Open access peer-reviewed chapter

Does COVID-19 Affect Adult Neurogenesis? A Neurochemical Perspective

Written By

Jayakumar Saikarthik, Ilango Saraswathi and Abdulrahman A. Al-Atram

Submitted: 11 September 2021 Reviewed: 11 October 2021 Published: 25 November 2021

DOI: 10.5772/intechopen.101179

From the Edited Volume

COVID-19, Neuroimmunology and Neural Function

Edited by Thomas Heinbockel and Robert Weissert

Chapter metrics overview

368 Chapter Downloads

View Full Metrics

Abstract

COVID-19 has been found to cause neuropsychiatric symptoms which indicate brain involvement. SARS-CoV-2 may enter the brain by damaging and penetrating olfactory mucosa and via other possible routes like damaged blood–brain barrier, and hematologic spread. With SARS-CoV-2 having a higher affinity to ACE2 receptors, brain regions that have higher ACE2 receptors like the hippocampus, are more vulnerable to the effect of the viral invasion. In addition, immune cell activation, an important feature of COVID-19, leads to cytokine storm which causes neurotoxicity, neuroinflammation, and neurodegeneration. Impaired adult neurogenesis is related to many psychiatric disorders including depression, bipolar disorder, anxiety disorder, schizophrenia, and PTSD. It is known to be related to the depletion of neurotransmitters, dopamine, serotonin, norepinephrine, GABA, and glutamate which play a major role in adult neurogenesis. A recent study reveals that SSRI which acts by increasing serotonin is proven beneficial in COVID-19 patients. Thus, the current chapter will discuss the impact of COVID-19 on adult neurogenesis with emphasis on the role of ACE2 and neurotransmitters.

Keywords

  • COVID-19
  • SARS-CoV-2
  • ACE2
  • adult neurogenesis
  • glutamate
  • monoaminergic neurotransmitters
  • GABA

1. Introduction

The last two decades have seen epidemic outbreaks by novel viruses including SARS, MERS, and influenza which shared certain commonalities such as a likely zoonotic origin, high mortality rates, and less available therapeutic methods to counteract them. The COVID-19 pandemic shows no signs of slowing down with affecting 223 countries, with 224,811,910 cases, and 4,633,797 death tolls till date [1]. With what history on earlier pandemics has made us understand and with the rapidly mutating nature of the SARS-CoV-2 virus, it is not unreasonable to say that the pandemic is here to stay, and the world must learn to co-exist with it. The first reported case of COVID-19 was found in Wuhan, China in December 2019. By March 2020, the disease had spread across the globe and had become a public health emergency. The WHO declared a pandemic state to the disease spread on March 11, 2020 [2]. With more than a year since the declaration of the pandemic, the scientific community has yet not developed a definitive anti-viral drug to combat the disease spread. Even though the advent of vaccination has set the pace in favour of global health, we have a long way to go to eradicate if at all suppress the disease spread.

SARS-CoV-2 is highly virulent and highly contagious with the R0 value of 3.77 [3]. Though it predominantly affects the respiratory system, other organ systems like the gastrointestinal system, heart, kidney, and central nervous system are also targeted by the virus. Fever, chills, cough, shortness of breath or breathing difficulty, sore throat, nasal congestion, diarrhoea, nausea, vomiting, generalised body aches are some of the common symptoms noted in patients infected with COVID-19 [4].

Neurological manifestations of COVID-19 include non-specific symptoms like headache, dizziness, fatigue, and myopathy and more specific symptoms like anosmia, ageusia, impaired consciousness, stroke, meningitis, acute transverse myelitis, and Guillian-Barre syndrome [5, 6]. More than one third of the individuals with COVID-19 were found to present with neurological symptoms [7, 8]. The presence of viral RNA in cerebrospinal fluid and the brain was observed in COVID-19 patients [9]. Preliminary in vitro studies have found that SARS-CoV-2 can replicate in neuronal cells [10]. A post-mortem study has found that 48% of the studied cases had human CoV RNA in the CNS that was detectable [11]. SARS-CoV-2 is found to exhibit organotropism for the nervous system and SARS-CoV and MERS-CoV which are closely related to SARS-CoV-2 have neuro-invasive potential. Hence, apart from the secondary impact on the brain as a result of systemic complications like coagulopathy and hypoxia, the direct effect of SARS-CoV-2 infection on the brain and spinal cord is plausible and is being thoroughly studied by researchers globally. The neuropsychiatric symptoms in COVID-19 could be attributed to a variety of factors apart from the direct effect of the virus on the brain like psychological distress due to social isolation, the novelty of the disease spread and pandemic, concerns about family and friends contracting the disease, social stigma, etc. [12]. This chapter will, however, focus on the direct effects of the SARS-CoV-2 virus on the brain which could be attributed to the pathophysiology of neuropsychiatric symptoms with a special focus on ACE2 and monoaminergic neurotransmitters.

Advertisement

2. SARS-CoV-2

Coronaviruses are the largest among RNA viruses. They have a crown-like spikes on their surface and hence the name. SARS-CoV-2 is the latest/seventh coronavirus to become pathogenic to humans. It belongs to the Coronaviridae family which includes four genera; α−, β−, γ−, and δ-CoV. Out of these human pathogens include HCoV- 229E, HCoV- NL63 [α − CoV] and OC43, and HKU1 [β − CoV] that in most cases cause mild self-limiting respiratory disease. γ − and δ-CoV strains mainly affect avian species [13]. SARS-CoV and MERS-CoV, causatives of SARS and MERS, are beta coronaviruses that caused up to 9.6% and 34.3% mortality rates which were responsible for earlier pandemics that resulted in a death toll of 812 and 866, respectively [14]. SARS-CoV-2 is more similar to SARS-CoV and MERS-CoV while being far more pathogenic and transmissible than the earlier known coronaviruses.

SARS-CoV-2 is a beta coronavirus that is positive-sense single-stranded RNA virus with 29–30 kb in size. It has four structural proteins and 16 non-structural proteins. Nucleocapsid protein [N], membrane protein [M], spike protein [S], and envelope protein [E] are the four structural proteins (Figure 1). The capsid of the genome is formed by N protein and the genome is further surrounded by an envelope that is made up of M, E, and S proteins. Like other coronaviruses, SARS-CoV-2 has enveloped with a crown-like spikes on its surface. It is the spike protein that is responsible for the variations in host specificity and tissue tropism of the different coronavirus. Spike protein is a type-I membrane glycoprotein and has two functional subunits S1 and S2 with different functional domains in the amino and carboxy terminal. S1 subunit contains the receptor-binding domain [RBD] and binds with the receptor in the host cells. S2 subunit fuses the membranes of the host cells and the virus. The entry of the virus into the host cell involves binding of the S protein [S1 subunit] to a specific cell receptor followed by priming of the S protein by proteases in the host cell. This leads to the fusion of the spike protein to the cell membrane which is mediated by the S2 subunit [15]. The specific cell receptor through which SARS-CoV-2 enters the host cell is the ACE2 receptor and the protease in the host cell that processes the spike protein to reveal the fusion peptide between S1 and S2 subunits facilitating its entry, is a TMPRSS2 serine protease, member of the hepsin/TMPRSS subfamily [16]. Another protein named furin or paired basic amino acid cleaving enzyme [PACE], a member of the subtilisin-like proprotein convertase family, mediates proteolytic cut of the S protein at S1-S2 boundary, is required for TMPRSS2 processing of S protein. Both TMPRSS2 and furin are essential for the entry of SARS-CoV-2 into the cell. The furin cleavage site in the S protein of SARS-CoV-2 is not found in SARS-CoV and other beta coronaviruses [17].

Advertisement

3. ACE2

ACE2 is a cell surface protein, a metalloproteinase and an ectoenzyme which is an obligatory receptor for SARS-CoV and SARS-CoV-2. The affinity of SARS-CoV-2 to ACE2 is ten times higher than that of SARS-CoV which partly explains its higher pathogenicity [18]. It was discovered in 2000 by two independent groups of researchers while searching for human ACE homologues [19, 20]. The gene for ACE2 in humans is located in Xp22 and has 18 exons, a majority of which are similar to the exons of the ACE gene [21]. Despite ACE2 exhibiting 42% sequence identify and 61% sequence similarity with ACE, the two enzymes show enormous variations (Table 1) [27].

ACEACE2
FormsExists as a 2-domain somatic form and a one domain testicular formExists as a single form
StructureTransmembrane ectoenzyme with two active sitesTransmembrane ectoenzyme with one active site
Enzymatic actionRemoves C-terminal dipeptide – peptidyl-dipeptidaseRemoves single amino acid from C-terminus – carboxypeptidase
Substrate specificityConverts Ang I to Ang IIConverts Ang I to Ang (1-9)
Does not cleave Ang IIConverts Ang II to Ang (1-7)
Converts Ang (1-9) to Ang (1-7)Does not cleave Ang (1-9)
Converts Ang (1-7) to Ang (1-5)Does not cleave Ang (1-7)
Does not cleave Ang AConverts Ang A to Alamandine
Hydrolyses bradykininDoes not cleave bradykinin
Does not cleave des-Arg9-bradykininHydrolyses des-Arg9-bradykinin
Action on amyloid proteinHydrolyses Aβ-43 to Aβ41Hydrolyses Aβ43 to Aβ42
Hydrolyses Aβ-42 to Aβ40Does not cleave Aβ-42
Localisation within cellsEqual distribution between apical and basolateral membranesLocalised on the apical membrane
Transports intestinal amino acidsNoTransports intestinal neutral amino acids
Shedding into plasmaUnidentified. May involve metalloproteinase and A DisintegrinBy A Disintegrin and Metalloprotease 17 (ADAM 17)
Response to ACE inhibitorInhibitedResistant, gets upregulated
Acts as a receptor to virusNoReceptor for SARS-CoV and SARS-CoV-2

Table 1.

The comparison between ACE and ACE2 is given in Table 1 [22, 23, 24, 25, 26].

Since the 20 years of its discovery, ACE2 was found to have a multitude of physiological and pathological functions based on its three fundamental actions viz. negative regulation of renin-angiotensin system [RAS], facilitation of amino acid transport in the intestine, and surface receptor for SARS-CoV and SARS-CoV-2. ACE2 is mainly expressed in the lungs, intestine, liver, heart, kidneys, testes, and brain. In the brain, it is expressed in neurons, astrocytes and oligodendrocytes, and in ventricles, substantia nigra, hypothalamus, hippocampus, middle temporal gyrus, posterior cingulate cortex, nuclei in pons—the nucleus of tractus solitarius and pre-Bötzinger complex and olfactory bulb [21, 28]. ACE2 expression is higher in astrocytes, astrocytic foot processes, pericytes, and endothelial cells which form the key components of the blood–brain barrier [29]. In the olfactory epithelium, its expression is higher in the supporting sustentacular cells than in olfactory sensory neurons [30]. The sites of ACE2 expression are given in Table 2 [31].

Vascular systemEndothelial cells, vascular smooth muscle cells, and migratory angiogenic cells
HeartCardiomyocytes, endothelial cells, pericytes, and epicardial adipose cells, and cardiofibroblasts
Skinsebaceous gland cells and basal epidermal layer
Kidneysglomerular endothelial cells, proximal tubule epithelial cells, bladder urothelial cells, luminal surface of tubular epithelial cells, and podocytes
Reproductive systemOvary, oocyte, uterus, vagina, and placenta of the female reproductive system
Adult Leydig cells and cells in the seminiferous ducts in the testis of the male reproductive system
LiverPerinuclear hepatocytes, cholangiocytes, epithelial cells of the bile duct
GutStratified epithelial cells of oesophagus, stomach, Intestinal epithelial cells, enterocytes of small intestine, absorptive enterocytes from the ileum, colon and rectum, and endothelial cells
PancreasAcinar cells and duct cells of the exocrine gland and alpha, beta, delta, and PP cells of islets of Langerhans
ThyroidGlandular cells
Oral cavityTongue, buccal mucosa, gingiva, leucocytes within the oral mucosa, non-keratinising squamous epithelium of the oral cavity – basal layer
Upper airwayCiliated epithelial cells, goblet cells
LungsPulmonary vasculature, type I and II alveolar epithelial cells, bronchiolar epithelial cells
EyesPigmented epithelial cells, photoreceptor cells, Müller glial cells
Central nervous systemNeurons, astrocytes, and oligodendrocytes, and in ventricles, substantia nigra, hypothalamus, hippocampus, middle temporal gyrus, posterior cingulate cortex, nuclei in pons – nucleus of tractus solitarius and pre-Bötzinger complex and olfactory bulb and cerebral vasculature and components of blood–brain barrier (astrocytes, astrocytic foot processes, pericytes, and endothelial cells)

Table 2.

Sites of ACE2 expression.

3.1 Structure

ACE2 is a type 1 integral membrane protein that includes a short cytoplasmic C-terminus, a transmembrane region, collectrin, and N-terminal ectodomain. Zinc-binding motifs, HEMGH forms the active site of the enzyme. N-terminal domain has a claw-shaped protease domain which is the binding site of receptor-binding domain [RBD] of SARS-CoV and SARS-CoV-2. N terminus is homologous to ACE and is a carboxypeptidase that metabolises peptides like angiotensin II, kinins, apelin-13, apelin-36, neurotensin 1–13, kinetensin, and morphins, and C terminus is homologous to collectrin which is involved in the trafficking of neutral amino acid transporter [B[o]AT1] in the intestinal epithelium [32].

Advertisement

4. Role of ACE2 in renin-angiotensin system [RAS]

Both ACE and ACE2 play a major role in maintaining renin-angiotensin system [RAS] homeostasis. ACE2 acts like a negative regulator of ACE in RAS. RAS involves a variety of proteins and enzymes. Angiotensinogen is an inactive precursor that gets cleaved by renin to form angiotensin I. ACE acts on angiotensin I to convert into angiotensin II [Ang II] while ACE2 converts Ang II to Ang [1-7]. Ang [1-7] then binds to Mas receptors and causes attenuation of the signal cascade that was activated by Ang II (Figure 2). Thus, ACE2 not only inactivates Ang II but also generates the antagonistic peptide Ang [1-7] [33]. Ang [1-7] can also be formed from Ang I by neutral endopeptidases and neprilysin, but the most effective pathway of Ang [1-7] generation is through ACE2 [34]. The conversion of Ang II to Ang [1-7] by ACE2 is 70 folds more efficient than the conversion of Ang I to Ang [1-9] by ACE2. Thus, under physiological conditions, ACE2 mainly forms Ang [1-7] than Ang [1-9] [34].

While Ang II, which acts via angiotensin 1/AT1 [primary mediator] and angiotensin 2/AT2 receptors is a potent vasoconstrictor, a pro-fibrotic, and a pro-inflammatory agent, Ang [1-7] acts via Mas receptors and has vasodilator, anti-apoptotic and anti-proliferative effect. Mas receptors are G protein-coupled receptors and in the brain, they are highly expressed in the dentate gyrus of the hippocampus, a site-specific for adult neurogenesis and in blood vessels [35]. The ACE2/Ang [1-7]/Mas receptor axis of the RAS is considered to be the protective arm of the renin-angiotensin system. A balance in ACE/ACE2 is critical which implies a balance between the pro-inflammatory pro-oxidative arm and the anti-inflammatory and anti-oxidative arm of RAS. An increase in ACE/ACE2 ratio was observed in many pathological conditions including cardiovascular pathology, renal dysfunction, pulmonary hypertension, in cigarette smokers, and Alzheimer’s disease [36, 37, 38, 39]. SARS-CoV-2 which enters the host cells via ACE2 also causes downregulation of ACE2 and the major targets of SARS-CoV-2 are those which express higher levels of ACE2 [26]. The fibrotic and inflammatory processes observed in various organs in COVID-19 patients could be attributed to the dysregulation of ACE2 and subsequently, RAS which is observed in endocrine, paracrine, and intracrine levels in several organs [40]. Dysregulation of RAS in the brain is associated with neuroinflammation and neurodegeneration [41].

Advertisement

5. Neurogenesis

The old dogma that the production of functional neurons does not occur in adult life was refuted when Altman and Das published evidence to support the continuation of neurogenesis in adult life in rodents [42]. Neurogenesis refers to the process of the generation of new neurons from neural stem cells. This process which plays a major role in brain development in embryonic life ceases to exist shortly after birth in the majority of brain areas except two. The subgranular zone [SGZ] of the dentate gyrus of the hippocampus and subventricular zone [SVZ], lining the lateral wall of the lateral ventricles are the two areas where neurogenesis persists well into adult life albeit declining slightly with ageing (Figure 3) [43, 44]. There is a complex microenvironment that nourishes and supports the neural progenitor cells and their progeny which is called the ‘neurogenic niche’. There are various trophic factors, blood vessels, supporting glial cells, and hormones in the neurogenic niche that help to control and enhance neurogenesis [45]. The newborn neurons mature and get integrated into neural circuits and are involved in a variety of functions including learning and memory like temporal and pattern separation, high-resolution memory, synaptic plasticity, fear conditioning and emotions, and olfaction [46]. Incidentally altered neurogenesis is implicated in several neuropsychiatric diseases like Alzheimer’s disease, Parkinson’s disease, depression, Huntington’s disease, and stroke, epilepsy, and demyelinating disease [46, 47].

Figure 1.

(a) Structure of ACE2 and SARS-CoV-2; (b) Interaction of spike protein and ACE2; (c) Shedding of ACE2 and entry of SARS-CoV-2 into the cell.

Figure 2.

Renin-Angiotensin System.

Figure 3.

Coronal section of the brain showing the sites of adult neurogenesis.

5.1 Stages of adult neurogenesis

The process of adult neurogenesis occurs in stages viz. maintenance of neural stem/progenitor cells [NPC] and proliferation of NPC, fate specification/commitment, differentiation, maturation, survival of immature neurons, and integration into neural circuitry. The defining abilities of NPC are self-replication and multipotency, that is, the ability to differentiate into multiple lineages of cells and in this case neurons, astrocytes, and oligodendrocytes [48]. There are different types of neural progenitor cells in SGZ and SVZ. Type-1 cells in SGZ, B-cells in SVZ, and radial glia-like cells in SGZ and SVZ are largely quiescent cells, which are similar to radial glia cells found during embryonic development and have a morphology similar to mature astrocytes. Type-2 cells in SGZ and C-cells in SVZ are small roundish cells that are highly proliferative, and they give rise to type-3 cells in SGZ and A-cells in SVZ which represent committed neuroblasts. The type-1/B-cells are multipotent and have unlimited self-renewal capacity which get activated by various factors and multiply to form highly proliferative transient intermediate progenitor cells [TIP] in the SGZ. In SVZ, the transit-amplifying cells [TAC] [type-2/C-cells] has the ability to differentiate into neurons. These divide to form neuroblasts or immature neurons [type-3/A-cells] which proceed to neuronal differentiation and forms newborn neurons that mature and get integrated into neural circuitry in the brain. It is pertinent to know that many of the newborn neurons perish and only 15–30% of immature neurons survive the maturation process. There are various factors that regulate this step and thereby the process of adult neurogenesis [49, 50, 51].

In SGZ, the NPCs form granule cells which are the principal excitatory cells of the dentate gyrus. Their axons form the mossy fibres extending to the CA3 region and their dendrites are in the molecular layer which receives connections from the entorhinal cortex. Immature neurons that are less than a week-old start to have neurite outgrowth and by one- or two-weeks axons can be observed in the hilus, and dendrites start to extend to the molecular layer without spines which being developed by around the 16th day. By 17 days, functional connections are formed by the axons [mossy fibres] with the CA3 pyramidal neurons [52]. They release glutamate as the neurotransmitter. After around 1 week of birth, the newborn granule cells receive GABAergic inputs and after 2 weeks receive glutamatergic inputs [53]. These immature neurons exhibit enhanced excitability by virtue of high input resistance and subthreshold calcium ion conductance which enables them to develop action potential with less excitatory currents. They also have a low threshold for induction of LTP [long-term potentiation] [54, 55]. Between 3 weeks and 2 months, there occurs a gradual increase in spine formation, dendritic arborisation and connection, boutons on CA3 neurons, and maturation of mossy fibres. By less than 2 months, the newborn neurons become functionally indistinguishable from fully mature granule cells [52].

In SVZ, restricted neural progenitor cells migrate along scaffolds maintained by specialised astrocytes via the rostral migratory stream [RMS] to reach the olfactory bulb. By 15–30 days, they differentiate into two types of interneurons, GABAergic granule neurons [95%] and GABA or dopaminergic periglomerular neurons [5%]. The newborn GABAergic granule neurons can become cells with dendrites that do not cross beyond the mitral cell layer and those with non-spiny dendrites that extend till the external plexiform layer. These interneurons mature and get integrated into olfactory network and start responding to olfactory signals [52].

There are various factors that regulate neurogenesis. These include intrinsic niche-derived intrinsic mechanisms and extrinsic systemic factors. The intrinsic factors that regulate adult neurogenesis are given in Table 3. There are extrinsic environmental cues and systemic factors that can positively and negatively affect adult neurogenesis like physical exercise, dietary intake, olfactory/hippocampal-dependent learning, environmental enrichment, ageing, stress, alcohol abuse, and certain inflammatory conditions [46, 56, 57, 58, 59].

Intrinsic factorsExamples
Neurotrophic factorsbrain-derived neurotrophic factor (BDNF), insulin-like growth factor-1 (IGF-1), nerve growth factor (NGF), glia-derived nerve factor (GDNF), fibroblast growth factor 2 (FGF-2), epidermal growth factor (EGF)
MorphogensNotch, sonic hedgehog (Shh), wingless ligands (Wnts), and bone morphogenic proteins (BMPs).
Inflammatory cytokinestissue necrosis factors α (TNFα), interleukin-6 (IL-6) and IL-1β
IL-4 and IL-10
Neurotransmittersgamma-aminobutyric acid (GABA), glutamate, dopamine, serotonin, norepinephrine, acetylcholine
HormonesGlucocorticoids, sex hormones, leptin, incretin
Epigenetic factorsmethyl-CpG-binding domain protein 1 (Mbd1), MYST family histone acetyltransferase Querkopf (Qkf), mixed-lineage leukaemia 1 (Mll1), polycomb complex protein (Bmi-1), histone deacetylase 2 (HDAC2), and microRNAs (miR124, 137, 184, 185, and 491-3p)
Transcriptional factorssex-determining region Y-box 2 (Sox2), Orphan nuclear receptor TLX, forkhead box O proteins (FoxOs), prospero homeobox 1 (Prox1), neurogenic differentiation1 (NeuroD1), Kruppel-like factor 9, cyclic AMP response element-binding protein (CREB), paired box protein (Pax6), and neurogenin 2 (Neurog2)

Table 3.

List of intrinsic factors that affect adult neurogenesis.

Advertisement

6. Entry of SARS-CoV-2 into the brain

There are different ways that are the possible pathway for the entry of SARS-CoV-2 into the brain. Some of the ways include olfactory transmucosal invasion, hematogenous dissemination, and neuronal retrograde dissemination [5]. The olfactory sensory neurons of the olfactory mucosa are bipolar neurons. The axons of the olfactory sensory neurons along the apical side project into the nasal cavity while that on the basal side merge into filia and protrudes into the olfactory bulb through the cribriform plate. Thus, the olfactory sensory neurons are in direct contact with the cerebrospinal fluid [60]. In the olfactory mucosa, ACE2 receptors are mainly found in the non-neuronal cells, sustentacular cells while their expression in the olfactory sensory neurons is less [30]. The blood vessels lining the olfactory mucosa express both ACE2 and TMPRSS2 protease receptors which help in the invasion of the SARS-CoV-2 virus and facilitate binding, replication, and accumulation of the virus [61, 62]. Studies have found that SARS-CoV-2 enters CNS through this neural-mucosal interface by infection of the olfactory neurons or by diffusion through channels formed by olfactory ensheathing cells in the olfactory mucosa [60, 63]. Following the olfactory transmucosal invasion, the virus passes along the olfactory tract via axonal transport, trans-synaptic transport, or microfusion to different areas of the brain linked with the olfactory tract [60, 64].

Recent studies have observed that SARS-CoV-2 RNA was found in brain regions that are not directly connected to olfactory mucosa like the cerebellum which shows that other forms/routes of viral entry into the brain are at play. Neuronal retrograde dissemination is the one where the virus may breach peripheral nerve terminals and take a trans-synaptic route to reach CNS. For instance, SARS-CoV-2 may invade peripheral chemoreceptors and may reach the cardiorespiratory centre in the brain stem [65] or through the gut-brain axis where the virus may enter the brain through enteric nerves [66]. In case of hematogenous dissemination, the virus after infecting the airways may breach the epithelial barrier and enter the bloodstream. Through systemic circulation, the virus may reach the cerebral circulation and could infect endothelial cells of blood–brain barrier or epithelial cells of the blood CSF barrier to reach the brain or via circumventricular organs which lack the blood–brain barrier [5]. Trojan horse mechanism is another way by which SARS-CoV-2 could reach the brain parenchyma. It is the process in which the virus infects leucocytes which get activated and disseminate to other tissues and cross blood–brain barrier [67].

Once SARS-CoV-2 enters the brain, it enters and infects the neurons, glial cells, and endothelial cells through ACE2 and replicates which leads to cell death. It causes damage to the blood–brain barrier which will increase its permeability and cause oedema, intracerebral bleeding, and neuronal death. The infected neurons can release inflammatory mediators that can activate other immune cells like mast cells, neurons, microglia, astrocytes, endothelial cells, and pericytes [68, 69].

Advertisement

7. Adult neurogenesis in COVID-19

Earlier studies show that survivors of critical illness have higher risk of developing neuropsychiatric consequences after discharge from the hospital. The prevalence of symptoms of depression, anxiety, and post-traumatic stress was found to be 29% [28, 29, 30, 31, 32, 33, 34], 34% [30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42], and 34% [27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50] in survivors of critical illness, respectively [70, 71, 72]. Impairment in memory, attention, and concentration was observed in SARS survivors 1 year after recovery [73]. Based on the knowledge from earlier infections by coronaviruses, SARS, and MERS, an increased risk of neuropsychiatric disorders like depression, anxiety, post-traumatic stress disorder, are possible in a long-term follow-up of patients recovered from COVID-19 [12].

Neuropsychiatric disorders that display impaired adult neurogenesis include major depressive disorder, Alzheimer’s disease, Parkinson’s disease, schizophrenia, and post-traumatic stress disorder. All of these correlate well with the reduction in hippocampal volume, cognitive deficits, and mood dysregulation [74]. A recent 3-month prospective study by Yiping Lu et al. conducted in COVID-19 recovered patients found that there was grey matter enlargement in olfactory cortices and hippocampus bilaterally [75]. Yiping Lu et al. also found that the grey matter volume of the hippocampus was negatively related to loss of smell during the disease phase [75]. Anosmia over a course of time in upper respiratory tract infections was found to be associated with a decrease in the grey matter volume [GMV] of the central olfactory system due to loss of stimulation while enlargement of GMV is observed during recovery [76]. Functional compensation in the form of enlarged neurons and an increase in the dendritic spine and compensatory enhanced neurogenesis are believed to be the reason behind GMV enlargement during recovery [77]. Loss of memory that persisted 3 months after the active infection in COVID-19 recovered patients was found to be negatively related to hippocampal grey matter volume [75]. Memory acquisition depends on newborn neurons and impairment in the acquisition of memory occurs due to inhibition of adult neurogenesis in the hippocampus [78, 79].

Anosmia is regarded as the key feature of COVID-19 which either occurs as an only symptom or in association with other signs and symptoms [80, 81]. Earlier studies show that any impairment in olfactory neurogenesis is associated with anosmia since neurogenesis in the olfactory epithelium and olfactory bulb is essential for the sense of smell [82, 83]. Dysfunction or atrophy of the olfactory bulb was observed in COVID-19 patients by recent studies done using brain imaging reports [84, 85]. Pathogenic changes in COVID-19 seem to cause loss of dopaminergic neurons, defects in the dopamine system, and exacerbate the clinical features of Parkinson’s disease [PD] [86, 87]. Anosmia is an important premotor symptom of PD which is not directly related to the neurodegenerative process in substantia nigra but appears to be related to defective adult neurogenesis [88, 89].

Understanding the process of adult neurogenesis in COVID-19 may reveal a critical role of the regenerative capacity of NPCs in combating the neuropsychiatric consequence of COVID-19. There are no studies or evidence to link COVID-19 with adult neurogenesis yet. Based on the factors like the presentation of neuropsychiatric symptoms in COVID-19, the occurrence of symptoms like anosmia, memory and cognitive deficits in COVID-19, the neuro-invasive potential of SARS-CoV-2, ACE2 expression in sites of adult neurogenesis, increased levels of pro-inflammatory cytokines like IL-6, Il-1β which inhibit adult neurogenesis and impact of earlier coronavirus infections, it might not be far-fetched to say that COVID-19 could have a possible impact on adult neurogenesis. There is a severe scarcity in research analysing the effect of SARS-CoV-2 infection on adult neurogenesis. The current chapter, which is speculative and based on a thorough literature search, discusses the possible changes in adult neurogenesis in COVID-19 emphasising the role of ACE2. If proven to be true in the future, the findings in this article will help in achieving early intervention to address the neuropsychiatric long-term consequence of COVID-19.

Advertisement

8. Role of ACE2 in adult neurogenesis in COVID-19

SARS-CoV-2 entry into the cell through ACE2 is followed by the downregulation of ACE2. A decrease in ACE2 will lead to dysregulation of RAS and various other complications. A recent study has found that ACE2 is expressed in young neurons and in human-induced pluripotent stem cell-derived neural progenitor cells [90]. ACE2 is found to have various neuroprotective functions. It converts neurotoxic amyloid protein Aβ into neuroprotective one in transgenic mice [91]. ACE2 activator, diminazene increased CREB, BDNF, glutamate, and nicotinic receptor and decreased the levels of apoptotic and inflammatory proteins in the AD model of D-galactose-ovariectomized rats [92]. All these factors play a major role in adult neurogenesis. ACE2 deficiency in mice was found to be accompanied by significantly impaired learning and memory [93]. Exercise-induced neurogenesis in the dentate gyrus was abolished in ACE2 deficient mice. Ang II, Ang [1-7], and Mas receptors were not found to be responsible and hence the mediator of this effect is not identified yet [94].

ACE2 expression is stronger in the enterocytes of the small intestine and colon, which is even higher than in the lungs. Neural ganglia cells in the colon of the enteric nervous system also express ACE2 receptors. Intestinal ACE2 plays a major role in the transport of neutral amino acids via B0AT1, neutral amino acid transporter. ACE2/B0AT1 complex regulates the composition and function of gut microbiota. ACE2 knockout animals showed lower levels of serum neutral amino acid levels like tryptophan, and impaired gut microbiota composition along with reduced expression of small intestinal antimicrobial peptides [95]. Enteric infection is an important presentation of COVID-19. Faeces of COVID-19 patients were found to have Viral mRNA [96, 97]. SARS-CoV-2 entry via the enteric route into host cell leads to ACE2 shedding due to S priming which may lead to gut microbiota dysbiosis [98]. Depletion of gut microbiota by prolonged antibiotic treatment resulted in impairment in cognitive function and hippocampal neurogenesis in adult mice [99]. The existence of a strong link between gut microbiota and the development of mental disorders, depression, and anxiety which are associated with impaired adult neurogenesis has been explored in recent studies [100].

Neuroinflammation directly impairs adult hippocampal neurogenesis. Pro-inflammatory cytokine IL-1β, IL-6, IFN-α causes a reduction in neural cell proliferation and suppresses adult hippocampal neurogenesis [101, 102, 103]. SARS-CoV-2 entry into the brain triggers an immune response by activating microglia, astrocytes, and other immune cells. This leads to increased production of cytokines in the brain. Cytokine storm which is a deadly hyperinflammatory response is considered to be a hallmark feature of COVID-19 pathogenesis [104]. Hypercytokinemia of IL-6, IL-10, and TNF-α was observed in COVID-19 patients. Increased levels of IL-6 correlate with mortality and the need for ventilator support [105, 106].

Thus, there are different possible mechanisms through which SARS-CoV-2 affects adult neurogenesis via ACE2. This chapter, however, will focus on the role of ACE2 in possible alterations in adult neurogenesis in COVID-19 via neurotransmitters.

Advertisement

9. ACE2 and neurotransmitters involved in adult neurogenesis

Neurotransmitter signalling is found to play a major role in the formation of new neurons in addition to its clear and indisputable role in communication between neurons. Starting from embryogenesis, neurotransmitters are involved in neuronal proliferation. In adult neurogenesis, they influence various steps including proliferation, differentiation, and migration. In addition to the direct action of neurotransmitters on adult neurogenesis, they also influence other factors that regulate neurogenesis like neurotrophic factors and growth factors [107].

9.1 ACE2 and serotonin

Serotonin is a crucial monoaminergic neurotransmitter that acts as a mood stabiliser and is associated with feelings of happiness, well-being, and contentedness. In the brain, it is synthesised by the Raphe nuclei neurons in the brain stem from tryptophan using neuron-specific tryptophan hydroxylase 2 enzymes. Vesicular monoamine transporter 2 [VMAT] packs the synthesised serotonin into vesicles. Serotonin transporters [SERT] re-uptake serotonin back to presynaptic neurons after its release, thereby regulating its extracellular levels [108]. The serotonergic fibres from raphe nuclei have projections throughout the brain and especially to the granule cells and interneurons of the dentate gyrus of the hippocampus. Serotonin is known to play a major regulatory role in adult hippocampal neurogenesis. Selective serotonin reuptake inhibitors [SRRI] are commonly used antidepressants that act by increasing serotonin levels in the brain causes clinical improvement associated with an increase in adult hippocampal neurogenesis characterised by increased neuronal proliferation and number of newborn neurons [109]. Malberg et al. in 2000 were the first to show that chronic treatment with fluoxetine improved adult hippocampal neurogenesis [109]. In the dentate gyrus, serotonin is known to promote neuronal development and its depletion was found to cause reduced dendritic spine density of granule cells [110, 111, 112, 113]. Chronic treatment with SSRI, fluoxetine was found to increase the survival of newborn neurons in the dentate gyrus [109, 114]. In stress models like inescapable stress, cold restraint stress in the animal model, fluoxetine administration was found to exhibit neurogenic and neuroprotective roles in the hippocampus [114, 115]. Accelerated synaptogenesis and increased long-term potentiation [LTP] in the hippocampus were also observed by long-term treatment by fluoxetine [116].

Recent studies have found that ACE2 plays a major role in the biosynthesis of serotonin [5HT]. The precursor for 5HT is an essential amino acid, tryptophan which can cross the blood–brain barrier and whose intestinal absorption was found to be reduced by 70% in case of ACE2 deficiency. Thus, ACE2 has an indirect modulatory role in 5HT synthesis in the brain [117]. There are recent studies that show that 5HT synthesis in the brain is dependent on ACE2, which acts by modulating 5HT metabolism and ACE2 deficiency leads to decreased serum tryptophan levels and decreased serotonin levels in the brain [94].

9.2 ACE2 and dopamine

Dopamine is involved in executive functions, volition, motor control, motivation, pleasure/reward, and attention/concentration [118]. The role and mechanism of action of dopamine in adult neurogenesis are not elucidated fully. Dopamine was found to modulate cell proliferation in the embryonic brain [119]. Hippocampus and sub-ventricular zone [SVZ] which are the neurogenic niche containing neural stem cells receive dopaminergic projections from the substantia nigra and ventral tegmental area. Dopamine receptors are also widely expressed in these two areas and play a regulatory role in adult neurogenesis and neural plasticity [120, 121]. Earlier studies show that depletion of dopamine in the rat model reduces both proliferation and survival of neural precursor cells in the sub-granular zone [SGZ] of the dentate gyrus [122, 123]. Dopaminergic denervation in substantia nigra caused a significant reduction in the proliferation of neural stem cells in SGZ and SVZ which was reversed by D2 receptor stimulation in rodents [123]. In humans, post-mortem studies have revealed that the number of neural precursor cells in SGZ and SVZ was reduced in patients with Parkinson’s disease [124]. Dopamine was also found to increase the type 2A early progenitor cell in the hippocampus of rodents via D1 like receptors [118]. Dopamine receptor agonist pramipexole increases the proliferation and survival of newborn neurons in SVZ, olfactory bulb [119].

RAS plays a major role in dopaminergic vulnerability through AT1 receptors. Dysregulation of RAS due to the downregulation of ACE2 induced by SARS-CoV-2 may increase the vulnerability of dopaminergic neurons and subsequently dopamine levels [125]. Interactions between dopamine and angiotensin receptors that are counterregulatory in nature are observed in substantia nigra and striatum [125]. The gene for ACE2 was found to coexpress and coregulate with that of dopa decarboxylase [DDC] in non-neuronal cells, which is a major enzyme of dopamine, serotonin, and histamine biosynthesis. DDC converts L-3,4-dihydroxyphenylalanine [L-DOPA] into dopamine which subsequently forms norepinephrine and epinephrine and L-5-hydroxytryptophan into serotonin. This coexpression and coregulation link between the genes for ACE2 and DDC gives rise to the possibility of a functional link between the actions of ACE2 and DDC [i.e.,] in the synthesis of Ang [1-7] and dopamine and serotonin mediated by ACE2 and DDC, respectively [126]. Following the infusion of Ang [1-7] in the hypothalamus of rats, brain dopamine levels increased which emphasises the link between ACE2 and DDC. SARS-CoV-2 induced downregulation of ACE2 could cause the decreased synthesis of serotonin and dopamine [94, 127].

The SARS-CoV-2 infection has been found to cause loss of dopaminergic neurons and deficits in the dopamine system [86, 128]. ACE2 expression is high in dopaminergic neurons and the downregulation of ACE2 by SARS-CoV-2 may cause depletion of dopaminergic neurons and dopamine levels. This is evident from the worsening of symptoms observed in COVID-19 patients with Parkinson’s disease [PD], requiring increased dopamine replacement therapy [129]. ACE2 deletion in the knockout mouse model caused a significant reduction in dopamine D1 mRNA expression in substantia nigra [130].

9.3 ACE2 and norepinephrine

Norepinephrine is an important catecholamine that is involved in alertness, arousal, sleep–wake cycle, memory storage, and emotions. It modulates various functions of the hippocampus like learning, memory, and mood. Noradrenergic axon terminals arising from the locus coeruleus densely innervate the neurogenic niche in the adult hippocampus [131]. Norepinephrine along with the other monoaminergic neurotransmitters plays a major role in adult neurogenesis. Norepinephrine was found to activate the stem cells and neural precursor cells via β3-adrenergic receptors where non-proliferating latent precursor cells develop the ability to respond to mitogens and generate neurospheres. It also increases the proliferation of early progenitor cells in the adult hippocampus via β2-adrenergic receptors [132, 133]. Depletion of norepinephrine significantly decreased the proliferation of progenitor cells of granule cells in the hippocampus [134]. Antidepressants that selectively increase norepinephrine were found to increase adult hippocampal neurogenesis [132].

Downregulation of ACE2 by SARS-CoV-2 may affect the activity of DDC due to the coexpression and coregulation between the genes for ACE2 and DDC. This could lead to a decrease in the biosynthesis of dopamine and subsequently norepinephrine [126].

9.4 ACE2 and glutamate and GABA

Glutamate is the predominant excitatory neurotransmitter of the CNS. It plays a vital role in both embryonic brain development and adult neurogenesis. Its extracellular levels are especially higher in the neurogenic niche when compared to other areas of the brain [135, 136]. It has trophic effects on the developing neurons before synapse formation like proliferation, migration, and maturation. It causes an increase in the proliferation of neural progenitor cells [NPC]. The NPCs express NMDA metabotropic glutamate receptors, stimulation of which caused increased intracellular calcium and activation of NeuroD1, proneural gene [137]. Glutamate signalling plays a positive role in maintaining the proliferation of NPCs and the survival rates of newborn neurons [137, 138].

Gamma-aminobutyric acid [GABA] is a principal inhibitory neurotransmitter in the CNS. It is produced from glutamate by the action of the enzymes glutamate decarboxylase GAD65 and GAD67 [139]. Dysfunction in the GABAergic system is implicated in major depressive disorder and anxiety [140]. However, in the developing brain, GABA exerts an excitatory effect, that is, GABA is excitatory in immature neurons. Tonic discharge from GABAergic neurons is necessary for maintaining the quiescent state of NPCs. The absence of GABAergic excitability will cause impairment in neuronal maturation and synapse formation while an excess of it over newborn neurons will lead to seizures [141]. In SGZ, GABA mediates depolarisation of progenitor cells which is involved in the incorporation of AMPA receptors in immature granule cells, which is critical for learning and formation of memory [142]. It has a negative influence on neuroblasts. It inhibits the proliferation and migration of neuroblasts. It also inhibits the proliferation of NPCs [143, 144, 145]. It also promotes the differentiation of hippocampal NPCs. GABAA receptor agonist, phenobarbital caused a reduction in NPC proliferation and increase in differentiation which resulted in an increased number of newborn neurons [146]. Thus, it plays crucial role in different stages of adult neurogenesis. GABA and glutamate signalling play a major role in adult neurogenesis. Selective activation of the receptor subtypes of GABA and glutamate expressed in NPCs plays a pivotal role in self-replication and fate commitment of the developing neurons into a particular progeny [147].

A recent study has found ACE2 to be located mainly in excitatory neurons of the brain and to a lesser extent in inhibitory neurons like GABAergic neurons [148]. This indicates that SARS-CoV-2 once enters the brain has the potential to access the glutamatergic and GABAergic neurons. The consequence of this is not known however, viral entry may trigger apoptotic pathways and cause excitatory-inhibitory imbalance, and lead to neuronal death [149]. Cytokine release from infected neurons and other activated microglia and astrocytes may also cause a decrease in glutamate and GABA [150]. These effects are implicated along with impaired adult neurogenesis in neurodegenerative diseases like Parkinson’s disease and Alzheimer’s disease. Seizure is one of the neurological symptoms in COVID-19 patients, in which an increase in glutamate levels and decrease in GABA levels in the cerebral cortex and hippocampus is an implicated mechanism [151]. This further emphasises the possible impact of SARS-CoV-2 on glutamate and GABA.

Thus, SARS-CoV-2 induced downregulation of ACE2 in COVID-19 is potentially detrimental to adult neurogenesis. ACE2 deficiency affects the levels and actions of the neurotransmitters serotonin, dopamine, norepinephrine, GABA, and glutamate which play crucial roles in adult neurogenesis.

Advertisement

10. Conclusion

SARS-CoV-2 has been found to have a high affinity to ACE2 receptors. Such high affinity has been linked to affect neurogenesis through a variety of mechanisms. The present chapter has clearly postulated the link between this deadly virus and its effect on monoaminergic neurotransmitters as well as GABA and glutamate which play a major role in adult neurogenesis. As ACE2 receptors are expressed in the hippocampus, decreased neurogenesis in this region could be one of the major factors behind the neuropsychiatric disorders associated with patients affected with COVID-19. Awareness and early intervention to prevent and treat long-term psychiatric consequences of COVID-19 are crucial. We should be aware of the possibility that in the long term, COVID-19 may be associated with cognitive and psychiatric disorders in those who recovered. Despite having a mild course of disease in children and adolescents, immunological response to the infection in this population may affect synaptic pruning which may lead to various issues that may not be immediately apparent. Insights into the various machinations of adult neurogenesis in COVID-19 can be used to engineer the process to help with the pathological changes in the brain inflicted by the disease.

Acknowledgments

The authors would like to thank the Deanship of Scientific Research, Majmaah University for the support of this chapter.

References

  1. 1. Worldometer. Worldometer Coronavirus. United States: Dadax Limited; 2021. Available from: https://www.worldometers.info/coronavirus/
  2. 2. WHO. Archived: WHO Timeline—COVID-19. Geneva: WHO; 2020 Available from: https://www.who.int/news/item/27-04-2020-who-timeline---covid-19
  3. 3. Wang S, Ye J, Kang Z, Peng H, Mackey V, Sun L. The COVID-19 pandemic and the potential treatment of the novel coronavirus SARS-CoV-2. American Journal of Translational Research. 2021;13(3):871
  4. 4. CDC. COVID-19; Symptoms of COVID-19. United Kingdom: CDC; 2020 Available from: https://www.cdc.gov/coronavirus/2019-ncov/symptoms-testing/symptoms.html [Accessed: 22 February 2021]
  5. 5. Pezzini A, Padovani A. Lifting the mask on neurological manifestations of COVID-19. Nature Reviews Neurology. 2020;16(11):636-644
  6. 6. Desai I, Manchanda R, Kumar N, Tiwari A, Kumar M. Neurological manifestations of coronavirus disease 2019: Exploring past to understand present. Neurological Sciences. 2021;42(3):773-785
  7. 7. Huang C, Wang Y, Li X, Ren L, Zhao J, Hu Y, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. The Lancet. 2020;395(10223):497-506
  8. 8. Cardona GC, Pájaro LDQ, Marzola IDQ, Villegas YR, Salazar LRM. Neurotropism of SARS-CoV 2: Mechanisms and manifestations. Journal of the Neurological Sciences. 2020;412:116824
  9. 9. Moriguchi T, Harii N, Goto J, Harada D, Sugawara H, Takamino J, et al. A first case of meningitis/encephalitis associated with SARS-Coronavirus-2. International Journal of Infectious Diseases. 2020;94:55-58
  10. 10. Chu H, Chan JF-W, Yuen TT-T, Shuai H, Yuan S, Wang Y, et al. Comparative tropism, replication kinetics, and cell damage profiling of SARS-CoV-2 and SARS-CoV with implications for clinical manifestations, transmissibility, and laboratory studies of COVID-19: An observational study. The Lancet Microbe. 2020;1(1):e14-e23
  11. 11. Arbour N, Day R, Newcombe J, Talbot PJ. Neuroinvasion by human respiratory coronaviruses. Journal of Virology. 2000;74(19):8913-8921
  12. 12. Rogers JP, Chesney E, Oliver D, Pollak TA, McGuire P, Fusar-Poli P, et al. Psychiatric and neuropsychiatric presentations associated with severe coronavirus infections: A systematic review and meta-analysis with comparison to the COVID-19 pandemic. The Lancet Psychiatry. 2020;7(7):611-627
  13. 13. Corman V, Lienau J, Witzenrath M. Coronaviruses as the cause of respiratory infections. Der Internist. 2019;60(11):1136-1145
  14. 14. Hewings-Martin Y. How do SARS and MERS compare with COVID-19? United Kingdom: MedicalNewsToday; 2020
  15. 15. Gallagher TM, Buchmeier MJ. Coronavirus spike proteins in viral entry and pathogenesis. Virology. 2001;279(2):371-374
  16. 16. Hoffmann M, Kleine-Weber H, Schroeder S, Krüger N, Herrler T, Erichsen S, et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell. 2020;181(2):271-80.e8
  17. 17. Coutard B, Valle C, de Lamballerie X, Canard B, Seidah NG, Decroly E. The spike glycoprotein of the new coronavirus 2019-nCoV contains a furin-like cleavage site absent in CoV of the same clade. Antiviral Research. 2020;176:104742
  18. 18. Wrapp D, Wang N, Corbett KS, Goldsmith JA, Hsieh C-L, Abiona O, et al. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation. Science. 2020;367(6483):1260-1263
  19. 19. Donoghue M, Hsieh F, Baronas E, Godbout K, Gosselin M, Stagliano N, et al. A novel angiotensin-converting enzyme–related carboxypeptidase [ACE2] converts angiotensin I to angiotensin 1-9. Circulation Research. 2000;87(5):e1-e9
  20. 20. Tipnis SR, Hooper NM, Hyde R, Karran E, Christie G, Turner AJ. A human homolog of angiotensin-converting enzyme: Cloning and functional expression as a captopril-insensitive carboxypeptidase. Journal of Biological Chemistry. 2000;275(43):33238-33243
  21. 21. Gheblawi M, Wang K, Viveiros A, Nguyen Q, Zhong J-C, Turner AJ, et al. Angiotensin-converting enzyme 2: SARS-CoV-2 receptor and regulator of the renin-angiotensin system: Celebrating the 20th anniversary of the discovery of ACE2. Circulation Research. 2020;126(10):1456-1474
  22. 22. Kalea AZ, Batlle D. Apelin and ACE2 in cardiovascular disease. Current Opinion in Investigational Drugs. 2010;11(3):273-282
  23. 23. Vickers C, Hales P, Kaushik V, Dick L, Gavin J, Tang J, et al. Hydrolysis of biological peptides by human angiotensin-converting enzyme-related carboxypeptidase. Journal of Biological Chemistry. 2002;277(17):14838-14843
  24. 24. Warner FJ, Lew RA, Smith AI, Lambert DW, Hooper NM, Turner AJ. Angiotensin-converting enzyme 2 [ACE2], but not ACE, is preferentially localized to the apical surface of polarized kidney cells. Journal of Biological Chemistry. 2005;280(47):39353-39362
  25. 25. Clarke NE, Turner AJ. Angiotensin-converting enzyme 2: The first decade. International Journal of Hypertension. 2012;2012:307315
  26. 26. Pagliaro P, Penna C. ACE/ACE2 ratio: A key also in 2019 coronavirus disease [Covid-19]? Frontiers in Medicine. 2020;7(335):1-5
  27. 27. Danilczyk U, Eriksson U, Crackower MA, Penninger JM. A story of two ACEs. Journal of Molecular Medicine. 2003;81(4):227-234
  28. 28. Walls AC, Park Y-J, Tortorici MA, Wall A, McGuire AT, Veesler D. Structure, function, and antigenicity of the SARS-CoV-2 spike glycoprotein. Cell. 2020;181(2):281-92.e6
  29. 29. Hernández VS, Zetter MA, Guerra EC, Hernández-Araiza I, Karuzin N, Hernández-Pérez OR, et al. ACE2 expression in rat brain: Implications for COVID-19 associated neurological manifestations. bioRxiv. 2021:2021.05.01.442293
  30. 30. Bilinska K, Jakubowska P, Von Bartheld CS, Butowt R. Expression of the SARS-CoV-2 entry proteins, ACE2 and TMPRSS2, in cells of the olfactory epithelium: Identification of cell types and trends with age. ACS Chemical Neuroscience. 2020;11(11):1555-1562
  31. 31. Hamming I, Timens W, Bulthuis M, Lely A, Gv N, van Goor H. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. The Journal of Pathology: A Journal of the Pathological Society of Great Britain and Ireland. 2004;203(2):631-637
  32. 32. Alenina N, Bader M. ACE2 in brain physiology and pathophysiology: Evidence from transgenic animal models. Neurochemical Research. 2019;44(6):1323-1329
  33. 33. Silva RAP, Chu Y, Miller JD, Mitchell IJ, Penninger JM, Faraci FM, et al. Impact of ACE2 deficiency and oxidative stress on cerebrovascular function with aging. Stroke. 2012;43(12):3358-3363
  34. 34. Rice GI, Thomas DA, Grant PJ, Turner AJ, Hooper NM. Evaluation of angiotensin-converting enzyme [ACE], its homologue ACE2 and neprilysin in angiotensin peptide metabolism. Biochemical Journal. 2004;383(1):45-51
  35. 35. Hellner K, Walther T, Schubert M, Albrecht D. Angiotensin-[1-7] enhances LTP in the hippocampus through the G-protein-coupled receptor Mas. Molecular and Cellular Neuroscience. 2005;29(3):427-435
  36. 36. Bernardi S, Toffoli B, Zennaro C, Tikellis C, Monticone S, Losurdo P, et al. High-salt diet increases glomerular ACE/ACE2 ratio leading to oxidative stress and kidney damage. Nephrology Dialysis Transplantation. 2012;27(5):1793-1800
  37. 37. Lavrentyev EN, Malik KU. High glucose-induced Nox1-derived superoxides downregulate PKC-βII, which subsequently decreases ACE2 expression and ANG [1-7] formation in rat VSMCs. American Journal of Physiology-Heart and Circulatory Physiology. 2009;296(1):H106-HH18
  38. 38. Yuan Y-M, Luo L, Guo Z, Yang M, Ye R-S, Luo C. Activation of renin–angiotensin–aldosterone system [RAAS] in the lung of smoking-induced pulmonary arterial hypertension [PAH] rats. Journal of the Renin-Angiotensin-Aldosterone System. 2015;16(2):249-253
  39. 39. Kehoe PG, Wong S, Mulhim NA, Palmer LE, Miners JS. Angiotensin-converting enzyme 2 is reduced in Alzheimer’s disease in association with increasing amyloid-β and tau pathology. Alzheimer’s Research & Therapy. 2016;8(1):1-10
  40. 40. Valenzuela R, Rodriguez-Perez AI, Costa-Besada MA, Rivas-Santisteban R, Garrido-Gil P, Lopez-Lopez A, et al. An ACE2/Mas-related receptor MrgE axis in dopaminergic neuron mitochondria. Redox Biology. 2021;46:102078
  41. 41. Labandeira-García JL, Garrido-Gil P, Rodriguez-Pallares J, Valenzuela R, Borrajo A, Rodríguez-Perez AI. Brain renin-angiotensin system and dopaminergic cell vulnerability. Frontiers in Neuroanatomy. 2014;8:67
  42. 42. Altman J, Das GD. Autoradiographic and histological evidence of postnatal hippocampal neurogenesis in rats. Journal of Comparative Neurology. 1965;124(3):319-335
  43. 43. Taverna E, Götz M, Huttner WB. The cell biology of neurogenesis: Toward an understanding of the development and evolution of the neocortex. Annual Review of Cell and Developmental Biology. 2014;30:465-502
  44. 44. Apple DM, Fonseca RS, Kokovay E. The role of adult neurogenesis in psychiatric and cognitive disorders. Brain Research. 2017;1655:270-276
  45. 45. Mu Y, Lee SW, Gage FH. Signaling in adult neurogenesis. Current Opinion in Neurobiology. 2010;20(4):416-423
  46. 46. Braun SM, Jessberger S. Adult neurogenesis: Mechanisms and functional significance. Development. 2014;141(10):1983-1986
  47. 47. Cameron HA, Glover LR. Adult neurogenesis: Beyond learning and memory. Annual Review of Psychology. 2015;66:53-81
  48. 48. Gage FH. Mammalian neural stem cells. Science. 2000;287(5457):1433-1438
  49. 49. Zhao C, Teng EM, Summers RG, Ming G-l, Gage FH. Distinct morphological stages of dentate granule neuron maturation in the adult mouse hippocampus. Journal of Neuroscience. 2006;26(1):3-11
  50. 50. Kempermann G, Jessberger S, Steiner B, Kronenberg G. Milestones of neuronal development in the adult hippocampus. Trends in Neurosciences. 2004;27(8):447-452
  51. 51. Toni N, Laplagne DA, Zhao C, Lombardi G, Ribak CE, Gage FH, et al. Neurons born in the adult dentate gyrus form functional synapses with target cells. Nature Neuroscience. 2008;11(8):901-907
  52. 52. Aimone JB, Jessberger S, Gage FH. Adult neurogenesis. Scholarpedia. 2007;2(2):2100
  53. 53. Ge S, Goh EL, Sailor KA, Kitabatake Y, Ming GL, Song H. GABA regulates synaptic integration of newly generated neurons in the adult brain. Nature. 2006;439(7076):589-593
  54. 54. Ge S, Sailor KA, Gl M, Song H. Synaptic integration and plasticity of new neurons in the adult hippocampus. The Journal of Physiology. 2008;586(16):3759-3765
  55. 55. Schmidt-Hieber C, Jonas P, Bischofberger J. Enhanced synaptic plasticity in newly generated granule cells of the adult hippocampus. Nature. 2004;429(6988):184-187
  56. 56. Alonso M, Viollet C, Gabellec M-M, Meas-Yedid V, Olivo-Marin J-C, Lledo P-M. Olfactory discrimination learning increases the survival of adult-born neurons in the olfactory bulb. Journal of Neuroscience. 2006;26(41):10508-10513
  57. 57. Gould E, Beylin A, Tanapat P, Reeves A, Shors TJ. Learning enhances adult neurogenesis in the hippocampal formation. Nature Neuroscience. 1999;2(3):260-265
  58. 58. Lugert S, Basak O, Knuckles P, Haussler U, Fabel K, Götz M, et al. Quiescent and active hippocampal neural stem cells with distinct morphologies respond selectively to physiological and pathological stimuli and aging. Cell Stem Cell. 2010;6(5):445-456
  59. 59. Murphy T, Dias GP, Thuret S. Effects of diet on brain plasticity in animal and human studies: Mind the gap. Neural Plasticity. 2014:1-32
  60. 60. Van Riel D, Verdijk R, Kuiken T. The olfactory nerve: A shortcut for influenza and other viral diseases into the central nervous system. The Journal of Pathology. 2015;235(2):277-287
  61. 61. Butowt R, Bilinska K. SARS-CoV-2: Olfaction, brain infection, and the urgent need for clinical samples allowing earlier virus detection. ACS Chemical Neuroscience. 2020;11(9):1200-1203
  62. 62. Sungnak W, Huang N, Bécavin C, Berg M, Queen R, Litvinukova M, et al. SARS-CoV-2 entry factors are highly expressed in nasal epithelial cells together with innate immune genes. Nature Medicine. 2020;26(5):681-687
  63. 63. Bohmwald K, Gálvez N, Ríos M, Kalergis A. Neurologic alterations due to respiratory virus infections. Frontiers in Cellular Neuroscience. 2018;12(386):2018
  64. 64. Meinhardt J, Radke J, Dittmayer C, Franz J, Thomas C, Mothes R, et al. Olfactory transmucosal SARS-CoV-2 invasion as a port of central nervous system entry in individuals with COVID-19. Nature Neuroscience. 2021;24(2):168-175
  65. 65. Li YC, Bai WZ, Hashikawa T. The neuroinvasive potential of SARS-CoV2 may play a role in the respiratory failure of COVID-19 patients. Journal of Medical Virology. 2020;92(6):552-555
  66. 66. Shi Y, Li Z, Yang C, Liu C. The role of gut-brain axis in SARA-CoV-2 neuroinvasion: Culprit or innocent bystander? Brain, Behavior, and Immunity. 2021;94:476-477
  67. 67. Desforges M, Le Coupanec A, Dubeau P, Bourgouin A, Lajoie L, Dubé M, et al. Human coronaviruses and other respiratory viruses: Underestimated opportunistic pathogens of the central nervous system? Viruses. 2020;12(1):14
  68. 68. Finsterer J, Stollberger C. Update on the neurology of COVID-19. Journal of Medical Virology. 2020;92(11):2316-2318
  69. 69. Li H, Xue Q, Xu X. Involvement of the nervous system in SARS-CoV-2 infection. Neurotoxicity Research. 2020;1:1-7
  70. 70. Rabiee A, Nikayin S, Hashem MD, Huang M, Dinglas VD, Bienvenu OJ, et al. Depressive symptoms after critical illness: A systematic review and meta-analysis. Critical Care Medicine. 2016;44(9):1744
  71. 71. Nikayin S, Rabiee A, Hashem MD, Huang M, Bienvenu OJ, Turnbull AE, et al. Anxiety symptoms in survivors of critical illness: A systematic review and meta-analysis. General Hospital Psychiatry. 2016;43:23-29
  72. 72. Parker AM, Sricharoenchai T, Raparla S, Schneck KW, Bienvenu OJ, Needham DM. Posttraumatic stress disorder in critical illness survivors: A metaanalysis. Critical Care Medicine. 2015;43(5):1121-1129
  73. 73. Hopkins RO, Weaver LK, Pope D, Orme JF Jr, Bigler ED, Larson-Lohr V. Neuropsychological sequelae and impaired health status in survivors of severe acute respiratory distress syndrome. American Journal of Respiratory and Critical Care Medicine. 1999;160(1):50-56
  74. 74. DeCarolis NA, Eisch AJ. Hippocampal neurogenesis as a target for the treatment of mental illness: A critical evaluation. Neuropharmacology. 2010;58(6):884-893
  75. 75. Lu Y, Li X, Geng D, Mei N, Wu P-Y, Huang C-C, et al. Cerebral micro-structural changes in COVID-19 patients–an MRI-based 3-month follow-up study. EClinicalMedicine. 2020;25:100484
  76. 76. Gellrich J, Han P, Manesse C, Betz A, Junghanns A, Raue C, et al. Brain volume changes in hyposmic patients before and after olfactory training. The Laryngoscope. 2018;128(7):1531-1536
  77. 77. Karstensen HG, Vestergaard M, Baaré WF, Skimminge A, Djurhuus B, Ellefsen B, et al. Congenital olfactory impairment is linked to cortical changes in prefrontal and limbic brain regions. Brain Imaging and Behavior. 2018;12(6):1569-1582
  78. 78. Recinto P, Samant ARH, Chavez G, Kim A, Yuan CJ, Soleiman M, et al. Levels of neural progenitors in the hippocampus predict memory impairment and relapse to drug seeking as a function of excessive methamphetamine self-administration. Neuropsychopharmacology. 2012;37(5):1275-1287
  79. 79. Misane I, Kruis A, Pieneman AW, Ögren SO, Stiedl O. GABAA receptor activation in the CA1 area of the dorsal hippocampus impairs consolidation of conditioned contextual fear in C57BL/6J mice. Behavioural Brain Research. 2013;238:160-169
  80. 80. Aziz M, Goyal H, Haghbin H, Lee-Smith WM, Gajendran M, Perisetti A. The association of “loss of smell” to COVID-19: A systematic review and meta-analysis. The American Journal of the Medical Sciences. 2020;361(2):216-225
  81. 81. Meng X, Deng Y, Dai Z, Meng Z. COVID-19 and anosmia: A review based on up-to-date knowledge. American Journal of Otolaryngology. 2020;41(5):102581
  82. 82. Boesveldt S, Postma EM, Boak D, Welge-Luessen A, Schöpf V, Mainland JD, et al. Anosmia—a clinical review. Chemical Senses. 2017;42(7):513-523
  83. 83. Lledo P-M, Valley M. Adult olfactory bulb neurogenesis. Cold Spring Harbor Perspectives in Biology. 2016;8(8):a018945
  84. 84. Chiu A, Fischbein N, Wintermark M, Zaharchuk G, Yun PT, Zeineh M. COVID-19-induced anosmia associated with olfactory bulb atrophy. Neuroradiology. 2021;63(1):147-148
  85. 85. Galougahi MK, Ghorbani J, Bakhshayeshkaram M, Naeini AS, Haseli S. Olfactory bulb magnetic resonance imaging in SARS-CoV-2-induced anosmia: The first report. Academic Radiology. 2020;27(6):892-893
  86. 86. Sulzer D, Antonini A, Leta V, Nordvig A, Smeyne RJ, Goldman JE, et al. COVID-19 and possible links with Parkinson’s disease and parkinsonism: From bench to bedside. npj. Parkinson's Disease. 2020;6(1):1-10
  87. 87. Brundin P, Nath A, Beckham JD. Is COVID-19 a perfect storm for Parkinson’s disease? Trends in Neurosciences. 2020;43(12):931-933
  88. 88. Marxreiter F, Regensburger M, Winkler J. Adult neurogenesis in Parkinson’s disease. Cellular and Molecular Life Sciences. 2013;70(3):459-473
  89. 89. Winner B, Kohl Z, Gage FH. Neurodegenerative disease and adult neurogenesis. European Journal of Neuroscience. 2011;33(6):1139-1151
  90. 90. Kase Y, Okano H. Expression of ACE2 and a viral virulence-regulating factor CCN family member 1 in human iPSC-derived neural cells: Implications for COVID-19-related CNS disorders. Inflammation and Regeneration. 2020;40(1):1-8
  91. 91. Liu S, Liu J, Miura Y, Tanabe C, Maeda T, Terayama Y, et al. Conversion of Aβ43 to Aβ40 by the successive action of angiotensin-converting enzyme 2 and angiotensin-converting enzyme. Journal of Neuroscience Research. 2014;92(9):1178-1186
  92. 92. Kamel AS, Abdelkader NF, Abd El-Rahman SS, Emara M, Zaki HF, Khattab MM. Stimulation of ACE2/ANG[1-7]/Mas axis by diminazene ameliorates Alzheimer’s disease in the D-Galactose-ovariectomized rat model: Role of PI3K/Akt pathway. Molecular Neurobiology. 2018;55(10):8188-8202
  93. 93. Wang X-L, Iwanami J, Min L-J, Tsukuda K, Nakaoka H, Bai H-Y, et al. Deficiency of angiotensin-converting enzyme 2 causes deterioration of cognitive function. NPJ Aging and Mechanisms of Disease. 2016;2(1):1-8
  94. 94. Klempin F, Mosienko V, Matthes S, Villela DC, Todiras M, Penninger JM, et al. Depletion of angiotensin-converting enzyme 2 reduces brain serotonin and impairs the running-induced neurogenic response. Cellular and Molecular Life Sciences. 2018;75(19):3625-3634
  95. 95. Hashimoto T, Perlot T, Rehman A, Trichereau J, Ishiguro H, Paolino M, et al. ACE2 links amino acid malnutrition to microbial ecology and intestinal inflammation. Nature. 2012;487(7408):477-481
  96. 96. Chen Y, Chen L, Deng Q, Zhang G, Wu K, Ni L, et al. The presence of SARS-CoV-2 RNA in the feces of COVID-19 patients. Journal of Medical Virology. 2020;92(7):833-840
  97. 97. Liang W, Feng Z, Rao S, Xiao C, Xue X, Lin Z, et al. Diarrhoea may be underestimated: A missing link in 2019 novel coronavirus. Gut. 2020;69(6):1141-1143
  98. 98. Viana SD, Nunes S, Reis F. ACE2 imbalance as a key player for the poor outcomes in COVID-19 patients with age-related comorbidities - Role of gut microbiota dysbiosis. Ageing Research Reviews. 2020;62:101123
  99. 99. Möhle L, Mattei D, Heimesaat Markus M, Bereswill S, Fischer A, Alutis M, et al. Ly6Chi monocytes provide a link between antibiotic-induced changes in Gut Microbiota and adult hippocampal neurogenesis. Cell Reports. 2016;15(9):1945-1956
  100. 100. Peirce JM, Alviña K. The role of inflammation and the gut microbiome in depression and anxiety. Journal of Neuroscience Research. 2019;97(10):1223-1241
  101. 101. Koo JW, Duman RS. IL-1β is an essential mediator of the antineurogenic and anhedonic effects of stress. Proceedings of the National Academy of Sciences. 2008;105(2):751-756
  102. 102. Borsini A, Alboni S, Horowitz MA, Tojo LM, Cannazza G, Su K-P, et al. Rescue of IL-1β-induced reduction of human neurogenesis by omega-3 fatty acids and antidepressants. Brain, Behavior, and Immunity. 2017;65:230-238
  103. 103. Borsini A, Cattaneo A, Malpighi C, Thuret S, Harrison NA, Consortium MI, et al. Interferon-alpha reduces human hippocampal neurogenesis and increases apoptosis via activation of distinct STAT1-dependent mechanisms. International Journal of Neuropsychopharmacology. 2018;21(2):187-200
  104. 104. Kempuraj D, Selvakumar GP, Ahmed ME, Raikwar SP, Thangavel R, Khan A, et al. COVID-19, mast cells, cytokine storm, psychological stress, and neuroinflammation. The Neuroscientist. 2020;26(5-6):402-414
  105. 105. Pedersen SF, Ho Y-C. SARS-CoV-2: A storm is raging. The Journal of Clinical Investigation. 2020;130(5):2202-2205
  106. 106. Vardhana SA, Wolchok JD. The many faces of the anti-COVID immune response. Journal of Experimental Medicine. 2020;217(6):1-10
  107. 107. Hagg T. From neurotransmitters to neurotrophic factors to neurogenesis. The Neuroscientist. 2009;15(1):20-27
  108. 108. Walther D, Peter JU, Bashammakh S, Hortnagl H, Voits M, Fink H, et al. Synthesis of serotonin by a second tryptophan hydroxylase isoform. Science. 2003;299:76
  109. 109. Malberg JE, Eisch AJ, Nestler EJ, Duman RS. Chronic antidepressant treatment increases neurogenesis in adult rat Hippocampus. The Journal of Neuroscience. 2000;20(24):9104-9110
  110. 110. Yan W, Wilson CC, Haring JH. 5-HT1a receptors mediate the neurotrophic effect of serotonin on developing dentate granule cells. Developmental Brain Research. 1997;98(2):185-190
  111. 111. Santarelli L, Saxe M, Gross C, Surget A, Battaglia F, Dulawa S, et al. Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science. 2003;301(5634):805-809
  112. 112. Banasr M, Hery M, Printemps R, Daszuta A. Serotonin-induced increases in adult cell proliferation and neurogenesis are mediated through different and common 5-HT receptor subtypes in the dentate gyrus and the subventricular zone. Neuropsychopharmacology. 2004;29(3):450-460
  113. 113. Brezun J, Daszuta A. Serotonergic reinnervation reverses lesion-induced decreases in PSA-NCAM labeling and proliferation of hippocampal cells in adult rats. Hippocampus. 2000;10(1):37-46
  114. 114. Malberg JE, Duman RS. Cell proliferation in adult hippocampus is decreased by inescapable stress: Reversal by fluoxetine treatment. Neuropsychopharmacology. 2003;28(9):1562-1571
  115. 115. Jayakumar S, Raghunath G, Ilango S, Vijayakumar J, Vijayaraghavan R. Effect of fluoxetine on the Hippocampus of Wistar Albino rats in cold restraint stress model. Journal of Clinical and Diagnostic Research. 2017;11(6):AF01-AAF6
  116. 116. Wang J-W, David DJ, Monckton JE, Battaglia F, Hen R. Chronic fluoxetine stimulates maturation and synaptic plasticity of adult-born hippocampal granule cells. Journal of Neuroscience. 2008;28(6):1374-1384
  117. 117. Singer D, Camargo SM, Ramadan T, Schäfer M, Mariotta L, Herzog B, et al. Defective intestinal amino acid absorption in Ace2 null mice. American Journal of Physiology-Gastrointestinal and Liver Physiology. 2012;303(6):G686-GG95
  118. 118. Takamura N, Nakagawa S, Masuda T, Boku S, Kato A, Song N, et al. The effect of dopamine on adult hippocampal neurogenesis. Progress in Neuro-Psychopharmacology and Biological Psychiatry. 2014;50:116-124
  119. 119. Winner B, Desplats P, Hagl C, Klucken J, Aigner R, Ploetz S, et al. Dopamine receptor activation promotes adult neurogenesis in an acute Parkinson model. Experimental Neurology. 2009;219(2):543-552
  120. 120. Curtis MA, Low VF, Faull RL. Neurogenesis and progenitor cells in the adult human brain: A comparison between hippocampal and subventricular progenitor proliferation. Developmental Neurobiology. 2012;72(7):990-1005
  121. 121. Diaz J, Ridray S, Mignon V, Griffon N, Schwartz J-C, Sokoloff P. Selective expression of dopamine D3 receptor mRNA in proliferative zones during embryonic development of the rat brain. Journal of Neuroscience. 1997;17(11):4282-4292
  122. 122. Khaindrava V, Salin P, Melon C, Ugrumov M, Kerkerian-Le-Goff L, Daszuta A. High frequency stimulation of the subthalamic nucleus impacts adult neurogenesis in a rat model of Parkinson's disease. Neurobiology of Disease. 2011;42(3):284-291
  123. 123. Höglinger GU, Rizk P, Muriel MP, Duyckaerts C, Oertel WH, Caille I, et al. Dopamine depletion impairs precursor cell proliferation in Parkinson disease. Nature Neuroscience. 2004;7(7):726-735
  124. 124. Borta A, Höglinger GU. Dopamine and adult neurogenesis. Journal of Neurochemistry. 2007;100(3):587-595
  125. 125. Villar-Cheda B, Dominguez-Meijide A, Valenzuela R, Granado N, Moratalla R, Labandeira-Garcia JL. Aging-related dysregulation of dopamine and angiotensin receptor interaction. Neurobiology of Aging. 2014;35(7):1726-1738
  126. 126. Nataf S. An alteration of the dopamine synthetic pathway is possibly involved in the pathophysiology of COVID-19. Journal of Medical Virology. 2020;92(10):1743-1744
  127. 127. Pawlak R, Napiorkowska-Pawlak D, Takada Y, Urano T, Nagai N, Ihara H, et al. The differential effect of angiotensin II and angiotensin 1-7 on norepinephrine, epinephrine, and dopamine concentrations in rat hypothalamus: The involvement of angiotensin receptors. Brain Research Bulletin. 2001;54(6):689-694
  128. 128. Merello M, Bhatia KP, Obeso JA. SARS-CoV-2 and the risk of Parkinson's disease: Facts and fantasy. The Lancet Neurology. 2021;20(2):94-95
  129. 129. Antonini A, Leta V, Teo J, Chaudhuri KR. Outcome of Parkinson’s disease patients affected by COVID-19. Movement Disorders. 2020;35(6):905-908
  130. 130. Rodriguez-Perez AI, Garrido-Gil P, Pedrosa MA, Garcia-Garrote M, Valenzuela R, Navarro G, et al. Angiotensin type 2 receptors: Role in aging and neuroinflammation in the substantia nigra. Brain, Behavior, and Immunity. 2020;87:256-271
  131. 131. Loy R, Milner TA. Sexual dimorphism in extent of axonal sprouting in rat hippocampus. Science. 1980;208(4449):1282-1284
  132. 132. Jhaveri DJ, Mackay EW, Hamlin AS, Marathe SV, Nandam LS, Vaidya VA, et al. Norepinephrine directly activates adult hippocampal precursors via β3-adrenergic receptors. Journal of Neuroscience. 2010;30(7):2795-2806
  133. 133. Masuda T, Nakagawa S, Boku S, Nishikawa H, Takamura N, Kato A, et al. Noradrenaline increases neural precursor cells derived from adult rat dentate gyrus through beta2 receptor. Progress in Neuro-Psychopharmacology and Biological Psychiatry. 2012;36(1):44-51
  134. 134. Kulkarni VA, Jha S, Vaidya VA. Depletion of norepinephrine decreases the proliferation, but does not influence the survival and differentiation, of granule cell progenitors in the adult rat hippocampus. European Journal of Neuroscience. 2002;16(10):2008-2012
  135. 135. Miranda-Contreras L, Ramırez-Martens L, Benı́tez-Diaz P, Pena-Contreras Z, Mendoza-Briceno R, Palacios-Prü E. Levels of amino acid neurotransmitters during mouse olfactory bulb neurogenesis and in histotypic olfactory bulb cultures. International Journal of Developmental Neuroscience. 2000;18(1):83-91
  136. 136. Bolteus AJ, Garganta C, Bordey A. Assays for measuring extracellular GABA levels and cell migration rate in acute slices. Brain Research Protocols. 2005;14(2):126-134
  137. 137. Castiglione M, Calafiore M, Costa L, Sortino MA, Nicoletti F, Copani A. Group I metabotropic glutamate receptors control proliferation, survival and differentiation of cultured neural progenitor cells isolated from the subventricular zone of adult mice. Neuropharmacology. 2008;55(4):560-567
  138. 138. Platel J-C, Dave KA, Gordon V, Lacar B, Rubio ME, Bordey A. NMDA receptors activated by subventricular zone astrocytic glutamate are critical for neuroblast survival prior to entering a synaptic network. Neuron. 2010;65(6):859-872
  139. 139. Erlander MG, Tillakaratne NJ, Feldblum S, Patel N, Tobin AJ. Two genes encode distinct glutamate decarboxylases. Neuron. 1991;7(1):91-100
  140. 140. Cryan JF, Slattery DA. GABAB receptors and depression: Current status. Advances in Pharmacology. 2010;58:427-451
  141. 141. Zhu L, Polley N, Mathews GC, Delpire E. NKCC1 and KCC2 prevent hyperexcitability in the mouse hippocampus. Epilepsy Research. 2008;79(2-3):201-212
  142. 142. Chancey JH, Adlaf EW, Sapp MC, Pugh PC, Wadiche JI, Overstreet-Wadiche LS. GABA depolarization is required for experience-dependent synapse unsilencing in adult-born neurons. Journal of Neuroscience. 2013;33(15):6614-6622
  143. 143. Liu X, Wang Q, Haydar TF, Bordey A. Nonsynaptic GABA signaling in postnatal subventricular zone controls proliferation of GFAP-expressing progenitors. Nature Neuroscience. 2005;8(9):1179-1187
  144. 144. Nguyen L, Malgrange B, Breuskin I, Bettendorff L, Moonen G, Belachew S, et al. Autocrine/paracrine activation of the GABAA receptor inhibits the proliferation of neurogenic polysialylated neural cell adhesion molecule-positive [PSA-NCAM+] precursor cells from postnatal striatum. Journal of Neuroscience. 2003;23(8):3278-3294
  145. 145. Bolteus AJ, Bordey A. GABA release and uptake regulate neuronal precursor migration in the postnatal subventricular zone. Journal of Neuroscience. 2004;24(35):7623-7631
  146. 146. Tozuka Y, Fukuda S, Namba T, Seki T, Hisatsune T. GABAergic excitation promotes neuronal differentiation in adult hippocampal progenitor cells. Neuron. 2005;47(6):803-815
  147. 147. Nakamichi N, Takarada T, Yoneda Y. Neurogenesis mediated by γ-aminobutyric acid and glutamate signaling. Journal of Pharmacological Sciences. 2009;110(2):133-149
  148. 148. Chen R, Wang K, Yu J, Howard D, French L, Chen Z, et al. The spatial and cell-type distribution of SARS-CoV-2 receptor ACE2 in the human and mouse brains. Frontiers in Neurology. 2021;11(1860):1-12
  149. 149. Ramani A, Müller L, Ostermann PN, Gabriel E, Abida-Islam P, Müller-Schiffmann A, et al. SARS-CoV-2 targets neurons of 3D human brain organoids. The EMBO Journal. 2020;39(20):e106230
  150. 150. Galic MA, Riazi K, Pittman QJ. Cytokines and brain excitability. Frontiers in Neuroendocrinology. 2012;33(1):116-125
  151. 151. Nikbakht F, Mohammadkhanizadeh A, Mohammadi E. How does the COVID-19 cause seizure and epilepsy in patients? The potential mechanisms. Multiple Sclerosis and Related Disorders. 2020;46:102535

Written By

Jayakumar Saikarthik, Ilango Saraswathi and Abdulrahman A. Al-Atram

Submitted: 11 September 2021 Reviewed: 11 October 2021 Published: 25 November 2021