Next Article in Journal
Genome-Wide Association Study Reveals Novel Powdery Mildew Resistance Loci in Bread Wheat
Next Article in Special Issue
Thymol as a Component of Chitosan Systems—Several New Applications in Medicine: A Comprehensive Review
Previous Article in Journal
Analysis of Auxin-Encoding Gene Family in Vigna radiata and It’s Cross-Species Expression Modulating Waterlogging Tolerance in Wild Vigna umbellata
Previous Article in Special Issue
Rowanberry—A Source of Bioactive Compounds and Their Biopharmaceutical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutics to Treat Psychiatric and Neurological Disorders: A Promising Perspective from Algerian Traditional Medicine

1
Laboratoire d’Obtention de Substances Thérapeutiques (LOST), Université Frères Mentouri-Constantine 1, Route de Ain El Bey, Constantine 25017, Algeria
2
Department of Drug Discovery and Biomedical Sciences, College of Pharmacy, University of South Carolina, Columbia, SC 29208, USA
*
Author to whom correspondence should be addressed.
Plants 2023, 12(22), 3860; https://doi.org/10.3390/plants12223860
Submission received: 29 September 2023 / Revised: 5 November 2023 / Accepted: 10 November 2023 / Published: 15 November 2023
(This article belongs to the Collection Feature Review Papers in Phytochemistry)

Abstract

:
Ancient people sought out drugs in nature to prevent, cure, and treat their diseases, including mental illnesses. Plants were their primary source for meeting their healthcare needs. In Algeria, folk medicine remains a fundamental part of the local intangible knowledge. This study aims to conduct a comprehensive ethnomedicinal investigation and documentation of medicinal plants and the different plant formulations traditionally used in Algeria for the treatment of pain, psychiatric, and neurological disorders. It also intends to improve the current knowledge of Algerian folk medicine. Several scientific databases were used to accomplish this work. Based on this investigation, we identified 82 plant species belonging to 69 genera and spanning 38 distinct botanical families used as remedies to treat various psychological and neurological conditions. Their traditional uses and methods of preparation, along with their phytochemical composition, main bioactive constituents, and toxicity were noted. Therefore, this review provides a new resource of information on Algerian medicinal plants used in the treatment and management of neurological and psychological diseases, which can be useful not only for the documentation and conservation of traditional knowledge, but also for conducting future phytochemical and pharmacological studies.

1. Introduction

Algeria is the largest country in Africa. It has a diverse range of landforms, including a Mediterranean coast that is separated from the Sahara Desert by a highlands plateau (Figure 1). Furthermore, Algeria experiences a diverse range of climates, spanning from a mild Mediterranean climate along the coast to highland winters marked by snowfall and desert regions characterized by scorching heat. This unique geographical positioning and climatic diversity have fostered the growth of a remarkably diverse flora encompassing an assortment of trees, shrubs, and herbs. Throughout ancient history, this rich and varied flora has consistently served as a primary source of raw materials for traditional medicinal practices [1,2,3,4,5].
Folk medicine comprises a reservoir of knowledge and practices utilized for the preservation of health and the prevention, diagnosis, enhancement, or treatment of both physical and mental ailments. This shared ethnopharmacological heritage has been transmitted across generations for millennia through both oral traditions and written records.
Indeed, Algerian ethnopharmacology predominantly represents a fusion of time-honored insights inherited from Islamic civilization and the pragmatic, empirical application of a diverse spectrum of substances, encompassing those originating from plants, animals, and inorganic sources [9]. For many centuries, the Islamic world was a center of scientific and medical enlightenment. This was due in part to the extensive network of trade routes that connected the East and West, including the famous Silk Road [10,11]. These routes facilitated the exchange of ideas, knowledge, and goods, allowing for the spread of scientific and medical advancements throughout the region (Figure 2). New disciplines emerged—algebra, trigonometry, and chemistry—as well as major advances in medicine, astronomy, engineering, and agriculture. Arabic texts replaced Greek as the fonts of wisdom [12]. Through their interactions with Asian traditions, including Chinese and Hindu influences, Arab pharmacists ingeniously merged medicinal plant knowledge, harmonizing their effects, and enhancing their palatability. Furthermore, indigenous medical wisdom, encompassing traditions, like traditional Chinese medicine, Ayurveda, and Tibetan practices, disseminated the essence of healing globally [10]. Islamic intellectual heritage has recognized mental health and played an active role in cultivating mental well-being, starting from the Prophet and inspiring Muslim scholars who followed. The Holy Quran is also considered a source of healing and comfort for mental and physical ailments [13,14]. Traditional Arabic and Islamic medicines are based on the concept of restoring balance to the body through diet, lifestyle, exercise, body cleansing, and the health of the mind, body, and spirit [15]. In Algeria, it is considered a form of alternative medicine that focuses on preventing and treating health conditions using methods and theories derived from the Holy Quran and Hadiths (Prophetic Medicine). These practices incorporate herbal medicines, spiritual therapies, dietary practices, mind–body practices, and manual techniques applied singularly or in combination to treat, diagnose, and prevent illnesses [15].
This review seeks to create an extensive catalog of medicinal plants available in Algeria, with a well-documented historical usage for treating, preventing, and managing mental and neuropsychiatric disorders. Additionally, it strives to provide a thorough insight into the potential therapeutic benefits linked to these plant species, serving as a foundation for guiding and motivating future phytochemical and pharmacological investigations.

2. Medicinal Plants Used for Mental Illnesses

Nervous system disorders include depressive disorders, anxiety disorders, bipolar disorder, schizophrenia, autism spectrum disorders, conduct disorder, attention-deficit hyperactivity disorder, eating disorders, idiopathic developmental intellectual disability, central nervous system conditions, like Alzheimer’s disease (AD) or Parkinson’s disease (PD), and a range of other mental disorders. Neurological disorders affect people of all ages, genders, education levels, and income levels in every country [17]. AD is the most common neurodegenerative disorder, according to the World Health Organization (WHO), and is one of the top 10 causes of death worldwide. Mental and neurological disorders affect approximately 907 million people worldwide [18]. In 2019, the global population of individuals affected by AD and other forms of dementia dramatically increased by a staggering 160.84% compared to the figures recorded in 1990, which were approximately 51.62 million. The number of global fatalities due to AD and other dementias rose from 0.56 million in 1990 to 1.62 million in 2019, nearly tripling over the span of three decades [19].
Traditional healing practices are indeed very important. The effectiveness of traditional treatment in psychotherapy has been demonstrated [20]. For example, the antipsychotic drug reserpine, an alkaloid isolated from the dried root of the Rauwolfia serpentina species, revolutionized the treatment of schizophrenia [21,22]. Research on natural psychoactive compounds has provided a wealth of information to the field of neuroscience [23,24]. Several medicinal plants have been shown to have beneficial effects on AD, PD, and depression [25,26]. Prescribed as a mild antidepressant, commercially available products of Hypericum perforatum are among the best-selling dietary supplements and are considered one of the most successful and effective herbal products in the world [27]. Another example is Janssen’s Alzheimer’s drug, Razadyne® ER (also called galantamine), an alkaloid that was originally isolated from the Galanthus species and later found in other Amaryllidaceae genera, including daffodil bulbs (Narcissus genus) [28].

3. Traditional Treatment of Mental Disorders in Algeria

Depression, anxiety (including panic attacks and panic disorder), schizophrenia, epilepsy, insomnia, migraines, headaches, and eating disorders are common mental illnesses in Algeria [29]. These illnesses are also impacted by the spread of drug addiction and alcoholism in Algerian society. In fact, the number of individuals struggling with a drug addiction in Algeria is estimated to exceed 300,000 [30]. In 2019, there were 983 deaths attributed to drug use disorders in Algeria, accounting for a rate of 2.28 per 100,000 of the population [29] Furthermore, neuropsychiatric disorders are estimated to contribute to 13.1% of the global burden of disease [29]. According to a prior study conducted in Algiers, the capital of Algeria, the prevalence of anxiety disorders was as high as 43%, with 13% of individuals exhibiting symptoms of post-traumatic stress disorder [30].
Mental and neuropsychiatric disorders result from the intricate interplay of biological, psychological, social, and cultural factors. In some of the most remote African regions, such as among the nomad Berbers who uphold ancient traditions, the process of healing relies on a mystical realm deeply ingrained within these communities [31]. In traditional Algerian society, possession by a djinn (spirit) is considered the most likely cause of disease. Therefore, healers use plants with magical properties to restore one’s health. Someone presenting with mental disorders or chronic pain is often believed to have been a victim of the evil eye (Ein) and/or magic (Sihr) [32]. These are popularly considered causes of both physical and mental illnesses, often motivated by envy and jealousy. The evil eye and sorcery are seen as sanctions against anyone who exceeds the limits set by society in terms of wealth, beauty, intelligence, and happiness [32]. In such cases, traditional healers play a crucial role in identifying and countering these spiritual afflictions, using their knowledge of herbal remedies and rituals to bring about healing and balance within the affected individuals [33].
The local population possesses valuable indigenous knowledge that has been passed down through generations [3,34,35,36]. Various plant parts, such as leaves, bark, fruit, buds, seeds, flowers, roots, and rhizomes, are employed for enhancing mental well-being and addressing various health issues [5,31]. Plant extracts are meticulously prepared through methods, like decoction, infusion, maceration, or tincture, and they can be consumed orally, applied topically, or administered through fumigation and vapor inhalation [31,34]. The historical use of numerous plant species has deep roots not only in Algeria but also in various other countries. Some, like Atropa belladonna (deadly nightshade) and Ferula communis L. (giant fennel), have associations with magic, while others, such as Commiphora myrrha (myrrh), hold significance in religious rituals [37]. Interestingly, species such as Lavandula spp. (lavender), Marrubium vulgare (white horehound), Mentha spp. (mint), Ocimum basilicum (basil), Origanum majorana (sweet marjoram), Rosmarinus officinalis (rosemary), Salvia officinalis (sage), Stachys arvensis (field woundwort), and Thymus spp. (thyme) were historically employed for protection against the evil eye [38] Decoctions made from Stachys recta, Melissa officinalis, or even Myrtus communis are traditionally employed for protection against negative influences or spirits [39] These herbal infusions have been historically and continue to be utilized for cleansing the face, hands, and wrists to alleviate various forms of pain, fear, and anxiety [39]. In Algeria, a rich diversity of 3139 plant species has been documented, of which 35% are from Mediterranean origins, with Euro-Mediterranean species accounting for 12% and North African species at 11%. Remarkably, 700 of these plant species are endemic [40]. The majority of plant species’ phytochemical and pharmacological properties remain undiscovered, and the realm of plant natural products remains largely unexplored. Only a handful of studies have explored the effects of Algerian medicinal plants on psychiatric and central nervous system disorders [41,42,43,44]. Figure 3 displays some common Algerian medicinal plants that have the potential to be used in the treatment of mental disorders.

4. Results

4.1. Medicinal Plant Diversity

The data for 82 plants traditionally utilized in the management of neurological and psychological disorders in Algeria were categorized based on their accepted Latin names, as per http://www.plantlist.org, and were grouped according to their respective botanical families (see Table 1). In the table, each plant entry includes information, such as the family name, common name(s), documented experimental evidence of activity (when available), the specific plant part(s) used, traditional methods of preparation, any identified active constituent(s) (if known), as well as the relevant data regarding interaction and toxicity studies.
In total, this study identified 82 plant species belonging to 69 genera and spanning 38 distinct botanical families (as outlined in Table 1).

4.2. Most Frequently Cited Plant Species

Many plants have been identified in ethnobotanical surveys, but only a limited number of pharmacological studies focusing on brain disorders in Algeria have been documented. This study uncovered a total of 82 medicinal species utilized for the treatment, management, and prevention of mental illnesses. These species are distributed among 69 genera and 38 families (refer to Table 1). The most frequently mentioned families included Lamiaceae, accounting for 17 species (21%), Asteraceae with 8 species (10%), Fabaceae with 5 species (6%), and both Apiaceae and Rosaceae, each comprising 4 species (5%). The remaining 33 families (53%) were represented by one to two species each (Figure 4).
Mentha, a genus within the Lamiaceae family, occupies a prominent position in the realm of botanical significance. With a substantial representation of 29% among genera of the Lamiaceae family (Table 1), it stands out as a noteworthy and extensively studied genus. The distinguishing characteristic of Mentha species lies in their aromatic nature, which is pervasive across various regions worldwide. It encompasses five distinct species: Mentha arvensis L., Mentha aquatica L., Mentha rotundifolia (L.) Huds., Mentha pulegium L., and Mentha x piperita L. Notably, peppermint (Mentha x piperita L.) is the most widely utilized among these species and stands as one of the most crucial commercially grown aromatic herbs globally. Peppermint essential oil possesses a distinctive sharp, cool, and invigorating aroma, and it carries significant pharmacological importance, primarily owing to its primary active component, menthol. Menthol exhibits a broad spectrum of biological activities and therapeutic potentials [115].
Peppermint is traditionally used in Algeria as a remedy for several conditions, including the common cold and headaches. It is also used as a sedative [34]. Its essential oil is commonly applied topically for relief from muscle pain and headaches. Many studies have shown that peppermint has a considerable effect on certain mental disorders, such as anxiety, stress, and insomnia [116,117]. It has been reported that the administration of a peppermint infusion can improve sleep quality and enhance memory [116].
Our results are consistent with previous studies that reports that the most commonly used plants in several Algerian regions belong mainly to the Lamiaceae, Asteraceae, and Apiaceae families [34,36,72]. A previous study conducted by A. Bouasla et al. revealed that Mentha piperita (Lamiaceae), Thymus vulgaris (Lamiaceae), Zingiber officinale (Zingiberaceae), Myrtus communis (Myrtaceae), Nigella sativa (Rutaceae), Chamaemelum nobile (Asteraceae), and Olea europaea (Oleaceae) were the species most frequently used to treat nervous system disorders in the region of Skikda in Northeastern Algeria [34].

4.3. Used Parts

It is evident that leaves are the most commonly utilized plant parts among local populations, not only in Algeria but also in numerous regions globally [36,118,119]. Our findings indicate that the most commonly employed plant parts are leaves, followed by the aerial parts, flowers, seeds, roots, and rhizomes. Additionally, other plant components, such as fruits, stems, barks, resin, capitulum, buds, and even the entire plant, were reported to be used, albeit at a lower frequency.

4.4. Method of Preparation

Our results unveil that the most prevalent methods for preparing plant remedies among local populations are decoction and infusion in water, consistent with the findings in other studies [34,36].

4.5. Pharmacognostic Investigations on the CNS System

The effects of medicinal plant preparations, their extracts, and isolated secondary metabolites on the central nervous system (CNS) have been extensively studied [120,121,122]. Numerous active principles from plants, such as polyphenols, polysaccharides, terpenoids, alkaloids, coumarins, tannins, and other natural bioactive ingredients, are used in the management of psychiatric and mental disorders [123,124,125]. For example, polyphenols, major groups of natural compounds that act as primary antioxidants [126,127], can protect DNA from oxidative damage, inhibit tumor cell growth, and possess anti-inflammatory and neuroprotective properties [128,129,130,131,132,133]. Flavonoids, including chalcones, flavones, flavonols, flavanones, isoflavonoids, anthocyanins, and proanthocyanidins, are widely distributed in the plant kingdom. These compounds exhibit various pharmacological effects, such as antioxidant, anti-inflammatory, anxiolytic, sedative, anticonvulsant, and analgesic properties, through their actions on the CNS [134]. Recent studies have shown that the regular consumption of flavonoid-rich foods can effectively enhance cognitive functions in humans [135,136,137]. Moreover, several flavonoids have been shown to prevent the progression of AD pathologies by reversing the cognitive deficits in various normal and transgenic animal models [135,137,138]. Essential oils are another example of active compounds derived from aromatic plants. Their primary constituents are vital agents well-known for their antimicrobial and antifungal properties, alongside their recognized roles in antiviral, antioxidant, anti-inflammatory, antidiabetic, and anticancer activities [139]. Several studies have demonstrated the relationship between the chemical composition and biological properties of essential oils and their applications in various commercial and pharmacological preparations to prevent, treat, and manage neurological disorders [140,141]. Moreover, numerous research works have proven that the inhalation of certain essential oils improves mood, enhances memory, and reduces stress [142,143,144].

4.6. In Vitro and In Vivo Pharmacological Evidence

Numerous plant crude extracts, fractions, and isolated secondary metabolites have undergone binding studies targeting CNS receptors, including dopamine, serotonin, cholinergic, GABA, opioids, and cannabinoid receptors [145]. Subsequently, a series of in vivo studies were conducted using appropriate animal models, primarily rodents, to obtain essential pharmacological and toxicological data before progressing to human clinical trials [146,147] These animal studies helped establish the basic safety and efficacy profiles of the substances under investigation, paving the way for further research on human subjects.
Roman chamomile (Chamaemelum nobile, Asteraceae) finds extensive application in traditional medicine globally, with a notable presence in Algeria, particularly in the Northeastern region. Chamaemelum nobile is renowned for its tranquilizing properties, serving as a mild sedative to alleviate nervousness, anxiety, and a range of disorders. It has been historically employed to address conditions, such as hysteria, insomnia, and various ailments, earning its place as one of the oldest known medicinal herbs [34]. Flavonoids, including apigenin, luteolin, quercetin, and patuletin, constitute significant components in Chamaemelum nobile (Figure 5). Interestingly, these flavonoids are not exclusive to Roman chamomile; they are also present in other plants, like Passiflora incarnata (passionflower) and Matricaria chamomilla (chamomile) [31,62]. Apigenin has garnered attention due to its reported binding affinity to gamma-aminobutyric acid (GABA)-benzodiazepine receptors in the brain. This interaction with GABA receptors may be responsible for the sedative effects associated with apigenin [148]. Such insights into the specific bioactive compounds in Chamaemelum nobile contribute to our understanding of its potential pharmacological properties and its historical use as a natural remedy for calming and soothing effects [149].
Artemisia absinthium is a perennial herbaceous plant from the Asteraceae family, native to North Africa, including Algeria. Traditionally, it has been valued for its diverse therapeutic properties and widely used in conventional medicine across the world. It is employed for treating various conditions, such as digestive problems, morning sickness, irregular menstrual cycles, typhoid, epilepsy, renal issues, bronchitis, and malaria, among others [150]. Additionally, in Algerian traditional medicine, Artemisia absinthium is known to have calming properties that can potentially help with anxiety, restlessness, and insomnia [31,34,57]. In addition to flavonoids and phenolic acids, the plant also contains other compounds, one of which is thujone. Thujone, a volatile monoterpene ketone, has been shown to interact with GABA receptors in the brain [151]. This interaction suggests that thujone may exhibit GABAergic activity, which potentially underlies its sedative and anxiolytic effects [152].
Valeriana officinalis (Caprifoliaceae) is one of the most widely recognized herbal sedatives worldwide. It has long been used in alternative medicine for the treatment of insomnia and other related neurological disorders [153]. Valeriana officinalis is also cultivated in Algeria and traditionally used for treating neurological disorders [154]. The hydroalcoholic extract of Valeriana officinalis interacts with ionotropic glutamate receptors, believed to be one of the reasons for its anxiolytic effect [155]. Valerianic acid, a common standard marker for Valeriana officinalis, has also been observed to interact with metabolic glutamine receptors, which partly explains its anxiolytic effects [156]. The secondary metabolites of the Valeriana genus include, among others, sterols, alkaloids, and terpenes. Among the terpenes, valepotriates and sesquiterpenoids are noteworthy [157]. However, it is likely that the effects result from a summation and synergy of all essential oils, along with the terpenes found in Valeriana officinalis [158,159]. Valeriana tuberosa L., a Mediterranean species naturally found in Algeria, is traditionally considered as a complement in the prevention and treatment of epilepsy, as well as for its antispasmodic, calming, and hypotensive properties [46].
Hypericum perforatum (St. John’s wort) from the family of Hypericaceae stands out as one of the most extensively studied due to its recognized anti-depressant effects. In a study presented by Simmen et al. [160], Hypericum perforatum extracts, fractions, and constituents were subjected to radioligand binding assays on opioid, serotonin, estrogen, histamine, neurokinin, and metabotropic glutamate and GABAA receptors. Hypericin exhibited the most pronounced binding inhibition with the human CRF1 receptor. Hyperforin hindered binding with opioid and serotonin (5-HT) receptors, while hypericin and pseudohypericin displayed comparatively weaker inhibitory effects. The biflavonoid I3,II8-biapigenin effectively restrained 3H-estradiol binding to the estrogen-alpha receptor. It is worth noting that the inhibition of 3H-muscimol binding to the GABAA receptor is likely attributed solely to the presence of GABA in the extract. The study suggests that the beneficial antidepressant effect of St. John’s wort may result from the additive or synergistic actions of several distinct compounds [160].
In a previous investigation, Simmen et al. explored the impact of Hypericum perforatum extracts, fractions, and individual components on the binding of various ligands to recombinant CNS receptors expressed through the Semliki Forest virus expression system [161]. Among these, a lipophilic fraction of Hypericum exhibited notably strong inhibitions of mu-, delta-, and kappa-opioid receptors, as well as 5-HT6 and 5-HT7 receptors. Furthermore, constituents of Hypericum, including naphthodianthrones, like hypericin and pseudohypericin, and the phloroglucinol hyperforin, displayed inhibitory effects on both opioid and serotonin receptors in the lower micromolar range. These findings provide support for the notion that multiple active components in Hypericum may work synergistically to contribute to its antidepressant effects in the central nervous system [161].
In another study on pure compounds isolated from Hypericum perforatum extracts, the in vitro pharmacologies of hypericin, pseudohypericin, and hyperforin, and several flavonoids were characterized at 42 biogenic amine receptors and transporters [162]. Amentoflavone demonstrated significant inhibitions of serotonin 5-HT1D, 5-HT2C, dopamine D3, and delta-opioid receptors, while hypericin displayed potent activities on dopamine D3 and D4, as well as adrenergic receptors. Notably, only the dopamine D1 receptor interacted with hyperforin [162]. The traditional use of the Hypericum species is not limited to depression, but is also used to treat wounds and burns, diarrhea, pain, and fever [163]. In fact, in addition to its traditional use in Algeria for treating depression and migraines, Hypericum perforatum is also reported to be used as an ointment, balsam, tincture to treat inflammation, antimicrobial agent, and wound healing agent [164]. Other Hypericum species found in Algeria have also been studied. For example, do Rego et al. found hyperfoliatin to be the metabolite responsible for the antidepressant effect of Algerian Hypericum perfoliatum, with the mechanism of action being associated with monoamine uptake inhibition [42]. From Hypericum afrum, another commonly used medicinal plant in Algeria, a series of flavonols, including quercetin and myricetin, were found to have neuroprotection properties and were identified as potent MAOs inhibitors [43]. Figure 6 shows a series of Hypericum metabolites with CNS activity.
Two intriguing herbaceous and aromatic genera commonly encountered in Algeria are Salvia spp and Melissa spp. These two genera belong to the Lamiaceae family and are primarily distributed in subtropical and tropical regions across the globe [165]. These two genera are renowned for their utilization as spices and their ethnopharmacological medicinal applications, with Salvia officinalis (Sage) and Melissa officinalis (Lemon Balm) being two prominent examples [166,167].
Salvia species are well-known for their abundance of essential oils and phenolic compounds, which encompass flavonoids and terpenes, particularly diterpenes. These compounds exhibit intriguing biological activities, notably an outstanding antibacterial and antioxidant potential [168,169,170]. Algeria hosts approximately 23 species of Salvia, with five of them being endemic [171]. In Algeria, Salvia officinalis is extensively employed in the northern regions for its traditional use in treating stress and insomnia, acting as a tranquilizer, and serving as a nervous tonic [34,72,172]. Recently, the neuroprotective effect of Algerian Salvia officinalis was investigated [173]. An aqueous extract of Salvia officinalis was administrated on rats exposed to an aluminum-induced neurodegeneration model. The study concluded that the extract of Salvia officinalis effectively reduced oxidative stress and improved biochemical parameters in the animals [173]. Other attractive Algerian Salvia species include Salvia microphylla, traditionally employed in Algeria for addressing memory loss and rheumatism [174].
On the other hand, Melissa officinalis, a Mediterranean medicinal plant, has been utilized for centuries to address CNS disorders, encompassing conditions such as depression, dementia, and various other ailments [175]. Melissa officinalis has been a valuable component of Algerian folk medicine, employed to address various neurological conditions, such as headaches, insomnia, migraines, nervousness, and depression [34,36]. Previous studies have shown that extracts of both Melissa officinalis and Salvia elegans exhibited a significant binding affinity to nicotinic and muscarinic acetylcholine receptors [176]. Melissa officinalis has also demonstrated potent neuroprotective effects [177].
The chemical composition of Melissa officinalis includes flavonoids, phenolic compounds, and terpenoids [166]. Notably, essential oils and phenolic acids, such as caffeic, ferulic, chlorogenic, rosmarinic, and p-coumaric acids, are believed to be the primary bioactive constituents, acting synergistically to impart its medicinal properties (Figure 7).
Thymus vulgaris (common thyme) and Rosmarinus officinalis (rosemary) are two aromatic herbs belonging to the Lamiaceae family, commonly found in Algeria [34,72]. They are widely cultivated and can also be found growing wild in various regions of the country. Both Thymus vulgaris and Rosmarinus officinalis have a rich history of traditional use for culinary and medicinal purposes [31,34].
Thymus vulgaris contains several bioactive compounds, notably thymol and carvacrol (see Figure 8) [31]. These two compounds are known for their potent antioxidant and anti-inflammatory properties. These properties have been linked to potential neuroprotective effects by reducing oxidative stress and inflammation, both of which are implicated in neurodegenerative diseases [178,179].
Thymol, at doses of 15 and 30 mg/kg, has been shown to increase the levels of central neurotransmitters and inhibit the expression of pro-inflammatory cytokines in a chronic unpredictable mild stress (CUMS) mouse model [180]. In a study conducted by Azizi et al. [181], thymol at doses of 0.5–2 mg/kg was shown to inhibit cognitive impairments caused by increased Aβ levels or cholinergic hypofunction in Aβ (25–35) or scopolamine-treated rats. This effect was attributed to thymol’s antioxidant, anti-inflammatory, and anticholinesterase properties. Lee et al. [182] demonstrated that thymol at concentrations of 0.39–25 mg/mL may inhibit H2O2-induced oxidative stress in PC-12 cells. Thymol at concentrations of 100 and 1000 mg/mL also inhibited both acetylcholinesterase (AChE) and butyrylcholinesterase (BChE) in a dose-dependent manner. Moreover, it has been reported that thymol at concentrations of 10–100 ppm, in combination with gamma-terpinene or para-cymene, attenuated cholinergic dysfunction, which is manifested in a plethora of neurodegenerative and psychiatric disorders, such as Alzheimer’s, Parkinson’s, and Huntington’s diseases, by enhancing the synaptic levels of acetylcholine (ACh) and the responsiveness of the nicotinic acetylcholine receptor (nAChR) in a Caenorhabditis elegans model [183]. According to the data presented by Zotti et al. [184], carvacrol is a brain-active molecule that influences neuronal activity through the modulation of neurotransmitters. Carvacrol was screened for pharmacological effects on the central nervous system and was found to present anxiolytic activity when administered orally to mice. Antidepressant effects were also observed in mice after carvacrol administration at doses of 12.5 to 50 mg/kg [184]. According to the researchers, the observed result was most likely due to an increase in dopamine levels. In a study conducted by Peters et al. [185], carvacrol was administered to mice after traumatic brain injury (TBI) and its effect on their functional recovery was followed for several weeks. The results showed that neurological recovery after TBI was significantly enhanced by the application of carvacrol. The authors found that neurological recovery after TBI was significantly enhanced by combining carvacrol with TRPC1 elimination [185].
Figure 8. Major components of Thymus numidicus Poir. and Thymus fontanesii Boiss. & Reut. essential oils from Algeria [186,187].
Figure 8. Major components of Thymus numidicus Poir. and Thymus fontanesii Boiss. & Reut. essential oils from Algeria [186,187].
Plants 12 03860 g008
Beyond its culinary appeal, rosemary has been treasured for its potential health benefits. It contains several bioactive compounds, including rosmarinic and carnosic acids, which possess antioxidant and anti-inflammatory properties [188]. These compounds have attracted interest in the field of natural medicine, and some studies have explored their potential neuroprotective effects on neurodegenerative diseases [188]. Rosemary’s aromatic properties have made it a popular choice for use in aromatherapy. Its essential oil is often used for relaxation, stress relief, and mental clarity [189]. In Algerian traditional medicine, rosemary has been used to support digestion, alleviate headaches, and promote overall well-being [36].
Rosmarinic acid, a polyphenolic compound found in high concentrations in rosemary, has been shown to have anti-epileptic activity by increasing the latency and decreasing the percentage of seizure incidents, reducing the levels of free radicals and DNA damage in a kindling CF-1 male mouse model of epilepsy induced by PTZ (rosmarinic acid at 1, 2, or 4 mg/kg b.w., i.p.) [190]. The administration of rosmarinic acid for 7 days at 5 and 10 mg/kg b.w./day led to the downregulation of mitogen-activated protein kinase phosphatase-1, the upregulation of BDNF, and the modulation of dopamine and corticosterone synthesis in a model of depression in mice with bupropion as a positive control [191]. In addition, rosmarinic acid has been shown to exhibit anti-tauopathy activity [192].
Lavandula, commonly known as lavender, is a fragrant flowering plant that belongs to the Lamiaceae family. It is native to countries bordering the Mediterranean Sea and is widely cultivated. Lavandula angustifolia (Lavandula officinalis), also known as true lavender, is the predominant species widely cultivated for its essential oil [193]. The plant’s essential oil finds extensive application in various industries, including perfumery, cosmetics, and aromatherapy, among others [193,194]. In Algeria, two Lavandula species, Lavandula angustifolia and Lavandula stoechas, hold a prominent position in traditional medicine [34,36]. Both of these plants are harnessed for their antispasmodic and wound disinfectant properties, making them valuable for addressing skin-related issues [72]. Additionally, these plants have been documented for their sedative and antidepressant attributes [72].
The local anesthetic effect of lavender oil and its major constituents, linalool and linalyl acetate (see Figure 9), has been reported in both in vivo and in vitro studies. An investigation of the effects of inhaled linalool on anxiety, aggressiveness, and social interaction in mice showed anxiolytic properties in the light/dark test, increased social interaction, and decreased aggressive behavior [195]. The anxiolytic effects of linalool odor were examined using the light/dark box test and the elevated plus maze (EPM). It was found that linalool odor has an anxiolytic effect without motor impairment in mice. Moreover, the effect was antagonized by flumazenil, indicating that the linalool odor-induced anxiolytic effect was mediated by γ-aminobutyric acid GABAergic transmission via benzodiazepine (BDZ)-responsive GABAA receptors [196]. These results provide information about the potential central neuronal mechanisms underlying the odor-induced anxiolytic effects and lay the foundation for exploring the clinical application of linalool odor in anxiety treatments [196].
Species of the Stachys genus, a member of the Lamiaceae family, are recognized for their medicinal properties and are particularly valued for their distinctive aroma and flavor [93]. In Algeria, this genus is represented by 14 species, including four that are endemic to the region [171]. The main secondary metabolites identified from species in this genus encompass flavonoids, iridoids, fatty acids, phenolic acids, and diterpenoids [93]. Extracts and essential oils derived from Stachys species have exhibited promising anti-Alzheimer’s disease effects, primarily attributed to their ability to inhibit cholinesterase enzymes [198,199]. Figure 10 illustrates the key components of the essential oil extracted from Stachys circinnata l’Her. sourced from Algeria.
The effects of a hydroalcoholic extract and essential oil of Stachys lavandulifolia Vahl, a species used as an anxiolytic and sedative in Iranian folk medicine, were investigated using the elevated plus-maze (EPM) model of anxiety [200]. The Stachys lavandulifolia extract or its essential oil was administered intraperitoneally to male TO mice at various doses, 30 min before the behavioral evaluation. The results suggest that the extract of Stachys lavandulifolia possesses anxiolytic effects with relatively lower sedative activity compared to diazepam [200].
Figure 10. Main components of the essential oil of Stachys circinnata l’Her. From Algeria [201].
Figure 10. Main components of the essential oil of Stachys circinnata l’Her. From Algeria [201].
Plants 12 03860 g010
Peganum harmala L. (Nitrariaceae), commonly known as Harmal, is a perennial herbaceous plant thriving in arid and semiarid regions across various parts of the world, including North Africa, the Eastern Mediterranean, the Middle East, India, and other regions [202,203]. This plant has been used in traditional medicine for decades. Various parts of Peganum Harmala, including its seeds, bark, and root, are highly regarded for treating a wide range of human ailments [202,203]. Peganum harmala exhibits a diverse array of medicinal properties encompassing cardiovascular, antidiabetic, antimicrobial, insecticidal, antineoplastic, antiproliferative, gastrointestinal, and neurologic effects [202,203].
Algerian folk medicine, in particular, has harnessed the potential of Peganum harmala to address nervous system disorders and manage various psychiatric conditions, including nervosity, anxiety, and severe pain relief [34,72,202]. Peganum harmala holds a significant place in the fumigation rituals practiced across various cultures, including Algeria, which is believed to ward off the evil eye and bring good luck [202]. These rituals typically involve the burning of its seeds [72]. This tradition extends beyond Algeria and is observed in several other regions as well [204]. The smoke produced from the burning of these seeds is believed to have disinfectant properties [205]. Its composition includes alkaloids, essential oils, fatty acids, flavonoids, and anthraquinones [206]. Among these constituents, beta-carboline alkaloids, like harmalol, harmaline, and harmine, stand out as the most significant (Figure 11). These compounds can stimulate the central nervous system either by inhibiting amine neurotransmitter metabolism or through a direct interaction with specific receptors [207,208].
It has been proven that Peganum harmala-derived beta-carboline alkaloids interact with a plethora of targets in the CNS, including GABA (gamma-aminobutyric acid), NMDA (N-methyl-D-aspartate), glutamate, dopaminergic, serotonergic, and imidazoline receptors, as well as several enzymes, such as AChE, monoamine oxidase, and DyRK1A (dual-specificity tyrosine-phosphorylation-regulated kinase 1A), along with others, inducing many pharmacological effects [156,209]. Due to the fact that beta-carbolines are strong inhibitors of monoamine oxidase, they are potential agents to treat some neurological and psychological conditions [207]. As examples, Harmane, a beta carboline alkaloid isolated from Peganum harmala, induced amnesia through a mechanism of interaction with the dopaminergic system [210]. In a recent study, the extract of Peganum harmala was found to be able to enhance glucagon-like peptide 1 (GLP-1) and help to restore insulin signaling avoiding the progression of the AD [211].
Passiflora genus (Passifloraceae) includes perennial plants found largely endemically on the America’s (North, Central, and South Americas) [212]. They are mostly distributed in the warm, temperate, and tropical regions of the world. The Passiflora genus is another example of using plants for the treatment of insomnia, anxiety, and various other disorders of the CNS since antiquity; in fact, several Passiflora species have been documented for their use in the North/Central America Mayan empire [213]. Passiflora genus have been propagated and can been found in the tropical and subtropical areas in the world, and indeed in Algeria these species are usually planted in gardens and pergolas, but some are reported to be used in Algerian traditional medicine [72]. For example, Passiflora edulis, commonly known as passion fruit, has been used in Algerian folk medicine in the region of Skikda in northeast Algeria, is use being reported to treat hypertension and asthma, as well as a tranquilizer [214]. Passiflora edulis is a well-known crop cultivated around the world and highlighted for its nutritional and health benefits [215]. Another notable example is Passiflora incarnata, a medicinal plant used in northeast Algeria due to its effects on anxiety as a tranquilizer [72]. The potential effects of this species for treating some diseases, including opiate withdrawal, insomnia, attention-deficit disorder, and epilepsy, have also been widely reported [216,217,218,219,220]. Those two plants, and in general all Passiflora species, are a good source of flavonoids, particularly rich in C-glycoside flavones, such as vitexin, isovitexin, orientin, and isoorientin; flavones, such as apigenin, chrysin, and luteolin; and flavonols, including kaempferol and quercetin [221]. This species of β-carbolines alkaloids also includes harman, harmol, harmine, harmalol, and harmaline [222,223]. The overall effect of these plants has been ruled as being synergistic, however some of the compounds mentioned above have been studied, particularly to treat CNS disorders [223,224]. Figure 11 shows a series of bioactive compounds isolated from the Passiflora species.
With the traditional knowledge of Passiflora incarnata’s historical use in treating cannabis or cannabis product addiction, Dhawan et al. [225] conducted a study to explore the effect of the benzoflavone moiety from the Passiflora incarnata Linneaus plant on combatting cannabis addiction. When benzoflavone was administered at dosages of 10 and 20 mg/kg twice daily in combination with cannabinoids, specifically delta-9-tetrahydrocannabinol (Δ9-THC) at 10 mg/kg twice daily, a noteworthy outcome was observed. This combination resulted in a significant reduction in the tolerance to and dependence on cannabinoids. The benzoflavone component extracted from Passiflora incarnata, when administered simultaneously with Δ9-THC, effectively thwarted the development of a dependence on cannabinoids in the studied mice. Notably, even a single acute administration of the benzoflavone compound at a dose of 20 mg/kg via the oral route significantly inhibited the manifestation of withdrawal effects on Δ9-THC-dependent mice [225].
In diverse animal models designed to assess the anti-anxiety effects of Passiflora incarnata, the extracts consistently demonstrated pronounced anti-anxiety effects at the test doses employed [226]. This activity can be attributed to the presence of various secondary metabolites in the plant, with chrysin being a key compound responsible for its anxiolytic effects. Chrysin exerts its influence by modulating γ-aminobutyric acid GABAA receptors. [226]. Both the extract and chrysin have been shown to display substantial anxiolytic activity in rodent models, as demonstrated through the elevated plus-maze (EPM) model of anxiety. The hypothesis suggests that chrysin mitigates anxiety by engaging with the GABAA receptor in laboratory rats, as substantiated by assessments involving the elevated plus-maze, corticosterone levels, and catecholamine assays [226].
Curcuma longa L., a plant belonging to the Zingiberaceae family, is commonly known as turmeric and is a popular spice in Algerian cuisine [227]. The local knowledge of turmeric in Algeria is primarily linked to its traditional use in Ayurvedic herbal medicine [227]. Its vibrant yellow color and unique flavor make it a favorite in many dishes. Beyond its culinary uses, turmeric has a long history of medicinal applications, dating back thousands of years, especially in traditional Indian and Chinese medicines [228]. The main active compound in turmeric is curcumin (Figure 12), which has garnered significant attention for its potential health benefits, including anti-inflammatory, antioxidant, and even anticancer effects [228].
Curcumin has been demonstrated to exert neuroprotective effects on a wide range of neurological disorders, including stroke, spinal cord injury, traumatic brain injury, AD, PD, multiple sclerosis, and epilepsy [229,230]. These neuroprotective properties of curcumin stem from its multifaceted attributes, including its role as an antioxidant, anti-inflammatory agent, anti-amyloidogenic compound, antidepressant, antidiabetic substance, and antiaging factor [231].
One of the most promising applications of curcumin in neurodegenerative disease therapy is its anti-amyloid property [232]. This property positions it as a highly promising compound for the treatment of various brain diseases characterized by amyloid accumulation. A recent review by Tang and Taghiglou [233] comprehensively summarized the latest developments in curcumin research related to AD. The review delved into the various mechanisms of action of curcumin in AD, encompassing its role in inhibiting Aβ and tau proteins, its capacity to bind copper, lower cholesterol levels, serve as an anti-inflammatory agent, modulate microglia, inhibit AChE, and function as an antioxidant. Furthermore, the review addressed topics concerning the bioavailability of curcumin and the current challenges encountered in curcumin therapy for AD patients [233].
Traditional medicine often incorporates at least one component or product derived from trees, such as fruits, flowers, bark, roots, seeds, resin, or oil [234,235].
Populus nigra L., a member of the Salicaceae family, is a cottonwood poplar species indigenous to North Africa, southern and central Europe, southwest and central Asia, and various other regions [236]. This tree thrives in numerous areas across Algeria, with its primary distribution occurring in the northern mountainous regions [237].
The therapeutic attributes of Populus nigra have been recognized since ancient times. This species is notable for its composition of phenolic glucosides, including salicin, poplin, and salicoside, as well as fatty acids, flavonoids, like flavones and flavanones, tannins, salicylates, and an essential oil containing sesquiterpenols, terpenes, and alkanes [5]. Notably, the buds of this tree contain the glucoside populoside benzoylosalicoside and are enriched with salicates, phenolic acids, and flavonoids [238].
Populus nigra buds are noted for their antiseptic properties and can also serve as a remedy for gout and hyperuricemia [238]. Furthermore, they have shown potential as a protective agent against neurodegenerative disorders [41]. Debbache-Benaida et al. conducted a study to investigate the effects of Populus nigra flower bud extract on brain modifications in mice subjected to aluminum-induced neurotoxicity. The study involved a battery of tests designed to assess various aspects of learning and memory functions, with the aim of evaluating the impact of the plant extract on AlCl3-induced neurotoxicity in mice [41]. Furthermore, the study examined the hypouricemic properties of the extract by measuring uric acid levels and liver xanthine oxidoreductase activity in both normal and hyperuricemic mice. The results demonstrated that the co-administration of Populus nigra extract (at a dose of 200 mg/kg) with AlCl3 (at a dose of 100 mg/kg/day), along with D-galactose (at a dose of 200 mg/kg/day) over a four-week period effectively counteracted the detrimental effects of AlCl3 by restoring all tested parameters. Populus nigra flower bud extract was proposed as a potential dietary supplement capable of mitigating the toxicity caused by AlCl3 and regulating serum uric acid levels [41,109].
Fraxinus (Oleaceae) is a genus of trees commonly referred to as ash trees, distributed across diverse regions of the world, with a presence in Algeria [5]. Algeria is home to the species Fraxinus angustifolia, which has a rich history of use in traditional medicine. Fraxinus angustifolia Vahl. has been used in traditional medicine for the relief of articular pain; prevention of gout; anti-inflammatory, antioxidant, diuretic, and digestive effects; and treatment of minor urinary infections [239]. It has also been suggested to have neuroprotective properties [240]. A range of bioactive phytochemicals, including secoiridoids, phenylethanoids, lignans, flavonoids, and coumarins, have been isolated from the Fraxinus plant [241]. Secondary metabolites, such as esculetin, esculin, fraxin, and fraxetin (Figure 13), and extracts from this plant have been found to present a range of biological activities, such as antioxidant, anti-inflammatory, antimicrobial, anticancer, hepatoprotective, antiallergic, skin regenerating, and diuretic [242,243]. Esculetin (coumarin) has been widely used in Chinese herbal medicine due to its pharmacological properties, such as antioxidant, anticancer, antibacterial, and anti-inflammatory [243].
In a recent study by Azib et al., the neuroprotective potential of Fraxinus angustifolia Vahl. bark extract was investigated both in vitro and in vivo. The study aimed to assess its efficacy in mitigating Aβ-aggregation and alleviating aluminum-induced neurotoxicity in mice [244]. The neuroprotective properties of the extract were assessed in various experimental setups. Initially, its efficacy against Aβ25-35 aggregation was determined by direct incubation with Aβ25-35, and the kinetics of aggregation were monitored using a spectrophotometer at 200 nm. Subsequently, the extract was subjected to an evaluation against Aβ25-35-induced cytotoxicity in PC12 cells, with the cell viability assessed through an MTT test [244]. Furthermore, Fraxinus angustifolia Vahl. Bark extract was tested across a range of concentrations (0.01–0.5 mg/mL) to gauge its impact on aluminum-activated lipid peroxidation in the synaptosomal membranes of mice. In vivo experiments were conducted on N.M.R.I. male mice to investigate its potential in countering aluminum-induced neurotoxicity. The results demonstrated that the extract exhibited a pronounced anti-aggregative effect against Aβ25-35 and displayed a dose-dependent protective effect on PC12 cells. Additionally, the extract showcased a significant inhibition of lipid peroxidation and effectively mitigated the detrimental effects of aluminum. This was evident through notable enhancements in locomotor activity, reduced anxiety, improved memory, and a decrease in histological alterations [244].
Juniperus (Cupressaceae) is a genus encompassing both trees and shrubs, widely recognized as junipers, with a longstanding history of traditional medicinal usage in Algeria. In Algerian flora, there are five notable species of junipers [1] Among these, two prominent species stand out: Juniperus oxycedrus L. and Juniperus phoenicea L.
Juniperus oxycedrus L., commonly referred to as Cade, is a species primarily distributed throughout the Mediterranean region [78]. In Algerian traditional medicine, it is revered for its diuretic, stimulative, and stomachic tonic properties [57]. Furthermore, it is highly valued for its effectiveness as a pulmonary and depurative disinfectant, in addition to its various other applications [5]. Juniperus phoenicea L. is another species indigenous to the Mediterranean region, including Algeria. In a study conducted to assess traditional knowledge and the utilization of medicinal plants by Algerian traditional healers, this species exhibited an exceptional fidelity value of 100% [36].
Juniperus species have garnered significant attention due to the diverse bioactivities associated with the extracts and secondary metabolites derived from these plants [245]. These bioactivities encompass a broad range of beneficial effects, including antimicrobial, antioxidant, antidiabetic, anti-inflammatory, anticonvulsant, analgesic, and cytotoxic activities [245]. Furthermore, the research findings highlight their potential as neuroprotective agents, suggesting promising attributes in supporting and safeguarding the health of the nervous system [66,246].
A study conducted by Gao et al. [247] revealed the vascular protective effects of totarol (Figure 14), a tricyclic phenolic diterpene abundantly present in various Juniperus species [248]. The study demonstrated that totarol activated the protein kinase B/heme oxygenase-1 (PKB/HO-1) pathway, leading to increased levels of superoxide dismutase (SOD) and the antioxidant glutathione (GSH). This, in turn, suppressed ischemia-induced brain injury. To simulate the conditions of patients with acute strokes, rats were administered totarol post-ischemia at doses of 1 and 10 µg/kg. The results revealed a significant reduction in the infarct volume compared to the untreated group. Furthermore, totarol treatment at both 1 and 10 µg/kg significantly enhanced the ischemia-induced response [247].

4.7. Clinical Trials and Therapeutic Applications

Clinical trials are research studies involving human participants, focusing on evaluating the efficacy and safety of new treatments, such as novel drugs or diets.
Medicinal plants continue to be a significant source of new drugs, drug leads, and chemical entities [249]. The research on the drug discovered in plants has led to the development of numerous bioactive agents, including those with antidiabetic, anticancer, anti-infectious, anti-withdrawal, antianxiety, and antidepression properties, among others [250]. However, despite the burgeoning body of in vitro and in vivo investigations into medicinal plant extracts and their active constituents, a dearth of comprehensive assessments pertaining to the safety and efficacy of herbal medicines persists, largely attributable to an array of challenges confronting researchers in the domain of herbal drug discovery. These challenges include difficulties in accessing expensive screening methods, technological limitations that hinder the full analysis of phytochemical complexity, and the significant time required to conduct a comprehensive phytochemical study [251,252]. In contrast, pharmaceutical companies have achieved remarkable success in drug discovery research, demonstrated by their rapid development of potential vaccines for the COVID-19 pandemic in 2020 [253,254].
In recent years, there has been a notable increase in studies related to clinical trials on medicinal plants and their main active components. In 2005, the World Health Organization (WHO) issued operational guidelines outlining the regulatory requirements to support clinical trials of herbal products [84].
In Algeria, conducting clinical trials on pharmaceutical products, herbal preparations, or natural bioactive compounds is severely restricted, and the process is subject to complex laws [255]. Numerous species of Hypericum have undergone thorough investigations to explore their pharmacological and phytochemical properties, particularly concerning their clinical efficacy in treating mild to moderate depression [256]. St. John’s wort (Hypericum perforatum) is a herbal product known for its clinically significant effects [257]. In a single-blind clinical study conducted by Dimpfel et al., the effects on the CNS of two commercially available extracts of St. John’s wort (Texx 300 and Jarsin 300) were compared with those of a placebo in a group of healthy young volunteers (n = 35) [258]. Both extracts showed a decrease in the cognitive potential P300, suggesting an enhanced mental performance. The observed neurophysiological changes were consistent with the proposed clinical efficacy. Moreover, quantitative EEG allowed for a discrimination between St. John’s wort extracts with respect to the time of effect and profile changes on the neuronal communication structure [258].
Melissa officinalis, also known as lemon balm, is traditionally used as a mild sedative, for memory enhancement, and for treating anxiety, depression, and insomnia [34,259]. It has been reported to have many biological activities, such as antioxidant, antimicrobial, antitumor, antiviral, antiallergic, anti-inflammatory, and flatulence-inhibiting effects [175]. The beneficial properties of Melissa officinalis can be related to bioactive compounds, such as terpenoids, alcohols, rosmarinic acid, and phenolic antioxidants, which are present in the plant [175]. In a clinical study conducted by Safari et al., Melissa officinalis extract demonstrated a significant impact on reducing depression and anxiety in patients with type 2 diabetes [260]. Another study evaluated samples of Melissa officinalis for their potential to inhibit human AChE and bind to cholinergic receptors [261]. Subsequently, the cognitive and mood effects of single doses of the most cholinergic active dried leaf extract were assessed in a randomized, placebo-controlled, double-blind, balanced crossover study. Twenty healthy, young participants received single doses of 600, 1000, and 1600 mg of encapsulated dried leaf, or a matching placebo, at 7-day intervals. Cognitive performance and mood were evaluated before the dose and at 1, 3, and 6 h post-dose using the Cognitive Drug Research computerized assessment battery and Bond–Lader visual analog scales, respectively. The most prominent cognitive and mood effects consistently observed at all post-dose time points were the enhancement of memory performance and an increase in feelings of calmness, particularly noticeable with the highest dose of 1600 mg. These findings suggest that doses of Melissa officinalis equal to or exceeding the maximum dosage utilized in the study have the potential to improve cognitive function and mood. As a result, they may hold promise as a valuable adjunct in the treatment of AD.
Furthermore, these results underscore the significance of considering that different preparations derived from the same plant may exhibit diverse characteristics based on the specific sample preparation methods employed.
In an 8-week randomized, double-blind clinical trial designed to assess the effectiveness of Melissa officinalis and Lavandula angustifolia in comparison to fluoxetine for treating mild to moderate depression, 45 adult outpatients meeting the criteria outlined in the Diagnostic and Statistical Manual of Mental Disorders, 5th edition (DSM-5), for major depression were randomly divided into three groups. These groups were assigned to daily doses of either M. officinalis (2 g), L. angustifolia (2 g), or fluoxetine (20 mg). Evaluations were conducted at weeks 0, 2, 4, and 8 using the Hamilton Rating Scale for Depression (HAM-D), which comprised 17 items [262]. The results indicate that both Melissa officinalis and Lavandula angustifolia have a similar effect to fluoxetine when treating mild to moderate depression (F = 0.131, df = 2.42, p = 0.877).
The selection of Salvia officinalis species for clinical trials was grounded in its well-established properties and supported by a substantial body of evidence derived from both in vitro and in vivo studies. These clinical trials have a primary objective of evaluating its potential beneficial impact on cognitive function, targeting both healthy individuals and patients coping with cognitive impairments, such as AD [263].
In a controlled, double-blind, crossover study, a total of 30 healthy participants visited the laboratory on three separate occasions, spaced 7 days apart. During each visit, they were administered a different treatment in a counterbalanced sequence, which included a placebo and two doses of dried sage leaf (300 and 600 mg). Mood assessments were conducted both before taking the treatment and at 1 and 4 h after ingestion. These mood assessments involved the completion of Bond–Lader mood scales and the State Trait Anxiety Inventory (STAI), both before and after engaging in a 20 min session of the Defined Intensity Stress Simulator (DISS) computerized multitasking battery [263]. In a correlated study, an extract derived from Salvia officinalis leaves demonstrated a dose-dependent inhibition of AChE and, to a more pronounced extent, butyrylcholinesterase in in vitro experiments. The lower dosage was observed to alleviate anxiety, while the higher dosage heightened feelings of “alertness”, “calmness”, and “contentedness”, as measured by the Bond–Lader mood scales. These findings provide further validation of previous research concerning the cholinesterase inhibitory characteristics of Salvia officinalis [263].
The well-known plant Passiflora incarnata, often used as a calming herb to treat anxiety and insomnia, was the subject of a clinical study conducted by Movafegh et al. [264]. In this study, the effectiveness of Passiflora incarnata in controlling anxiety in preoperative patients was investigated. The plant extract was consumed as tea and showed positive effects on the quality of sleep [264]. In the study, 60 patients were randomized into two groups to orally receive Passiflora incarnata (500 mg, Passipy™ IranDarouk) (n = 30) or placebo (n = 30) as premedication, 90 min before surgery. A numerical rating scale (NRS) was used for each patient to assess anxiety and sedation before and 10, 30, 60, and 90 min after premedication. The results revealed that the NRS anxiety scores were significantly lower in the Passiflora group compared to the control group (p < 0.001). Importantly, the study suggested that the administration of oral Passiflora incarnata as a premedication effectively reduced anxiety without inducing sedation in outpatient surgery [264].

4.8. Toxicological Evidence

The widespread availability of plants renders them easily accessible for various applications, with the majority of these plants generally regarded as safe. Nonetheless, it is crucial to acknowledge that certain plants can elicit adverse secondary effects, including harmful toxic reactions. In specific instances, the utilization of particular plants or specific plant components can even prove fatal. The toxic properties of these plants emanate from the presence of active compounds distributed throughout the entire plant or in one or more of its constituent parts, such as the leaves, fruits, or roots. These compounds are associated with specific organ toxicity in both humans and animals, leading to a spectrum of disorders that can range from mild to severe, and in extreme cases may culminate in fatality.
Various plants and their active constituents, as outlined in Table 1, have been linked to toxicity. For instance, wormwood (Artemisia absinthium L.) was banned in numerous countries during the early 20th century due to its neuropsychiatric toxicity. This toxic effect can be attributed to thujones, which are volatile monoterpene ketones. Thujones make up approximately 0.25–1.32% of the entire herb and 3–12% of the essential oil. The neurotoxic mechanism of thujones has undergone extensive investigation in animal studies, cultured neuronal cells, and studies involving expressed receptors [151]. Among thujones, α-thujone is roughly two-to-three-times more potent than β-thujone in modulating the GABA-gated chloride channel. It is worth noting that the neuronal effects of thujone were found to be fully reversible [151]. Thujone has been documented as toxic to the brain, liver, and kidney cells, and its ingestion in excessive amounts can lead to various adverse effects, including convulsions (muscle spasms), wheezing, restlessness, anxiety, insomnia, vomiting, vertigo, rapid heart rate, kidney damage, epileptic seizures, and even psychedelic effects [151].
Ferula communis L. is a common Mediterranean plant of the Apiaceae family. It has traditionally been used to treat anxiety and hysteria [50,265]. The rhizomes of this plant, known as Al-kalakh in Arabic, are used locally as a traditional remedy for the treatment of skin infections [50,52]. In vitro and in vivo studies have shown that the plant’s crude extracts and isolated components possess various pharmacological properties, including antidiabetic, antimicrobial, antiproliferative, and cytotoxic activities [265]. The toxicity of Ferula has been well-documented in North Africa for an extended period of time [50]. Incidents of poisoning have been frequently reported in various animal species, particularly sheep, cattle, pigs, horses, and goats, often occurring after several days of grazing in the fields. Additionally, cases of Ferula-induced toxicity have been documented in humans [265]. Exposure to Ferula communis can lead to a hemorrhagic syndrome known as ferulosys [266]. Active compounds, notably the prenylated coumarin derivatives ferulenol and ferprenin (Figure 15), have been identified as inhibitors of the vitamin K recycling process in liver microsomes, specifically targeting the enzymatic system known as vitamin K epoxide reductase (VKORC1) [267]. The inhibition of VKORC1 by these 4-hydroxycoumarin derivatives limits the availability of KH2 for the carboxylation reaction, resulting in the partial carboxylation of vitamin K-dependent blood-clotting factors II, VII, IX, and X. The specific inhibition of VKORC1 leads to the cessation of clotting factors’ activation, ultimately culminating in a fatal hemorrhage [267].
A study conducted by Lahmar et al. [268] reported on Ferula communis intoxication in goats. They determined the acute LD50 of ferulenol in albino mice to be 2100 mg/kg bw after a single oral administration and 319 mg/kg bw after an intraperitoneal administration. Notably, ferulenol exhibited a higher LD50 compared to warfarin, indicating lower acute toxicity. Three days following the ferulenol administration, the dosed animals exhibited hypoprothrombinemia, accompanied by both internal and external hemorrhages resembling the symptoms described in ferulosis and experimental anti-vitamin K poisonings [268].
Merino sheep in a flock located on the southwestern slopes of New South Wales experienced intoxication, serving as yet another instance of neurological toxicity in animals resulting from the consumption of specific plant species while grazing in the fields [269]. After consuming Stachys arvensis, a plant belonging to the Lamiaceae family, these animals developed a neurological disorder, with numerous fatalities linked to the outbreak. The impacted sheep exhibited symptoms consistent with mild degenerative myelopathy and peripheral neuropathy. Additionally, deficiencies in vitamins A and E were identified within the affected flock. Clinical examinations and post-mortem analyses were conducted on the sheep displaying clinical signs. The research concluded that grazing on Stachys arvensis is at times linked to a neurological locomotor disorder in sheep [269].
Peganum harmala L. is a plant with a broad distribution across the Mediterranean region, including Algeria. It is commonly utilized in traditional medicine for its sedative properties and various other purposes, but its use comes with the inherent risk of overdose and poisoning for users [203]. The intoxication caused by Peganum harmala L. seeds has been widely reported in both animals and humans [270,271]. Ingesting high quantities of Peganum harmala can lead to various clinical manifestations of intoxication, including digestive disorders, bradycardia, neurological symptoms, such as euphoria, hallucinations, generalized tremors, and, in severe cases, it can lead to paralysis, central nervous system depression, dyspnea, and arterial hypotension [272]. The main alkaloid components found in Peganum harmala seeds consist of the beta-carboline alkaloids harmine, harmaline, and tetrahydroharmine [203]. These compounds are also detectable in other plants, like Passiflora incarnata, and are notably present in the psychoactive brew called Ayahuasca [273]. Ayahuasca is a hallucinogenic beverage with roots in indigenous Amazonian traditions, where it has been employed in religious ceremonies and therapeutic rituals [274]. The toxicokinetics and toxicodynamics of the psychoactive alkaloids harmine, harmaline, and tetrahydroharmine have undergone extensive examination. Comprehensive documentation exists on the psychological, physiological, and toxic effects resulting from their simultaneous consumption [275].
Datura stramonium and Atropa belladonna belong to the Solanaceae family and have a rich historical background as hallucinogenic substances [276]. These plants have also played a significant role in traditional medicine and have been linked to practices involving sorcery, witchcraft, and magico-religious rituals. It is important to note that both of these plants contain the active hallucinogenic compounds atropine and scopolamine [277]. Tropane alkaloids are important natural products that are abundantly present in the Solanaceae family and include the anticholinergic drugs atropine, hyoscyamine, and scopolamine (Figure 16) [278].
Instances of Datura intoxication can arise from various causes, including medication overdoses, the misuse of edible plants, intentional abuse for hallucinogenic purposes, criminal activities, such as homicide or robbery, and accidental poisoning from contaminated food [279]. Reported symptoms of toxicity encompass dizziness, dry mouth, flushed skin, palpitations, nausea, drowsiness, tachycardia, blurred vision, mydriasis (dilated pupils), hyperthermia, disorientation, vomiting, agitation, delirium, urine retention, hypertension, and even coma [279]. The belladonna alkaloids (atropine, hyoscyamine, and scopolamine) present in the roots, leaves, and fruits of these plants pose a substantial risk, particularly to infants, children, and adolescents, and can result in an anticholinergic toxidrome [280]. Cases of anticholinergic poisoning stemming from contamination with belladonna alkaloids have been reported in various food items, such as commercially purchased Paraguay tea, hamburgers, and honey [281]. Additionally, there have been other documented instances of tropane alkaloid-related intoxications, including a significant epidemic in New York and the eastern United States, attributed to heroin contamination with scopolamine [282].
Figure 16. Toxic tropane alkaloids identified in Datura stramonium growing in Algeria [283,284].
Figure 16. Toxic tropane alkaloids identified in Datura stramonium growing in Algeria [283,284].
Plants 12 03860 g016
While essential oils have a wide range of applications, they should be used with caution due to their potential toxic effects, such as phototoxicity or photosensitivity. In some cases, high doses of certain essential oils can lead to hepatotoxicity, nephrotoxicity, and neurotoxicity [285]. For instance, previous studies recommend that oregano essential oil should never be applied topically to mucous membranes in concentrations higher than 1% due to its potential for skin irritation and even a burning sensation [286]. The main compounds in oregano essential oil, carvacrol and thymol (Figure 8), can be toxic to the liver, kidneys, and nervous system if consumed in excessive quantities. Furthermore, carvacrol and thymol have demonstrated in vivo mutagenicity and genotoxicity [287,288].
The essential oils derived from the Juniperus species exhibit a diverse range of pharmacological properties. Nevertheless, it is crucial to highlight that there have been numerous reported cases of intoxication associated with cade oil, obtained from the branches of Juniperus oxycedrus L. (Cupressaceae) [289,290,291]. Phenol, a constituent of cade oil, can exert adverse effects on various organ systems, including the central and peripheral nervous systems, cardiovascular system, liver and biliary systems, skin, and respiratory system [291].
In a study aimed at assessing the toxicity of constituents found in cade essential oil, Juniperus oxycedrus L., researchers identified eight structurally related compounds and an additional 16 previously known cade oil constituents that were toxic to adult American house dust mites (AHDMs) [292]. Among these compounds, methyl-eugenol (LD50, 5.82 µg/cm2) and guaiacol (LD50, 8.24 µg/cm2) demonstrated the highest toxicity against the mites. It is worth noting that the toxicity of these compounds, as well as benzyl benzoate, do not exhibit significant differences. Additionally, eugenol, m-cresol, and nerolidol (LD50, 12.52–19.52 µg/cm2) exhibited notably high toxicity levels compared to N,N-diethyl-3-methylbenzamide (DEET) (LD50, 37.67 µg/cm2). When cade oil was applied experimentally as sprays at concentrations of 30 or 40 mg/L, it resulted in 96% and 100% mortality rates among the mites, respectively. In contrast, the application of permethrin (cis:trans, 25:75) at a concentration of 2.5 g/L via spray treatment only yielded a 17% mortality rate [292].

5. Materials and Methods

Information regarding the traditional use of medicinal plants for the management of mental illnesses and central nervous system (CNS) disorders in Algeria was gathered through an extensive literature search. This encompassed sourcing data from books, e-books, encyclopedias detailing medicinal plant uses, as well as Master’s and PhD theses focused on the subject matter. Additionally, various scientific databases, including Scifinder®, Scopus, Scholars, PubMed, Medline, Web of Science, the World Health Organization (WHO), SFE (Societe Francaise d’Ethnopharmacologie), and Google Scholar, were meticulously scrutinized for historical and ethno-pharmacological information, biological and pharmacological actions.
The search parameters were confined to specific keywords, including “treatment of mental illnesses with medicinal plants”, “Algeria”, “North Africa”, “Algerian medicinal plants”, “pharmacopeia”, “neuropsychiatric disorders”, “in vitro study”, “in vivo study”, “clinical trials of medicinal plants”, “secondary metabolites and CNS interaction”, and “natural drug interactions and toxicity”. All identified papers underwent a thorough inspection for relevance, and additional sources were identified by examining reference lists.

6. Conclusions and Prospects

Algeria’s vast and diverse territory, enriched with abundant biodiversity and a variety of landscapes, stands as a testament to its unique geographical location and historical interactions with diverse civilizations and cultures. This historical tapestry has profoundly influenced the development of Algeria’s pharmacopeia, creating a rich repository of knowledge regarding the medicinal properties of countless herbs. The indigenous wisdom held by Algeria’s local population concerning medicinal plants is an invaluable treasure trove of insights.
Nonetheless, there exists an urgent imperative to systematically document the medicinal applications of Algerian plants. This urgency stems from the alarming rate of natural habitat degradation, driven by factors like overexploitation, wildfires, anthropogenic activities, and the fading expertise of traditional healers. Together, these factors contribute to the unfortunate erosion of precious traditional knowledge.
To address this pressing need, securing funding for dedicated research initiatives becomes paramount. These initiatives should focus on conducting comprehensive investigations into specific plant species native to Algeria. The primary objective of these studies would be to assess their potential effectiveness in treating brain disorders and neurodegenerative diseases, as well as to pinpoint the bioactive compounds responsible for their pharmacological effects.
Looking ahead, our primary focus should center on the meticulous selection of the most promising plants, particularly those demonstrating evident effects on the central nervous system yet lacking sufficient scientific scrutiny to date. By delving into thorough pharmacological examinations of these selected plants, we can unlock the secrets of their bioactive molecules and elucidate their mechanisms of action. This approach not only safeguards Algeria’s rich herbal heritage but also holds promise for discovering novel treatments for neurological ailments, ultimately benefiting both local communities and the broader scientific community.

Author Contributions

Conceptualization-writing—original draft preparation, F.L. (Farida Larit); writing—review and editing, F.L. (Farida Larit) and F.L. (Francisco León); visualization-reviewing, F.L. (Farida Larit) and F.L. (Francisco León); supervision, F.L. (Farida Larit) and F.L. (Francisco León). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grant number 5P20GM109091 from the National Institute of General Medical Sciences (NIGMS), a component of the National Institutes of Health (NIH); its contents are solely the responsibility of the authors and do not necessarily represent the official view of the NIGMS or the NIH.

Data Availability Statement

Data are contained within the article.

Acknowledgments

F.L. thanks the College of Pharmacy, University of South Carolina, for the start-up funds.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Quezel, P.; Santa, S. New Flora of Algeria and Southern Desert Regions; Centre National de la Recherche Scientifique: Paris, France, 1962. [Google Scholar]
  2. Ozenda, P. Flore du Sahara Septentrional et Central; Centre National de la Recherche Scientifique: Paris, France, 1958. [Google Scholar]
  3. Benchelah, A.-C.; Bouziane, H.; Maka, M. Fleurs du Sahara, arbres et arbustes, voyage au coeur de leurs usages avec les Touaregs du Tassili. Phytothérapie 2004, 2, 191–197. [Google Scholar] [CrossRef]
  4. Bellakhdar, J. Plantes Médicinales au Maghreb et Soins de Base. Précis de Phytothérapie Moderne; Editions le Fennec: Casablanca, Morocco, 2006; 386p. [Google Scholar]
  5. Baba Aissa, F. Encyclopédie des Plantes Utiles; EDAS: Leonia, NJ, USA, 1999. [Google Scholar]
  6. Allen, C.D.; Macalady, A.K.; Chenchouni, H.; Bachelet, D.; McDowell, N.; Vennetier, M.; Kitzberger, T.; Rigling, A.; Breshears, D.D.; Hogg, E.T. A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. For. Ecol. Manag. 2010, 259, 660–684. [Google Scholar] [CrossRef]
  7. Amrani, S.; Seaward, M.R.; Sipman, H.J.; Feuerer, T. Lichenological exploration of Algeria II: Checklist of lichenized, lichenicolous and allied fungi. Herzogia 2018, 31, 817–892. [Google Scholar] [CrossRef]
  8. Cercleux, A.-L.; Comãnescu, L.; Costachie, S.; Demeter, T.; Dobre, R.; Grecu, F.; Hachemi, K.; Cristian, I.; Kanapyanov, T.; Mihai, B. Cinq Continents Revue Roumaine de Géographie. 2018, Volume 17, pp. 2247–2290. Available online: http://cinqcontinents.geo.unibuc.ro/ (accessed on 5 November 2023).
  9. Tahri, D.; Elhouiti, F.; Ouinten, M.; Yousfi, M. Historical perspective of Algerian pharmacological knowledge. Adv. Tradit. Med. 2020, 20, 279–290. [Google Scholar] [CrossRef]
  10. Yoeli-Tlalim, R. Reorienting Histories of Medicine: Encounters Along the Silk Roads; Bloomsbury Publishing: London, UK, 2021. [Google Scholar]
  11. Bustinza, V.P.d. How Early Islamic Science Advanced Medicine. National Geographic HISTORY. 2016. Available online: https://www.nationalgeographic.com/history/history-magazine/article/muslim-medicine-scientific-discovery-islam (accessed on 23 October 2023).
  12. Al-Khalili, J. The greatest scientific advances from the Muslim world. The Guardian. 2010. Available online: https://www.theguardian.com/science/2010/feb/01/islamic-science (accessed on 23 October 2023).
  13. Abdekhoda, M.; Ranjbaran, F. The Holy Quran and Treatment of Mental and Physical Diseases. Pastor. Psychol. 2022, 71, 423–435. [Google Scholar] [CrossRef]
  14. Awaad, R.; Elsayed, D.; Helal, H. Holistic healing: Islam’s legacy of mental health. Yaqeen Inst. Islam. Res. 2021. Available online: https://yaqeeninstitute.ca/read/paper/holistic-healing-islams-legacy-ofmental-health. (accessed on 23 October 2023).
  15. AlRawi, S.N.; Khidir, A.; Elnashar, M.S.; Abdelrahim, H.A.; Killawi, A.K.; Hammoud, M.M.; Fetters, M.D. Traditional Arabic & Islamic medicine: Validation and empirical assessment of a conceptual model in Qatar. BMC Complement. Altern. Med. 2017, 17, 1–10. [Google Scholar]
  16. Available online: https://www.ixl.com/social-studies/grade-7/the-ancient-silk-road-geography-and-transportation (accessed on 5 November 2023).
  17. Feigin, V.L.; Nichols, E.; Alam, T.; Bannick, M.S.; Beghi, E.; Blake, N.; Culpepper, W.J.; Dorsey, E.R.; Elbaz, A.; Ellenbogen, R.G. Global, regional, and national burden of neurological disorders, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2019, 18, 459–480. [Google Scholar] [CrossRef]
  18. WHO. Mental Disorders. 2022. Available online: https://www.who.int/news-room/fact-sheets/detail/mental-disorders (accessed on 23 October 2023).
  19. Ding, C.; Wu, Y.; Chen, X.; Chen, Y.; Wu, Z.; Lin, Z.; Kang, D.; Fang, W.; Chen, F. Global, regional, and national burden and attributable risk factors of neurological disorders: The Global Burden of Disease study 1990–2019. Front. Public Health 2022, 10, 952161. [Google Scholar] [CrossRef]
  20. Metzner, R. Hallucinogenic drugs and plants in psychotherapy and shamanism. J. Psychoact. Drugs 1998, 30, 333–341. [Google Scholar] [CrossRef]
  21. Moore, J.; Martin, E. Trial of reserpine in treatment of schizophrenia. Br. Med. J. 1957, 1, 8. [Google Scholar] [CrossRef] [PubMed]
  22. Lobay, D. Rauwolfia in the treatment of hypertension. Integr. Med. A Clin. J. 2015, 14, 40. [Google Scholar]
  23. Prisinzano, T. Natural products as tools for neuroscience: Discovery and development of novel agents to treat drug abuse. J. Nat. Prod. 2009, 72, 581–587. [Google Scholar] [CrossRef]
  24. Blaylock, R.L.; Maroon, J. Natural plant products and extracts that reduce immunoexcitotoxicity-associated neurodegeneration and promote repair within the central nervous system. Surg. Neurol. Int. 2012, 3, 19. [Google Scholar] [CrossRef] [PubMed]
  25. Bui, T.T.; Nguyen, T.H. Natural product for the treatment of Alzheimer’s disease. J. Basic Clin. Physiol. Pharmacol. 2017, 28, 413–423. [Google Scholar] [CrossRef]
  26. Hritcu, L.; Ionita, R.; Postu, P.A.; Gupta, G.K.; Turkez, H.; Lima, T.C.; Carvalho, C.U.S.; de Sousa, D.P. Antidepressant Flavonoids and Their Relationship with Oxidative Stress. Oxidat. Med. Cell. Longev. 2017, 2017, 5762172. [Google Scholar] [CrossRef]
  27. Oliveira, A.I.; Pinho, C.; Sarmento, B.; Dias, A.C. Neuroprotective activity of Hypericum perforatum and its major components. Front. Plant Sci. 2016, 7, 1004. [Google Scholar] [CrossRef]
  28. Taqui, R.; Debnath, M.; Ahmed, S.; Ghosh, A. Advances on plant extracts and phytocompounds with acetylcholinesterase inhibition activity for possible treatment of Alzheimer’s disease. Phytomed. Plus 2022, 2, 100184. [Google Scholar] [CrossRef]
  29. WHO. Mental Health Atlas 2020 Country Profile: Algeria. Available online: https://www.who.int/publications/m/item/mental-health-atlas-dza-2020-country-profile (accessed on 23 October 2023).
  30. Kacha, F. La psychiatrie en Algérie. l’Inf. Psychiatr. 2005, 81, 145–148. [Google Scholar]
  31. Iserin, P.; Masson, M.; Restellini, J.-P. LAROUSSE des Plantes Médicinales; Larousse: Paris, France, 2007. [Google Scholar]
  32. Al-Issa, I. psychiatry in Algeria. Psychiatr. Bull. 1989, 13, 240–245. [Google Scholar] [CrossRef]
  33. Dein, S. Traditional Healers and Global Mental Health. In Innovations in Global Mental Health; Springer: Cham, Switzerland, 2021; pp. 807–817. [Google Scholar]
  34. Bouasla, A.; Bouasla, I. Ethnobotanical survey of medicinal plants in northeastern of Algeria. Phytomedicine 2017, 36, 68–81. [Google Scholar] [CrossRef]
  35. Hassaïne, S.; Benmalek, S. Medicinal plants traditionally used in the Algerian Sahara: An ethnobotanical study. Vegetos 2023, 36, 400–426. [Google Scholar] [CrossRef]
  36. Belhouala, K.; Benarba, B. Medicinal plants used by traditional healers in Algeria: A multiregional ethnobotanical study. Front. Pharmacol. 2021, 12, 760492. [Google Scholar] [CrossRef] [PubMed]
  37. Mann, J. Magic, Murder and Medicine; Oxford University Press: Oxford, UK, 1994. [Google Scholar]
  38. Mulas, M. Traditional uses of Labiatae in the Mediterranean area. Acta Hortic. 2006, 723, 25–32. [Google Scholar] [CrossRef]
  39. Cornara, L.; La Rocca, A.; Marsili, S.; Mariotti, M. Traditional uses of plants in the Eastern Riviera (Liguria, Italy). J. Ethnopharmacol. 2009, 125, 16–30. [Google Scholar] [CrossRef]
  40. EAGL. Related Expert Assessment Group: Maghreb Algeria. Available online: https://iucngreenlist.org/country/algeria/ (accessed on 24 October 2023).
  41. Debbache-Benaida, N.; Berboucha, M.; Ayouni, K.; Atmani, D.; Nassima, C.; Boudaoud, H.; Djebli, N.; Atmani, D. Anti-hyperuricemic and neuroprotective effects of Populus nigra L. (Saliacaceae) flower buds used in Algerian folk medicine. J. Pharm. Pharmacogn. Res. 2018, 6, 471–482. [Google Scholar]
  42. Do Rego, J.-C.; Benkiki, N.; Chosson, E.; Kabouche, Z.; Seguin, E.; Costentin, J. Antidepressant-like effect of hyperfoliatin, a polyisoprenylated phloroglucinol derivative from Hypericum perfoliatum (Clusiaceae) is associated with an inhibition of neuronal monoamines uptake. Eur. J. Pharmacol. 2007, 569, 197–203. [Google Scholar] [CrossRef]
  43. Larit, F.; Elokely, K.M.; Chaurasiya, N.D.; Benyahia, S.; Nael, M.A.; León, F.; Abu-Darwish, M.S.; Efferth, T.; Wang, Y.-H.; Belouahem-Abed, D. Inhibition of human monoamine oxidase A and B by flavonoids isolated from two Algerian medicinal plants. Phytomedicine 2018, 40, 27–36. [Google Scholar] [CrossRef]
  44. Sakhri, F.Z.; Adachi, N.; Zerizer, S.; Ohashi, Y.; Ikemoto, H.; Tsukada, M.; Kabouche, Z.; Hisamitsu, T.; Sunagawa, M. Behavioral and neurological improvement by Cydonia oblonga fruit extract in chronic immobilization stress rats. Phytother. Res. 2021, 35, 2074–2084. [Google Scholar] [CrossRef]
  45. Al-Snafi, A.E. Constituents and pharmacology of Narcissus tazetta. IOSR J. Pharm. 2020, 10, 44–53. [Google Scholar]
  46. Choukry, K.T. Neuf Espèces Végétales Anti-Epileptiques de la Flore d’Algérie. Alger. J. Nat. Prod. 2019, 7, 706–713. [Google Scholar]
  47. López, S.; Bastida, J.; Viladomat, F.; Codina, C. Acetylcholinesterase inhibitory activity of some Amaryllidaceae alkaloids and Narcissus extracts. Life Sci. 2002, 71, 2521–2529. [Google Scholar] [CrossRef]
  48. Azib, L.; Debbache-Benaida, N.; Da Costa, G.; Atmani-Kilani, D.; Saidene, N.; Ayouni, K.; Richard, T.; Atmani, D. Pistacia lentiscus L. leaves extract and its major phenolic compounds reverse aluminium-induced neurotoxicity in mice. Ind. Crops Prod. 2019, 137, 576–584. [Google Scholar] [CrossRef]
  49. Taïbi, K.; Abderrahim, L.A.; Ferhat, K.; Betta, S.; Taïbi, F.; Bouraada, F.; Boussaid, M. Ethnopharmacological study of natural products used for traditional cancer therapy in Algeria. Saudi Pharm. J. 2020, 28, 1451–1465. [Google Scholar] [CrossRef]
  50. Hammiche, V.; Merad, R.; Azzouz, M.; Goetz, P. Plantes Toxiques à Usage Médicinal du Pourtour Méditerranéen; Springer: Paris, France, 2013. [Google Scholar]
  51. Küpeli Akkol, E.; Tatlı Çankaya, I.; Şeker Karatoprak, G.; Carpar, E.; Sobarzo-Sánchez, E.; Capasso, R. Natural compounds as medical strategies in the prevention and treatment of psychiatric disorders seen in neurological diseases. Front. Pharmacol. 2021, 12, 669638. [Google Scholar] [CrossRef]
  52. Tagliapietra, S.; Aragno, M.; Ugazio, G.; Nano, G. Experimental studies on the toxicity of some compounds isolated from Ferula communis in the rat. Res. Commun. Chem. Pathol. Pharmacol. 1989, 66, 333–336. [Google Scholar] [PubMed]
  53. Mahendra, P.; Bisht, S. Ferula asafoetida: Traditional uses and pharmacological activity. Pharmacogn. Rev. 2012, 6, 141. [Google Scholar] [CrossRef]
  54. Chevallier, A. The Encyclopedia of Medicinal Plants; Dorling Kindersley Limited: London, UK, 2001. [Google Scholar]
  55. Motti, R.; de Falco, B. Traditional herbal remedies used for managing anxiety and insomnia in Italy: An ethnopharmacological overview. Horticulturae 2021, 7, 523. [Google Scholar] [CrossRef]
  56. Elyebdri, N.; Boumediou, A.; Addoun, S. Ethnobotanical study on the usage of toxic plants in traditional medicine in the city center of Tlemcen, Algeria. Int. J. Pharmacol. Pharm. Sci. 2017, 11, 642–646. [Google Scholar]
  57. Boudjelal, A.; Henchiri, C.; Sari, M.; Sarri, D.; Hendel, N.; Benkhaled, A.; Ruberto, G. Herbalists and wild medicinal plants in M’Sila (North Algeria): An ethnopharmacology survey. J. Ethnopharmacol. 2013, 148, 395–402. [Google Scholar] [CrossRef]
  58. Desai, C. Meyler’s side effects of drugs: The international encyclopedia of adverse drug reactions and interactions. Indian J. Pharmacol. 2016, 48, 224. [Google Scholar]
  59. Bellakhdar, J. Contribution à l’Étude de la Pharmacopée Traditionnelle au Maroc: La Situation Actuelle, les Produits, les Sources du Savoir (Enquête Ethnopharmacologique de Terrain Réalisée de 1969 à 1992); Université Paul Verlaine-Metz: Metz, France, 1997. [Google Scholar]
  60. Lefranc, M.E. Étude Botanique, Chimique et Toxicologique sur l’Atractylis Gummifera. Bull. Soc. Bot. Fr. 1866, 13, 146–157. [Google Scholar] [CrossRef]
  61. Bouabid, K.; Lamchouri, F.; Toufik, H.; Faouzi, M.E.A. Phytochemical investigation, in vitro and in vivo antioxidant properties of aqueous and organic extracts of toxic plant: Atractylis gummifera L. J. Ethnopharmacol. 2020, 253, 112640. [Google Scholar] [CrossRef] [PubMed]
  62. Shaaban, M.; El-Hagrassi, A.M.; Osman, A.F.; Soltan, M.M. Bioactive compounds from Matricaria chamomilla: Structure identification, in vitro antiproliferative, antimigratory, antiangiogenic, and antiadenoviral activities. Z. Für Nat. C 2022, 77, 85–94. [Google Scholar] [CrossRef] [PubMed]
  63. Hamel, T.; Sadou, S.; Seridi, R.; Boukhdir, S.; Boulemtafes, A. Pratique traditionnelle d’utilisation des plantes médicinales dans la population de la péninsule de l’edough (nord-est algérien). Ethnopharmacologia 2018, 59, 65–70. [Google Scholar]
  64. Baziz, K.; Maougal, R.A.A. An ethnobotanical survey of spontaneous plants used in traditional medicine in the region of Aures, Algeria. Eur. J. Ecol. 2021, 6, 14669. [Google Scholar] [CrossRef]
  65. Meddour, R.; Sahar, O.; Ouyessad, M. Ethnobotanical survey on medicinal plants in the Djurdjura National Park and its influence area, Algeria. Ethnobot. Res. Appl. 2020, 20, 1–25. [Google Scholar]
  66. Tavares, L.; McDougall, G.J.; Fortalezas, S.; Stewart, D.; Ferreira, R.B.; Santos, C.N. The neuroprotective potential of phenolic-enriched fractions from four Juniperus species found in Portugal. Food Chem. 2012, 135, 562–570. [Google Scholar] [CrossRef]
  67. Dabaghzadeh, F.; Sharififar, F.; Ahmadzadeh, A.-M.; Karami-Mohajeri, S. The effects of Berberis vulgaris L. root extract on the opiate withdrawal syndrome and psychological factors: A randomized double-blind clinical trial. J. Basic Clin. Physiol. Pharmacol. 2021, 34, 465–472. [Google Scholar] [CrossRef]
  68. Imenshahidi, M.; Qaredashi, R.; Hashemzaei, M.; Hosseinzadeh, H. Inhibitory effect of Berberis vulgaris aqueous extract on acquisition and reinstatement effects of morphine in conditioned place preferences (CPP) in mice. Jundishapur J. Nat. Pharm. Prod. 2014, 9, e16145. [Google Scholar] [CrossRef]
  69. Meliani, N.; Dib, M.E.A.; Allali, H.; Tabti, B. Hypoglycaemic effect of Berberis vulgaris L. in normal and streptozotocin-induced diabetic rats. Asian Pac. J. Trop. Biomed. 2011, 1, 468–471. [Google Scholar] [CrossRef] [PubMed]
  70. Rahimi-Madiseh, M.; Lorigoini, Z.; Zamani-Gharaghoshi, H.; Rafieian-Kopaei, M. Berberis vulgaris: Specifications and traditional uses. Iran. J. Basic Med. Sci. 2017, 20, 569. [Google Scholar]
  71. Merzouki, A.; Ed-Derfoufi, F.; Mesa, J.M. Contribution to the knowledge of Rifian traditional medicine. II: Folk medicine in Ksar Lakbir district (NW Morocco). Fitoterapia 2000, 71, 278–307. [Google Scholar] [CrossRef] [PubMed]
  72. Ouelbani, R.; Bensari, S.; Mouas, T.N.; Khelifi, D. Ethnobotanical investigations on plants used in folk medicine in the regions of Constantine and Mila (North-East of Algeria). J. Ethnopharmacol. 2016, 194, 196–218. [Google Scholar] [CrossRef] [PubMed]
  73. Azzi, R.; Djaziri, R.; Lahfa, F.; Sekkal, F.Z.; Benmehdi, H.; Belkacem, N. Ethnopharmacological survey of medicinal plants used in the traditional treatment of diabetes mellitus in the North Western and South Western Algeria. J. Med. Plants Res. 2012, 6, 2041–2050. [Google Scholar]
  74. Madia, V.N.; De Angelis, M.; De Vita, D.; Messore, A.; De Leo, A.; Ialongo, D.; Tudino, V.; Saccoliti, F.; De Chiara, G.; Garzoli, S. Investigation of Commiphora myrrha (Nees) Engl. oil and its main components for antiviral activity. Pharmaceuticals 2021, 14, 243. [Google Scholar] [CrossRef]
  75. Akkol, E.K.; Ilhan, M.; Karpuz, B.; Genç, Y.; Sobarzo-Sánchez, E. Sedative and anxiolytic activities of Opuntia ficus indica (L.) Mill.: An experimental assessment in mice. Molecules 2020, 25, 1844. [Google Scholar] [CrossRef]
  76. Sharma, V.; Katiyar, A.; Agrawal, R. Glycyrrhiza glabra: Chemistry and pharmacological activity. In Sweeteners; Springer: Cham, Switzerland, 2016; pp. 1–14. [Google Scholar]
  77. Lemus-Mondaca, R.; Marin, J.; Rivas, J.; Sanhueza, L.; Soto, Y.; Vera, N.; Puente-Díaz, L. Pumpkin seeds (Cucurbita maxima). A review of functional attributes and by-products. Rev. Chil. Nutr 2019, 46, 783–791. [Google Scholar] [CrossRef]
  78. Benítez, G.; González-Tejero, M.; Molero-Mesa, J. Pharmaceutical ethnobotany in the western part of Granada province (southern Spain): Ethnopharmacological synthesis. J. Ethnopharmacol. 2010, 129, 87–105. [Google Scholar] [CrossRef]
  79. Pistelli, L.; D’angiolillo, F.; Morelli, E.; Basso, B.; Rosellini, I.; Posarelli, M.; Barbafieri, M. Response of spontaneous plants from an ex-mining site of Elba island (Tuscany, Italy) to metal (loid) contamination. Environ. Sci. Pollut. Res. 2017, 24, 7809–7820. [Google Scholar] [CrossRef]
  80. Gürağaç Dereli, F.T.; Khan, H.; Sobarzo-Sánchez, E.; Akkol, E.K. Antidepressant Potential of Lotus corniculatus L. subsp. corniculatus: An Ethnobotany Based Approach. Molecules 2020, 25, 1299. [Google Scholar] [CrossRef]
  81. Morris, J.B.; Tonnis, B.D.; Wang, M.L. Variability for Sennoside A and B concentrations in eight Senna species. Ind. Crops Prod. 2019, 139, 111489. [Google Scholar] [CrossRef]
  82. Rosmalena, R.; Senlia, A.O.; Hanafi, M.; Artanti, N.; Eldafira, E.; Handayani, S.I.; Lotulung, P.D.; Hartati, S.; Elya, B.; Zulfa, A. Phytochemical, Antioxidant and Antidiabetic properties of Senna alexandrina Leaf Extract. Res. J. Pharm. Technol. 2022, 15, 5835–5840. [Google Scholar] [CrossRef]
  83. Bouyahya, A.; El Omari, N.; Elmenyiy, N.; Guaouguaou, F.-E.; Balahbib, A.; El-Shazly, M.; Chamkhi, I. Ethnomedicinal use, phytochemistry, pharmacology, and toxicology of Ajuga iva (L.) schreb. J. Ethnopharmacol. 2020, 258, 112875. [Google Scholar] [CrossRef]
  84. Wang, X.Q. Operational guidance: Information needed to support clinical trials of herbal products. Chin. J. Clin. Pharmacol. Ther. 2005, 12, 582. [Google Scholar]
  85. Wei, H.; Kong, S.; Jayaraman, V.; Selvaraj, D.; Soundararajan, P.; Manivannan, A. Mentha arvensis and Mentha × piperita-Vital Herbs with Myriads of Pharmaceutical Benefits. Horticulturae 2023, 9, 224. [Google Scholar] [CrossRef]
  86. Arnold, H.-J.; Gulumian, M. Pharmacopoeia of traditional medicine in Venda. J. Ethnopharmacol. 1984, 12, 35–74. [Google Scholar] [CrossRef]
  87. Pooley, R.A. Fundamental physics of MR imaging. Radiographics 2005, 25, 1087–1099. [Google Scholar] [CrossRef]
  88. Boualam, K.; Bouhaddou, N.; Sobeh, M.; Tabyaoui, M.; Taghzouti, K. Mentha rotundifolia (L.) Huds. aqueous extract attenuates H2O2 induced oxidative stress and neurotoxicity. Front. Neurosci. 2023, 17, 1121029. [Google Scholar] [CrossRef]
  89. González-Tejero, M.; Casares-Porcel, M.; Sánchez-Rojas, C.; Ramiro-Gutiérrez, J.; Molero-Mesa, J.; Pieroni, A.; Giusti, M.; Censorii, E.; De Pasquale, C.; Della, A. Medicinal plants in the Mediterranean area: Synthesis of the results of the project Rubia. J. Ethnopharmacol. 2008, 116, 341–357. [Google Scholar] [CrossRef]
  90. Benarba, B.; Belabid, L.; Righi, K.; amine Bekkar, A.; Elouissi, M.; Khaldi, A.; Hamimed, A. Ethnobotanical study of medicinal plants used by traditional healers in Mascara (North West of Algeria). J. Ethnopharmacol. 2015, 175, 626–637. [Google Scholar] [CrossRef] [PubMed]
  91. Farida, S.H.M.; Ghorbani, A.; Ajani, Y.; Sadr, M.; Mozaffarian, V. Ethnobotanical applications and their correspondence with phylogeny in Apiaceae-Apioideae. Res. J. Pharmacogn. 2018, 5, 79–97. [Google Scholar]
  92. Piozzi, F.; Bruno, M. Diterpenoids from roots and aerial parts of the genus Stachys. Rec. Nat. Prod. 2011, 5, 1. [Google Scholar]
  93. Tomou, E.-M.; Barda, C.; Skaltsa, H. Genus Stachys: A review of traditional uses, phytochemistry and bioactivity. Medicines 2020, 7, 63. [Google Scholar] [CrossRef] [PubMed]
  94. Uritu, C.M.; Mihai, C.T.; Stanciu, G.-D.; Dodi, G.; Alexa-Stratulat, T.; Luca, A.; Leon-Constantin, M.-M.; Stefanescu, R.; Bild, V.; Melnic, S. Medicinal plants of the family Lamiaceae in pain therapy: A review. Pain Res. Manag. 2018, 2018, 7801543. [Google Scholar] [CrossRef]
  95. Kemassi, A.; Darem, S.; Cherif, R.; Boual, Z.; Sadine, S.E.; Aggoune, M.S.; Ould El Hadj-Khelil, A.; Ould El Hadj, M. Recherche et identification de quelques plantes médicinales à caractère hypoglycémiant de la pharmacopée traditionnelle des communautés de la vallée du M’Zab (Sahara septentrional Est Algérien). J. Adv. Res. Sci. Technol. 2014, 1, 1–5. [Google Scholar]
  96. Telli, A.; Esnault, M.-A.; Khelil, A.O.E.H. An ethnopharmacological survey of plants used in traditional diabetes treatment in south-eastern Algeria (Ouargla province). J. Arid Environ. 2016, 127, 82–92. [Google Scholar] [CrossRef]
  97. Mehdi, A.; Lamiae, B.; Samira, B.; Ramchoun, M.; Abdelouahed, K.; Tamas, F.; Hicham, B. Pomegranate (Punica granatum L.) Attenuates Neuroinflammation Involved in Neurodegenerative Diseases. Foods 2022, 11, 2570. [Google Scholar] [CrossRef]
  98. Belwal, T.; Nabavi, S.M.; Nabavi, S.F.; Dehpour, A.R.; Shirooie, S. Naturally Occurring Chemicals against Alzheimer’s Disease; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar]
  99. Fatma, B.; Fatiha, M.; Elattafia, B.; Noureddine, D. Phytochemical and antimicrobial study of the seeds and leaves of Peganum harmala L. against urinary tract infection pathogens. Asian Pac. J. Trop. Dis. 2016, 6, 822–826. [Google Scholar] [CrossRef]
  100. Belkhir, M.; Rebai, O.; Dhaouadi, K.; Sioud, B.; Amri, M.; Fattouch, S. Antioxidant and antimicrobial activities of Tunisian azarole (Crataegus azarolus L.) leaves and fruit pulp/peel polyphenolic extracts. Int. J. Food Prop. 2013, 16, 1380–1393. [Google Scholar] [CrossRef]
  101. Ampofo, O. Plants That Heal; World Health Organization: Geneva, Switzerland, 1977; pp. 26–30. [Google Scholar]
  102. El Omari, N.; Guaouguaou, F.E.; El Menyiy, N.; Benali, T.; Aanniz, T.; Chamkhi, I.; Balahbib, A.; Taha, D.; Shariati, M.A.; Zengin, G. Phytochemical and biological activities of Pinus halepensis mill., and their ethnomedicinal use. J. Ethnopharmacol. 2021, 268, 113661. [Google Scholar] [CrossRef]
  103. Mitra, S.; Anjum, J.; Muni, M.; Das, R.; Rauf, A.; Islam, F.; Emran, T.B.; Semwal, P.; Hemeg, H.A.; Alhumaydhi, F.A. Exploring the journey of emodin as a potential neuroprotective agent: Novel therapeutic insights with molecular mechanism of action. Biomed. Pharmacother. 2022, 149, 112877. [Google Scholar] [CrossRef]
  104. Chen, J.; Liu, X.; Li, Z.; Qi, A.; Yao, P.; Zhou, Z.; Dong, T.T.; Tsim, K.W. A review of dietary Ziziphus jujuba fruit (Jujube): Developing health food supplements for brain protection. Evid.-Based Complement. Altern. Med. 2017, 2017, 3019568. [Google Scholar] [CrossRef]
  105. Seong, N.-W.; Oh, W.-J.; Kim, I.-S.; Kim, S.-J.; Seo, J.-E.; Park, C.-E.; Kim, D.-Y.; Ko, J.-W.; Kim, J.-C. Efficacy and local irritation evaluation of Eriobotrya japonica leaf ethanol extract. Lab. Anim. Res. 2019, 35, 1–10. [Google Scholar] [CrossRef] [PubMed]
  106. Sagar, N.; Pareek, S.; Bhardwaj, R.; Vyas, N. Bioactive compounds of loquat (Eriobotrya japonica (Thunb.) L.). In Bioactive Compounds in Underutilized Fruits and Nuts; Springer: Cham, Switzerland, 2020; pp. 123–143. [Google Scholar]
  107. Szewczyk, A.; Grabowski, M.; Zych, D. Ruta chalepensis L. In Vitro Cultures as a Source of Bioactive Furanocoumarins and Furoquinoline Alkaloids. Life 2023, 13, 457. [Google Scholar] [CrossRef]
  108. Al-Said, M.S.; Tariq, M.; Al-Yahya, M.; Rafatullah, S.; Ginnawi, O.; Ageel, A. Studies on Ruta chalepensis, an ancient medicinal herb still used in traditional medicine. J. Ethnopharmacol. 1990, 28, 305–312. [Google Scholar] [CrossRef] [PubMed]
  109. Debbache, N.; Atmani, D.; Atmani, D. Chemical analysis and biological activities of Populus nigra, flower buds extracts as source of propolis in Algeria. Ind. Crops Prod. 2014, 53, 85–92. [Google Scholar] [CrossRef]
  110. Ali, M.; Alhazmi, H.A.; Ansari, S.; Hussain, A.; Ahmad, S.; Alam, M.S.; Ali, M.S.; El-Sharkawy, K.A.; Hakeem, K.R. Tamarix aphylla (L.) Karst. Phytochemical and bioactive profile compilations of less discussed but effective naturally growing Saudi plant. In Plant and Human Health, Volume 3: Pharmacology and Therapeutic Uses; Springer: Cham, Switzerland, 2019; pp. 343–352. [Google Scholar]
  111. Bano, S.; Sharif, A.; Akhtar, B.; Abdel-Daim, M.M.; Akhtar, M.F.; Ali, F.L. Mechanistic insights on the possible protective role of polyphenols extracted from Tamarix aphylla aerial parts against sodium arsenite-induced hepatotoxicity in rats. Environ. Sci. Pollut. Res. 2023, 30, 16565–16578. [Google Scholar] [CrossRef]
  112. Khan, A.W.; Khan, A.-u.; Ahmed, T. Anticonvulsant, anxiolytic, and sedative activities of Verbena officinalis. Front. Pharmacol. 2016, 7, 499. [Google Scholar] [CrossRef] [PubMed]
  113. Kubica, P.; Szopa, A.; Dominiak, J.; Luczkiewicz, M.; Ekiert, H. Verbena officinalis (Common Vervain)—A review on the investigations of this medicinally important plant species. Planta Med. 2020, 86, 1241–1257. [Google Scholar] [CrossRef]
  114. Bekara, A.; Amazouz, A.; Douma, T.B. Evaluating the antidepressant Effect of Verbena officinalis L. (Vervain) aqueous extract in adult rats. Basic Clin. Neurosci. 2020, 11, 91–98. [Google Scholar] [CrossRef]
  115. Zhao, H.; Ren, S.; Yang, H.; Tang, S.; Guo, C.; Liu, M.; Tao, Q.; Ming, T.; Xu, H. Peppermint essential oil: Its phytochemistry, biological activity, pharmacological effect and application. Biomed. Pharmacother. 2022, 154, 113559. [Google Scholar] [CrossRef]
  116. Abdelhalim, A.R. The effect of Mentha piperita L. on the mental health issues of university students: A pilot study. J. Pharm. Pharmacogn. Res. 2021, 9, 49–57. [Google Scholar] [CrossRef]
  117. Mahdavikian, S.; Rezaei, M.; Modarresi, M.; Khatony, A. Comparing the effect of aromatherapy with peppermint and lavender on the sleep quality of cardiac patients: A randomized controlled trial. Sleep Sci. Pract. 2020, 4, 1–8. [Google Scholar] [CrossRef]
  118. Gumisiriza, H.; Birungi, G.; Olet, E.A.; Sesaazi, C.D. Medicinal plant species used by local communities around queen elizabeth national park, maramagambo central forest reserve and ihimbo central forest reserve, south western Uganda. J. Ethnopharmacol. 2019, 239, 111926. [Google Scholar] [CrossRef]
  119. Tsobou, R.; Mapongmetsem, P.M.; Van Damme, P. Medicinal plants used for treating reproductive health care problems in Cameroon, Central Africa. Econ. Bot. 2016, 70, 145–159. [Google Scholar] [CrossRef]
  120. Das, G.; Shin, H.-S.; Tundis, R.; Gonçalves, S.; Tantengco, O.A.G.; Campos, M.G.; Acquaviva, R.; Malfa, G.A.; Romano, A.; Robles, J.A.H. Plant species of sub-family Valerianaceae—A review on its effect on the central nervous system. Plants 2021, 10, 846. [Google Scholar] [CrossRef]
  121. Ortiz, Y.T.; McMahon, L.R.; Wilkerson, J.L. Medicinal cannabis and central nervous system disorders. Front. Pharmacol. 2022, 13, 881810. [Google Scholar] [CrossRef]
  122. Perera, P.K.; Meedeniya, A.C.B.; Chamikara, N.H.A. Traditional Medicinal Plants of Sri Lanka and Their Derivatives of Benefit to the Nervous System. In Medicinal Herbs and Fungi Neurotoxicity vs. Neuroprotection; Springer: Singapore, 2021; pp. 315–346. [Google Scholar]
  123. Morris, G.; Gamage, E.; Travica, N.; Berk, M.; Jacka, F.N.; O’Neil, A.; Puri, B.K.; Carvalho, A.F.; Bortolasci, C.C.; Walder, K. Polyphenols as adjunctive treatments in psychiatric and neurodegenerative disorders: Efficacy, mechanisms of action, and factors influencing inter-individual response. Free Radic. Biol. Med. 2021, 172, 101–122. [Google Scholar] [CrossRef]
  124. Weston-Green, K.; Clunas, H.; Jimenez Naranjo, C. A review of the potential use of pinene and linalool as terpene-based medicines for brain health: Discovering novel therapeutics in the flavours and fragrances of cannabis. Front. Psychiatry 2021, 12, 583211. [Google Scholar] [CrossRef]
  125. Trebaticka, J.; Ďuračková, Z. Psychiatric disorders and polyphenols: Can they be helpful in therapy? Oxidative Med. Cell. Longev. 2015, 2015, 248529. [Google Scholar] [CrossRef] [PubMed]
  126. Balasundram, N.; Sundram, K.; Samman, S. Phenolic compounds in plants and agri-industrial by-products: Antioxidant activity, occurrence, and potential uses. Food Chem. 2006, 99, 191–203. [Google Scholar] [CrossRef]
  127. Robards, K.; Prenzler, P.D.; Tucker, G.; Swatsitang, P.; Glover, W. Phenolic compounds and their role in oxidative processes in fruits. Food Chem. 1999, 66, 401–436. [Google Scholar] [CrossRef]
  128. Li, Z.; Zhao, T.; Shi, M.; Wei, Y.; Huang, X.; Shen, J.; Zhang, X.; Xie, Z.; Huang, P.; Yuan, K. Polyphenols: Natural food grade biomolecules for treating neurodegenerative diseases from a multi-target perspective. Front. Nutr. 2023, 10, 1139558. [Google Scholar] [CrossRef]
  129. Hussain, T.; Tan, B.; Yin, Y.; Blachier, F.; Tossou, M.C.; Rahu, N. Oxidative stress and inflammation: What polyphenols can do for us? Oxidative Med. Cell. Longev. 2016, 2016, 7432797. [Google Scholar] [CrossRef]
  130. Cháirez-Ramírez, M.H.; de la Cruz-López, K.G.; García-Carrancá, A. Polyphenols as antitumor agents targeting key players in cancer-driving signaling pathways. Front. Pharmacol. 2021, 12, 710304. [Google Scholar] [CrossRef]
  131. Rahman, M.M.; Rahaman, M.S.; Islam, M.R.; Rahman, F.; Mithi, F.M.; Alqahtani, T.; Almikhlafi, M.A.; Alghamdi, S.Q.; Alruwaili, A.S.; Hossain, M.S. Role of phenolic compounds in human disease: Current knowledge and future prospects. Molecules 2021, 27, 233. [Google Scholar] [CrossRef]
  132. Zhang, X.-L.; Guo, Y.-S.; Wang, C.-H.; Li, G.-Q.; Xu, J.-J.; Chung, H.Y.; Ye, W.-C.; Li, Y.-L.; Wang, G.-C. Phenolic compounds from Origanum vulgare and their antioxidant and antiviral activities. Food Chem. 2014, 152, 300–306. [Google Scholar] [CrossRef]
  133. Gutiérrez-Grijalva, E.; Picos-Salas, M.; Leyva-López, N.; Criollo-Mendoza, M.; Vazquez-Olivo, G.; Heredia, J.J.P. Flavonoids and phenolic acids from oregano: Occurrence, biological activity and health benefits. Plants 2018, 7, 2. [Google Scholar] [CrossRef] [PubMed]
  134. Calis, Z.; Mogulkoc, R.; Baltaci, A.K. The roles of flavonols/flavonoids in neurodegeneration and neuroinflammation. Mini Rev. Med. Chem. 2020, 20, 1475–1488. [Google Scholar] [CrossRef]
  135. Macready, A.L.; Kennedy, O.B.; Ellis, J.A.; Williams, C.M.; Spencer, J.P.; Butler, L.T. Flavonoids and cognitive function: A review of human randomized controlled trial studies and recommendations for future studies. Genes Nutr. 2009, 4, 227–242. [Google Scholar] [CrossRef] [PubMed]
  136. Socci, V.; Tempesta, D.; Desideri, G.; De Gennaro, L.; Ferrara, M. Enhancing human cognition with cocoa flavonoids. Front. Nutr. 2017, 4, 19. [Google Scholar] [CrossRef]
  137. Bakoyiannis, I.; Daskalopoulou, A.; Pergialiotis, V.; Perrea, D. Phytochemicals and cognitive health: Are flavonoids doing the trick? Biomed. Pharmacother. 2019, 109, 1488–1497. [Google Scholar] [CrossRef] [PubMed]
  138. Spencer, J.P. The impact of fruit flavonoids on memory and cognition. Br. J. Nutr. 2010, 104, S40–S47. [Google Scholar] [CrossRef] [PubMed]
  139. Sharifi-Rad, J.; Sureda, A.; Tenore, G.C.; Daglia, M.; Sharifi-Rad, M.; Valussi, M.; Tundis, R.; Sharifi-Rad, M.; Loizzo, M.R.; Ademiluyi, A.O. Biological activities of essential oils: From plant chemoecology to traditional healing systems. Molecules 2017, 22, 70. [Google Scholar] [CrossRef] [PubMed]
  140. Sattayakhom, A.; Wichit, S.; Koomhin, P. The Effects of Essential Oils on the Nervous System: A Scoping Review. Molecules 2023, 28, 3771. [Google Scholar] [CrossRef]
  141. Pai, S.R.; Sonkamble, V.V.; Wagh, N.S. Essential oils as effective agents against neurological disorders. In Plant-Derived Bioactives Production, Properties and Therapeutic Applications; Springer: Singapore, 2020; pp. 409–433. [Google Scholar]
  142. Lehrner, J.; Marwinski, G.; Lehr, S.; Johren, P.; Deecke, L. Ambient odors of orange and lavender reduce anxiety and improve mood in a dental office. Physiol. Behav. 2005, 86, 92–95. [Google Scholar] [CrossRef]
  143. Moss, M.; Cook, J.; Wesnes, K.; Duckett, P. Aromas of rosemary and lavender essential oils differentially affect cognition and mood in healthy adults. Int. J. Neurosci. 2003, 113, 15–38. [Google Scholar] [CrossRef]
  144. Bikmoradi, A.; Seifi, Z.; Poorolajal, J.; Araghchian, M.; Safiaryan, R.; Oshvandi, K. Effect of inhalation aromatherapy with lavender essential oil on stress and vital signs in patients undergoing coronary artery bypass surgery: A single-blinded randomized clinical trial. Complement. Ther. Med. 2015, 23, 331–338. [Google Scholar] [CrossRef] [PubMed]
  145. Lundstrom, K.; Pham, H.T.; Dinh, L.D. Interaction of plant extracts with central nervous system receptors. Medicines 2017, 4, 12. [Google Scholar] [CrossRef]
  146. Du Sert, N.P.; Ahluwalia, A.; Alam, S.; Avey, M.T.; Baker, M.; Browne, W.J.; Clark, A.; Cuthill, I.C.; Dirnagl, U.; Emerson, M. Reporting animal research: Explanation and elaboration for the ARRIVE guidelines 2.0. PLoS Biol. 2020, 18, e30004112020. [Google Scholar]
  147. Prior, H.; Haworth, R.; Labram, B.; Roberts, R.; Wolfreys, A.; Sewell, F. Justification for species selection for pharmaceutical toxicity studies. Toxicol. Res. 2020, 9, 758–770. [Google Scholar] [CrossRef] [PubMed]
  148. Salehi, B.; Venditti, A.; Sharifi-Rad, M.; Kręgiel, D.; Sharifi-Rad, J.; Durazzo, A.; Lucarini, M.; Santini, A.; Souto, E.B.; Novellino, E. The therapeutic potential of apigenin. Int. J. Mol. Sci. 2019, 20, 1305. [Google Scholar] [CrossRef] [PubMed]
  149. Avallone, R.; Zanoli, P.; Corsi, L.; Cannazza, G.; Baraldi, M. Benzodiazepine-like compounds and GABA in flower heads of Matricaria chamomilla. Phytother. Res. 1996, 10, 177–179. [Google Scholar]
  150. Bisht, D.; Kumar, D.; Kumar, D.; Dua, K.; Chellappan, D.K. Phytochemistry and pharmacological activity of the genus artemisia. Arch. Pharmacal. Res. 2021, 44, 439–474. [Google Scholar] [CrossRef] [PubMed]
  151. Zámboriné Németh, É.; Thi Nguyen, H. Thujone, a widely debated volatile compound: What do we know about it? Phytochem. Rev. 2020, 19, 405–423. [Google Scholar] [CrossRef]
  152. Rivera, E.; Cid, M.P.; Zunino, P.; Baiardi, G.; Salvatierra, N.A. Central α-and β-thujone: Similar anxiogenic-like effects and differential modulation on GABAA receptors in neonatal chicks. Brain Res. 2014, 1555, 28–35. [Google Scholar] [CrossRef] [PubMed]
  153. Bent, S.; Padula, A.; Moore, D.; Patterson, M.; Mehling, W. Valerian for sleep: A systematic review and meta-analysis. Am. J. Med. 2006, 119, 1005–1012. [Google Scholar] [CrossRef]
  154. Khalfa, H.; Rebbas, K.; Miara, M.; Bendif, H.; Boufissiou, A.; Souilah, N.; Daoud, N.; Peroni, A. Diversity and Traditional Use Value of Medicinal Plants in Bou Saada District of M’Sila Province, South East Algeria. J. Biodivers. Conserv. Bioresour. Manag. 2022, 8, 61–78. [Google Scholar] [CrossRef]
  155. Valle-Mojica, D.; Lisa, M.; Cordero-Hernández, J.M.; González-Medina, G.; Ramos-Vélez, I.; Berríos-Cartagena, N.; Torres-Hernández, B.A.; Ortíz, J.G. Aqueous and ethanolic Valeriana officinalis extracts change the binding of ligands to glutamate receptors. Evid.-Based Complement. Altern. Med. 2011, 2011, 891819. [Google Scholar]
  156. Becker, A.; Felgentreff, F.; Schröder, H.; Meier, B.; Brattström, A. The anxiolytic effects of a Valerian extract is based on valerenic acid. BMC Complement. Altern. Med. 2014, 14, 1–5. [Google Scholar] [CrossRef]
  157. Singh, B.; Sharma, R.A. Secondary Metabolites of Medicinal Plants, 4 Volume Set: Ethnopharmacological Properties, Biological Activity and Production Strategies; John Wiley & Sons: Hoboken, NJ, USA, 2020. [Google Scholar]
  158. Valle-Mojica, D.; Lisa, M.; Ayala-Marín, Y.M.; Ortiz-Sanchez, C.M.; Torres-Hernández, B.A.; Abdalla-Mukhaimer, S.; Ortiz, J.G. Selective interactions of Valeriana officinalis extracts and valerenic acid with [3H] glutamate binding to rat synaptic membranes. Evid.-Based Complement. Altern. Med. 2011, 2011, 403591. [Google Scholar]
  159. Wang, W.; Wang, Y.; Guo, Q.; Li, H.; Wang, Z.; Li, J.; Li, T.; Tang, T.; Wang, Y.; Jia, Y. Valerian essential oil for treating insomnia via the serotonergic synapse pathway. Front. Nutr. 2022, 9, 927434. [Google Scholar] [CrossRef] [PubMed]
  160. Simmen, U.; Higelin, J.; Berger-Büter, K.; Schaffner, W.; Lundstrom, K. Neurochemical studies with St. John’s wort in vitro. Pharmacopsychiatry 2001, 34, 137–142. [Google Scholar] [CrossRef]
  161. Simmen, U.; Burkard, W.; Berger, K.; Schaffner, W.; Lundstrom, K. Extracts and constituents of Hypericum perforatum inhibit the binding of various ligands to recombinant receptors expressed with the Semliki Forest virus system. J. Recept. Signal Transduct. 1999, 19, 59–74. [Google Scholar] [CrossRef]
  162. Butterweck, V.; Nahrstedt, A.; Evans, J.; Hufeisen, S.; Rauser, L.; Savage, J.; Popadak, B.; Ernsberger, P.; Roth, B.L. In vitro receptor screening of pure constituents of St. John’s wort reveals novel interactions with a number of GPCRs. Psychopharmacology 2002, 162, 193–202. [Google Scholar] [CrossRef]
  163. Caldeira, G.I.; Gouveia, L.P.; Serrano, R.; Silva, O.D. Hypericum Genus as a Natural Source for Biologically Active Compounds. Plants 2022, 11, 2509. [Google Scholar] [CrossRef]
  164. Marrelli, M.; Statti, G.; Conforti, F.; Menichini, F. New potential pharmaceutical applications of hypericum species. Mini Rev. Med. Chem. 2016, 16, 710–720. [Google Scholar] [CrossRef]
  165. Wei, Y.; Wang, Q.; Huang, Y. Species diversity and distribution of Salvia (Lamiaceae). Biodivers. Sci. 2015, 23, 3. [Google Scholar]
  166. Petrisor, G.; Motelica, L.; Craciun, L.N.; Oprea, O.C.; Ficai, D.; Ficai, A. Melissa officinalis: Composition, pharmacological effects and derived release systems—A review. Int. J. Mol. Sci. 2022, 23, 3591. [Google Scholar] [CrossRef]
  167. Porres-Martínez, M.; González-Burgos, E.; Emilia Carretero, M.; Pilar Gómez-Serranillos, M. Salvia spp.: An Updated on Antioxidant Activity and Pharmacological Uses. In Salvia Biotechnology; Springer: Cham, Switzerland, 2017; pp. 151–177. [Google Scholar]
  168. Kabouche, A.; Kabouche, Z.; Öztürk, M.; Kolak, U.; Topçu, G. Antioxidant abietane diterpenoids from Salvia barrelieri. Food Chem. 2007, 102, 1281–1287. [Google Scholar] [CrossRef]
  169. Kabouche, A.; Boutaghane, N.; Kabouche, Z.; Seguin, E.; Tillequin, F.; Benlabed, K. Components and antibacterial activity of the roots of Salvia jaminiana. Fitoterapia 2005, 76, 450–452. [Google Scholar] [CrossRef] [PubMed]
  170. Sharifi-Rad, M.; Ozcelik, B.; Altın, G.; Daşkaya-Dikmen, C.; Martorell, M.; Ramírez-Alarcón, K.; Alarcón-Zapata, P.; Morais-Braga, M.F.B.; Carneiro, J.N.; Leal, A.L.A.B. Salvia spp. plants-from farm to food applications and phytopharmacotherapy. Trends Food Sci. Technol. 2018, 80, 242–263. [Google Scholar] [CrossRef]
  171. Quézel, P.; Santa, S. Nouvelle Flore de l’Algérie et des Régions Désertiques Méridionales; Éditions du Centre National de la Recherche Scientifique: Paris, France, 1962. [Google Scholar]
  172. Ghorbani, A.; Esmaeilizadeh, M. Pharmacological properties of Salvia officinalis and its components. J. Tradit. Complement. Med. 2017, 7, 433–440. [Google Scholar] [CrossRef]
  173. Boussadia, A.; Kharoubi, O.; Lahouel, Z.; Benglia, A.; Aoues, A. Effect of aqueous Salvia officinalis extract on Aluminum chloride-induced neurotoxicity in female rats. Int. J. Pharm. Res. Allied Sci. 2020, 9, 139–150. [Google Scholar]
  174. Chouit, H.; Touafek, O.; Brada, M.; Benssouici, C.; Fauconnier, M.-L.; Hattab, M.E. GC-MS analysis and biological activities of Algerian Salvia microphylla essential oils. J. Mex. Chem. Soc. 2021, 65, 582–601. [Google Scholar] [CrossRef]
  175. Sharifi-Rad, J.; Quispe, C.; Herrera-Bravo, J.; Akram, M.; Abbaass, W.; Semwal, P.; Painuli, S.; Konovalov, D.A.; Alfred, M.A.; Kumar, N.V.A. Phytochemical constituents, biological activities, and health-promoting effects of the Melissa officinalis. Oxidative Med. Cell. Longev. 2021, 2021, 1–20. [Google Scholar] [CrossRef]
  176. Wake, G.; Court, J.; Pickering, A.; Lewis, R.; Wilkins, R.; Perry, E. CNS acetylcholine receptor activity in European medicinal plants traditionally used to improve failing memory. J. Ethnopharmacol. 2000, 69, 105–114. [Google Scholar] [CrossRef]
  177. Iranshahi, M.; Javadi, B. Neurological and Neuroprotective effects of Melissa officinalis L. Navid No 2019, 22, 60–71. [Google Scholar]
  178. Capatina, L.; Todirascu-Ciornea, E.; Napoli, E.M.; Ruberto, G.; Hritcu, L.; Dumitru, G. Thymus vulgaris essential oil protects zebrafish against cognitive dysfunction by regulating cholinergic and antioxidants systems. Antioxidants 2020, 9, 1083. [Google Scholar] [CrossRef] [PubMed]
  179. Javed, H.; Azimullah, S.; Meeran, M.N.; Ansari, S.A.; Ojha, S. Neuroprotective effects of thymol, a dietary monoterpene against dopaminergic neurodegeneration in rotenone-induced rat model of Parkinson’s disease. Int. J. Mol. Sci. 2019, 20, 1538. [Google Scholar] [CrossRef] [PubMed]
  180. Deng, X.-Y.; Li, H.-Y.; Chen, J.-J.; Li, R.-P.; Qu, R.; Fu, Q.; Ma, S.-P. Thymol produces an antidepressant-like effect in a chronic unpredictable mild stress model of depression in mice. Behav. Brain Res. 2015, 291, 12–19. [Google Scholar] [CrossRef] [PubMed]
  181. Azizi, A.; Yan, F.; Honermeier, B. Herbage yield, essential oil content and composition of three oregano (Origanum vulgare L.) populations as affected by soil moisture regimes and nitrogen supply. Ind. Crops Prod. 2009, 29, 554–561. [Google Scholar] [CrossRef]
  182. Lee, B.H.; Nam, T.G.; Park, W.J.; Kang, H.; Heo, H.J.; Chung, D.K.; Kim, G.H.; Kim, D.-O. Antioxidative and neuroprotective effects of volatile components in essential oils from Chrysanthemum indicum Linné flowers. Food Sci. 2015, 24, 717–723. [Google Scholar] [CrossRef]
  183. Sammi, S.R.; Trivedi, S.; Rath, S.K.; Nagar, A.; Tandon, S.; Kalra, A.; Pandey, R. 1-Methyl-4-propan-2-ylbenzene from Thymus vulgaris attenuates cholinergic dysfunction. Mol. Neurobiol. 2017, 54, 5468–5481. [Google Scholar] [CrossRef]
  184. Zotti, M.; Colaianna, M.; Morgese, M.G.; Tucci, P.; Schiavone, S.; Avato, P.; Trabace, L. Carvacrol: From Ancient Flavoring to Neuromodulatory Agent. Molecules 2013, 18, 6161. [Google Scholar] [CrossRef]
  185. Peters, M.; Trembovler, V.; Alexandrovich, A.; Parnas, M.; Birnbaumer, L.; Minke, B.; Shohami, E. Carvacrol together with TRPC1 elimination improve functional recovery after traumatic brain injury in mice. J. Neurotrauma 2012, 29, 2831–2834. [Google Scholar] [CrossRef]
  186. Ghannadi, A.; Sajjadi, S.E.; Kabouche, A. Thymus fontanesii Boiss. & Reut.-A potential source of thymol-rich essential oil in North Africa. Z. Für Nat. C 2004, 59, 187–189. [Google Scholar]
  187. Kabouche, A.; Kabouche, Z.; Bruneau, C. Analysis of the essential oil of Thymus numidicus (Poiret) from Algeria. Flavour Fragr. J. 2005, 20, 235–236. [Google Scholar] [CrossRef]
  188. Rahbardar, M.G.; Hosseinzadeh, H. Therapeutic effects of rosemary (Rosmarinus officinalis L.) and its active constituents on nervous system disorders. Iran. J. Basic Med. Sci. 2020, 23, 1100. [Google Scholar]
  189. Enwright, P.; Blank, S.; Wells, B.M.; Nightingale, L.M.; Torgerud, S. Effect of lavender and rosemary aromatherapy on test anxiety in chiropractic students. J. Chiropr. Educ. 2023, 37, 26–32. [Google Scholar] [CrossRef]
  190. Coelho, V.R.; Vieira, C.G.; de Souza, L.P.; Moysés, F.; Basso, C.; Papke, D.K.M.; Pires, T.R.; Siqueira, I.R.; Picada, J.N.; Pereira, P. Antiepileptogenic, antioxidant and genotoxic evaluation of rosmarinic acid and its metabolite caffeic acid in mice. Life Sci. 2015, 122, 65–71. [Google Scholar] [CrossRef] [PubMed]
  191. Kondo, S.; El Omri, A.; Han, J.; Isoda, H. Antidepressant-like effects of rosmarinic acid through mitogen-activated protein kinase phosphatase-1 and brain-derived neurotrophic factor modulation. J. Funct. Foods 2015, 14, 758–766. [Google Scholar] [CrossRef]
  192. Shan, Y.; Wang, D.-D.; Xu, Y.-X.; Wang, C.; Cao, L.; Liu, Y.-S.; Zhu, C.-Q. Aging as a precipitating factor in chronic restraint stress-induced tau aggregation pathology, and the protective effects of rosmarinic acid. J. Alzheimer’s Dis. 2016, 49, 829–844. [Google Scholar] [CrossRef]
  193. Crișan, I.; Ona, A.; Vârban, D.; Muntean, L.; Vârban, R.; Stoie, A.; Mihăiescu, T.; Morea, A. Current trends for lavender (Lavandula angustifolia Mill.) crops and products with emphasis on essential oil quality. Plants 2023, 12, 357. [Google Scholar] [CrossRef]
  194. Batiha, G.E.-S.; Teibo, J.O.; Wasef, L.; Shaheen, H.M.; Akomolafe, A.P.; Teibo, T.K.A.; Al-Kuraishy, H.M.; Al-Garbeeb, A.I.; Alexiou, A.; Papadakis, M. A review of the bioactive components and pharmacological properties of Lavandula species. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2023, 396, 877–900. [Google Scholar] [CrossRef] [PubMed]
  195. Linck, V.M.; da Silva, A.L.; Figueiro, M.; Caramao, E.B.; Moreno, P.R.H.; Elisabetsky, E. Effects of inhaled Linalool in anxiety, social interaction and aggressive behavior in mice. Phytomedicine 2010, 17, 679–683. [Google Scholar] [CrossRef]
  196. Harada, H.; Kashiwadani, H.; Kanmura, Y.; Kuwaki, T. Linalool odor-induced anxiolytic effects in mice. Front. Behav. Neurosci. 2018, 12, 241. [Google Scholar] [CrossRef]
  197. Boughendjioua, H. Composition chimique et activité antibactérienne de l’huile essentielle de Lavandula officinalis cultivées dans la région de Skikda-Algérie. Chemical composition and antibacterial activity of essential oil of Lavandula officinalis grown in the region of Skikda-Algeria. Bull. Soc. R. Sci. Liège 2017, 86, 88–95. [Google Scholar]
  198. Ferhat, M.; Erol, E.; Beladjila, K.A.; Çetintaş, Y.; Duru, M.E.; Öztürk, M.; Kabouche, A.; Kabouche, Z. Antioxidant, anticholinesterase and antibacterial activities of Stachys guyoniana and Mentha aquatica. Pharm. Biol. 2017, 55, 324–329. [Google Scholar] [CrossRef]
  199. Bahadori, M.B.; Maggi, F.; Zengin, G.; Asghari, B.; Eskandani, M. Essential oils of hedgenettles (Stachys inflata, S. lavandulifolia, and S. byzantina) have antioxidant, anti-Alzheimer, antidiabetic, and anti-obesity potential: A comparative study. Ind. Crops Prod. 2020, 145, 112089. [Google Scholar] [CrossRef]
  200. Rabbani, M.; Sajjadi, S.; Zarei, H. Anxiolytic effects of Stachys lavandulifolia Vahl on the elevated plus-maze model of anxiety in mice. J. Ethnopharmacol. 2003, 89, 271–276. [Google Scholar] [CrossRef]
  201. Laggoune, S.; Kabouche, A.; Kabouche, Z.; El-Azzouny, A. Analysis of the Essential Oil of Stachys circinnata l’Her. from Algeria. J. Essential Oil Res. 2009, 21, 67–68. [Google Scholar] [CrossRef]
  202. Mamadalieva, N.; Ashour, M.; Mamedov, N. Peganum harmala: Phytochemistry, traditional uses, and biological activities. In Biodiversity, Conservation and Sustainability in Asia: Volume 2: Prospects and Challenges in South and Middle Asia; Springer: Cham, Switzerland, 2022; pp. 721–744. [Google Scholar]
  203. Li, L.-N. Peganum harmala L.: A Review of Botany, Traditional Use, Phytochemistry, Pharmacology, Quality marker, and Toxicity. Comb. Chem. HighThroughput Screen. 2023; e-pub ahead of print. [Google Scholar]
  204. Shahrajabian, M.H.; Sun, W.; Cheng, Q. Improving health benefits with considering traditional and modern health benefits of Peganum harmala. Clin. Phytosci. 2021, 7, 1–9. [Google Scholar] [CrossRef]
  205. Filban, F.; Ravanbakhsh, M.; Poormohammadi, A.; Khaghani, S.; Sadeghi-Nejad, B.; Neisi, A.; Goudarzi, G. Antimicrobial properties of Peganum harmala L. seeds’ smoke in indoors: Applications and prospects. Environ. Monit. Assess. 2022, 194, 1–13. [Google Scholar] [CrossRef]
  206. Apostolico, I.; Aliberti, L.; Caputo, L.; De Feo, V.; Fratianni, F.; Nazzaro, F.; Souza, L.F.; Khadhr, M. Chemical composition, antibacterial and phytotoxic activities of Peganum harmala seed essential oils from five different localities in Northern Africa. Molecules 2016, 21, 1235. [Google Scholar] [CrossRef] [PubMed]
  207. Ayipo, Y.O.; Mordi, M.N.; Mustapha, M.; Damodaran, T. Neuropharmacological potentials of β-carboline alkaloids for neuropsychiatric disorders. Eur. J. Pharmacol. 2021, 893, 173837. [Google Scholar] [CrossRef]
  208. Shatarat, A.T.; Abuhamdah, S.; Alefishat, E.; Al-Essa, M.K.; Altaweel, R.; Mohammed, F.; Badran, D.; Jafar, H. Effects of beta-carboline alkaloids of Peganum harmala on induced rat ileum contractions. Pharmacogn. J. 2020, 12, 260–265. [Google Scholar] [CrossRef]
  209. Herraiz, T.; González, D.; Ancín-Azpilicueta, C.; Arán, V.J.; Guillén, H. β-Carboline alkaloids in Peganum harmala and inhibition of human monoamine oxidase (MAO). Food Chem. Toxicol. 2010, 48, 839–845. [Google Scholar] [CrossRef]
  210. Nasehi, M.; Hasanvand, S.; Khakpai, F.; Zarrindast, M.-R. The effect of CA1 dopaminergic system on amnesia induced by harmane in mice. Acta Neurol. Belg. 2019, 119, 369–377. [Google Scholar] [CrossRef]
  211. Saleh, R.A.; Eissa, T.F.; Abdallah, D.M.; Saad, M.A.; El-Abhar, H.S. Peganum harmala enhanced GLP-1 and restored insulin signaling to alleviate AlCl3-induced Alzheimer-like pathology model. Sci. Rep. 2021, 11, 12040. [Google Scholar] [CrossRef] [PubMed]
  212. Cerqueira-Silva, C.B.M.; Jesus, O.N.; Santos, E.S.; Corrêa, R.X.; Souza, A.P. Genetic breeding and diversity of the genus Passiflora: Progress and perspectives in molecular and genetic studies. Int. J. Mol. Sci. 2014, 15, 14122–14152. [Google Scholar] [CrossRef]
  213. Castañeda, R.; Cáceres, A.; Velásquez, D.; Rodríguez, C.; Morales, D.; Castillo, A. Medicinal plants used in traditional Mayan medicine for the treatment of central nervous system disorders: An overview. J. Ethnopharmacol. 2022, 283, 114746. [Google Scholar] [CrossRef] [PubMed]
  214. Souilah, N.; Miara, M.D.; Bendif, H.; Medjroubi, K.; Snorek, J. Traditional Ethnobotanical Knowledge on Medicinal Plants Used by the Populations in Central Russikada (Northeastern Algeria). J. Herbs Spices Med. Plants 2022, 28, 15–35. [Google Scholar] [CrossRef]
  215. He, X.; Luan, F.; Yang, Y.; Wang, Z.; Zhao, Z.; Fang, J.; Wang, M.; Zuo, M.; Li, Y. Passiflora edulis: An insight into current researches on phytochemistry and pharmacology. Front. Pharmacol. 2020, 11, 617. [Google Scholar] [CrossRef] [PubMed]
  216. Akhondzadeh, S.; Naghavi, H.; Vazirian, M.; Shayeganpour, A.; Rashidi, H.; Khani, M. Passionflower in the treatment of generalized anxiety: A pilot double-blind randomized controlled trial with oxazepam. J. Clin. Pharm. Ther. 2001, 26, 363–367. [Google Scholar] [CrossRef] [PubMed]
  217. Akhondzadeh, S.; Kashani, L.; Mobaseri, M.; Hosseini, S.; Nikzad, S.; Khani, M. Passionflower in the treatment of opiates withdrawal: A double-blind randomized controlled trial. J. Clin. Pharm. Ther. 2001, 26, 369–373. [Google Scholar] [CrossRef]
  218. Reginatto, F.H.; De-Paris, F.; Petry, R.D.; Quevedo, J.; Ortega, G.G.; Gosmann, G.; Schenkel, E.P. Evaluation of anxiolytic activity of spray dried powders of two South Brazilian Passiflora species. Phytother. Res. Int. J. Devoted Pharmacol. Toxicol. Eval. Nat. Prod. Deriv. 2006, 20, 348–351. [Google Scholar]
  219. Wheatley, D. Medicinal plants for insomnia: A review of their pharmacology, efficacy and tolerability. J. Psychopharmacol. 2005, 19, 414–421. [Google Scholar] [CrossRef]
  220. Singh, B.; Singh, D.; Goel, R.K. Dual protective effect of Passiflora incarnata in epilepsy and associated postictal depression. J. Ethnopharmacol. 2012, 139, 273–279. [Google Scholar] [CrossRef]
  221. Ożarowski, M.; Karpiński, T.M. Extracts and flavonoids of Passiflora species as promising anti-inflammatory and antioxidant substances. Curr. Pharm. Des. 2021, 27, 2582–2604. [Google Scholar] [CrossRef] [PubMed]
  222. Frye, A.; Haustein, C. Extraction, identification, and quantification of harmala alkaloids in three species of Passiflora. Am. J. Undergrad. Res. 2007, 6, 19–26. [Google Scholar] [CrossRef]
  223. Fonseca, L.R.d.; Rodrigues, R.d.A.; Ramos, A.d.S.; da Cruz, J.D.; Ferreira, J.L.P.; Silva, J.R.d.A.; Amaral, A.C.F. Herbal medicinal products from Passiflora for anxiety: An unexploited potential. Sci. World J. 2020, 2020, 6598434. [Google Scholar] [CrossRef] [PubMed]
  224. Patel, S.; Verma, N.; Gauthaman, K. Passiflora incarnata Linn: A review on morphology, phytochemistry and pharmacological aspects. Pharmacogn. Rev. 2009, 3, 186. [Google Scholar]
  225. Dhawan, K.; Kumar, S.; Sharma, A. Suppression of alcohol-cessation-oriented hyper-anxiety by the benzoflavone moiety of Passiflora incarnata Linneaus in mice. J. Ethnopharmacol. 2002, 81, 239–244. [Google Scholar] [CrossRef] [PubMed]
  226. Brown, E.; Hurd, N.S.; McCall, S.; Ceremuga, T.E. Evaluation of the anxiolytic effects of chrysin, a Passiflora incarnata extract, in the laboratory rat. AANA J. 2007, 75, 333–337. [Google Scholar]
  227. Bouzabata, A.; Boukhari, A. Variation in the Traditional Knowledge of Curcuma longa L. in North-Eastern Algeria. Int. J. Pharmacol. Pharm. Sci. 2014, 8, 1227–1231. [Google Scholar]
  228. El-Saadony, M.T.; Yang, T.; Korma, S.A.; Sitohy, M.; El-Mageed, A.; Taia, A.; Selim, S.; Al Jaouni, S.K.; Salem, H.M.; Mahmmod, Y. Impacts of turmeric and its principal bioactive curcumin on human health: Pharmaceutical, medicinal, and food applications: A comprehensive review. Front. Nutr. 2023, 9, 1040259. [Google Scholar] [CrossRef]
  229. Farooqui, A.A. Curcumin in Neurological Disorders. In Curcumin for Neurological and Psychiatric Disorders; Elsevier: Amsterdam, The Netherlands, 2019; pp. 45–62. [Google Scholar]
  230. Fan, F.; Lei, M. Mechanisms underlying curcumin-induced neuroprotection in cerebral ischemia. Front. Pharmacol. 2022, 13, 893118. [Google Scholar] [CrossRef]
  231. Ramaholimihaso, T.; Bouazzaoui, F.; Kaladjian, A. Curcumin in depression: Potential mechanisms of action and current evidence—A narrative review. Front. Psychiatry 2020, 11, 572533. [Google Scholar] [CrossRef]
  232. den Haan, J.; Morrema, T.H.; Rozemuller, A.J.; Bouwman, F.H.; Hoozemans, J.J. Different curcumin forms selectively bind fibrillar amyloid beta in post mortem Alzheimer’s disease brains: Implications for in-vivo diagnostics. Acta Neuropathol. Commun. 2018, 6, 1–12. [Google Scholar] [CrossRef]
  233. Tang, M.; Taghibiglou, C. The mechanisms of action of curcumin in Alzheimer’s disease. J. Alzheimer’s Dis. 2017, 58, 1003–1016. [Google Scholar] [CrossRef]
  234. Abate, L.; Tadesse, M.G.; Bachheti, A.; Bachheti, R.K. Traditional and phytochemical bases of herbs, shrubs, climbers, and trees from Ethiopia for their anticancer response. BioMed Res. Int. 2022, 2022, 1589877. [Google Scholar] [CrossRef]
  235. Malík, M.; Tlustoš, P. Nootropic Herbs, Shrubs, and Trees as Potential Cognitive Enhancers. Plants 2023, 12, 1364. [Google Scholar] [CrossRef]
  236. Häne, K.; Dobbertin, M.K. Le frêne, arbre aux mille vertus. La Forêt 2006, 59, 20–21. [Google Scholar]
  237. Beloued, A. Plantes médicinales d’Algérie; Office des Publications Universitaires: Algiers, Algeria, 1998; 277p. [Google Scholar]
  238. Kis, B.; Avram, S.; Pavel, I.Z.; Lombrea, A.; Buda, V.; Dehelean, C.A.; Soica, C.; Yerer, M.B.; Bojin, F.; Folescu, R. Recent advances regarding the phytochemical and therapeutic uses of Populus nigra L. buds. Plants 2020, 9, 1464. [Google Scholar] [CrossRef]
  239. Zouaghi, N.; Houda Bensiradj, N.E.; Cavaleiro, C.; Nadjemi, B.; Trari, M. Phytochemical Study and Antibacterial Effects of Fraxinus angustifolia Vahl (Algeria): Experimental and Computational Investigations. Waste Biomass Valorizat. 2021, 12, 3605–3616. [Google Scholar] [CrossRef]
  240. Bouguellid, G.; Russo, C.; Lavorgna, M.; Piscitelli, C.; Ayouni, K.; Wilson, E.; Kim, H.K.; Verpoorte, R.; Choi, Y.H.; Kilani-Atmani, D. Antimutagenic, antigenotoxic and antiproliferative activities of Fraxinus angustifolia Vahl. leaves and stem bark extracts and their phytochemical composition. PLoS ONE 2020, 15, e02306902020. [Google Scholar] [CrossRef]
  241. Sarfraz, I.; Rasul, A.; Jabeen, F.; Younis, T.; Zahoor, M.K.; Arshad, M.; Ali, M. Fraxinus: A plant with versatile pharmacological and biological activities. Evid.-Based Complement. Altern. Med. 2017, 2017, 4269868. [Google Scholar] [CrossRef]
  242. Wang, Z.; Zhu, W.; Liu, H.; Wu, G.; Song, M.; Yang, B.; Yang, D.; Wang, Q.; Kuang, H. Simultaneous determination of aesculin, aesculetin, fraxetin, fraxin and polydatin in beagle dog plasma by UPLC-ESI-MS/MS and its application in a pharmacokinetic study after oral administration extracts of Ledum palustre L. Molecules 2018, 23, 2285. [Google Scholar] [CrossRef]
  243. Liang, C.; Ju, W.; Pei, S.; Tang, Y.; Xiao, Y. Pharmacological activities and synthesis of esculetin and its derivatives: A mini-review. Molecules 2017, 22, 387. [Google Scholar] [CrossRef] [PubMed]
  244. Azib, L.; Debbache-Benaida, N.; Da Costa, G.; Atmani-Kilani, D.; Saidene, N.; Bouguellid, G.; Ourabah, A.; Krisa, S.; Richard, T.; Atmani, D. Neuroprotective effects of Fraxinus angustifolia Vahl. bark extract against Alzheimer’s disease. J. Chem. Neuroanat. 2020, 109, 101848. [Google Scholar] [CrossRef] [PubMed]
  245. Tavares, W.R.; Seca, A.M. The current status of the pharmaceutical potential of Juniperus L. metabolites. Medicines 2018, 5, 81. [Google Scholar] [CrossRef] [PubMed]
  246. Bais, S.; Gill, N.S.; Rana, N.; Shandil, S. A phytopharmacological review on a medicinal plant: Juniperus communis. Int. Sch. Res. Not. 2014, 2014, 634723. [Google Scholar] [CrossRef]
  247. Gao, Y.; Xu, X.; Chang, S.; Wang, Y.; Xu, Y.; Ran, S.; Huang, Z.; Li, P.; Li, J.; Zhang, L. Totarol prevents neuronal injury in vitro and ameliorates brain ischemic stroke: Potential roles of Akt activation and HO-1 induction. Toxicol. Appl. Pharmacol. 2015, 289, 142–154. [Google Scholar] [CrossRef]
  248. Seca, A.M.; Pinto, D.; Silva, A.M. The current status of bioactive metabolites from the genus Juniperus. Bioact. Phytochem. Perspect. Mod. Med. 2015, 3, 365–407. [Google Scholar]
  249. Mohan, V.; Doss, A.; Tresina, P. Ethnomedicinal Plants with Therapeutic Properties; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  250. Goyal, M.R.; Suleria, H.A.R.; Ayeleso, A.O.; Joel, T.J.; Panda, S.K. The Therapeutic Properties of Medicinal Plants: Health-Rejuvenating Bioactive Compounds of Native Flora; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  251. Karbwang, J.; Crawley, F.P.; Na-Bangchang, K.; Maramba-Lazarte, C. Herbal medicine development: Methodologies, challenges, and issues. Evid.-Based Complement. Altern. Med. 2019, 2019, 4935786. [Google Scholar] [CrossRef]
  252. Singh, D.B.; Pathak, R.K.; Rai, D. From traditional herbal medicine to rational drug discovery: Strategies, challenges, and future perspectives. Rev. Bras. Farmacogn. 2022, 32, 147–159. [Google Scholar] [CrossRef]
  253. Kashte, S.; Gulbake, A.; El-Amin, S.F., III; Gupta, A. COVID-19 vaccines: Rapid development, implications, challenges and future prospects. Hum. Cell 2021, 34, 711–733. [Google Scholar] [CrossRef]
  254. Kuter, B.J.; Offit, P.A.; Poland, G.A. The development of COVID-19 vaccines in the United States: Why and how so fast? Vaccine 2021, 39, 2491. [Google Scholar] [CrossRef]
  255. Ministère. Arrêté Fixant les Éléments du Dossier de Demande d’Agrément de l’Établissement Pharmaceutique d’Éxploitation. Available online: www.miph.gov.dz (accessed on 25 October 2023).
  256. Laakmann, G.; Schüle, C.; Baghai, T.; St. Kieser, M. John’s wort in mild to moderate depression: The relevance of hyperforin for the clinical efficacy. Pharmacopsychiatry 1998, 31, 54–59. [Google Scholar] [CrossRef]
  257. Henderson, L.; Yue, Q.; Bergquist, C.; Gerden, B.; Arlett, P. St John’s wort (Hypericum perforatum): Drug interactions and clinical outcomes. Br. J. Clin. Pharmacol. 2002, 54, 349–356. [Google Scholar] [CrossRef] [PubMed]
  258. Dimpfel, W.; Todorova, A.; Vonderheid-Guth, B. Pharmacodynamic properties of St. John’s wort—A single blind neurophysiological study in healthy subjects comparing two commercial preparations. Eur. J. Med. Res. 1999, 4, 303–312. [Google Scholar]
  259. Iserin, P.; Masson, M.; Restellini, J.; Ybert, E.; De Laage de Meux, A.; Moulard, F.; Zha, E.; De la Roque, R.; De la Roque, O.; Vican, P. Larousse des Plantes Médicinales Identification, Préparation, Soins; Larousse: Paris, France, 2001; Volume 15. [Google Scholar]
  260. Safari, M.; Asadi, A.; Aryaeian, N.; Huseini, H.F.; Jazayeri, S.; Malek, M.; Hosseini, A.F. The effects of Melissa officinalis on depression and anxiety in type 2 diabetes patients with depression: A randomized double-blinded placebo-controlled clinical trial. BMC Complement. Med. Ther. 2023, 23, 1–10. [Google Scholar] [CrossRef] [PubMed]
  261. Kennedy, D.O.; Wake, G.; Savelev, S.; Tildesley, N.T.; Perry, E.K.; Wesnes, K.A.; Scholey, A. Modulation of mood and cognitive performance following acute administration of single doses of Melissa officinalis (Lemon balm) with human CNS nicotinic and muscarinic receptor-binding properties. Neuropsychopharmacology 2003, 28, 1871–1881. [Google Scholar] [CrossRef] [PubMed]
  262. Araj-Khodaei, M.; Noorbala, A.A.; Yarani, R.; Emadi, F.; Emaratkar, E.; Faghihzadeh, S.; Parsian, Z.; Alijaniha, F.; Kamalinejad, M.; Naseri, M. A double-blind, randomized pilot study for comparison of Melissa officinalis L. and Lavandula angustifolia Mill. with Fluoxetine for the treatment of depression. BMC Complement. Med. Ther. 2020, 20, 1–9. [Google Scholar] [CrossRef]
  263. Kennedy, D.O.; Pace, S.; Haskell, C.; Okello, E.J.; Milne, A.; Scholey, A.B. Effects of cholinesterase inhibiting sage (Salvia officinalis) on mood, anxiety and performance on a psychological stressor battery. Neuropsychopharmacology 2006, 31, 845–852. [Google Scholar] [CrossRef]
  264. Movafegh, A.; Alizadeh, R.; Hajimohamadi, F.; Esfehani, F.; Nejatfar, M. Preoperative oral Passiflora incarnata reduces anxiety in ambulatory surgery patients: A double-blind, placebo-controlled study. Anesth. Analg. 2008, 106, 1728–1732. [Google Scholar] [CrossRef]
  265. Akaberi, M.; Iranshahy, M.; Iranshahi, M. Review of the traditional uses, phytochemistry, pharmacology and toxicology of giant fennel (Ferula communis L. subsp. communis). Iran. J. Basic Med. Sci. 2015, 18, 1050. [Google Scholar]
  266. Appendino, G.; Jakupovic, J.; Alloatti, S.; Ballero, M. Daucane esters from Ferula arrigonii. Phytochemistry 1997, 45, 1639–1643. [Google Scholar] [CrossRef]
  267. Louvet, M.-S.; Gault, G.; Lefebvre, S.; Popowycz, F.; Boulven, M.; Besse, S.; Benoit, E.; Lattard, V.; Grancher, D. Comparative inhibitory effect of prenylated coumarins, ferulenol and ferprenin, contained in the ‘poisonous chemotype’of Ferula communis on mammal liver microsomal VKORC1 activity. Phytochemistry 2015, 118, 124–130. [Google Scholar] [CrossRef]
  268. Lahmar, R.; Mahjoub, T.; Fourel, I.; Gault, G.; Grancher, D. Giant fennel (Ferula communis L.) intoxication in goats in Tunisia. Vet. Rec. Case Rep. 2018, 6, e000738. [Google Scholar] [CrossRef]
  269. Philbey, A.; Hawker, A.; Evers, J. A neurological locomotor disorder in sheep grazing Stachys arvensis. Aust. Vet. J. 2001, 79, 427–430. [Google Scholar] [CrossRef]
  270. Berdai, M.A.; Labib, S.; Harandou, M. Peganum harmala L. intoxication in a pregnant woman. Case Rep. Emerg. Med. 2014, 2014, 783236. [Google Scholar] [PubMed]
  271. Dehiri, M.; Diafat, A.; Fatmi, W.; Bouaziz, F.; Khalil, R.; Bahloul, A. Toxicity evaluation of Algerian Peganum harmala seed hydromethanolic extract. Toxicol. Environ. Health Sci. 2022, 14, 351–359. [Google Scholar] [CrossRef]
  272. Djafer, R.; Akil Dahdouh, S.; Boukachabia, R.; Megueddem, M. À propos d’un cas d’intoxication mortelle par le harmel (Peganum harmala L.). Phytothérapie 2017, 15, 288–289. [Google Scholar] [CrossRef]
  273. Simão, A.Y.; Gonçalves, J.; Gradillas, A.; García, A.; Restolho, J.; Fernández, N.; Rodilla, J.M.; Barroso, M.; Duarte, A.P.; Cristóvão, A.C. Evaluation of the cytotoxicity of ayahuasca beverages. Molecules 2020, 25, 5594. [Google Scholar] [CrossRef]
  274. Hamill, J.; Hallak, J.; Dursun, S.M.; Baker, G. Ayahuasca: Psychological and physiologic effects, pharmacology and potential uses in addiction and mental illness. Curr. Neuropharmacol. 2019, 17, 108–128. [Google Scholar] [CrossRef]
  275. Brito-da-Costa, A.M.; Dias-da-Silva, D.; Gomes, N.G.; Dinis-Oliveira, R.J.; Madureira-Carvalho, Á. Toxicokinetics and toxicodynamics of ayahuasca alkaloids N, N-dimethyltryptamine (DMT), harmine, harmaline and tetrahydroharmine: Clinical and forensic impact. Pharmaceuticals 2020, 13, 334. [Google Scholar] [CrossRef]
  276. Fatur, K. “Hexing herbs” in ethnobotanical perspective: A historical review of the uses of anticholinergic Solanaceae plants in Europe. Econ. Bot. 2020, 74, 140–158. [Google Scholar] [CrossRef]
  277. Houmani, Z.; Cosson, L.; Corbineau, F.; Côme, D. Etude de la teneur en hyoscyamine et scopolamine d’une population sauvage de Datura stramonium L. en Algérie. Acta Bot. Gall. 1994, 141, 61–66. [Google Scholar] [CrossRef]
  278. Grynkiewicz, G.; Gadzikowska, M. Tropane alkaloids as medicinally useful natural products and their synthetic derivatives as new drugs. Pharmacol. Rep. 2008, 60, 439. [Google Scholar] [PubMed]
  279. Chang, S.-S.; Wu, M.-L.; Deng, J.-F.; Lee, C.-C.; Chin, T.-F.; Liao, S.-J. Poisoning by Datura leaves used as edible wild vegetables. Vet. Hum. Toxicol. 1999, 41, 242–245. [Google Scholar] [PubMed]
  280. Kwakye, G.F.; Jiménez, J.; Jiménez, J.A.; Aschner, M. Atropa belladonna neurotoxicity: Implications to neurological disorders. Food Chem. Toxicol. 2018, 116, 346–353. [Google Scholar] [CrossRef] [PubMed]
  281. Meggs, W.; Weisman, R. Anticholinergic poisoning associated with an herbal tea—New York City, 1994. MMWR Morb. Mortal. Wkly. Rep. 1995, 44, 193–195. [Google Scholar]
  282. Control, C.f.D.; Prevention. Scopolamine poisoning among heroin users—New York City, Newark, Philadelphia, and Baltimore, 1995 and 1996. MMWR Morb. Mortal. Wkly. Rep. 1996, 45, 457–460. [Google Scholar]
  283. Bouzidi, A.; Mahdeb, N.; Kara, N. Toxicity studies of alkaloids of seeds of Datura stramonium and synthesis alkaloids in male rats. J. Med. Plants Res. 2011, 5, 3421–3431. [Google Scholar]
  284. Benouadah, Z.; Mahdeb, N.; Bouzidi, A. Evaluation of acute and sub-acute toxicity of alkaloids from Datura stramonium sp. in mice. Int. J. Pharmacogn. Phytochem. Res. 2016, 8, 1759–1766. [Google Scholar]
  285. Nath, S.S.; Pandey, C.; Roy, D. A near fatal case of high dose peppermint oil ingestion-Lessons learnt. Indian J. Anaesth. 2012, 56, 582. [Google Scholar] [CrossRef]
  286. Singletary, K. Oregano: Overview of the literature on health benefits. Nutr. Today 2010, 45, 129–138. [Google Scholar] [CrossRef]
  287. Llana-Ruiz-Cabello, M.; Gutiérrez-Praena, D.; Puerto, M.; Pichardo, S.; Jos, Á.; Cameán, A.M. In vitro pro-oxidant/antioxidant role of carvacrol, thymol and their mixture in the intestinal Caco-2 cell line. Toxicol. Vitr. 2015, 29, 647–656. [Google Scholar] [CrossRef] [PubMed]
  288. Azirak, S.; Rencuzogullari, E. The in vivo genotoxic effects of carvacrol and thymol in rat bone marrow cells. Environ. Toxicol. Int. J. 2008, 23, 728–735. [Google Scholar] [CrossRef]
  289. Achour, S.; Abourazzak, S.; Mokhtari, A.; Soulaymani, A.; Soulaymani, R.; Hida, M. Juniper tar (cade oil) poisoning in new born after a cutaneous application. Case Rep. 2011, 2011, bcr0720114427. [Google Scholar] [CrossRef]
  290. Azizi, M.; El Kaouini, A.; Lafkih, M.A.; Bouayed, M.Z.; Bkiyar, H.; Housni, B. A rare case of cade oil poisoning complicated by acute pancreatitis and acute tubular necrosis. Ann. Med. Surg. 2022, 76, 103562. [Google Scholar] [CrossRef] [PubMed]
  291. Skalli, S.; Chebat, A.; Badrane, N.; Bencheikh, R.S. Side effects of cade oil in Morocco: An analysis of reports in the Moroccan herbal products database from 2004 to 2012. Food Chem. Toxicol. 2014, 64, 81–85. [Google Scholar] [CrossRef] [PubMed]
  292. Kim, J.-R.; Perumalsamy, H.; Shin, H.M.; Lee, S.-G.; Ahn, Y.-J. Toxicity of Juniperus oxycedrus oil constituents and related compounds and the efficacy of oil spray formulations to Dermatophagoides farinae (Acari: Pyroglyphidae). Exp. Appl. Acarol. 2017, 73, 385–399. [Google Scholar] [CrossRef]
Figure 1. Major biogeographical/bioclimatic regions of Algeria [6,7,8].
Figure 1. Major biogeographical/bioclimatic regions of Algeria [6,7,8].
Plants 12 03860 g001
Figure 2. Routes of the Silk Road [16].
Figure 2. Routes of the Silk Road [16].
Plants 12 03860 g002
Figure 3. Algerian medicinal plants used to treat mental disorders. (A) Pistacia lentiscus L., (B) Myrtus communis L., (C) Mentha piperita L., (D) Pinus halepensis Mill., (E) Crataegus oxyacantha L., (F) Crataegus azarolus L., (G) Olea europaea L., (H) Hypericum perforatum L., (I) Salvia officinalis L.
Figure 3. Algerian medicinal plants used to treat mental disorders. (A) Pistacia lentiscus L., (B) Myrtus communis L., (C) Mentha piperita L., (D) Pinus halepensis Mill., (E) Crataegus oxyacantha L., (F) Crataegus azarolus L., (G) Olea europaea L., (H) Hypericum perforatum L., (I) Salvia officinalis L.
Plants 12 03860 g003
Figure 4. Distribution of the different reported species among the botanical families.
Figure 4. Distribution of the different reported species among the botanical families.
Plants 12 03860 g004
Figure 5. Selected flavonoids found in Roman chamomile.
Figure 5. Selected flavonoids found in Roman chamomile.
Plants 12 03860 g005
Figure 6. Selected secondary metabolites isolated from Hypericum species.
Figure 6. Selected secondary metabolites isolated from Hypericum species.
Plants 12 03860 g006
Figure 7. Bioactive phenolic acids from Melissa officinalis.
Figure 7. Bioactive phenolic acids from Melissa officinalis.
Plants 12 03860 g007
Figure 9. Some principle constituents of Lavandula officinalis essential oil grown in northeastern Algeria [197].
Figure 9. Some principle constituents of Lavandula officinalis essential oil grown in northeastern Algeria [197].
Plants 12 03860 g009
Figure 11. Selected secondary metabolites isolated from Passiflora and Peganum harmala species.
Figure 11. Selected secondary metabolites isolated from Passiflora and Peganum harmala species.
Plants 12 03860 g011
Figure 12. Curcumin, the main active molecule of Curcuma longa L.
Figure 12. Curcumin, the main active molecule of Curcuma longa L.
Plants 12 03860 g012
Figure 13. Selected hydroxycoumarins isolated from Fraxinus species.
Figure 13. Selected hydroxycoumarins isolated from Fraxinus species.
Plants 12 03860 g013
Figure 14. Totarol, a molecule found in the Juniperus species.
Figure 14. Totarol, a molecule found in the Juniperus species.
Plants 12 03860 g014
Figure 15. Toxic components present in Ferula communis L.
Figure 15. Toxic components present in Ferula communis L.
Plants 12 03860 g015
Table 1. Medicinal plants used in the management of mental illnesses in Algeria.
Table 1. Medicinal plants used in the management of mental illnesses in Algeria.
S/NoPlant NameFamilyCommon NameLocal Algerian Name(s)Use for Mental DisordersPlant Part/s UsedTraditional Preparation and Administration MethodsIdentified Active Constituent(s)Other Relevant Phytoconstituents Identified in the PlantInteraction/ToxicityReferences
1Narcissus tazetta L.AmaryllidaceaeNarcisse à bouquetNardjesEpilepsy, memorigenic,
hysteria,
AD.
Roots,
flowers, bulbs.
Infusion.Alkaloids (galanthamine, lycorine, homolycorine, tazettine, narciclasine).Flavonoids, saponins, tannins, cardiac glycosides, essential oil, steroids, terpenoids, anthraquinones. [45,46,47]
2Pistacia lentiscus L.AnacardiaceaePistachier lentisqueDharwMemory.Leaves, fruits, resin, essential oil.Decoction,
infusion,
fruits can be naturally eaten raw.
Tannins, essential oil (monoterpenes), triterpenes.Polyphenols, phytosterols, flavonoids, triglycerides, tocopherols, carotenoids.Toxic at higher doses.[34,48,49,50]
3Coriandrum sativum L.ApiaceaeCoriandreKousborEpilepsy, nervous tension,
tranquilizer, migraine.
Essential oil, leaves, seeds.Decoction, infusion,
can be naturally chewed.
Essential oil (coriandrol, pinenes, terpinenes, borneol, linalool, geraniol).Aromatic acids,
Isocoumarins,
polyphenols.
[46,51]
4Ferula communis L.ApiaceaeLa FéruleFessoukh,
Kelekh
Anxiety, anti-hysteria.Gum resin (latex), roots,
leaves,
stems
Decoction
infusion
powder
cataplasm
fumigation.
Coumarins, sesquiterpene prenylated coumarins (ferulenol).Daucane esters, phenylpropanoids, phenolic compounds.Toxic
(ferulenol, 4-hydroxycoumarin derivatives, ferprenin).
[5,50,52]
5Ferula assa-foetida L.ApiaceaeFérule persiqueHantitEpilepsy, tranquilizer, stimulant to the brain and nerves.Oleo-gum resin,
aerial parts,
seeds,
roots, young shoots and
leaves.
Decoction, powder.Essential oil,
sesquiterpene, coumarins (foetidin).
Disulfides, ferulic acid, valeric acid. [53,54]
6Pimpinella anisum L.ApiaceaeL’anis vertHabet h’lawa,
Yansoune
Insomnia.Essential oil,
seeds.
Infusion.Essential oil (anethol, methyl chavicol) furanocoumarins, flavonoids.Sterols, proteins, fatty acid, terpenes.May be toxic
under certain conditions.
[55,56]
7Anacyclus pyrethrum (L.) Lag.AsteraceaePyrèthre d’AfriqueAgargarha,
Kantass
Epilepsy, paralysis, seizures, depression, anxiety.Roots, essential oil.Decoction, chew, lozenge, powder.Essential oil, anacycline, inulin.Pellitorine. [5,54]
8Artemisia absinthium L.AsteraceaeL’absintheChadjrat meriemInsomnia, slightly antidepressant.Aerial parts,
leaves.
Infusion,
maceration,
decoction.
Essential oil, terpenes, azulenes, thujone, sesquiterpene lactones (artabasine, absinthin).Phenolic compounds, lignans.Essential oil constituents are highly toxic (alpha-thujone and beta-thujone).[34,54,57,58]
9Artemisia herba alba Asso.AsteraceaeL’armoiseChihAD, epilepsy, depression, neuroinflammation.Aerial parts.Infusion,
decoction.
Essential oil, herbalbin, cis-chryanthenyl acetate, flavonoids (hispidulin and cirsilineol), monoterpenes, sesquiterpene lactones.Coumarins, tannins.Toxic at over dose.[5,34,50,57]
10Carlina gummifera (L.) Less.
Syn. Atractylis gummifera L.
AsteraceaeLe chardon à gluAddadEpilepsy, seizure management, mania.Capitulum, leaves, roots.Decoction.Polyphenols, tannins.Diterpenoid glucosides.Roots are highly toxic (mortal).[36,50,59,60,61]
11Chamaemelum Nobile syn. Anthemis nobilis L.Asteraceae.La Camomille romaineBabounejAnxiety.Flowers,
essential oil.
Infusion, decoction.Essential oil (angelic acid esters, chamazulenes), sesquiterpene lactones (nobilin), flavonoids, coumarins.Polyphenols. [5,34,54]
12Matricaria chamomilla L.AsteraceaeLa Camomille sauvageBabounejMigraines, insomnia.Capitulum.Infusion.Essential oil (alpha-bisabolol, chamazulene), flavonoids, coumarins, tannins.Polyphenols. [36,62]
13Scolymus hispanicus L.AsteraceaeChardon
d’ Espagne
Zernich,
Guernina
Different neurological
conditions.
Roots, stems, leaves, flowers.Infusion.Flavonoids, tannins.Phenolics. [63,64,65,66]
14Silybum marianum (L.) GaertnerAsteraceaeChardon marieChoukDepression.Seeds.Dried seeds, decoction, tincture.Silymarin, silybin A, silybin B, isosilybin A, isosilybin B, silychristin A, silydianin, taxifolin.Polyphenols, essential oil, tannins. [5,63]
15Berberis vulgaris L.BerberidaceaeL’épine-vinetteOud ghrissSedative, morphine addiction.Leaves, fruits, stem bark, roots.Raw, decoction, infusion.Berberine. Alkaloids (berbamine, jateorrhizine, palmatine, oxycanthine).Toxic in higher doses.[54,67,68,69,70]
16Lepidium sativum L.BrassicaceaeCresson alénoisHab err-chadInsomnia, memory.Seeds.Powder.Tannins, vitamins, minerals.Flavonoids, carbohydrates, phenolics, alkaloids, proteins, saponins, lipids. [71,72,73]
17Commiphora myrrha
(Nees) Engl.
BurseraceaeMyrrhEl-morraMemory impairment,
tranquilizer, anxiety.
Gum resin,
seeds.
Infusion, powder.Essential oil, sesquiterpenes, furanosesquiterpenes, polysaccharides, tannins.Proteins and long-chain aliphatic derivatives, steroids, sterols, terpenes. [5,74]
18Boswellia sacra Flueck.BurseraceaeOlibanLoubeneCNS disorders, AD, depression, mental fatigue, stress.Gum resin,
stems.
Powder, infusion, maceration, mastication, decoction, fumigation.Boswellic acids. Essential oil, phenols, terpenoids, uronic acids, steroids, tannins. [5,54]
19Opuntia ficus-indica (L.) Mill.CactaceaeFiguier de BarbarieTine chawki, El-hendiHeadache, dizziness.Leaves.Decoction.Flavonoids, polyphenols.Polysaccharides, sterols, omega-3 fatty acid. [31,36,75]
20Humulus lupulus L.CannabaceaeHoublonJenjelHeadache. Leaves.Raw: topical.Flavonoids, polyphenols. Sesquiterpenoids, diterpenoids, triterpenoids. [36,76]
21Cucurbita maxima DuchesneCucurbitaceaePotironElkaraeaMigraine.Seeds.Decoction: inhalation.Phytosterols (cucurbitacin), vitamins (tocopherols
and carotenoids), unsaturated
fatty acids.
Mineral salts (zinc, selenium). [31,36,77]
22Juniperus phoenicea L.CupressaceaeGenévrier de PhénicieAaraarDifferent neurological conditions.Aerial parts,
berries.
Decoction, infusion,
tablet.
Essential oil, flavonoids,
terpenoids, diterpenes (totarol), lignans, tannins.
Phenyl propanoids and furanone glycosides, sugars, fatty acids, organic acids, sterols.Toxic in higher doses.[5,57]
23Juniperus oxycedrus L.CupressaceaeGenévrier cadeTtagaDifferent neurological conditions.Aerial parts,
berries.
Infusion, decoction, ext. essential oil,
tablet.
Cade oil.Essential oil, phenolic compounds, sesquiterpenes, cresol.Cade oil is toxic in excessive amounts.[5,78]
24Bituminaria bituminosa (L.) C. H. Stirt.FabaceaeHerbe au bitume, Trèfle bitumineuxAdna,
Menita
Epilepsy.Whole plant.Infusion.Phenylpropanoids, coumarins, furanocoumarins, pterocarpans, flavonoids, isoflavones (daidzein and genistein), meroterpenes, sesquiterpenes.Chalcones, phenols, phenolic cinnamates, phenylpropenes, sterols, terpenes, tocopherols, benzofurans, fatty acids.Toxic in excessive amounts.[66,79]
25Glycyrrhiza glabra L.FabaceaeLa réglisseArk-essousHead problems,
psychosis.
Roots.Decoction: oral/topical.Triterpene saponins (glycyrrhizin), isoflavones, phytosterols, coumarins, polysaccharides, asparagine.Pectins, simple sugars, amino acids, mineral salts, essential oil, gum, protein, resin, volatile oils, tannins, glycosides. [31,36,76]
26Lotus corniculatus L.FabaceaeLotier corniculéLotus el karniInsomnia,
depression, tranquilizer,
neurological and psychological disorders.
Leaves,
aerial parts.
Infusion, decoction.Flavonoids, tannins.Alkaloids, terpenes, fatty acids. [63,66,80]
27Senna alexandrina Mill.FabaceaeLe sénéSena-mekkiHead problems, psychosis.Leaves.Decoction, topical/oral. Sennosides A and B, dianthrones, anthrone.Polyphenols. [36,81,82]
28Trigonella foenum-graecum L.FabaceaeLe fenugrecHelbaAnxiety.Seeds.Infusion, raw,
cataplasm.
Essential oil, alkaloids, flavonoids.Saponins, proteins, vitamins, minerals, carbohydrates. [34,36]
29Hypericum perforatum L. HypericaceaeSt. John’s worth,
Millepertuis
Mesmoun, BerslounaEpilepsy, anxiety, depression, neurosedative, nervousness.Flower heads.Infusion.Phenolic compounds (hyperforin), naphthodianthrones (hypericin), proanthocyanins, essential oil.Flavonoids, amentoflavones, carotenoids, catechic tannins. [46,54]
30Crocus sativus L.IridaceaeSafran cultivéZaafranMemory and learning, insomnia, tranquilizer,
neurodegenerative diseases, mild-to-moderate depression, anxiety, headache.
Stamen.Infusion, powder, raw.Essential oil, crocetin glucosides (crocins), carotenoids, vitamins (B1, B2).Flavonoids.Toxic at higher doses.[5,54,71,72,78]
31Ajuga iva (L.) SchrebLamiaceaeBugle ivetteChengouraCNS
diseases,
memory, mental nervousness.
Whole plant,
aerial parts,
leaves,
flowers,
roots.
Infusion, decoction, powder,
cataplasm.
Iridoids, diterpenes, phytoecdysone, caffeic acids.Steroids, terpenoids, flavonoids, fatty acids. [63,72,83]
32Lavandula stoechas L.LamiaceaeLavande papillon, lavande à toupet,
lavande stéchade
HelhalAnxiety, depression.Leaves, stems, flowers.Decoction, infusion,
ext. essential oil.
Essential oil (terpenes, linalyl acetate, linalool, cineole, limonene), tannins, coumarins, flavonoids.Triterpenes, alcohols, ketones, polyphenols. [31]
33Lavandula officinalis Chaix
syn. Lavandula angustifolia Mill.
LamiaceaeLavande vraie, lavande à feuilles étroitesKhzamaStress, anxiety, nervousness, depression, insomnia, tranquilizer.Leaves, stems, flowers.Decoction, infusion,
ext. essential oil.
Essential oil (linalyl acetate, cineol, linalool, borneol), flavonoids, tannins, coumarins.Polyphenols. [31]
34Melissa officinalis L.LamiaceaeMelisseMelissaDepression, stress,
anxiety, nervousness, insomnia, nervous tonic.
Aerial parts.Decoction, infusion,
ext. essential oil, cataplasm, ointment.
Essential oil (citral, caryophyllene, linalool, citronellal), flavonoids, triterpenes, tannins.Terpenoids, polyphenols.The essential oil has moderate toxicity at higher doses.[34,36,54,84]
35Mentha arvensis L.LamiaceaeMenthe des champsNaanaHead problems, anxiety, psychosis, insomnia.Aerial parts.Raw.Menthol, menthone.Isomenthone, neomenthol, limonene, methyl acetate, piperitone, beta-caryophyllene, alpha-pinene, beta-pinene, tannins, flavonoids. [54,85]
36Mentha aquatica L.LamiaceaeMenthe aquatiqueHbak elmaaAnxiety, hypochondria.Aerial part.Raw.Menthol, menthone.Flavonoids (luteolin, menthoside), phenols, triterpenes. [36,54]
37Mentha rotundifolia (L.) Huds.LamiaceaeMenthe à feuilles rondes, Menthe du NilTimarssatPain, mental illnesses,
removal of “curses” and protect from “evil spirits”.
Aerial part.Decoction.Menthol, menthone, pulegone.Phenolic acids, flavonoids. [86,87,88]
38Mentha pulegium L. LamiaceaeMenthe pouliotFliouInsomnia, headache, migraine. Aerial part.Decoction.Pulegone, isopulegone, menthol.Terpenes, tannins. The essential oil is highly toxic.[31,54]
39Mentha x piperita L.LamiaceaeLa MentheNaanaTranquilizer, anxiety.Aerial parts.Infusion, capsules,
essential oil: topical.
Essential oil (menthol, menthone), flavonoids (luteolin, menthoside).Phenols, triterpenes.[31]
40Ocimum basilicum L.LamiaceaeBasilicHbakEpilepsy, depression, anxiety, sleeping disorders, insomnia. mental exhaustion, migraine.Leaves,
flowering tops.
Decoction, powder,
ext. essential oil.
Essential oil (linalool, methyl-chavicol, methyl cinnamate, cineol).Terpenes. [54,72,89,90]
41Origanum majorana L.LamiaceaeMarjolaineMerdakoucheAnxiety, insomnia, nervousness.Aerial parts,
ext. Essential oil.
Infusion.Essential oil (sabinene hydrate, linalool, carvacrol), caffeic acid, rosmarinic acid, flavonoids.Triterpenes. [31]
42Origanum vulgare L.LamiaceaeOriganZaatarAnxiety, sleeping disorders.Aerial parts,
ext. essential oil.
Infusion.Essential oil (thymol, carvacrol, borneol, linalool, beta-bisabolene, caryophyllene), flavonoids.Phenolic acids, tannins. [31]
43Rosmarinus officinalis L.LamiaceaeRomarinLeklilMemory, slightly anti-depressive, psychical stimulant, nervousness.Leaves, flowers,
aerial parts.
Infusion, decoction, maceration.Essential oil (borneol, camphor, camphene, cineol), flavonoids (apigenin, diosmin), rosmarinic acid, alkaloids(rosmaricin).Tannins, diterpenes.Toxic at higher doses.[34,54,72]
44Salvia Officinalis L.LamiaceaeSaugeMiramiaStress, tranquilizer, nervous tonic, nervousness, memory, insomnia, agitation in AD.Leaves, stems.Raw leaves, infusion, Decoction, maceration.Essential oil, phenolic compounds (coumarins, flavonoids, tannins).Alkaloids, carbohydrate, fatty acids, glycosidic derivatives, poly acetylenes, steroids, terpenoids.Toxic in case of prolonged use or at high doses.[34,54,72,91]
45Stachys arvensis (L.) L.LamiaceaeÉpiaire des champsChatraAntidepressant, anxiety.Capitulum.Decoction.Hydroxyflavone-allosylglucosides.Flavonoids, iridoids, fatty acids, phenolic acids, diterpenoids. [38,63,92,93]
46Stachys officinalis L.
Syn. Betonica officinalis L.
Lamiaceaela BétoineChatraNervous system stimulant, stress, insomnia, anxiety, tranquilizer, memory deficit, lightly sedative.Aerial parts.Tea of dried leaves,
extract of flowers.
Stachydrine, betonicine, betaine, choline.Iridoids, flavonoid, tannins, fatty acids, phenolic acids, diterpenoids, phenols, hydroxycinnamic acid derivatives, sesquiterpenes. [54,92,93,94]
47Thymus vulgaris L.LamiaceaeThymZaatarAnxiety, phobia.Aerial parts,
ext. essential oil.
Infusion,
decoction.
Essential oil (thymol, carvacrol, linalool, cineol), flavonoids.Phenolic acids, saponins, tannins. [34,54,95,96]
48Cinnamomum camphora (L.)LauraceaeCamphrierKafourMigraine.Wax.Infusion: topical.Camphor. Essential oil (camphor, safrole, eugenol, terpineol), lignans.Should be used externally only.[31,36]
49Cinnamomum verum J. PreslLauraceaeCannelierKarfaMigraine.Bark.Raw: oral/topical.Essential oil (cinnamaldehyde: 65–75%). Polyphenols (vanillic acid, caffeic acid, gallic acid, p-coumaric acid, ferulic acid, cinnamic acid, proanthocyanidins A and B, kaempferol,), tannins (phlobatannins), coumarins, mucilages, eugenol. [31]
50Lawsonia inermis L.
Syn. Lawsonia alba
LythraceaeHennéHennaAnxiety, hypochondria.LeavesRaw: topical.Naphtoquinones (lawsone).Polyphenols, coumarins, flavonoids, tannins, alkaloids, terpenoids, sterols, carbohydrates, proteins, fatty acids. [31]
51Punica granatum L. LythraceaeGrenadierRommaneHeadache.Fruit peels, fruits.Decoction: topical.Flavonoids, ellagitannins, alcaloids.Triterpens, polyphenols.Peels may be toxic if consumed in excess.[31,97]
52Myristica fragrans Houtt.MyristicaceaeMuscadierDjouz-ettibHead problems, psychosis.Seeds.Raw: topical.Essential oil (α-pinene, β-pinene, alpha terpinene, beta terpinene, myristicin, elemicin, safrole).Fixed oil (myristicin).Toxic overdose may cause CNS excitation with anxiety/fear.[98]
53Syzygium aromaticum (L.) Merr. & L.M.Perry.
syn. Eugenia caryophyllata
MyrtaceaeGirofleK’rounfelMental asthenia, loss of memory, migraine.Flower buds,
leaves, stems.
Infusion,
essential oil: topical.
Essential oil (eugenol, eugenol acetyl, methyl salicylate, pinene, vanillin).Gum, tannins. [54]
54Myrtus communis L.MyrtaceaeLe Myrte communRihaneAnxiety, tranquilizer.Leaves and flowers,
essential oil,
fruits may be eaten raw or dried.
Infusion, decoction.Essential oil (alpha pinene, cineol, myrtenol), flavonoids, tannins.Phenolic compounds. [34,54]
55Peganum harmala L.NitrariaceaeHarmelHarmalAnxiety, memory, nervous system disorders, Parkinson’s agitation.Seeds, roots.Infusion, decoction, fumigation.Alkaloids (harmine, harmaline, harmalol).Saponins, tannins, glycosides, terpenoids, steroids.Toxic.[34,54,99]
56Fraxinus angustifolia Vahl.OleaceaeFrêne à feuilles étroitesDardarAD, different nervous system conditions.Barks,
leaves,
grains.
Infusion.Tannins, flavonoids, coumarins, essential oil, resin, malic acid.Saccharides, minerals, vitamins. [5,63,66,100]
57Olea europaea L.OleaceaeOliveZitounMemory, anxiety, head problems, psychosis,
depression.
Leaves,
fruits, fruit oil.
Infusion, raw.Oleuropein, oleuropeoside, monounsaturated fatty acids, tyrosol fatty acid esters, hydroxytyrosols, oleic acid, tocopherols.Flavonoids, tannins, saponines, saccharides, triterpenes, vitamins, minerals. [5,54]
58Papaver rhoeas L.PapaveraceaeCoquelicotKhashkhash,
thekouche
Insomnia, stress;
tranquilizer,
nervousness.
Flowers.Infusion.Alkaloids (papaverine, rheadine, isorheadine, rhoeagenine).Anthocyanins, mucilage, tannins.Toxic.[54,72]
59Passiflora incarnata L.
PassifloraceaePassifloreNouar al saaEpilepsy,
depression, insomnia, anxiety, nervousness, hysteria, stress, tranquilizer.
Leaves,
flowers.
Infusion, compressed.Indole alkaloids (harmane), glycosidic flavonoids (vitexin, isovitexin, orientin), flavonoids (luteolin, chrysin, kaempferol, apigenin), glycosides (passiflorine), alkaloids (harmine, harmaline), palmitic acid, myristic acid.Maltol, cyanogenetic glucosides (gynocardin). [54,72,101]
60Pinus halepensis Mill.PinaceaePin maritimeSnoubar,
Z’koukou (fruits)
Memory deficits, tranquilizer, strengthen the nervous system.Leaves,
buds, bark seeds, resin.
Decoction,
infusion,
chewed,
ext. ointment, cataplasm, powder, essential oil.
Essential oil.Terpenoids, phenolic acids, flavonoids, fatty acids, steroids, aldehydes, ketones. [34,66,72,102]
61Avena sativa L.PoaceaeAvoineChofanDepression, asthenia, stress, insomnia, nervous fatigue,
stimulates the
nervous system.
Seeds, dried stems.Infusion, tincture.Saponins, alkaloids, trigonelline, silicic acid.Proteins, vitamins B, minerals. [54]
62Nigella sativa L.RanunculaceaeNigelleSanouj,
Zrarâ
Anxiety.Seeds.Decoction, infusion,
ext. powder, essential oil.
Fatty acids (linoleic acid, oleic acid), saponins, essential oil.Alkaloids (Nigellimine N-oxide, nigellidine, nigellicine), carbohydrates, proteins, minerals, tannins.Toxic (melanthin).[34,54]
63Rhamnus alaternus L.RhamnaceaeLe Nerprun alaterneMelilesseNeuroprotective.Barks, leaves, fruits.Powder, decoction,
infusion.
Polyphenols, anthraquinones (emodin), anthrone, anthranols.Tannins, anthocyanins, alkaloids. [54,63,66,72,89,103]
64Zizyphus lotus L.RhamnaceaeLe jujubier sauvageSedraNeuroprotective, promoting memory and learning, insomnia, forgetfulness, hypnotic sedative, anxiolytic.Leaves, fruits, roots.Infusion, decoction, powder, fruits may be eaten raw or dried.Tannins, vitamins, flavonoids, polyphenols, polysaccarides. Alkaloids, minerals. [54,57,72,104]
65Crataegus oxyacantha L.RosaceaeAubepineBoukhrourou,
Bou m’kherry
Epilepsy, loss of memory, insomnia, neurosedative,
sleep disorders.
Flowers,
fruits.
Infusion decoction, tincture,
compressed.
Flavonoids (rutin, quercetin), flavones (vitexin, orientin, rhamnosylvitexin), triterpenes, proanthocyanes, polyphenols, tannins, coumarins, saponins, alkaloids (nicotine).Amines, anthocyans, phenolic acids, triterpenic acids, sitosterols, purines. [5,54,72]
66Crataegus azarolus L.RosaceaeAzerolierZaarourAnxiety, stress, psychical disorders, insomnia, nervous tonic.Leaves, flowers,
fruits.
Extracts, tincture, fruit can be eaten raw, cooked, or as preserves.Essential oil, tannins, amino acids, proanthocyanidins, flavonoids.Sugars (fructose, glucose et rhamnose), vitamins. [5,72,89,100]
67Eriobotrya japonica (Thunb.) LindlRosaceaeLe néflier du JaponNiflaHeadache, dizziness.LeavesDecoctionQuercetin, ursolic acid, oleanolic acid, tannins, chlorogenic acid, caffeoylquinic acid.Polyphenols, flavonoids, carotenoids, triterpenoids. [105,106]
68Rosa canina L.RosaceaeÉglantierNasrine, OuardzeroubTranquilizer, anxiety, depression.Fruits (rose hips), leaves,
flowers.
Tea infusion.Vitamins (C, A, B1, B2, P, K), flavonoids, tannins, citric acid, carotenoids, essential oil, D-sorbitol.Polyphenols, carotenoids, carbohydrates, fatty acids. [54,72,89]
69Galium verum L.RubiaceaeCaille-laitFouaouaEpilepsy.Aerial parts, flowers.Infusion.Iridoids (asperulosides), flavonoids, anthraquinones, alcanes.Phenolic compounds, tannins, saponins, triterpenes, essential oil, wax, pigments, vitamin C. [54,66]
70Citrus limon (L.) Osbeck.RutaceaeCitronnierLaymouneDizziness.Fruits.Decoction.Essential oil, terpenes (limonene), flavonoids (hesperidin), vitamins, mucilage.Sesquiterpenes, aldehydes (citral), coumarins. [31,36]
71Ruta chalepensis L.RutaceaeRue de ChalepFidjelHeadache, mental disorders.Aerial partsDecoctionCoumarins (furanocoumarins, dihydrofuranocoumarins), furoquinoline alkaloids. Flavonoids, essential oil. [107,108]
72Populus Nigra L.SalicaceaePeuplier noirSafsafNeurodegenerative disorders.Leaves, flower buds,
barks.
Infusion, tincture, powder, ointment.Tannins, flavonoids, saccharides, essential oil.Phenolic compounds, terpenoids. [5,73,109]
73Santalum album L.SantalaceaeSantal blancSandelMigraine.Bark, fruits.Decoction: topical/oral.Essential oil, tannins.Resin. [31]
74Atropa belladonna L.SolanaceaeBelladoneSitt ElhossnParkinson’s disease, neurological
disorders,
tranquilizer.
Leaves, roots.Tincture, decoction.Tropane alkaloids.Nicotine, flavonoids, coumarins.Toxic (effects includes pupil dilatation, confusion, hallucination).[34,54,72]
75Datura stramonium L.SolanaceaeStramoineTatura,
Mendj
Parkinson’s disease, stress.Leaves, capitulum, seeds.Tincture, decoction,
fumigation,
powder.
Tropane alkaloids (hyoscyamine, hyoscine). Flavonoids, withanolides, coumarins, tannins, minerals.Highly toxic.[54,72]
76Tamarix aphylla (L.) H.KarstTamaricaceaeTamarix aphyllaTahtahHeadache.Leaves.Decoction.Flavonoids, polyphenols, tannins.Catechins, triterpenoids. [110,111]
77Tilia cordata Mill.TiliaceaeTilleulZayzafouneEpilepsy,
neurosedative,
sleep disorders, stress, anxiety.
Flowers.Infusion.Flavonoids (quercetin, rhamnoside, kaempferol).Tannins, essential oil, mucilage. [54,72,89]
78Valeriana tuberosa L.ValerianaceaeValérianeSenbel,
nardine
Epilepsy, stress, anxiety, insomnia.Roots and
rhizome.
Infusion, decoction.Essential oil (bornyl acetate, beta-caryophyllene), iridoids (valepotriate), alkaloidsTannins, flavonoids. [5,46]
79Valeriana officinalis L.ValerianaceaeValérianeSenbelEpilepsy, stress, anxiety, insomnia, nervousness.Roots and rhizome.Infusion, decoction, tincture, powder, compressed.Essential oil (bornyl acetate, beta-caryophyllene), iridoids (valepotriate), alkaloids.Tannins, flavonoids. [5,54,72,90]
80Verbena officinalis L.VerbenaceaeVerveineLouisaTranquilizer, lightly sedative, anxiolytic, anticonvulsant,
lightly antidepressive, insomnia, anxiety, mental fatigue.
Aerial parts.Tincture, infusion, powder.Iridoids (verbenone, verbenaline), Essential oil.Mucilage, tannins, flavonoids, phenolic acids. [54,112,113,114]
81Zingiber officinale L.ZigiberaceaeGingembreZanjabilAnxiety, head problems, psychosis.Rhizome.Infusion, maceration, capsules, tincture,
ext. essential oil.
Essential oil, sesquiterpenes, oleoresin, phenols (gingerol).Phenolic compounds. [34,54]
82Curcuma longa L.ZigiberaceaeCurcumaKorkoumAnxiety, hypochondria.Rhizome.Decoction, powder, cataplasm, tincture.Essential oil, zingiberene, turmerone, curcuminoids (curcumin), resin.Phenolic compounds, caffeic acid derivatives. [31,34,36]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Larit, F.; León, F. Therapeutics to Treat Psychiatric and Neurological Disorders: A Promising Perspective from Algerian Traditional Medicine. Plants 2023, 12, 3860. https://doi.org/10.3390/plants12223860

AMA Style

Larit F, León F. Therapeutics to Treat Psychiatric and Neurological Disorders: A Promising Perspective from Algerian Traditional Medicine. Plants. 2023; 12(22):3860. https://doi.org/10.3390/plants12223860

Chicago/Turabian Style

Larit, Farida, and Francisco León. 2023. "Therapeutics to Treat Psychiatric and Neurological Disorders: A Promising Perspective from Algerian Traditional Medicine" Plants 12, no. 22: 3860. https://doi.org/10.3390/plants12223860

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