. Malter, M.P.; Helmstaedter, C.; Urbach, H.; Vincent, A.; Bien, C.G. Antibodies to glutamic acid decarboxylase define a form of limbic encephalitis. Ann. Neurol. 2010, 67, 470–478. [Google Scholar] [CrossRef] [PubMed]
  136. Chang, T.; Alexopoulos, H.; Pettingill, P.; McMenamin, M.; Deacon, R.; Erdelyi, F.; Szabo, G.; Buckley, C.J.; Vincent, A. Immunization against GAD induces antibody binding to GAD-independent antigens and brainstem GABAergic neuronal loss. PLoS ONE 2013, 8, e72921. [Google Scholar] [CrossRef] [Green Version]
  137. Haberlandt, E.; Bast, T.; Ebner, A.; Holthausen, H.; Kluger, G.; Kravljanac, R.; Kroll-Seger, J.; Kurlemann, G.; Makowski, C.; Rostasy, K.; et al. Limbic encephalitis in children and adolescents. Arch. Dis. Child. 2011, 96, 186–191. [Google Scholar] [CrossRef]
  138. Baizabal-Carvallo, J.F. The neurological syndromes associated with glutamic acid decarboxylase antibodies. J. Autoimmun. 2019, 101, 35–47. [Google Scholar] [CrossRef]
  139. Gagnon, M.M.; Savard, M. Limbic Encephalitis Associated With GAD65 Antibodies: Brief Review of the Relevant literature. Can. J. Neurol. Sci. 2016, 43, 486–493. [Google Scholar] [CrossRef]
  140. Hansen, N.; Widman, G.; Witt, J.A.; Wagner, J.; Becker, A.J.; Elger, C.E.; Helmstaedter, C. Seizure control and cognitive improvement via immunotherapy in late onset epilepsy patients with paraneoplastic versus GAD65 autoantibody-associated limbic encephalitis. Epilepsy Behav. 2016, 65, 18–24. [Google Scholar] [CrossRef]
  141. Liu, B.; Zhou, Y.; Meng, L.; Skinner, H. A Survival Case of Super-refractory Status Epilepticus due to Glutamic Acid Decarboxylase Antibodies-associated Limbic Encephalitis. Cureus 2018, 10, e3125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Boronat, A.; Sabater, L.; Saiz, A.; Dalmau, J.; Graus, F. GABA(B) receptor antibodies in limbic encephalitis and anti-GAD-associated neurologic disorders. Neurology 2011, 76, 795–800. [Google Scholar] [CrossRef] [PubMed]
  143. Jarius, S.; Ringelstein, M.; Haas, J.; Serysheva, I.I.; Komorowski, L.; Fechner, K.; Wandinger, K.P.; Albrecht, P.; Hefter, H.; Moser, A.; et al. Inositol 1,4,5-trisphosphate receptor type 1 autoantibodies in paraneoplastic and non-paraneoplastic peripheral neuropathy. J. Neuroinflamm. 2016, 13, 278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Berger, B.; Bischler, P.; Dersch, R.; Hottenrott, T.; Rauer, S.; Stich, O. “Non-classical” paraneoplastic neurological syndromes associated with well-characterized antineuronal antibodies as compared to “classical” syndromes—More frequent than expected. J. Neurol. Sci. 2015, 352, 58–61. [Google Scholar] [CrossRef]
  145. Sun, X.; Tan, J.; Sun, H.; Liu, Y.; Guan, W.; Jia, J.; Wang, Z. Anti-SOX1 Antibodies in Paraneoplastic Neurological Syndrome. J. Clin. Neurol. 2020, 16, 530–546. [Google Scholar] [CrossRef]
  146. Yamamoto, T.; Tsuji, S. Anti-Ma2-associated encephalitis and paraneoplastic limbic encephalitis. Brain Nerve 2010, 62, 838–851. [Google Scholar]
  147. Shen, K.; Xu, Y.; Guan, H.; Zhong, W.; Chen, M.; Zhao, J.; Li, L.; Wang, M. Paraneoplastic limbic encephalitis associated with lung cancer. Sci. Rep. 2018, 8, 6792. [Google Scholar] [CrossRef] [Green Version]
  148. Ortega Suero, G.; Sola-Valls, N.; Escudero, D.; Saiz, A.; Graus, F. Anti-Ma and anti-Ma2-associated paraneoplastic neurological syndromes. Neurologia 2018, 33, 18–27. [Google Scholar] [CrossRef]
  149. Kohler, W.; Ehrlich, S.; Dohmen, C.; Haubitz, M.; Hoffmann, F.; Schmidt, S.; Klingel, R.; Kraft, A.; Neumann-Haefelin, T.; Topka, H.; et al. Tryptophan immunoadsorption for the treatment of autoimmune encephalitis. Eur. J. Neurol. 2015, 22, 203–206. [Google Scholar] [CrossRef]
  150. Heine, J.; Ly, L.T.; Lieker, I.; Slowinski, T.; Finke, C.; Pruss, H.; Harms, L. Immunoadsorption or plasma exchange in the treatment of autoimmune encephalitis: A pilot study. J. Neurol. 2016, 263, 2395–2402. [Google Scholar] [CrossRef]
  151. Dogan Onugoren, M.; Golombeck, K.S.; Bien, C.; Abu-Tair, M.; Brand, M.; Bulla-Hellwig, M.; Lohmann, H.; Munstermann, D.; Pavenstadt, H.; Tholking, G.; et al. Immunoadsorption therapy in autoimmune encephalitides. Neurol. Neuroimmunol. Neuroinflamm. 2016, 3, e207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Rossling, R.; Pruss, H. Apheresis in Autoimmune Encephalitis and Autoimmune Dementia. J. Clin. Med. 2020, 9, 2683. [Google Scholar] [CrossRef] [PubMed]
  153. Dalakas, M.C. Neurological complications of immune checkpoint inhibitors: What happens when you ‘take the brakes off’ the immune system. Ther. Adv. Neurol. Disord. 2018, 11, 1756286418799864. [Google Scholar] [CrossRef] [Green Version]
  154. Barbagallo, M.; Vitaliti, G.; Pavone, P.; Romano, C.; Lubrano, R.; Falsaperla, R. Pediatric Autoimmune Encephalitis. J. Pediatr. Neurosci. 2017, 12, 130–134. [Google Scholar] [CrossRef] [PubMed]
  155. Shapiro-Shelef, M.; Calame, K. Regulation of plasma-cell development. Nat. Rev. Immunol. 2005, 5, 230–242. [Google Scholar] [CrossRef]
  156. Kosmidis, M.L.; Dalakas, M.C. Practical considerations on the use of rituximab in autoimmune neurological disorders. Ther. Adv. Neurol. Disord. 2010, 3, 93–105. [Google Scholar] [CrossRef] [Green Version]
  157. Hoftberger, R.; van Sonderen, A.; Leypoldt, F.; Houghton, D.; Geschwind, M.; Gelfand, J.; Paredes, M.; Sabater, L.; Saiz, A.; Titulaer, M.J.; et al. Encephalitis and AMPA receptor antibodies: Novel findings in a case series of 22 patients. Neurology 2015, 84, 2403–2412. [Google Scholar] [CrossRef] [Green Version]
  158. Brummaier, T.; Pohanka, E.; Studnicka-Benke, A.; Pieringer, H. Using cyclophosphamide in inflammatory rheumatic diseases. Eur. J. Intern. Med. 2013, 24, 590–596. [Google Scholar] [CrossRef]
  159. Kanter, I.C.; Huttner, H.B.; Staykov, D.; Biermann, T.; Struffert, T.; Kerling, F.; Hilz, M.J.; Schellinger, P.D.; Schwab, S.; Bardutzky, J. Cyclophosphamide for anti-GAD antibody-positive refractory status epilepticus. Epilepsia 2008, 49, 914–920. [Google Scholar] [CrossRef]
  160. Gea-Banacloche, J.C. Rituximab-associated infections. Semin. Hematol. 2010, 47, 187–198. [Google Scholar] [CrossRef] [Green Version]
  161. Mihara, M.; Kasutani, K.; Okazaki, M.; Nakamura, A.; Kawai, S.; Sugimoto, M.; Matsumoto, Y.; Ohsugi, Y. Tocilizumab inhibits signal transduction mediated by both mIL-6R and sIL-6R, but not by the receptors of other members of IL-6 cytokine family. Int. Immunopharmacol. 2005, 5, 1731–1740. [Google Scholar] [CrossRef]
  162. Chavele, K.M.; Merry, E.; Ehrenstein, M.R. Cutting edge: Circulating plasmablasts induce the differentiation of human T follicular helper cells via IL-6 production. J. Immunol. 2015, 194, 2482–2485. [Google Scholar] [CrossRef] [Green Version]
  163. Lee, W.J.; Lee, S.T.; Moon, J.; Sunwoo, J.S.; Byun, J.I.; Lim, J.A.; Kim, T.J.; Shin, Y.W.; Lee, K.J.; Jun, J.S.; et al. Tocilizumab in Autoimmune Encephalitis Refractory to Rituximab: An Institutional Cohort Study. Neurotherapeutics 2016, 13, 824–832. [Google Scholar] [CrossRef] [Green Version]
  164. Benucci, M.; Tramacere, L.; Infantino, M.; Manfredi, M.; Grossi, V.; Damiani, A.; Gobbi, F.L.; Piccininni, M.; Zaccara, G.; Cincotta, M. Efficacy of Tocilizumab in Limbic Encephalitis with Anti-CASPR2 Antibodies. Case Rep. Neurol. Med. 2020, 2020, 5697670. [Google Scholar] [CrossRef]
  165. Randell, R.L.; Adams, A.V.; Van Mater, H. Tocilizumab in Refractory Autoimmune Encephalitis: A Series of Pediatric Cases. Pediatr. Neurol. 2018, 86, 66–68. [Google Scholar] [CrossRef]
  166. Scheibe, F.; Ostendorf, L.; Reincke, S.M.; Pruss, H.; von Brunneck, A.C.; Kohnlein, M.; Alexander, T.; Meisel, C.; Meisel, A. Daratumumab treatment for therapy-refractory anti-CASPR2 encephalitis. J. Neurol. 2020, 267, 317–323. [Google Scholar] [CrossRef]
  167. Nosadini, M.; Mohammad, S.S.; Toldo, I.; Sartori, S.; Dale, R.C. Mycophenolate mofetil, azathioprine and methotrexate usage in paediatric anti-NMDAR encephalitis: A systematic literature review. Eur. J. Paediatr. Neurol. 2019, 23, 7–18. [Google Scholar] [CrossRef] [Green Version]
  168. Dale, R.C.; Gorman, M.P.; Lim, M. Autoimmune encephalitis in children: Clinical phenomenology, therapeutics, and emerging challenges. Curr. Opin. Neurol. 2017, 30, 334–344. [Google Scholar] [CrossRef]
  169. Dutra, L.A.; Abrantes, F.; Toso, F.F.; Pedroso, J.L.; Barsottini, O.G.P.; Hoftberger, R. Autoimmune encephalitis: A review of diagnosis and treatment. Arq. Neuropsiquiatr. 2018, 76, 41–49. [Google Scholar] [CrossRef] [Green Version]
  170. Argyropoulos, G.P.D.; Moore, L.; Loane, C.; Roca-Fernandez, A.; Lage-Martinez, C.; Gurau, O.; Irani, S.R.; Zeman, A.; Butler, C.R. Pathologic tearfulness after limbic encephalitis: A novel disorder and its neural basis. Neurology 2020, 94, e1320–e1335. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Antibodies and their pathogenic effects in limbic encephalitis. The pathologic effects of antibodies include blocking of receptors or ion channels, disruption of interaction with neighboring molecules, and crosslinking and internalization of receptors from cell surface. (1) CASPR2 and contactin2 antibodies inhibit the interaction between these proteins and reduce the clustering and surface expression of VGKC; (2) LGI1 antibody disrupts interaction between protein components such as LGI1 to ADAM22/23, downregulates VGKC and reduces AMPAR clustering and synaptic transmission; (3) AMPAR antibody causes cross-linking and receptor internalization; (4) NMDAR* antibody causes cross-linking and receptor internalization; (5) mGluR5 antibody causes decreased synaptic mGluR5 cluster density; (6) GAD antibodies target mostly the GAD65 subunits but also the GAD67 subunits, disrupting GABAergic signaling; (7) GABA-BR antibody prevents ligand binding to the receptor and alters receptor function; (8) GlyR antibody probably acts as antagonist to disrupt receptor function. Ab: antibody; ADAM: a disintegrin and metalloproteinase; AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate receptor; CASPR2: contactin associated protein-like 2; GABA-AR: γ-aminobutyric acid-A receptor; GABA-BR: γ-aminobutyric acid-B receptor, GAD: glutamic acid decarboxylase; GlyR: glycine receptor; LGI1: leucine-rich, glioma inactivated 1; mGluR5: metabotropic glutamate receptor 5; NMDAR: N-methyl-D-asparate receptor; PSD-95: postsynaptic density protein 95; VGKC: voltage-gated potassium channels.
Figure 1. Antibodies and their pathogenic effects in limbic encephalitis. The pathologic effects of antibodies include blocking of receptors or ion channels, disruption of interaction with neighboring molecules, and crosslinking and internalization of receptors from cell surface. (1) CASPR2 and contactin2 antibodies inhibit the interaction between these proteins and reduce the clustering and surface expression of VGKC; (2) LGI1 antibody disrupts interaction between protein components such as LGI1 to ADAM22/23, downregulates VGKC and reduces AMPAR clustering and synaptic transmission; (3) AMPAR antibody causes cross-linking and receptor internalization; (4) NMDAR* antibody causes cross-linking and receptor internalization; (5) mGluR5 antibody causes decreased synaptic mGluR5 cluster density; (6) GAD antibodies target mostly the GAD65 subunits but also the GAD67 subunits, disrupting GABAergic signaling; (7) GABA-BR antibody prevents ligand binding to the receptor and alters receptor function; (8) GlyR antibody probably acts as antagonist to disrupt receptor function. Ab: antibody; ADAM: a disintegrin and metalloproteinase; AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate receptor; CASPR2: contactin associated protein-like 2; GABA-AR: γ-aminobutyric acid-A receptor; GABA-BR: γ-aminobutyric acid-B receptor, GAD: glutamic acid decarboxylase; GlyR: glycine receptor; LGI1: leucine-rich, glioma inactivated 1; mGluR5: metabotropic glutamate receptor 5; NMDAR: N-methyl-D-asparate receptor; PSD-95: postsynaptic density protein 95; VGKC: voltage-gated potassium channels.
Ijms 22 00389 g001
Figure 2. Diagnostic algorithm of limbic encephalitis. NMDAR antibody (*) generally causes autoimmune encephalitis but rarely causes limbic encephalitis. Nevertheless, it should be included in neuronal antibody survey for limbic encephalitis due to clinical similarity. GAD (**) is a synaptic enzyme.
Figure 2. Diagnostic algorithm of limbic encephalitis. NMDAR antibody (*) generally causes autoimmune encephalitis but rarely causes limbic encephalitis. Nevertheless, it should be included in neuronal antibody survey for limbic encephalitis due to clinical similarity. GAD (**) is a synaptic enzyme.
Ijms 22 00389 g002
Table 1. Autoantibodies associated with limbic encephalitis.
Table 1. Autoantibodies associated with limbic encephalitis.
SymptomsTumor AssociationImagePrognosis
Cell-surface
LGI1LE, seizure including FBDS, hyponatremia, paroxysmal dizziness spells, pain, dysautonomia<10%
SCLC, thymoma
Normal or nonspecific changes; T2-mesiotemporal hyperintensity; T1-bright basal ganglia in FBDSAlmost fair response to immunotherapy but high relapse rate (15–35%)
CASPR2LE, Morvan syndrome, neuromyotonia, seizure, neuropathic pain, dysautonomia10–40% (44% for LGI1 and CASPR2 dual seropositivity)
thymoma
Normal or nonspecific changes; T2 mesiotemporal hyperintensity Favorable response to immunotherapy but high relapse rate (35%)
AMPARLE, confusion, psychiatric symptoms, seizure50–70%
lung, breast, thymoma
T2 mesiotemporal hyperintensityFavorable response to immunotherapy but common relapse
GABA-BRLE, seizure, dysautonomia, movement disorder, rapidly progressive dementia50–60%
SCLC
T2 mesiotemporal hyperintensityFrequent co-expression with other Abs; poor prognosis with concurrent tumor or convulsive SE; relapse rate (20%)
mGluR5LE, Ophelia syndrome seizures, movement disorders50%
Hodgkin lymphoma
Limbic and extra-limbic (thalamus, pons, frontal or parieto-occipital cortex, cerebellum)Complete or partial recovery to immunotherapy
DPPXLE, BE, diarrhea, CNS hyperexcitability, PERM, dysautonomia<10–30%
B cell tumor
Normal or nonspecific T2/FLAIR white matter abnormalitiesChronic and second-line immunotherapy frequently required, relapse rate (23%)
Synaptic
GADLE, SPS, PERM, seizure, CA, oculomotor dysfunction, diabetes<15% (higher with coexisting Abs, esp. GABA-BR Ab)
lung, thymoma
T2 mesiotemporal hyperintensity70% had partial improvement with immunotherapy
Intracellular
MaLE, BE, diencephalic encephalitis, seizure, CS90%
testicular, lung and pleural
NonspecificFavorable in anti-Ma2 but poorer in anti-Ma
HuLE, CS, BE, dysautonomia,
sensory neuropathy
>90%
SCLC
NonspecificPoor
RiLE, CS, BE, OMS, movement disorders>90%
woman-breast, man-lung and bladder
NonspecificCommon co-expression with other Abs
CRMP5LE, encephalomyelitis, CS, SPS>90%
SCLC, thymoma
Normal or T2 mesiotemporal hyperintensity Poor
SOX1LE, LEMS, CS, neuropathy, LEMS30–60%
SCLC
Normal or T2 mesiotemporal hyperintensity Common co-expression with other Abs; poor response to treatment
AK5LENo dataHippocampal atrophyPoor response to treatment
Abs: antibodies; AK5: adenylate kinase 5; AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate receptor; BE: brainstem encephalitis; CASPR2: contactin-associated protein-like 2; CRMP5: collapsin response-mediator protein-5; CNS: central nervous system; CS: cerebellar syndrome including paraneoplastic cerebellar degeneration; DPPX: dipeptidyl peptidase-like protein 6; FBDS: faciobrachial dystonic seizures; GABA-BR: γ-aminobutyric acid B receptor; GAD: glutamic acid decarboxylase; LE: limbic encephalitis; LEMS: Lambert-Eaton myasthenic syndrome; LGI1: leucine-rich, glioma inactivated 1; mGluR5: metabotropic glutamate receptor subtype 5; OMS: opsoclonus-myoclonus syndrome; PERM: progressive encephalomyelitis with rigidity and myoclonus; SCLC: small-cell lung cancer; SE: status epilepticus; SOX1: sex-determining region Y (SRY)-box 1; SPS: stiff-person syndrome.
Table 2. Immunotherapy for limbic encephalitis.
Table 2. Immunotherapy for limbic encephalitis.
Dose & DurationMechanismNote
First Line
MethylprednisoloneAdult: 1 g daily
Children: 30 mg/kg/day (max. 1 g)
for 3–5 days
Inhibits NF-κB → anti-inflammation30 mg/kg/day (max. 1 g) once monthly for 3–6 months
IVIg2 g/kg in 2 or 5 days Neutralizes Abs and cytokines, decreases B cells, inhibits complement activation, modulates regulatory T cells1 g/kg monthly for 3 months or longer
Should not be given immediately prior to plasmapheresis
Plasmapheresis
(PE or IA)
5–7 exchanges in 10–14 days Remove Ab
Second Line
Rituximab375 mg/m2 weekly for 4 weeks, or 750 mg/m2 (max. 1000 mg/dose) for two doses 2 weeks apart Anti-CD20 Ab → depletion of B cells and plasmablasts
Cyclophosphamide750–1000 mg/m2 (max. 1000–1500 mg/dose) monthly for 3–6 months Akylating agents inhibiting DNA synthesis → suppress B and T cellsCause infertility in repeated doses
Tocilizumab8–12 mg/kg (max. 800 mg) monthly for 6 monthsAnti-IL6 receptor Ab → inhibits B and T cells
Daratumumab16 mg/kg weekly in cycle 1–8, every two weeks in cycle 9–13anti-CD38 Ab → depletion of plasma cellsOne case of anti-CASPR2 encephalitis
Bortezomib1.3 mg/m2 on day 1, 4, 8, and 11 of a 21-day cycle, total 3 cyclesproteasome inhibitor → depletion of plasma cellsClinical trial (NCT03993262)
Maintenance therapy
Prednisolone 1–2 mg/kg/day once daily or divided for 4 weeks; tapered over several weeks to months Inhibits NF-κB → anti-inflammation
Mycophenolate mofetil (MMF)Initial 300 mg/m2/day, target 600 mg/m2/day, 1–1.5 g/day, twice dailyInhibits purine nucleotides → inhibits B and T cells
Azathioprine (AZA)1–2.5 mg/kg/day (max.150 mg/day), once dailyInhibits purine synthesis → inhibits B and T cells
Methotrexate (MTX)
Oral: 10 mg/m2 weekly
Intrathecal: 10 mg weekly for 4 weeks
Inhibits NF-κB → anti-inflammation
Table 3. Characteristic symptoms and laboratory findings of limbic encephalitis.
Table 3. Characteristic symptoms and laboratory findings of limbic encephalitis.
SymptomsAntibodies
SeizureGABA-BR (~100%), LGI1 (80–100%), GAD (60–100%), CASPR2 (75%), NMDAR*(70%), AMPAR (33%)
FBDS in LGI1 (33–67%)
Paroxysmal dizzy spellsLGI1 (14%)
Cramps or neuropathic painCASPR2, LGI1
Movement disordersNMDAR *, CRMP5, Hu, GlyR, GABA-BR, DPPX, CASPR2, Ri
DysautonomiaNMDAR *, LGI1, CASPR2, GABA-BR, DPPX, GlyR
GI symptoms (diarrhea)DPPX
Laboratory findingsAntibodies
HyponatremiaGLI1 (~50%), CASPR2 (~25%)
HyperglycemiaGAD
CSF pleocytosis + OCBGABA-BR, GAD, mGluR5
NMDAR antibody (*), a frequently detected antibody in autoimmune encephalitis, should also be included in antibody survey for limbic encephalitis due to clinical similarity. AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; CASPR2: contactin-associated protein-like 2; CSF: cerebrospinal fluid; DPPX: dipeptidyl peptidase-like protein 6; FBDS: faciobrachial dystonic seizures; GABA-BR: γ-aminobutyric acid B receptor; GAD: glutamic acid decarboxylase; GI: gastrointestinal; GlyR: glycine receptor; LGI1: leucine-rich, glioma inactivated 1; NMDAR: N-methyl-D-aspartate receptor; OCB: oligoclonal band.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kao, Y.-C.; Lin, M.-I.; Weng, W.-C.; Lee, W.-T. Neuropsychiatric Disorders Due to Limbic Encephalitis: Immunologic Aspect. Int. J. Mol. Sci. 2021, 22, 389. https://doi.org/10.3390/ijms22010389

AMA Style

Kao Y-C, Lin M-I, Weng W-C, Lee W-T. Neuropsychiatric Disorders Due to Limbic Encephalitis: Immunologic Aspect. International Journal of Molecular Sciences. 2021; 22(1):389. https://doi.org/10.3390/ijms22010389

Chicago/Turabian Style

Kao, Yu-Chia, Ming-I Lin, Wen-Chin Weng, and Wang-Tso Lee. 2021. "Neuropsychiatric Disorders Due to Limbic Encephalitis: Immunologic Aspect" International Journal of Molecular Sciences 22, no. 1: 389. https://doi.org/10.3390/ijms22010389

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop