Next Article in Journal
Glyoxalase Goes Green: The Expanding Roles of Glyoxalase in Plants
Next Article in Special Issue
Hydroxytyrosol and Cytoprotection: A Projection for Clinical Interventions
Previous Article in Journal
Brain-Derived Neurotrophic Factor Mediated Perfluorooctane Sulfonate Induced-Neurotoxicity via Epigenetics Regulation in SK-N-SH Cells
Previous Article in Special Issue
Fisetin Protects PC12 Cells from Tunicamycin-Mediated Cell Death via Reactive Oxygen Species Scavenging and Modulation of Nrf2-Driven Gene Expression, SIRT1 and MAPK Signaling in PC12 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Health Effects of Psidium guajava L. Leaves: An Overview of the Last Decade

1
Department of Analytical Chemistry, Faculty of Sciences, University of Granada, Avd. Fuentenueva s/n, 18071 Granada, Spain
2
Functional Food Research and Development Center, Health Science Technological Park, Avd. Del Conocimiento, Bioregion Building, 18100 Granada, Spain
3
Department of Nutrition and Food Science, University of Granada, Campus of Cartuja, 18071 Granada, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(4), 897; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18040897
Submission received: 13 March 2017 / Revised: 17 April 2017 / Accepted: 19 April 2017 / Published: 24 April 2017

Abstract

:
Today, there is increasing interest in discovering new bioactive compounds derived from ethnomedicine. Preparations of guava (Psidium guajava L.) leaves have traditionally been used to manage several diseases. The pharmacological research in vitro as well as in vivo has been widely used to demonstrate the potential of the extracts from the leaves for the co-treatment of different ailments with high prevalence worldwide, upholding the traditional medicine in cases such as diabetes mellitus, cardiovascular diseases, cancer, and parasitic infections. Moreover, the biological activity has been attributed to the bioactive composition of the leaves, to some specific phytochemical subclasses, or even to individual compounds. Phenolic compounds in guava leaves have been credited with regulating blood-glucose levels. Thus, the aim of the present review was to compile results from in vitro and in vivo studies carried out with guava leaves over the last decade, relating the effects to their clinical applications in order to focus further research for finding individual bioactive compounds. Some food applications (guava tea and supplementary feed for aquaculture) and some clinical, in vitro, and in vivo outcomes are also included.

Graphical Abstract

1. Introduction

Ethnomedicine, which refers to the study of traditional medical practice, is an integral part of the culture and the interpretation of health by indigenous populations in many parts of the world [1]. For example, Indian Ayurveda and traditional Chinese medicine are among the most enduring folk medicines still practiced. These systems try to promote health and improve the quality of life, with therapies based on the use of indigenous drugs of natural origin [2]. Given that plants have been widely used as herbal medicines, several approaches are now being carried out to discover new bioactive compounds [3].
Psidium guajava L., popularly known as guava, is a small tree belonging to the myrtle family (Myrtaceae) [4]. Native to tropical areas from southern Mexico to northern South America, guava trees have been grown by many other countries having tropical and subtropical climates, thus allowing production around the world [5]. Traditionally, preparations of the leaves have been used in folk medicine in several countries, mainly as anti-diarrheal remedy [6]. Moreover, other several uses have been described elsewhere on all continents, with the exception of Europe [6,7,8]. Figure 1 summarizes the main traditional uses of guava leaves in the main producer countries. Depending upon the illness, the application of the remedy is either oral or topical. The consumption of decoction, infusion, and boiled preparations is the most common way to overcome several disorders, such as rheumatism, diarrhea, diabetes mellitus, and cough, in India, China, Pakistan, and Bangladesh [6,7,8,9], while in Southeast Asia the decoction is used as gargle for mouth ulcers [6,8,9] and as anti-bactericidal in Nigeria [8,9]. For skin and wound applications, poultice is externally used in Mexico, Brazil, Philippines, and Nigeria [6,7,8,9]. In addition, chewing stick is used for oral care in Nigeria [9].
Currently, there is increasing interest in studying of plants regarding their chemical components of bioactive compounds, their effects on several diseases, and their use for human health as functional foods and/or nutraceuticals [10]. In recent years, guava leaves tea and some complementary guava products are available in several shops in Japan as well as on the Internet [11], because guava leaf phenolic compounds have been claimed to be food for specified health use (FOSHU), since they have beneficial health effects related to the modulation of blood–sugar level [12]. Thus, the aim of this review was to summarize the biological activities, in vitro and in vivo, studied in the last decade on P. guajava L. leaves, relating them to the international classification of diseases provided by the World Health Organization. In addition, the beneficial effects of some applications of guava leaves are also been examined. For this purpose, a comprehensive review of the literature from 2004 to 2016 was done, although more recent studies have also been included. Reviewed journals, websites, books, and several databases as “Scopus”, “Google Scholar”, “PubMed”, and “ScienceDirect”, were used to compile them. To ensure that relevant works are included, terms such as “Psidium guajava”, “guava”, “leaves”, “in vitro”, “in vivo”, “clinical”, “trial”, “food application”, and those related with the diseases such as “bacteria”, “cancer”, “blood”, “glycaemia”, and “oral”, among others were matched in the search. Only complete available works published in English, Spanish, and Portuguese have been included.

2. Pharmacological Properties

2.1. In Vitro Studies

2.1.1. Infectious and Parasitic Diseases

Aqueous and organic extracts of guava leaves have been demonstrated to have antibacterial activity due to an inhibitory effect against antibiotics-resistant clinical isolates of Staphylococcus aureus strains [13,14]. Despite using the same diffusion method, differences are noticed in their inhibition zones, as shown in Table 1, probably due to extraction method or the dose assayed. A methanol extract exerted antibacterial effects, preventing the growth of different strains from several bacteria such as Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, Proteus spp., and Shigella spp. [15]. Furthermore, different extracts of the leaves such as aqueous, acetone–water, methanolic, spray-dried extracts, and the essential oil, showed potential inhibitory activity against Gram-positive and Gram-negative bacteria and fungi [16,17,18,19,20]. In these works, it is noticeable that Gram-positive bacteria exhibited greater inhibition zones and minimum inhibitory concentrations (MICs) than Gram-negative. Concerning the anti-fungal activity, less inhibition than bacteria is reported [16,17], except for Candida krusei and Candida glabrata which provided higher inhibition [18], and for Aspergillus spp. for which no activity was found [16] (Table 1). Moreover, Bezerra et al. [21] evaluated the effect of guava leaves on different bacterial strains, concluding that the synergistic action between the leaves and various antibiotics boosted its anti-bacterial activity. This effect was also observed by Betoni et al. [22] with target drugs for the protein synthesis, cell-wall synthesis, and folic acid. However, the latter did not find synergic effect with gentamicin, perhaps because the time of maceration was lower than the time used by Bezerra et al. [21], and also the solvent was different (Table 1).
Metwally et al. [23] associated the antimicrobial activity against some bacteria and fungi with five flavonoids isolated from the leaves. This effect was also related to the concentration of tannins in the leaves [24] and to the content of gallic acid and catechin [19]. Additionally, the activity against bacterial and fungal pathogens was traced to betulinic acid and lupeol [25]. In fact, these works are focused on the activity of these compounds, rather than on the effect of the whole extract of the leaves.
In addition, leaf acetone extract of P. guajava has also exhibited moderate acaricidal and insecticidal activities causing the dead of Hippobosca maculata adult fly [26].
Furthermore, Adeyemi et al. [27] suggested that an ethanol extract from the leaves function as a trypanocide agent, since its inhibition of Trypanosoma brucei brucei growth proved similar to that of the reference drugs. Kaushik et al. [28] proposed the leaves as an anti-malaria agent, due to their inhibitory activity and the resistance indices. Furthermore, the effect of guava leaf essential oil against toxoplasmosis caused by the growth of Toxoplasma gondii were reported [29]. Additionally, guava leaves were proposed for the treatment of diarrhea caused by enteric pathogens, since it showed significant inhibitory activity against Vibrio cholerae and V. parahemolyticus, Aeromonas hydrophila, Escherichia coli, Shigella spp. and Salmonella spp. [30,31,32]. It is suppose that the same plant origin and similar extraction procedure makes that these works show comparable inhibition zones for the bacteria tested [30,31], in contrast to the leaves of India and Bangladesh, where MIC values did not show any concordance [31,32] (Table 1). In addition, a reduction was described for S. flexneri and V. cholera invasion and for their adherence to the human laryngeal epithelial cells, and for the production of E. coli heat labile toxin and cholera toxin, as well as their binding to ganglioside monosialic acid enzyme [33]. Moreover, other studies also demonstrated the antimicrobial effect of some bacteria that cause gastrointestinal disorders by different methods [34,35]. In contrast to previous results [20,31], no inhibition of the hydrodistillation and n-hexane extract was found against E. coli Salmonella spp. [31] (Table 1).
Furthermore, guava leaf tea helped control of the growth of influenza viruses, including oseltamivir-resistant strains, via the prevention of viral entry into host cells, probably due to the presence of flavonols [36].

2.1.2. Neoplasms

All the results published regarding anti-cancer properties have been summarized in Table 2.
Kawakami et al. [37] evaluated the anti-proliferative activity of guava leaf extract in human-colon adenocarcinoma cell line (COLO320DMA). These authors found that the extract depressed the proliferation rate due to the presence of quercetin and quercetin glycosides. Moreover, different extracts were tested on three cancer cell lines (cervical cancer (HeLa), breast cancer (MDA-MB-231), and osteosarcoma (MG-63)). The extracts showed no anti-proliferative activity towards HeLa cells, although they displayed activity against MDA-MB-231 and MG-63, the ether extract being the most effective, followed by methanol and water extracts. However, ether and methanol extracts presented a cytotoxic effect on non-malignant cell Madine Darby canine kidney (MDCK) [38]. In contrast, an ethanol extract from the stem and leaves reported significant anti-tumor activity on HeLa and colorectal carcinoma (RKO-AS45-1), whereas its effect was less significant for a lung fibroblast cell line (Wi-26VA4) [39]. This difference could be due to the origin of the leaves, compounds in the steam, or even to the extraction method selected. In this context, an organic guava leaf extract provided molecular evidence of cytotoxic or anti-tumor activity in human breast carcinoma benign cells (MCF-7) and also in murine fibrosarcoma (L929sA) [40]. A fact worthy to comment is that the difference noticed in the cytotoxic effect on MDA-MB-231 cell line might be because the extraction differs [38,40]. Furthermore, the aqueous extract of budding guava leaves displayed an anti-tumor effect against human prostate epithelial (PZ-HPV-7) and carcinoma (DU-145) cells in view of the cell-killing-rate coefficients, as well as anti-angiogenesis and anti-migration activities, respectively [41,42].
Regarding the bioactivity of terpenes from guava, an enriched mixture of guajadial, psidial A, and psiguadial A and B proved anti-proliferative effect for nine human cancer lines: leukemia (K-562), breast (MCF-7), resistant ovarian cancer (NCI/ADR-RES), lung (NCI-H460), melanoma (UACC-62), prostate (PC-3), colon (HT-29), ovarian (OVCAR-3), and kidney (786-0) [43]. The apoptotic effect of β-caryophyllene oxide (CPO) on MCF-7 and PC-3 cell lines was also demonstrated because of its ability to interfere with multiple signaling cascades involved in tumor genesis [44]. Moreover, the essential oil from guava leaves exerted an anti-proliferative effect on human-mouth epidermal carcinoma (KB) and murine leukemia (P388) cell lines [45], while a hexane fraction of the leaves showed a cytotoxic effect against leukemia (Kasumi-1) cancer-cell line at higher half maximal inhibitory concentration (IC50), probably due to a less concentration of the bioactive compounds of the leaves [46]. Finally, cytotoxic and apoptotic effect in PC-3 cells and apoptotic effect in LNCaP cells was reported. The lack of cytotoxic effect in LNCaP might be because the cell growth is androgen-dependent, while in PC-3 is androgen-independent. [47]. Comparing these data with those reported by Park et al. [44], high concentration is needed for causing cell death, and a weak effect is found on early apoptotic cell. The main difference between these works is the composition of the extract, so it could be concluded that an antagonist effect is produced amongst the isolated compounds by Ryu et al. [47].

2.1.3. Diseases of the Blood and Immune System

A fermented guava leaf extract was tested in mouse macrophage (RAW 264.7) cells. The results confirmed its potential to decrease the expression of lipopolysaccharide-inducible nitric oxide synthase and cyclooxygenase-2 proteins level, two pro-inflammatory mediators, through the down-regulation of nuclear factor-κB transcriptional activity (NF-κB) [48]. This biological activity was also reported in other works [40,49,50]. Briefly, Jang et al. [49] evaluating the prostaglandin E2 production found that the inhibitory effect was highly correlated to the total phenolic content. Kaileh et al. [40] suggested that the suppression of the nuclear factor-κB could be at the transcriptional level because of the lack of binding between nuclear factor-κB and DNA in murine fibrosarcoma (L929sA) and two breast-cancer cell lines (MDA-MB231 and MCF7). At the same time, Jang et al. [50] found that the lipopolysaccharide-induced production of nitric oxide and prostaglandin E2 was due to the ability of guava leaf extract to suppress phosphorylation in protein expression. Moreover, Sen et al. [51] verified the inhibition of nuclear factor-κB activation in Labeo rohita head-kidney macrophages by the flavonoid fraction of guava leaf extract and Jang et al. [52] improved the inhibition of lipopolysaccharide-induced prostaglandin E2 and nitric oxide production by optimizing of the extraction conditions. Furthermore, methanol and ethanol leaf extracts also showed the inhibition of hypotonicity-induced lysis of erythrocyte membrane [53]. Meanwhile, Laily et al. [54] suggested the use of guava leaves as immune-stimulant agent because they modulated the lymphocyte proliferation response.
The results for this activity, confirm the potential of guava leaves as an anti-inflammatory treatment and as immune-system stimulatory agent. As is shown in Table 3, a general trend is reported in every work, although the differences noticed in the data are probably due to the different extraction method and to the doses assayed, or even to the harvesting time of the leaves. However, the mechanism should be further studied since two different pathways are suggested for the down-regulation of NF-κB.

2.1.4. Endocrine and Metabolic Diseases

Several works have focused on elucidating the anti-diabetic compounds present in guava leaves (Table 4). Although the origin of the leaves remains different, the presence of these compounds has demonstrated the hypoglycemic effect of the leaves via different assays. However, the main mode of action seems to be due to an inhibition of the enzymes related to this activity.
The anti-glycative potential of the guava leaves was evaluated, with the conclusion that the extract inhibited, in vitro, the formation of advanced glycation end-products formation [55]. Moreover, the aqueous guava leaf extract, in an albumin/glucose model system, also exerted the same effect and indeed inhibited Amadori products. Gallic acid, catechin and quercetin exhibited over 80% inhibitory effects whereas ferulic acid showed no activity [56]. In another study, seven pure flavonoid compounds (quercetin, kaempferol, guaijaverin, avicularin, myricetin, hyperin, and apigenin) showed strong inhibitory activities against sucrase, maltase, and α-amylase, and a clear synergistic effect against α-glucosidase [57]. Moreover, Deguchi and Miyazaki [58] suggested that the component that inhibited the in vitro activities of α-glucosidase enzymes in guava extract was a polymerized polyphenol. In addition, polysaccharides from guava leaves also exhibited α-glucosidase inhibition [59].
Eidenberger et al. [60] demonstrated the dose-dependent inhibition of guava leaf ethanol extracts on dipeptidyl-peptidase-IV due to the individual flavonol-glycosides: peltatoside, hyperoside, methylquercetin hexoside, isoquercitrin, quercetin/morin pentoside, guaijaverin, and quercetin/morin pentoside. Additionally, the individual flavonol-glycosides found in the guava extract reported no significant differences compared with the uptake of the whole guava extract into epithelial cells (CaCo-2) [60]. In the same cell line, the inhibition of fructose uptake was also tested by Lee et al. [61], who confirmed that catechin and quercetin contributed to the inhibition of glucose transporters. In addition, the enhancement of aqueous guava leaf extract was investigated with regard to glucose uptake in rat clone 9 hepatocytes. Moreover, quercetin was proposed as the active compound responsible for promoting glucose uptake in liver cells and contributing to the alleviation of hypoglycemia in diabetes [62]. Furthermore, Basha and Kumari [63] also estimated the glucose uptake of different extracts. The methanol extract of guava leaves was found to be the most efficient in lowering glucose levels. Basha et al. [64] demonstrated the ability of guavanoic-acid-mediated gold nanoparticles to inhibit the protein tyrosine phosphatase 1B activity.
Indeed, a guava leaf ethanol extract was tested in pre-adipocyte cell line (3T3-L1), which showed its ability to inhibit adipocyte differentiation via down-regulation of adipogenic transcription factors and markers, and hence may prevent obesity in vivo [65]. To evaluate the potential of the leaves on glucose uptake and glycogen synthesis, an aqueous extract was used in insulin-resistant mouse (FL83B) cells. The results confirmed the improved expression and phosphorylation of insulin signaling-related proteins, promoting glycogen synthesis and glycolysis pathways. In fact, this work provides new insights into the mechanisms through which the guava extract improves insulin resistance in the hepatocytes [66]. In the same cell line, vescalagin was postulated as the active component that may alleviate the insulin resistance in mouse hepatocytes [67].
In this sense, the latest study made in L6 myoblasts and myotubes cells confirmed that the glucose uptake recruitment followed a wortmannin-dependent pathway. In addition, guava leaves also inhibited aldose reductase activity, up-regulated gene- and protein-level expression of several insulin receptors and also improved cellular-level glucose uptake [68].

2.1.5. Diseases of the Circulatory System

Cardiovascular disorders have been related to the endothelial cell damage that causes atherosclerosis. In this sense, extracts from budding guava leaves demonstrated a protective, in vitro, effect in bovine aortal endothelial cells, delaying low-density lipoprotein oxidation and preventing oxidized low-density lipoprotein cytotoxicity [69]. A similar effect was also noted in human umbilical-vein endothelial cell due to the ability of saving cell-viability reduction, suppressing reactive oxygen species production and nitric oxide release, as well as inhibiting the expression of NF-κB [70]. Moreover, budding guava leaves also showed their ability as an anticoagulant in plasma, since they reduced thrombin clotting time and inhibited the activity of antithrombin III. Thus, they could help to reduce the development of cardiovascular complications [71].
In addition, flavonoids and phenolic acids in the leaves could contribute to the prevention and amelioration of gout and hypertension, since, in rat-tissues homogenates, they inhibit the activity of two enzymes related to the development of both diseases (xanthine oxidase and angiotensin 1-converting enzymes) [72].

2.1.6. Diseases of the Digestive System

Guaijaverin, isolated from guava leaves, displayed high inhibitory activity against Streptococcus mutans. In fact, guaijaverin exhibited its ability as an anti-plaque agent, becoming an alternative for oral care [73]. Furthermore, guava leaves showed greater bactericidal effect on early (Streptococcus sanguinis) and late (S. mutans) colonizers compared to Mangifera indica L. and Mentha piperita L. leaves, whereas, when they are compared with the plant extract mixture, the effect is slightly lower. By contrast, guava leaves showed similar and higher anti-adherence effect than the plant mixture [74]. In another study, the whole extract was tested on the cell-surface hydrophobicity of selected early settlers and primary colonizers of dental plaque, showing its ability to alter and disturb the surface characteristics of the agents, making them less adherent [75,76,77], and also delayed in the generation of dental biofilm by targeting growth, adherence, and co-aggregation [78]. This property could be due to the presence of flavonoids and tannins detected in P. guajava [79]. Shekar et al. [80] also confirmed the use of the leaves as anti-plaque agents against Streptococcus mutans, S. sanguinis, and S. salivarius. Kwamin et al. [81] discovered the effectiveness of guava leaf extract in the leukotoxin neutralization of Aggregatibacter actinomycetemcomitans, leading it to be considered as a possible agent for the treatment of aggressive forms of periodontitis. In addition, extracts rich in guava flavonoids have demonstrated their potential for preventing dental caries due to the growth inhibition of the oral flora [82]. Moreover, its soothing of toothache has been verified based on the analgesic, anti-inflammatory, and anti-microbial activity properties [83] and it has been reviewed positively as an adjutant for treating periodontal disease [84].
Concerning the liver disorders, the cytotoxic and hepato-protective effects of guava leaves were reported. Studies carried out in clone 9 cells treated with different extracts of the leaves showed that only ethanol and acetone extracts tend to have cytotoxicity effect at high concentrations. Moreover, the ethanol extract showed hepato-protective activity, although the hot-water extract reported greater effect and lower cytotoxicity [85].
Table 5 compiles the methodology followed and the results reported in the present works. It is important to keep in mind that the origin, the selection of the extraction method or solvent, and the concentration of the extract tested generally provide different data. For example, comparing data for inhibition zones, best results are noticed at long maceration time in acetone, which seems to be a better extracting solvent than ethanol [77,78,80,82]. Hydrophobicity depends on the origin of the leaves, the extraction method, and the concentration of the extract tested, and it also depends on the lipophilic (index > 70%) or hydrophilic nature of the strain [73,75,79]. Finally, minimum inhibitory concentration relies on all factors.

2.1.7. Diseases of the Skin and Subcutaneous Tissue

Qa’dan et al. [86] described the antimicrobial effect of a leaf extract against the main developer of acne lesions, Propionibacterium acnes, and other organisms isolated from acne lesions. The antimicrobial activity was also displayed against pathogenic bacteria associated with wound, skin, and soft-tissue infections [87]. Furthermore, antifungal properties have also been studied by Padrón-Márquez et al. [88]. The acetone and methanol extracts displayed relevant activity against dermatophytic fungi, and thus could be considered as new agents against skin disease. Furthermore, phenols from the leaves were tested on human-skin fibroblast cells and showed antifungal properties [89].
In addition, the tyrosinase inhibitory activities of 4 different parts (branch, fruit, leaf, and seed) of guava, extracted with acetone, ethanol, methanol, and water were tested by You et al. [90] who reported that the ethanol extract from the leaves reached the highest activity. Therefore, the leaves might be appropriate for both boosting the whitening of skin and inhibiting browning. In addition, in a human keratinocyte cell line, an ethyl acetate extract showed a positive effect on atopic dermatitis via the inhibition of cytokine-induced Th2 chemokine expression [91].
Lee et al. [92] carried out the first electrophysiological study based on ultraviolet (UV)-induced melanogenesis with guava leaves. The authors suggested the use of guava leaves for both direct and indirect prevention of skin melanogenesis caused by UV radiation. In fact they demonstrate that methanolic guava leaves extract inhibits tyrosinase, that is the key enzyme in melanin synthesis, and ORAI1 channel that has shown to be associated with UV-induced melanogenesis.

2.1.8. Other Activities Related to Several Diseases

An aqueous guava extract showed its ability to decrease the radiolabeling of blood constituent due to an antioxidant action and/or because it alters the membrane structures involved in ion transport into cells [93]. Guava leaves also have been demonstrated to possess anti-allergic effects in rat mast (RBL-2H3) cell line by the inhibition of degranulation and cytokine production, as well as blocking high-affinity immunoglobulin E-receptor signaling [94].

2.2. In Vivo Studies

2.2.1. Infectious and Parasitic Diseases

After checking the effect of guava leaf extract, in vitro, against Aeromonas hydrophila, in vivo experiments were carried out in tilapia (Oreochromis niloticus), indicating the potential use of P. guajava as environmentally friendly antibiotic [95]. The leaves also had anti-viral and anti-bacterial activity towards shrimp (Penaeus monodon) pathogens such as yellow-head virus, white spot syndrome virus, and Vibrio harvey. In addition, guava leaf extract improved the activities of prophenoloxidase and nitric oxide synthase in serum, and of superoxide dismutase, acid phosphatase, alkaline phosphatase, and lysozyme in serum and hepatopancreas [96].
Furthermore, guava leaves have been suggested for managing sleeping sickness, since they exhibited trypanocidal effect in albino rats [97]; the extract ameliorate the tissue-lipid peroxidation associated to trypanosomosis, as well as raising the level of the glutathione concentration [98]. The leaves also showed anti-malarial effect in BALB/c mice infected with Plasmodium berghei via parasitemia suppression [99]. Moreover, guava leaves are also recommended for treating infectious diarrhea since they prevented intestinal colonization of Citrobacter rodentium in Swiss albino mice [100]. In chicks, guava leaf extract enabled the control of diarrhea produced by E. coli and reduced the severity of its symptomatology [101]. In mice, the improvement of cholera symptoms caused by V. cholerae, a human pathogen, was also confirmed by Shittu et al. [102].
In addition, anti-helminthic properties towards gastro-intestinal nematodes have been found, as a result of the presence of condensed tannins in the guava plant, which raised the levels of hemoglobin, packed cell volume, total protein, globulin, glucose, and calcium, and lowered the levels of blood urea [103].
All the results published regarding in vivo anti-bacterial properties have been summarized in Table 6.

2.2.2. Neoplasms

Only one study is available on the anti-tumor effect that could be related to the phenolic composition of guava leaves. An ethanol extract of the leaves was administrated to B6 mice after inoculation of melanoma cells. The results suggested that the extract had a vaccine effect, but not a therapeutic effect, against tumors through by depressing T regulatory cells [104].
Moreover, the meroterpene-enriched fraction of guava leaves, containing guajadial, psidial A, and psiguadial A and B, was evaluated in vivo in a solid Ehrlich murine breast-adenocarcinoma model. The results suggested that these compounds may act as phytoestrogens, presenting tissue-specific antagonistic and agonistic activity on estrogen receptors [43]. These data partially confirmed the results in vitro obtained by Ryu et al. [47].

2.2.3. Diseases of the Blood and Immune System

Among blood diseases, anemia indicates a failure in the immune system. In this sense, guava extract presented an anti-anemic effect in trypanosomosis-infected Wistar rats, improving the values of hemoglobin, packed cell volume, red-blood cell counts, mean corpuscular volume, and mean concentration hemoglobin count while decreasing white-blood cell and neutrophil levels [105]. Moreover, the same trend in the hematological analyses was also recorded in mice. After the administration of guava leaf extract, no alterations on the erythron were detected [106]. Nevertheless, results differ because subjects under study are different, also the duration of the treatment, the extraction method and the dose assayed (Table 7).
The anti-inflammatory response of the leaves was dose-dependent in induced hyperalgesia in Sprague-Dawley rats, decreasing in paw-withdrawal latency, and significantly improving the survival rate of mice with lethal endotoxemia [50]. Moreover, the anti-inflammatory activity of aqueous and acetone–water extracts of the leaves was also confirmed in Swiss mice by reducing the amount of leukocyte migration. The acetone–water extract also exhibited peripheral analgesic activity, probably by blocking the effect or the release of endogenous substances that excite pain-nerve endings [19]. The analgesic effect in albino rats was also reported. The ethanol extract reduced the writhing response [107], and a jumping response was found after the administration of a distilled extract (combination of methanol and aqueous extracts) [108]. In this case, the writhing response for both Swiss mice and Wistar rats seems to be comparable, although the dose assayed is completely different (Table 7).

2.2.4. Endocrine and Metabolic Diseases

Guava leaves have shown their potential against one of the diseases with the highest incidence level worldwide, diabetes mellitus, and also towards biochemical changes caused by the disease. In spite of being leaves from different countries, treatments in different subjects or even different data, the same trend is followed in these works (Table 8).
The effect of aqueous guava leaf extract was investigated in rabbits, fed a high-cholesterol diet. Treatment with guava leaves reduced the plasma-cholesterol level, caused a remarkable spike in high-density lipoprotein, a dip in low-density lipoprotein levels, and significantly reduced the associated hyperglycemia. In addition, the extract showed hypolipidemic and hypoglycemic potentials in hypercholesterolemic rabbits [109]. Furthermore, guava leaves reduced oxidative stress induced by hypercholesterolemia in rats [110].
In addition, the anti-diabetic effect was also evaluated in Leprdb/Leprdb mice and significant blood-glucose-lowering effects were observed. In addition, histological analysis revealed a significant reduction in the number of lipid droplets, which, furthermore, at least in part, could be mediated via the inhibition of protein tyrosine phosphatase 1B [111].
In streptozotocin-induced diabetic rats, the administration of oral doses of aqueous and ethanol extracts from guava leaves could alter the Ca:Mg ratio [112]; however, in low-dose streptozotocin and nicotinamide-induced Sprague-Dawley diabetic rats, long-term administration of guava leaf extracts raised the plasma-insulin level, the glucose utilization, and the activity of hepatic enzymes [113]. Moreover, the leaves also lowered blood glucose levels and decreased protein glycation [55].
In agreement with the above, a lower blood-glucose level was also reported in alloxan-induced diabetic rats. Additionally, no side effects were observed in certain liver enzymes (alkaline phosphatase and aspartate aminotransferase) whereas alanine aminotransferase activity declined [114]. In alloxan-induced diabetic rats, a decrease was also found in blood glucose, total cholesterol, triglycerides, low-density lipoprotein cholesterol, very low-density lipoprotein cholesterol, and a significant increase in high-density lipoprotein cholesterol after 21 days of treatment with guava leaf ethanolic extract [115].
Among the works that evaluated only biochemical parameters, guava leaf extract promoted changes due to an alteration on the activity of alkaline phosphatase, aspartate aminotransferase, alanine aminotransferase, and acid phosphatase in the kidney, liver, and serum [106,116]. In addition, Adeyemi and Akanji [117] evaluated the effect of daily administration of guava leaves, demonstrating the alteration of the serum homeostasis and the pathological variations in rat tissues.

2.2.5. Diseases of the Circulatory System

Ademiluyi et al. [118] assessed the lipid peroxidation in rats after checking the antihypertensive effect, in vitro, of red and white guava leaves. The work concluded that the activity may be related to rosmarinic acid, eugenol, carvacrol, catechin, and caffeic acid since they were the major constituents of their extracts. In addition, this activity was supported by the biphasic and contractile effect on rat vascular smooth muscles [119,120].
In addition, atherosclerosis development was reduced in apoE-knockout mice by guava leaf extracts. In fact, the effect was connected to the presence of ethyl gallate and quercetin [121,122]. In streptozotocin-induced diabetic rats, vascular reactivity to vasoconstrictor agents was reduced, as was vessel atherosclerosis [112]. Furthermore, Soman et al. [123] found that an ethyl acetate fraction of guava leaves reduced cardiac hypertrophy in streptozotocin-induced diabetic rats due to an anti-glycative effect.

2.2.6. Diseases of the Digestive System

In the digestive system, formed by the gastrointestinal tract plus the group organs necessary for the digestion, guava leaves have demonstrated activity towards different parts.
On the one hand, the leaves have shown the ability to protect the stomach against ulceration by inhibiting gastric lesions, reducing gastric secretory volume, and acid secretion, and raising the gastric pH [124,125,126]. This anti-ulcer activity, resulting from the protection of the mucosa, was related to the flavonoids in the leaves [127]. Despite of the subject employed for the assay, similar data are reported in these works (Table 9). The anti-diarrheal activity of guava leaf aqueous extract was evaluated on experimentally induced diarrhea in rodents. The extract performed in the same way as the control drugs, offering protection, inhibiting intestinal transit, and delaying gastric emptying [128]. Another study attributed this activity to a dual action between the antimicrobial effect and the reduction in gastrointestinal motility ability of the extract [129]. In rabbits, the anti-spasmodic effects were connected to a calcium channel blocking activity, which explains the inhibitory effect on gut motility. The anti-diarrheal protection was also tested in mice [130]. As is shown in Table 9, the anti-diarrheal activity is dose-dependent, although the protection varied depending on the subjects.
On the other hand, guava leaves exhibited hepato-protective effect due to the reduction of serum parameters of hepatic enzymes markers and histopathological alterations in the acute liver damage induced in rats [131,132,133,134,135]. Here, a dose-dependent effect is also found. However, decoction of the leaves seems to be the best option for the extraction of the compounds that exhibited this activity (Table 9).

2.2.7. Diseases of the Skin and Subcutaneous Tissue

Guava leaves have been suggested as a therapeutic agent to control pruritus in atopic dermatitis. The improvement of the skin lesions was due to a reduction in serum immunoglobulin E level and in the eczematous symptoms [136]. Moreover, the epithelium was repaired with connective tissue and absence or moderate presence of inflammatory cells by the leaves. As a result, the leaves exhibited wound healing properties [137]. Furthermore, guava leaf extract was tested on rat skin, and exhibited inhibitory activity towards an active cutaneous anaphylaxis reaction [138].

2.2.8. Other Activities Related to Several Diseases

Triterpenoids from guava leaves were suggested as a potential therapeutic approach for treating diabetic peripheral neuropathy, as they enhanced physical functions and offered neuronal protection towards the suppression of the expression of pro-inflammatory cytokines [139]. In addition, the leaves can act as radio modulators for cancer patients because by preventing DNA damage and apoptosis. [140], as well as protective agents by restoring the normal values of sperm viability, sperm count, sperm motility, and sperm-head abnormality caused by caffeine-induced spermatotoxicity [141].
Moreover, the consumption of guava leaf tea was evaluated, in vivo, in the inhibition of cytochrome P450 (CYP) 3A-mediated drug metabolism by the interaction between guava tea and several drugs [11,142]. Matsuda et al. [11] investigated the consequence of the ingestion of guava tea for two weeks in rats, and the effect with an anxiolytic drug. The short-term consumption of the tea had little effect on the assays performed. This weak influence was due to the absence of interaction between the tea and midazolam in the metabolism studied. In addition, two in vivo studies were made in rats, to evaluate the interaction of guava leaf tea with an anti-coagulant drug (warfarin) [142]. Kaneko et al. [141] suggested that because the tea contained compounds that block the affinity between the enzyme and phenolic compounds of the tea, long-term administration showed a low probability of causing drug-metabolizing enzymes. Moreover, short-term administration revealed that the tea did not interfere with coagulation, meaning that the tea consumption did not alter the pharmacological effect and displayed no side effects.

2.3. Clinical Trials

To test the effect of guava leaf extract, several randomized clinical trials have been conducted during the last two decades, although only two studies are available in the last decade. One of the studies consisted of evaluating the effect of guava leaf extract pills on primary dysmenorrhea disorder. For this, 197 women were divided into four groups, and each received a different dosage: 3 and 6 mg extract/day, 300 mg placebo/day and 1200 mg ibuprofen/day. The administration took place over five days during three consecutive cycles. The results demonstrated that 6 mg extract/day alleviated menstrual pain and could replace the use of medicaments like ibuprofen. In fact, guava leaves could be used as a broad-spectrum phyto-drug and not only as an anti-spasmodic agent [143]. Furthermore, Deguchi and Miyazaki [58] reviewed several works regarding the effect of the intake of a commercial guava leaf tea (Bansoureicha®, Yakult Honsha, Tokyo, Japan) on different pathologies of diabetes mellitus illness such as the influence on postprandial blood glucose, on insulin resistance and on hypertriglyceridemia and hypercholesterolemia. The authors concluded that the ingestion of guava leaf tea can ameliorate the symptoms of diabetes mellitus and that it could be used as an alimentotherapy.

3. Other Applications

Further applications found with guava leaves are listed below: firstly, to prepare gelatin beads with marine-fish gelatin for various applications such as medicine, and the food and pharmaceutical industries [144]. Secondly, Giri et al. [145] suggested guava leaves as supplementary feed for the fish species Labeo rohita, due to the immune-stimulatory effect. The same conclusion was reached by Fawole et al. [146] in L. rohita. Thirdly, Gobi et al. [147] reported that guava leaf powder, mixed with a commercial diet, strengthened the immunological response of Oreochromis mossambicus, and recommended the leaves as feed complement in aquaculture.

4. Conclusions

Traditional claims generally require experimental research to establish their effectiveness. In this regard, ethnomedicine applications of Psidium guajava L. leaves have been verified by several researches over the last decade against many disorders, demonstrating its potential in the treatment of the most common worldwide diseases. In addition, the effects of the leaves have been related to individual compounds such as quercetin, catechin, vescalagin, gallic acid, peltatoside, hyperoside, isoquercitrin, and guaijaverin.
Future prospects should be aimed at investigating the biodiversity of guava and/or the purification of the different compounds present in guava leaves in order to obtain functional ingredients for further uses as alternative agents in natural therapeutic approaches.

Acknowledgments

The author Vito Verardo thanks the Spanish Ministry of Economy and Competitiveness (MINECO) for “Ramony Cajal” post-doctoral contract.

Author Contributions

Elixabet Díaz-de-Cerio contributed to the literature review and manuscript redaction; Vito Verardo and Ana María Gómez-Caravaca contributed to the conception of the idea and framework writing; and Alberto Fernández-Gutiérrez and Antonio Segura-Carretero supervised the progress of work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anyinam, C. Ecology and ethnomedicine: Exploring links between current environmental crisis and indigenous medical practices. Soc. Sci. Med. 1995, 40, 321–329. [Google Scholar] [CrossRef]
  2. Patwardhan, B.; Warude, D.; Pushpangadan, P.; Bhatt, N. Ayurveda and traditional Chinese medicine: A comparative overview. Evid. Based Complement. Altern. Med. 2005, 2, 465–473. [Google Scholar] [CrossRef] [PubMed]
  3. Fabricant, D.S.; Farnsworth, N.R. The value of plants used in traditional medicine for drug discovery. Environ. Health Perspect. 2001, 109, 69–75. [Google Scholar] [CrossRef] [PubMed]
  4. Morton, J.F. Fruits of Warm Climates; Creative Resource Systems, Inc.: Winterville, NC, USA, 1987. [Google Scholar]
  5. Salazar, D.M.; Melgarejo, P.; Martínez, R.; Martínez, J.J.; Hernández, F.; Burguera, M. Phenological stages of the guava tree (Psidium guajava L.). Sci. Hortic. 2006, 108, 157–161. [Google Scholar] [CrossRef]
  6. Gutiérrez, R.M.P.; Mitchell, S.; Solis, R.V. Psidium guajava: A review of its traditional uses, phytochemistry and pharmacology. J. Ethnopharmacol. 2008, 117, 1–27. [Google Scholar] [CrossRef] [PubMed]
  7. Shruthi, S.D.; Roshan, A.; Sharma, S.; Sunita, S. A review on the medicinal plant Psidium guajava Linn. (Myrtaceae). J. Drug Deliv. Ther. 2013, 3, 162–168. [Google Scholar]
  8. Morais-Braga, M.F.B.; Carneiro, J.N.P.; Machado, A.J.T.; dos Santos, A.T.L.; Sales, D.L.; Lima, L.F.; Figueredo, F.G.; Coutinho, H.D.M. Psidium guajava L., from ethnobiology to scientific evaluation: Elucidating bioactivity against pathogenic microorganisms. J. Ethnopharmacol. 2016, 194, 1140–1152. [Google Scholar] [CrossRef] [PubMed]
  9. Sanda, K.A.; Grema, H.A.; Geidam, Y.A.; Bukar-Kolo, Y.M. Pharmacological aspects of P. guajava: An update. Int. J. Pharmacol. 2011, 7, 316–324. [Google Scholar]
  10. Bernal, J.; Mendiola, J.A.; Ibáñez, E.; Cifuentes, A. Advanced analysis of nutraceuticals. J. Pharm. Biomed. Anal. 2011, 55, 758–774. [Google Scholar] [CrossRef] [PubMed]
  11. Matsuda, K.; Nishimura, Y.; Kurata, N.; Iwase, M.; Yasuhara, H. Effects of continuous ingestion of herbal teas on intestinal CYP3A in the rat. J. Pharmacol. Sci. 2007, 103, 214–221. [Google Scholar] [CrossRef] [PubMed]
  12. Arai, S.; Yasuoka, A.; Abe, K. Functional food science and food for specified health use policy in Japan: State of the art. Curr. Opin. Lipidol. 2008, 19, 69–73. [Google Scholar] [CrossRef] [PubMed]
  13. Milyani, R. Inhibitory effect of some plant extracts on clinical isolates of Staphylococcus aureus. Afr. J. Microbiol. Res. 2012, 6, 6517–6524. [Google Scholar] [CrossRef]
  14. Anas, K.; Jayasree, P.R.; Vijayakumar, T.; Manish Kumar, P.R. In vitro antibacterial activity of Psidium guajava Linn. leaf extract on clinical isolates of multidrug resistant Staphylococcus aureus. Indian J. Exp. Biol. 2008, 46, 41–46. [Google Scholar] [PubMed]
  15. Chah, K.F.; Eze, C.A.; Emuelosi, C.E.; Esimone, C.O. Antibacterial and wound healing properties of methanolic extracts of some Nigerian medicinal plants. J. Ethnopharmacol. 2006, 104, 164–167. [Google Scholar] [CrossRef] [PubMed]
  16. Nair, R.; Chanda, S. In Vitro antimicrobial activity of Psidium guajava L. leaf extracts against clinically important pathogenic microbial strains. Braz. J. Microbiol. 2007, 38, 452–458. [Google Scholar] [CrossRef]
  17. Dhiman, A.; Nanda, A.; Narasimhan, B. In vitro antimicrobial activity of methanolic leaf extract of Psidium guajava L. J. Pharm. Bioallied Sci. 2011, 3, 226–229. [Google Scholar] [CrossRef] [PubMed]
  18. Fernandes, M.R.V.; Dias, A.L.T.; Carvalho, R.R.; Souza, C.R.F.; Oliveira, W.P. Antioxidant and antimicrobial activities of Psidium guajava L. spray dried extracts. Ind. Crops Prod. 2014, 60, 39–44. [Google Scholar] [CrossRef]
  19. De Araújo, A.A.; Soares, L.A.L.; Assunção Ferreira, M.R.; de Souza Neto, M.A.; da Silva, G.R.; de Araújo, R.F.; Guerra, G.C.B.; de Melo, M.C.N. Quantification of polyphenols and evaluation of antimicrobial, analgesic and anti-inflammatory activities of aqueous and acetone-water extracts of Libidibia ferrea, Parapiptadenia rigida and Psidium guajava. J. Ethnopharmacol. 2014, 156, 88–96. [Google Scholar] [CrossRef] [PubMed]
  20. Nisha, K.; Darshana, M.; Madhu, G.; Bhupendra, M.K. GC-MS Analysis and anti-microbial activity of Psidium guajava (leaves) grown in Malva region of India. Int. J. Drug Dev. Res. 2011, 3, 237–245. [Google Scholar]
  21. Bezerra Morais-Braga, M.F.; Lima Sales, D.; dos Santos Silva, F.; Pereira Chaves, T.; de Carvalho Nilo Bitu, V.; Torres Avilez, W.M.; Ribeiro-Filho, J.; Douglas Melo Coutinho, H. Psidium guajava L. and Psidium brownianum Mart ex DC. potentiate the effect of antibiotics against Gram-positive and Gram-negative bacteria. Eur. J. Integr. Med. 2016, 8, 683–687. [Google Scholar] [CrossRef]
  22. Betoni, J.E.C.; Passarelli Mantovani, R.; Nunes Barbosa, L.; Di Stasi, L.C.; Fernandes Junior, A. Synergism between plant extract and antimicrobial drugs used on Staphylococcus aureus diseases. Mem. Inst. Oswaldo Cruz 2006, 101, 387–390. [Google Scholar] [CrossRef] [PubMed]
  23. Metwally, A.M.; Omar, A.A.; Harraz, F.M.; El Sohafy, S.M. Phytochemical investigation and antimicrobial activity of Psidium guajava L. leaves. Pharmacogn. Mag. 2010, 6, 212–218. [Google Scholar] [PubMed]
  24. Mailoa, M.N.; Mahendradatta, M.; Laga, A.; Djide, N. Antimicrobial activities of tannins extract from guava leaves (Psidium guajava L.) on pathogens microbial. Int. J. Sci. Technol. Res. 2014, 3, 236–241. [Google Scholar]
  25. Ghosh, P.; Mandal, A.; Chakraborty, P.; Rasul, M.G.; Chakraborty, M.; Saha, A. Triterpenoids from Psidium guajava with biocidal activity. Indian J. Pharm. Sci. 2010, 72, 504–507. [Google Scholar] [CrossRef] [PubMed]
  26. Zahir, A.A.; Rahuman, A.A.; Bagavan, A.; Santhoshkumar, T.; Mohamed, R.R.; Kamaraj, C.; Rajakumar, G.; Elango, G.; Jayaseelan, C.; Marimuthu, S. Evaluation of botanical extracts against Haemaphysalis bispinosa Neumann and Hippobosca maculata Leach. Parasitol. Res. 2010, 107, 585–592. [Google Scholar] [CrossRef] [PubMed]
  27. Adeyemi, O.S.; Sykes, M.L.; Akanji, M.A.; Avery, V.M. Anti-trypanosomal and cytotoxic activity of ethanolic extracts of Psidium guajava leaves in Alamar Blue based assays. Vet. Arh. 2011, 81, 623–633. [Google Scholar]
  28. Kaushik, N.K.; Bagavan, A.; Rahuman, A.A.; Zahir, A.A.; Kamaraj, C.; Elango, G.; Jayaseelan, C.; Kirthi, A.V.; Santhoshkumar, T.; Marimuthu, S.; et al. Evaluation of antiplasmodial activity of medicinal plants from North Indian Buchpora and South Indian Eastern Ghats. Malar. J. 2015, 14, 65. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, W.C.; Mahmud, R.; Noordin, R.; Pillai Piaru, S.; Perumal, S.; Ismail, S. Free radicals scavenging activity, cytotoxicity and anti-parasitic activity of essential oil of Psidium guajava L. leaves against Toxoplasma gondii. J. Essent. Oil Bear. Plants 2013, 16, 32–38. [Google Scholar] [CrossRef]
  30. Chanu, T.R.; Pai, V.; Chakraborty, R.; Raju, B.; Lobo, R.; Ballal, M. Screening for antidiarrheal activity of Psidium guajava: A possible alternative in the treatment against diarrhea causing enteric pathogens. J. Chem. Pharm. Res. 2011, 3, 961–967. [Google Scholar]
  31. Thiyagarajan, S.; Jamal, A. Evaluation of Lethal Activity of Psidium guajava Linn. extracts on bacterial pathogens causing diarrheal infections. Int. J. Res. Ayurveda Pharm. 2015, 6, 111–117. [Google Scholar] [CrossRef]
  32. Rahim, N.; Gomes, D.J.; Watanabe, H.; Rahman, S.R.; Chomvarin, C.; Endtz, H.P.; Alam, M. Antibacterial activity of Psidium guajava leaf and bark against multidrug-resistant Vibrio cholerae: Implication for cholera control. Jpn. J. Infect. Dis. 2010, 63, 271–274. [Google Scholar] [PubMed]
  33. Birdi, T.; Daswani, P.; Brijesh, S.; Tetali, P.; Natu, A.; Antia, N. Newer insights into the mechanism of action of Psidium guajava L. leaves in infectious diarrhoea. BMC Complement. Altern. Med. 2010, 10, 33. [Google Scholar] [CrossRef] [PubMed]
  34. Gonçalves, F.A.; Andrade Neto, M.; Bezerra, J.N.S.; Macrae, A.; De Sousa, O.V.; Fonteles-Filho, A.A.; Vieira, R.H.S.D.F. Antibacterial activity of guava, Psidium guajava Linnaeus, leaf extracts on diarrhea-causing enteric bacteria isolated from seabob shrimp, Xiphopenaeus kroyeri (Heller). Rev. Inst. Med. Trop. Sao Paulo 2008, 50, 11–15. [Google Scholar] [CrossRef] [PubMed]
  35. Nwinyi, O.; Chinedu, S.N.; Ajani, O.O. Evaluation of antibacterial activity of Pisidium guajava and Gongronema latifolium. J. Med. Plants Res. 2008, 2, 189–192. [Google Scholar]
  36. Sriwilaijaroen, N.; Fukumoto, S.; Kumagai, K.; Hiramatsu, H.; Odagiri, T.; Tashiro, M.; Suzuki, Y. Antiviral effects of Psidium guajava Linn. (guava) tea on the growth of clinical isolated H1N1 viruses: Its role in viral hemagglutination and neuraminidase inhibition. Antivir. Res. 2012, 94, 139–146. [Google Scholar] [CrossRef] [PubMed]
  37. Kawakami, Y.; Nakamura, T.; Hosokawa, T.; Suzuki-Yamamoto, T.; Yamashita, H.; Kimoto, M.; Tsuji, H.; Yoshida, H.; Hada, T.; Takahashi, Y. Antiproliferative activity of guava leaf extract via inhibition of prostaglandin endoperoxide H synthase isoforms. Prostaglandins Leukot. Essent. Fatty Acids 2009, 80, 239–245. [Google Scholar] [CrossRef] [PubMed]
  38. Sul´ain, M.D.; Zazali, K.E.; Ahmad, N. Screening on anti-proliferative activity of Psidium guajava Leaves extract towards selected cancer cell lines. J. US China Med. Sci. 2012, 9, 30–37. [Google Scholar]
  39. Vieira Braga, T.; Gonçalves Rodrigues das Dores, R.; Soncin Ramos, C.; Gontijo Evangelista, F.C.; Márcia da Silva Tinoco, L.; de Pilla Varotti, F.; Carvalho, M.D.G.; de Paula Sabino, A. Antioxidant, antibacterial and antitumor activity of ethanolic extract of the Psidium guajava leaves. Am. J. Plant Sci. 2014, 5, 3492–3500. [Google Scholar] [CrossRef]
  40. Kaileh, M.; Vanden Berghe, W.; Boone, E.; Essawi, T.; Haegeman, G. Screening of indigenous Palestinian medicinal plants for potential anti-inflammatory and cytotoxic activity. J. Ethnopharmacol. 2007, 113, 510–516. [Google Scholar] [CrossRef] [PubMed]
  41. Peng, C.C.; Peng, C.H.; Chen, K.C.; Hsieh, C.L.; Peng, R.Y. The aqueous soluble polyphenolic fraction of Psidium guajava leaves exhibits potent anti-angiogenesis and anti-migration actions on DU145 cells. Evid. Based Complement. Altern. Med. 2011, 2011, 219069. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, K.C.; Hsieh, C.L.; Huang, K.D.; Ker, Y.B.; Chyau, C.C.; Peng, R.Y. Anticancer activity of rhamnoallosan against DU-145 cells is kinetically complementary to coexisting polyphenolics in Psidium guajava budding leaves. J. Agric. Food Chem. 2009, 57, 6114–6122. [Google Scholar] [CrossRef] [PubMed]
  43. Rizzo, L.Y.; Longato, G.B.; Ruiz, A.L.T.G.; Tinti, S.V.; Possenti, A.; Vendramini-costa, D.B. In vitro, in vivo and in silico analysis of the anticancer and estrogen-like activity of guava leaf extracts. Curr. Med. Chem. 2014, 21, 2322–2330. [Google Scholar] [CrossRef] [PubMed]
  44. Park, K.R.; Nam, D.; Yun, H.M.; Lee, S.G.; Jang, H.J.; Sethi, G.; Cho, S.K.; Ahn, K.S. β-Caryophyllene oxide inhibits growth and induces apoptosis through the suppression of PI3K/AKT/mTOR/S6K1 pathways and ROS-mediated MAPKs activation. Cancer Lett. 2011, 312, 178–188. [Google Scholar] [CrossRef] [PubMed]
  45. Manosroi, J.; Dhumtanom, P.; Manosroi, A. Anti-proliferative activity of essential oil extracted from Thai medicinal plants on KB and P388 cell lines. Cancer Lett. 2006, 235, 114–120. [Google Scholar] [CrossRef] [PubMed]
  46. Levy, A.S.; Carley, S. Cytotoxic activity of hexane extracts of Psidium guajava L. (myrtaceae) and Cassia alata L. (caesalpineaceae) in kasumi-1 and OV2008 cancer cell lines. Trop. J. Pharm. Res. 2012, 11, 201–207. [Google Scholar] [CrossRef]
  47. Ryu, N.H.; Park, K.R.; Kim, S.M.; Yun, H.M.; Nam, D.; Lee, S.G.; Jang, H.J.; Ahn, K.S.; Kim, S.-H.; Shim, B.S.; et al. A hexane fraction of guava leaves (Psidium guajava L.) induces anticancer activity by suppressing AKT/mammalian target of rapamycin/ribosomal p70 S6 kinase in human prostate cancer cells. J. Med. Food 2011, 15, 231–241. [Google Scholar] [CrossRef] [PubMed]
  48. Choi, S.Y.; Hwang, J.H.; Park, S.Y.; Jin, Y.J.; Ko, H.C.; Moon, S.W.; Kim, S.J. Fermented guava leaf extract inhibits LPS-induced COX-2 and iNOS expression in Mouse macrophage cells by inhibition of transcription factor NF-kappaB. Phyther. Res. 2008, 22, 1030–1034. [Google Scholar] [CrossRef] [PubMed]
  49. Jang, M.; Jeong, S.W.; Cho, S.K.; Ahn, K.S.; Kim, B.K.; Kim, J.C. Anti-inflammatory effects of 4 medicinal plant extracts in lipopolysaccharide-induced RAW 264.7 cells. Food Sci. Biotechnol. 2013, 22, 213–220. [Google Scholar] [CrossRef]
  50. Jang, M.; Jeong, S.-W.; Cho, S.K.; Ahn, K.S.; Lee, J.H.; Yang, D.C.; Kim, J.-C. Anti-Inflammatory Effects of an Ethanolic Extract of Guava (Psidium guajava L.) Leaves In Vitro and In Vivo. J. Med. Food 2014, 17, 678–685. [Google Scholar] [CrossRef] [PubMed]
  51. Sen, S.S.; Sukumaran, V.; Giri, S.S.; Park, S.C. Flavonoid fraction of guava leaf extract attenuates lipopolysaccharide-induced inflammatory response via blocking of NF-κB signalling pathway in Labeo rohita macrophages. Fish Shellfish Immunol. 2015, 47, 85–92. [Google Scholar] [CrossRef] [PubMed]
  52. Jang, M.; Jeong, S.-W.; Cho, S.K.; Yang, H.J.; Yoon, D.-S.; Kim, J.-C.; Park, K.-H. Improvement in the anti-inflammatory activity of guava (Psidium guajava L.) leaf extracts through optimization of extraction conditions. J. Funct. Foods 2014, 10, 161–168. [Google Scholar] [CrossRef]
  53. Madduluri, S.; Sitaram, B. In vitro evaluation of anti inflammatory activity of methanolic and ethanolic leaf extracts of five indigenous plants in South India. Int. J. PharmTech Res. 2014, 6, 569–574. [Google Scholar]
  54. Laily, N.; Kusumaningtyas, R.W.; Sukarti, I.; Rini, M.R.D.K. The potency of guava Psidium guajava (L.) leaves as a Functional immunostimulatory ingredient. Procedia Chem. 2015, 14, 301–307. [Google Scholar] [CrossRef]
  55. Soman, S.; Rauf, A.A.; Indira, M.; Rajamanickam, C. Antioxidant and antiglycative potential of ethyl acetate fraction of Psidium guajava leaf extract in streptozotocin-induced diabetic rats. Plant Foods Hum. Nutr. 2010, 65, 386–391. [Google Scholar] [CrossRef] [PubMed]
  56. Wu, J.-W.; Hsieh, C.-L.; Wang, H.-Y.; Chen, H.-Y. Inhibitory effects of guava (Psidium guajava L.) leaf extracts and its active compounds on the glycation process of protein. Food Chem. 2009, 113, 78–84. [Google Scholar] [CrossRef]
  57. Wang, H.; Du, Y.-J.; Song, H.-C. α-Glucosidase and α-amylase inhibitory activities of guava leaves. Food Chem. 2010, 123, 6–13. [Google Scholar] [CrossRef]
  58. Deguchi, Y.; Miyazaki, K. Anti-hyperglycemic and anti-hyperlipidemic effects of guava leaf extract. Nutr. Metab. Lond. 2010, 7, 1–10. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, Z.; Kong, F.; Ni, H.; Mo, Z.; Wan, J.B.; Hua, D.; Yan, C. Structural characterization, α-glucosidase inhibitory and DPPH scavenging activities of polysaccharides from guava. Carbohydr. Polym. 2016, 144, 106–114. [Google Scholar] [CrossRef] [PubMed]
  60. Eidenberger, T.; Selg, M.; Krennhuber, K. Inhibition of dipeptidyl peptidase activity by flavonol glycosides of guava (Psidium guajava L.): A key to the beneficial effects of guava in type II diabetes mellitus. Fitoterapia 2013, 89, 74–79. [Google Scholar] [CrossRef] [PubMed]
  61. Lee, Y.; Lim, Y.; Kwon, O. Selected phytochemicals and culinary plant extracts inhibit fructose uptake in caco-2 cells. Molecules 2015, 20, 17393–17404. [Google Scholar] [CrossRef] [PubMed]
  62. Cheng, F.-C.; Shen, S.-C.; Wu, J.S.-B. Effect of guava (Psidium guajava L.) leaf extract on glucose uptake in rat hepatocytes. J. Food Sci. 2009, 74, H132–H138. [Google Scholar] [CrossRef] [PubMed]
  63. Basha, S.K.; Kumari, V.S. In vitro antidiabetic activity of Psidium guajava leaves extracts. Asian Pac. J. Trop. Dis. 2012, 2 (Suppl. S1), 98–100. [Google Scholar] [CrossRef]
  64. Basha, S.K.; Govindaraju, K.; Manikandan, R.; Ahn, J.S.; Bae, E.Y.; Singaravelu, G. Phytochemical mediated gold nanoparticles and their PTP 1B inhibitory activity. Colloids Surf. B Biointerfaces 2010, 75, 405–409. [Google Scholar] [CrossRef] [PubMed]
  65. Yoshitomi, H.; Qin, L.; Liu, T.; Gao, M. Guava leaf extracts inhibit 3T3-L1 adipocyte differentiation via activating AMPK. J. Nutr. Ther. 2012, 1, 107–113. [Google Scholar]
  66. Liu, C.-W.; Wang, Y.-C.; Hsieh, C.-C.; Lu, H.-C.; Chiang, W.-D. Guava (Psidium guajava Linn.) leaf extract promotes glucose uptake and glycogen accumulation by modulating the insulin signaling pathway in high-glucose-induced insulin-resistant mouse FL83B cells. Process Biochem. 2015, 50, 1128–1135. [Google Scholar] [CrossRef]
  67. Chang, W.-C.; Shen, S.-C. Effect of water extracts from edible myrtaceae plants on uptake of 2-(n-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)-2-deoxyglucose in TNF-α-treated FL83B mouse hepatocytes. Phytother. Res. 2013, 27, 236–243. [Google Scholar] [CrossRef] [PubMed]
  68. Anand, S.; Arasakumari, M.; Prabu, P.; Amalraj, A.J. Anti-diabetic and aldose reductase inhibitory potential of Psidium guajava by in vitro analysis. Int. J. Pharm. Pharm. Sci. 2016, 8, 271–276. [Google Scholar] [CrossRef]
  69. Owen, P.L.; Matainaho, T.; Sirois, M.; Johns, T. Endothelial cytoprotection from oxidized LDL by some crude melanesian plant extracts is not related to their antioxidant capacity. J. Biochem. Mol. Toxicol. 2007, 21, 231–242. [Google Scholar] [CrossRef] [PubMed]
  70. Hsieh, C.L.; Huang, C.N.; Lin, Y.C.; Peng, R.Y. Molecular action mechanism against apoptosis by aqueous extract from guava budding leaves elucidated with human umbilical vein endothelial cell (HUVEC) model. J. Agric. Food Chem. 2007, 55, 8523–8533. [Google Scholar] [CrossRef] [PubMed]
  71. Hsieh, C.-L.; Lin, Y.-C.; Yen, G.-C.; Chen, H.-Y. Preventive effects of guava (Psidium guajava L.) leaves and its active compounds against α-dicarbonyl compounds-induced blood coagulation. Food Chem. 2007, 103, 528–535. [Google Scholar] [CrossRef]
  72. Anyachukwu Irondi, E.; Olalekan Agboola, S.; Oboh, G.; Augusti Boligon, A.; Linde Athayde, M.; Shode, F.O. Guava leaves polyphenolics-rich extract inhibits vital enzymes implicated in gout and hypertension in vitro. J. Intercult. Ethnopharmacol. 2016, 5, 122–130. [Google Scholar] [CrossRef] [PubMed]
  73. Prabu, G.R.; Gnanamani, A.; Sadulla, S. Guaijaverin—A plant flavonoid as potential antiplaque agent against Streptococcus mutans. J. Appl. Microbiol. 2006, 101, 487–495. [Google Scholar] [CrossRef] [PubMed]
  74. Shafiei, Z.; Haji Abdul Rahim, Z.; Philip, K.; Thurairajah, N. Antibacterial and anti-adherence effects of a plant extract mixture (PEM) and its individual constituent extracts (Psidium sp., Mangifera sp., and Mentha sp.) on single- and dual-species biofilms. PeerJ 2016, 4, e2519. [Google Scholar] [CrossRef] [PubMed]
  75. Razak, F.A.; Othman, R.Y.; Rahim, Z.H.A. The effect of Piper betle and Psidium guajava extracts on the cell-surface hydrophobicity of selected early settlers of dental plaque. J. Oral Sci. 2006, 48, 71–75. [Google Scholar] [CrossRef] [PubMed]
  76. Fathilah, A.R.; Rahim, Z.H.A.; Othman, Y.; Yusoff, M. Bacteriostatic effect of Piper betle and Psidium guajava extracts on dental plaque bacteria. Pak. J. Biol. Sci. 2009, 12, 518–521. [Google Scholar] [CrossRef] [PubMed]
  77. Chandrashekar, B.R.; Nagarajappa, R.; Singh, R.; Thakur, R. An in vitro study on the anti-microbial effi cacy of ten herbal extracts on primary plaque colonizers. J. Young Pharm. 2014, 6, 33–39. [Google Scholar] [CrossRef]
  78. John, N.R.; Gala, V.C.; Sawant, C.S. Inhibitory effects of plant extracts on multi-species dental biofilm formation in vitro. Int. J. Pharm. Bio Sci. 2013, 4, 487–495. [Google Scholar]
  79. John, N.R.; Gala, V.C.; Bhagwat, A.M.; Sawant, C.S. HPTLC analysis of eclipta prostrata and Psidium guajava extracts and their effect on cell-surface hydrophobicity of a consortium of dental plaque isolates. Int. J. Pharm. Pharm. Sci. 2013, 5, 1–6. [Google Scholar]
  80. Shekar, C.; Nagarajappa, R.; Singh, R.; Thakur, R. Antimicrobial efficacy of Acacia nilotica, Murraya koenigii L. Sprengel, Eucalyptus hybrid, and Psidium guajava on primary plaque colonizers: An in vitro comparison between hot and cold extraction process. J. Indian Soc. Periodontol. 2015, 19, 174–179. [Google Scholar] [CrossRef] [PubMed]
  81. Kwamin, F.; Gref, R.; Haubek, D.; Johansson, A. Interactions of extracts from selected chewing stick sources with Aggregatibacter actinomycetemcomitans. BMC Res. Notes 2012, 5, 203. [Google Scholar] [CrossRef] [PubMed]
  82. Thalikunnil, S.T.; Ashok, A.; Sukesh, K. Screening of psidium gaujava for effective phytomedicines and study on its antibacterial effect against dental caries bacteria. Int. J. Pharm. Pharm. Sci. 2012, 4, 400–401. [Google Scholar]
  83. Jayakumari, S.; Anbu, J.; Ravichandiran, V.; Nithya, S.; Anjana, A.; Sudharani, D. Evaluation of toothache activity of methanolic extract and its various fractions from the leaves Psidium guajava Linn. Int. J. Pharma. Bio Sci. 2012, 3, 238–249. [Google Scholar]
  84. Ravi, K.; Divyashree, P. Psidium guajava: A review on its potential as an adjunct in treating periodontal disease. Pharmacogn. Rev. 2014, 8, 96–100. [Google Scholar] [PubMed]
  85. Chen, H.-H. Hepatoprotective effect of guava (Psidium guajava L.) Leaf extracts on ethanol-induced injury on clone 9 rat liver cells. Food Nutr. Sci. 2011, 2, 983–988. [Google Scholar] [CrossRef]
  86. Qa’dan, F.; Thewaini, A.-J.; Ali, D.A.; Afifi, R.; Elkhawad, A.; Matalka, K.Z. The antimicrobial activities of Psidium guajava and Juglans regia leaf extracts to acne-developing organisms. Am. J. Chin. Med. 2005, 33, 197–204. [Google Scholar] [CrossRef] [PubMed]
  87. Abubakar, E.M. The use of Psidium guajava Linn. in treating wound, skin and soft tissue infections. Sci. Res. Essay 2009, 4, 605–611. [Google Scholar]
  88. Padrón-Márquez, B.; Viveros-Valdez, E.; Oranday-Cárdenas, A.; Carranza-Rosales, P. Antifungal activity of Psidium guajava organic extracts against dermatophytic fungi. J. Med. Plants Res. 2012, 6, 5435–5438. [Google Scholar] [CrossRef]
  89. Suwanmanee, S.; Kitisin, T.; Luplertlop, N. In vitro screening of 10 edible thai plants for potential antifungal properties. Evid. Based Complement. Altern. Med. 2014, 2014, 138587. [Google Scholar] [CrossRef] [PubMed]
  90. You, D.-H.; Park, J.-W.; Yuk, H.-G.; Lee, S.-C. Antioxidant and tyrosinase inhibitory activities of different parts of guava (Psidium guajava L.). Food Sci. Biotechnol. 2011, 20, 1095–1100. [Google Scholar] [CrossRef]
  91. Han, E.H.; Hwang, Y.P.; Choi, J.H.; Yang, J.H.; Seo, J.K.; Chung, Y.C.; Jeong, H.G. Psidium guajava extract inhibits thymus and activation-regulated chemokine (TARC/CCL17) production in human keratinocytes by inducing heme oxygenase-1 and blocking NF-κB and STAT1 activation. Environ. Toxicol. Pharmacol. 2011, 32, 136–145. [Google Scholar] [CrossRef] [PubMed]
  92. Lee, D.-U.; Yeon Weon, K.; Nam, D.-Y.; Hyun Nam, J.; Kyung Kim, W. Skin protective effect of guava leaves against UV-induced melanogenesis via inhibition of ORAI1 channel and tyrosinase activity. Exp. Dermatol. 2016, 25, 977–982. [Google Scholar] [CrossRef] [PubMed]
  93. Abreu, P.R.C.; Almeida, M.C.; Bernardo, R.M.; Bernardo, L.C.; Brito, L.C.; Garcia, E.A.C.; Fonseca, A.S.; Bernardo-Filho, M. Guava extract (Psidium guajava) alters the labelling of blood constituents with technetium-99m. J. Zhejiang Univ. Sci. B 2006, 7, 429–435. [Google Scholar] [CrossRef] [PubMed]
  94. Han, E.H.; Hwang, Y.P.; Kim, H.G.; Park, J.H.; Choi, J.H.; Im, J.H.; Khanal, T.; Park, B.H.; Yang, J.H.; Choi, J.M.; et al. Ethyl acetate extract of Psidium guajava inhibits IgE-mediated allergic responses by blocking FcεRI signaling. Food Chem. Toxicol. 2011, 49, 100–108. [Google Scholar] [CrossRef] [PubMed]
  95. Pachanawan, A.; Phumkhachorn, P.; Rattanachaikunsopon, P. Potential of Psidium guajava supplemented fish diets in controlling Aeromonas hydrophila infection in tilapia (Oreochromis niloticus). J. Biosci. Bioeng. 2008, 106, 419–424. [Google Scholar] [CrossRef] [PubMed]
  96. Yin, X.L.; Li, Z.J.; Yang, K.; Lin, H.Z.; Guo, Z.X. Effect of guava leaves on growth and the non-specific immune response of Penaeus monodon. Fish Shellfish Immunol. 2014, 40, 190–196. [Google Scholar] [CrossRef] [PubMed]
  97. Adeyemi, S.O.; Akanji, M.A.; Oguntoye, S.A. Ethanolic leaf extract of Psidium guajava: Phyto-chemical and trypanocidal activity in rats infected with Trypanosoma brucei brucei. J. Med. Plants Res. 2009, 3, 420–423. [Google Scholar]
  98. Akanji, M.A.; Adeyemi, O.S.; Oguntoye, S.O.; Sulyman, F. Psidium guajava extract reduces trypanosomosis associated lipid peroxidation and raises glutathione concentrations in infected animals. EXCLI J. 2009, 8, 148–154. [Google Scholar]
  99. Rajendran, C.; Begam, M.; Kumar, D.; Baruah, I.; Gogoi, H.K.; Srivastava, R.B.; Veer, V. Antiplasmodial activity of certain medicinal plants against chloroquine resistant Plasmodium berghei infected white albino BALB/c mice. J. Parasit. Dis. 2014, 38, 148–152. [Google Scholar] [CrossRef] [PubMed]
  100. Gupta, P.; Birdi, T. Psidium guajava leaf extract prevents intestinal colonization of Citrobacter rodentium in the mouse model. J. Ayurveda Integr. Med. 2015, 6, 50. [Google Scholar] [PubMed]
  101. Geidam, Y.A.; Ambali, A.G.; Onyeyili, P.A.; Tijjani, M.B.; Gambo, H.I.; Gulani, I.A. Antibacterial efficacy of ethyl acetate fraction of Psidium guajava leaf aqueous extract on experimental Escherichia coli (O78) infection in chickens. Vet. World 2015, 8, 358–362. [Google Scholar] [CrossRef] [PubMed]
  102. Shittu, O.B.; Ajayi, O.L.; Bankole, S.O.; Popoola, T.O.S. Intestinal ameliorative effects of traditional Ogi-tutu, Vernonia amygdalina and Psidium guajava in mice infected with Vibrio cholera. Afr. Health Sci. 2016, 16, 620–628. [Google Scholar] [CrossRef] [PubMed]
  103. Jan, O.Q.; Kamili, N.; Ashraf, A.; Iqbal, A.; Sharma, R.K.; Rastogi, A. Haematobiochemical parameters of goats fed tannin rich Psidium guajava and Carissa spinarum against Haemonchus contortus infection in India. J. Parasit. Dis. 2013, 39, 1–8. [Google Scholar] [CrossRef] [PubMed]
  104. Seo, N.; Ito, T.; Wang, N.; Yao, X.; Tokura, Y.; Furukawa, F.; Takigawa, M.; Kitanaka, S. Anti-allergic Psidium guajava extracts exert an antitumor effect by inhibition of T regulatory cells and resultant augmentation of Th1 cells. Anticancer Res. 2005, 25, 3763–3770. [Google Scholar] [PubMed]
  105. Adeyemi, O.S.; Akanji, M.A.; Ekanem, J.T. Anti-anaemic properties of the ethanolic extracts of Psidium guajava in Trypanosoma brucei brucei Infected Rats. Res. J. Pharmacol. 2010, 4, 74–77. [Google Scholar] [CrossRef]
  106. Udem, S.C.; Anyanwu, M.U.; Obidike, R.I.; Udem, N.D. The effects of Psidium guajava Linn. (Myrtaceae) leaf chloroform extract on some hematological and biochemical parameters in mice. Comp. Clin. Pathol. 2011, 20, 47–51. [Google Scholar] [CrossRef]
  107. Raj, V.B.; Rao, R.M.; Kumar, R.K.; Srinivas, K. Analgesic effect of ethanol extracted leaves of Psidium guajava leaves in animal models. Res. J. Pharm. Biol. Chem. Sci. 2015, 6, 1796–1801. [Google Scholar]
  108. Stalin D., J. A study on the analgesic property of methanolic and aqueous extracts of dried leaf of Psidium guajava Linn. Int. J. Adv. Res. Pharm. Bio Sci. 2013, 3, 1–6. [Google Scholar]
  109. Akinloye, O.; Akinmoladun, A.C.; Farombi, E.O. Modulatory effect of Psidium guajava linn and ocimum gratissimum Linn on lipid profile and selected biochemical indices in rabbits fed high cholesterol diet. J. Complement. Integr. Med. 2010, 7. [Google Scholar] [CrossRef]
  110. Mesquita Freire, J.; Patto de Abreu, C.M.; da Silveira Duarte, S.M.; Borges Araújo de Paula, F.; Ribeiro Lima, A. Evaluation of the protective effect of guava fruits and leaves on oxidative stress. Acta Sci. Biol. Sci. 2014, 36, 35–40. [Google Scholar]
  111. Oh, W.K.; Lee, C.H.; Lee, M.S.; Bae, E.Y.; Sohn, C.B.; Oh, H.; Kim, B.Y.; Ahn, J.S. Antidiabetic effects of extracts from. Psidium Guajava 2005, 96, 411–415. [Google Scholar]
  112. Mansoori Bahrani, A.H.; Zaheri, H.; Soltani, N.; Kharazmi, F. Effect of the administration of Psidium guava leaves on blood glucose, lipid profiles and sensitivity of the vascular mesenteric bed to Phenylephrine in streptozotocin-induced diabetic rats. J. Diabetes Mellit. 2012, 2, 138–145. [Google Scholar] [CrossRef]
  113. Shen, S.-C.; Cheng, F.-C.; Wu, N.-J. Effect of guava (Psidium guajava Linn.) leaf soluble solids on glucose metabolism in type 2 diabetic rats. Phytother. Res. 2008, 22, 1458–1464. [Google Scholar] [CrossRef] [PubMed]
  114. Ogueri, C.C.; Elekwa, I.; Ude, V.C.; Ugbogu, A.E. Effect of aqueous extract of guava (Psidium guajava) leaf on blood glucose and liver enzymes in alloxan induced diabetic rats. Br. J. Pharm. Res. 2014, 4, 1079–1087. [Google Scholar] [CrossRef]
  115. Shakeera Banu, M.; Sujatha, K.; Sridharan, G.; Manikandan, R. Antihyperglycemic and antihyperlipidemic potentials of Psidium guajava in alloxan-induced diabetic rats. Asian J. Pharm. Clin. Res. 2013, 6, 88–89. [Google Scholar]
  116. Adeyemi, O.S.; Akanji, M.A. Biochemical changes in the kidney and liver of rats following administration of ethanolic extract of Psidium guajava leaves. Hum. Exp. Toxicol. 2011, 30, 1266–1274. [Google Scholar] [CrossRef] [PubMed]
  117. Adeyemi, O.S.; Akanji, M.A. Psidium guajava leaf extract: Effects on rat serum homeostasis and tissue morphology. Comp. Clin. Pathol. 2012, 21, 401–407. [Google Scholar] [CrossRef]
  118. Ademiluyi, A.O.; Oboh, G.; Ogunsuyi, O.B.; Oloruntoba, F.M. A comparative study on antihypertensive and antioxidant properties of phenolic extracts from fruit and leaf of some guava (Psidium guajava L.) varieties. Comp. Clin. Pathol. 2015, 25, 363–374. [Google Scholar] [CrossRef]
  119. Chiwororo, W.D.H.; Ojewole, J.A.O. Biphasic effect of Psidium guajava Linn. (Myrtaceae) leaf aqueous extract on rat isolated vascular smooth muscles. J. Smooth Muscle Res. 2008, 44, 217–229. [Google Scholar] [CrossRef] [PubMed]
  120. Olatunji-Bello, I.I.; Odusanya, A.J.; Raji, I.; Ladipo, C.O. Contractile effect of the aqueous extract of Psidium guajava leaves on aortic rings in rat. Fitoterapia 2007, 78, 241–243. [Google Scholar] [CrossRef] [PubMed]
  121. Kawakami, Y.; Hosokawa, T.; Morinaka, T.; Irino, S.; Hirano, S.; Kobayashi, H.; Yoshioka, A.; Suzuki-Yamamoto, T.; Yokoro, M.; Kimoto, M.; et al. Antiatherogenic effect of guava leaf extracts inhibiting leucocyte-type 12-lipoxygenase activity. Food Chem. 2012, 131, 1069–1075. [Google Scholar] [CrossRef]
  122. Takahashi, Y.; Otsuki, A.; Mori, Y.; Kawakami, Y.; Ito, H. Inhibition of leukocyte-type 12-lipoxygenase by guava tea leaves prevents development of atherosclerosis. Food Chem. 2015, 186, 2–5. [Google Scholar] [CrossRef] [PubMed]
  123. Soman, S.; Rajamanickam, C.; Rauf, A.A.; Indira, M. Beneficial effects of Psidium guajava leaf extract on diabetic myocardium. Exp. Toxicol. Pathol. 2013, 65, 91–95. [Google Scholar] [CrossRef] [PubMed]
  124. Livingston Raja, N.R.; Sundar, K. Psidium guajava Linn confers gastro protective effects on rats. Eur. Rev. Med. Pharmacol. Sci. 2012, 16, 151–156. [Google Scholar] [PubMed]
  125. Tende, J.A.; Eze, E.D.; Tende, Y.A.; Onaadepo, O.; Shaibu, A. Anti-ulcerogenic activity of guava (Psidium guajava) leaves extract in rats. Ann. Exp. Biol. 2013, 1, 6–9. [Google Scholar]
  126. Umana Uduak, E.; Timbuak, J.A.; Musa, S.A.; Ikyembe, D.T.; Abdurrashid, S.; Hamman, W.O. Ulceroprotective effect of methanol extract of Psidium guajava leaves on ethanol induced gastric ulcer in adult wistar rats. Asian J. Med. Sci. 2012, 4, 75–78. [Google Scholar]
  127. Jayakumari, S.; Anbu, J.; Ravichandiran, V.; Anjana, A.; Siva Kumar, G.M.; Maharaj, S. Antiulcerogenic and free radical scavenging activity of flavonoid fraction of Psidium guajava Linn leaves. Int. J. Pharm. Pharm. Sci. 2012, 4, 170–174. [Google Scholar]
  128. Ojewole, J.A.O.; Awe, E.O.; Chiwororo, W.D.H. Antidiarrhoeal activity of Psidium guajava Linn. (Myrtaceae) leaf aqueous extract in rodents. J. Smooth Muscle Res. 2008, 44, 195–207. [Google Scholar] [CrossRef] [PubMed]
  129. Ezekwesili, J.O.; Nkemdilim, U.U.; Okeke, C.U. Mechanism of antidiarrhoeal effect of ethanolic extract of Psidium guajava leaves. Biokemistri 2010, 22, 85–90. [Google Scholar]
  130. Shah, A.J.; Begum, S.; Hassan, S.I.; Ali, S.N.; Siddiqui, B.S.; Gilani, A.-H. Pharmacological basis for the medicinal use of Psidium guajava leave in hyperactive gut disorders. Bangladesh J. Pharmacol. 2011, 6, 100–105. [Google Scholar] [CrossRef]
  131. Roy, C.K.; Kamath, J.V.; Asad, M. Hepatoprotective activity of Psidium guajava Linn. leaf extract. Indian J. Exp. Biol. 2006, 44, 305–311. [Google Scholar] [PubMed]
  132. D’Mello, P.; Rana, M. Hepatoprotective activity of Psidium guajava extract and its phospholipid complex in paracetamol induced hepatic damage in rats. Int. J. Phytomed. 2010, 2, 85–93. [Google Scholar]
  133. Taju, G.; Jayanthi, M.; Majeed, S.A. Evaluation of hepatoprotective and antioxidant activity of Psidium guajava leaf extract against acetaminophen induced liver injury in rats. Int. J. Toxicol. Appl. Pharmacol. 2011, 1, 13–20. [Google Scholar]
  134. Osaman, M.; Ahamad, M.; Mahfouz, S.; Elaby, S. Biochemical studies on the hepatoprotective effects of pomegranate and guava ethanol extracts. N. Y. Sci. J. 2011, 4, 1–15. [Google Scholar]
  135. Mohamed, E.A.K. Hepatoprotective effect of aqueous leaves extract of Psidium guajava and Zizyphus spina-christi against paracetamol induced hepatotoxicity in rats. J. Appl. Sci. Res. 2012, 8, 2800–2806. [Google Scholar]
  136. Choi, J.H.; Park, B.H.; Kim, H.G.; Hwang, Y.P.; Han, E.H.; Jin, S.W.; Seo, J.K.; Chung, Y.C.; Jeong, H.G. Inhibitory effect of Psidium guajava water extract in the development of 2,4-dinitrochlorobenzene-induced atopic dermatitis in NC/Nga mice. Food Chem. Toxicol. 2012, 50, 2923–2929. [Google Scholar] [CrossRef] [PubMed]
  137. Porta Santos Fernandes, K.; Kalil Bussadori, S.; Martins Marques, M.; Sumie Wadt Yamashita, N.; Bach, E.; Domingues Martins, M. Healing and cytotoxic effects of Psidium guajava (Myrtaceae) leaf extracts. Braz. J. Oral Sci. 2010, 9, 9–14. [Google Scholar]
  138. Baroroh, H.N.; Utami, E.D. Harwoko inhibitory effect of ethanolic extract of Psidium guajava leaves in rat active cutaneus anaphylaxis reaction. Int. J. Pharm. Clin. Res. 2016, 8, 1–5. [Google Scholar]
  139. Wang, X.; Ye, K.; Lv, Y.; Wei, S.; Li, X.; Ma, J.; Zhang, X.; Ye, C. Ameliorative effect and underlying mechanisms of total triterpenoids from Psidium guajava Linn (myrtaceae) leaf on high-fat streptozotocin-induced diabetic peripheral neuropathy in rats. Trop. J. Pharm. Res. 2016, 15, 327–333. [Google Scholar] [CrossRef]
  140. Kumar, A.; Kumarchandra, R.; Rai, R.; Rao, B. Radiomodulatory role of Psidium guajava leaf extracts against X-ray induced genotoxicity, oxidative stress and apoptosis in albino wistar rat model. J. Appl. Pharm. Sci. 2016, 6, 58–65. [Google Scholar] [CrossRef]
  141. Ekaluo, U.B.; Ikpeme, E.V.; Uno, U.U.; Umeh, S.O.; Erem, F.A. Protective role of aqueous guava leaf extract against caffeine induced spermatotoxicity in albino rats. Res. J. Med. Plant 2016, 10, 98–105. [Google Scholar] [CrossRef]
  142. Kaneko, K.; Suzuki, K.; Iwadate-Iwata, E.; Kato, I.; Uchida, K.; Onoue, M. Evaluation of food-drug interaction of guava leaf tea. Phyther. Res. 2013, 27, 299–305. [Google Scholar] [CrossRef] [PubMed]
  143. Vladislavovna Doubova, S.; Reyes Morales, H.; Flores Hernández, S.; Martínez-García, M.D.C.; González de Cossío Ortiz, M.; Chávez Soto, M.A.; Rivera Arce, E.; Lozoya, X. Effect of a Psidii guajavae folium extract in the treatment of primary dysmenorrhea: A randomized clinical trial. J. Ethnopharmacol. 2007, 110, 305–310. [Google Scholar] [CrossRef] [PubMed]
  144. Ravindranath, D.; Nayana, O.V.; Thomas, L. Antimicrobial plant extract incorporated gelatin beads for potential application in pharmaceutical industry. Biotechnol. Res. 2016, 2, 11–14. [Google Scholar]
  145. Giri, S.S.; Sen, S.S.; Chi, C.; Kim, H.J.; Yun, S.; Park, S.C.; Sukumaran, V. Effect of guava leaves on the growth performance and cytokine gene expression of Labeo rohita and its susceptibility to Aeromonas hydrophila infection. Fish Shellfish Immunol. 2015, 46, 217–224. [Google Scholar] [CrossRef] [PubMed]
  146. Fawole, F.J.; Sahu, N.P.; Pal, A.K.; Ravindran, A. Haemato-immunological response of Labeo rohita (Hamilton) fingerlings fed leaf extracts and challenged by Aeromonas hydrophila. Aquac. Res. 2016, 47, 3788–3799. [Google Scholar] [CrossRef]
  147. Gobi, N.; Ramya, C.; Vaseeharan, B.; Malaikozhundan, B.; Vijayakumar, S.; Kadarkarai, M.; Benelli, G. Oreochromis mossambicus diet supplementation with Psidium guajava leaf extracts enhance growth, immune, antioxidant response and resistance to Aeromonas hydrophila. Fish Shellfish Immunol. 2016, 58, 572–583. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Main traditional uses of guava leaves in the principal producer countries.
Figure 1. Main traditional uses of guava leaves in the principal producer countries.
Ijms 18 00897 g001
Table 1. In vitro assays against infectious and parasitic diseases.
Table 1. In vitro assays against infectious and parasitic diseases.
OriginExtraction MethodMajor ConstituentMicroorganismsAssayMain ResultsRef.
Saudi ArabiaDecoction (30 min)-Staphylococcus aureus strainsAgar well diffusion assayAt 200 µL: iz ≤ 30 mm.[13]
IndiaSoxhlet with MeOH (12 h), maceration in H2O (4 h)-S. aureus strainsAgar well diffusion assay, time-kill of bacterial cell, SDS-PAGE analysis, and cellular toxicity to human erythrocytes assaysAt 20 mg/L: iz ≤ 20 mm, MIC: 25 µg/mL (MeOH) and 7.5 mg/mL (H2O). MBC: 1.25 and 12.5 mg/mL, respectively, 10 h to kill bacteria, ↑ degradation of protein, no hemolysis.[14]
NigeriaMaceration in MeOH (48 h)-S. aureus, Escherichia coli, Pseudomonas aeruginosa, Proteus spp., and Shigella spp.Agar well diffusion assayAt 20 mg/mL: iz ≤ 18 mm; 81.8% prevention growth.[15]
IndiaMaceration with agitation in MeOH, Ac, and DMF (12 h)-G-p and G-n bacteria and fungi (91 clinically important strains)Disc diffusion assayAt 25 mg/mL: against g-p 70% MeOH > 80% Ac > 50% DE, ↓ 76.36% g-n bacteria. Fungi 56% Ac > 38% ME > 31% DMF. No activity against Citrobacter spp., Alcaligenes fecalis, and Aspergillus spp.[16]
IndiaSoxhlet with MeOH (4 h)Phytochemical screening: mainly flavonoid-glycosides and tanninsBacteria (Bacillus subtilis, S. aureus and E. coli), and fungi (Candida albicans and Aspergillus niger)Paper disc diffusion assayAt 50 μg/mL: iz ≤ 12.6 mm and 10 mm for bacterial and fungi strains, respectively. E. coli: MIC 0.78 μg/mL, MBC 50 μg/mL, and MFC 12.5 μg/mL.[17]
BrazilMaceration with stirring in EtOH:H2O 70% (v/v) (50 °C, 1 h)TPC: 25.93 (% m/m, dry base), TFC: 23.48 (mg/g, dry base)Fungi (C. albicans, Candida krusei, and Candida glabrata), G-p (S. aureus) and G-n (E. coli and P. aeruginosa)Microdilution assayMIC = 80–100 µg/mL (C. krusei, C. glabrata and S. aureus) and MBC, MFC ≤ 250–1000 µg/mL (the others).[18]
BrazilTurbo-extraction with water or Ac:H2O 70% (v/v) (20 min)Gallic acid: 0.065 µg/g, Catechin: 1.04 µg/gG-p strains (S. aureus, Staphylococcus epidermidis, and Enterococcus faecalis) G-n (E. coli, Salmonella enteritidis, Shigella flexneri, and Klebsiella pneumoniae)Agar-diffusion and microdilution assaysAt 5 mg/mL: iz ≤ 20 mm, MIC = 39 μg/mL (S. epidermis), MIC < 600 μg/mL (the others).[19]
IndiaSoxhlet with n-hexaneMethyl 2,6,10-trimethyltridecanoate (28.86%) and Methyl octadecanoate (22.18%)G-p: S. aureus, Streptoccocus faecalis, Bacillus subtillis, Lactobacillus spp., Enterococcus aerogenes, Acinetobacter spp. G-n: E. coli, Proteus vulgari, Enterobacter aerogenes, Salmonella typhimurium, P. aeruginosa, and K. pneumoniaeAgar well diffusion assayAt 80 µL: iz ≤ 27 mm, MIC = 3–10 µL.[20]
BrazilMaceration in EtOH:H2O 70% (v/v) (72 h)-E. coli, P. aeruginosa, and S. AureusMicrodilution assayOnly S. aureus (MIC = 256 mg/mL). Synergic effect with ciprofloxacin and gentamicin at 1024 mg/mL.[21]
BrazilMaceration in MeOH:H2O 70% (v/v) (48h)-S. aureus strainsDisc diffusion assayMIC 90% = 0.52 mg/mL, at 131.75 mg/mL synergic effect with tetracycline, chloramphenicol, erythromycin, vancomycin, oxacillin, cephalothin, ampicillin, cefoxitin, cotrimoxazole.[22]
EgyptMaceration in EtOH:H2O 50% (v/v)Quercetin, avicularin, guajaverin, isoquercitrin, hyperinS. aureus, E. coli, P. aeruginosa, and C. albicansAgar well diffusion assayS. aureus: ↑ iz quercetin (28 mm), MIC (μg/mL) guajaverin (0.09–0.19) < avicularin (0.09–0.38) < quercetin (1.25) for all the microorganism tested.[23]
IndonesiaMaceration in EtOH:H2O 30% (v/v) (3 days)Tannins (2.35 mg/g)E. coli, P. aureginosa, S. aureus, A. niger and C. AlbicansPaper disc diffusion methodiz ≤ 15 mm.[24]
IndiaSoxhlet with toluene (72 h)Betulinic acid and lupeolFungi: Calletotricheme camellie, Fussarium equisitae, Alterneria alternate, Curvularia eragrostidies, and Colletrichum Gleosproides. Bacteria: E. Coli, B. Subtillis, S. aureus, and EnterobactorSlide germination methodBacteria: MIC < 100–200 μg/mL, fungi: MIC < 2.5–10 μg/mL.[25]
IndiaSoxhlet with Ac (8 h)-H. bispinosa Neumann (Acarina: Ixodidae) and H. maculata Leach (Diptera: Hippoboscidae)Antiparasitic activity method of FAO (2004)At 3 mg/mL: mortality 100% H. maculate, 78% H. bispinosa, parasite dead H. maculata (LC50 = 0.646 mg/mL).[26]
NigeriaMaceration with agitation in EtOH:H2O 20% and 80% (v/v) (24 h)-Trypanosoma brucei brucei and HEK293Alamar Blue assaysAt 238.10 μg/mL: IC50 (T. b. brucei) = 6.3 μg/mL and 48.9 μg/mL for 80% and 20% extracts, respectively, IC50 (HEK293) 30.1 and 24.16%, respectively.[27]
IndiaSoxhlet with ethyl acetate and MeOH (8 h)-Plasmodium falciparum strainsSYBR green assayIC50 9–18 μg/mL, resistance indices = 0.6 and 1.4 in MeOH and ethyl acetate, respectively.[28]
MalaysiaHydrodistillation (3 h)-Toxoplasma gondiiMTT assay with Vero cellsAt 200 µg/mL: No cytotoxic effect (EC50 = 37.54 µg/mL), anti-parasitic activity (EC50 of 3.94 µg/mL).[29]
IndiaSoxhlet with EtOH, and maceration in H2O (6 days)-E. coli, Shigella spp., Salmonella spp., Aeromonas spp., S. aureus, and Candida spp.Agar well diffusion assayH2O: iz ≤ 30 mm (max C. albicans). EtOH: iz ≤ 31 mm (max Aeromonas hydrophila).[30]
IndiaSoxhlet with EtOH:H2O 70% (v/v), MeOH, ethyl acetate, and H2OPhytochemical analysis: tannins, saponins, flavonoids, terpenoids, sugarsE. coli, Salmonella spp., and Vibrio choleraeAgar well diffusion assayAt 1000 µg/mL: iz ≤ 30 mm. MeOH: MIC (100%) > 250 µg/mL. EtOH:H2O:MICs (38–65%) > 500 µg/mL and > 750 µg/mL. Ethyl acetate and H2O: MICs > 750 µg/mL.[31]
BangladeshMaceration in H2O and MeOH:H2O 75% (v/v) (48 h)-V. choleraAgar well diffusion assayMICs = 1250 µg/mL (H2O), 850 µg/mL for (MeOH:H2O). Antibacterial resistance to trimethoprim/sulfomethoxazole, furazolidone, tetracycline, and erythromycin.[32]
IndiaDecoctionMajor component: quercetin (2 mg/g)E. coli (heat labile (HLT) and cholera toxin (CT)), V. cholerae, Shigella flexneriMicrotitre plate based assay. Assays for bacterial colonization (adherence and invasion) and enterotoxinsAt 2.7 mg/mL: (EC50 = 0.98 (S. flexneri) and 2.88% (V. cholerae). ↓ adherence and invasion to epithelial cells (EC50 = 0.37–1.25% and 0.04–0.25%, respectively). The effect on adherence is not due to quercetin and the invasion is lower than with the extract. ↓ Production of HLT and CT (EC50 = 1.03 and 2.69%) and binding to glioside monosialic acid enzyme (EC50 = 0.06 and 2.51%).[33]
BrazilSoxhlet with n-hexane, ethyl acetate, MeOH, H2O (24 h)-S. aureus, Salmonella spp., and E. coliDisc diffusion methodAt 1938 µg/disc: iz = 7.00–11.25 mm (Soxhlet), and 11–18 mm (H2O). No inhibition to E. coli (H2O) and Salmonella spp. (hexane and ethyl acetate).[34]
NigeriaSoxhlet with EtOH:H2O 60% (v/v) (5 h), and H2O (3 h)-E. coli and S. aureusAgar well diffusion assayAt 10 mg/mL: H2O: iz = 9–16 mm and 8–11 mm, MICs = 5 and 2.5 mg/mL (E. coli and S. aureus, respectively). EtOH:H2O: iz 12–21 and 11–14 mm, MICs = 1.25 and 0.625 mg/mL, respectively.[35]
JapanInfusion (8 min)Tannin content: 1.11 mg/mLH1N1 virus strains19-h Influenza growth inhibition assayAt 0.4 mg/mL: inhibition growth (IC50 = 0.05–0.42%).[36]
Acetone (Ac); N,N-dimethylformamide (DMF); dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE); effective concentration (EC50); inhibition zone (iz); inhibitory concentration (IC50); lethal concentration (LC50); minimum bactericidal concentration (MBC); minimum fungicidal concentration (MFC); minimum inhibitory concentration (MIC); total flavonoid content (TFC); total phenolic content (TPC); Tetrazolium (MTT); ↑ increases the affect; ↓ decreases the effect.
Table 2. In vitro studies against neoplasm.
Table 2. In vitro studies against neoplasm.
OriginExtraction MethodMajor ConstituentCell LineAssayMain ResultsRef.
JapanMaceration in EtOH:H2O 50% (v/v)TPC: 71 g/100 gHuman colon adenocarcinoma (COLO320DMA)Cyclooxygenase and cell proliferation assaysAt 1 mg/mL: ↓ human cyclooxygenase activity (IC50 55 and 560 µg/mL PGHS-1 and 2, respectively), ↓ IC50 5.1 µg/mL (PGSH) and 4.5 µg/mL (cyclooxygenase).[37]
At 100 µg/mL: Quercetin ↓ IC50 = 5.3 (PGSH-1) and 250 µg/mL (PGSH-2), ↓ DNA synthesis rate.
MalaysiaSoxhlet with ether, MeOH, and H2O-Cervical cancer (HeLa), breast cancer (MDA-MB-231) and osteosarcoma (MG-63). Control: non-malignant Madin-Darby canine kidney (MDCK)Methylene blue assayAt 10 mg/mL: HeLa: No anti-proliferative activity.[38]
MDA-MB-231: IC50 ether extract (4.2 µg/mL) > MeOH (18.6 µg/mL) > H2O (55.7 µg/mL).
MG-63: same order (IC50 of 5.42, 23.25, and 61.88 µg/mL, respectively).
MDCK: cytotoxic effect of ether and MeOH extract (IC50 = 5.03 and 11.55 µg/mL, respectively).
BrazilMaceration in EtOHTPC: 766.08 mg/g, TFC: 118.90 mg/gHeLa, colorectal carcinoma (RKO-AS45-1), and lung fibroblast (Wi-26VA4)MTT assayAt 1 mg/mL: IC50 = 15.6 µg/mL (HeLa), 21.2 (RKO) µg/mL, and 68.9 µg/mL (Wi-26VA4).[39]
PalestineMaceration in DCM:MeOH 50% (v/v) (24 h)-Murine fibrosarcoma (L929sA), and human breast cancer (MDA-MB-231 and MCF-7)MTT assayIC50 = 55 µg/mL (L929sA), 820 µg/mL (MCF7 cells), no cytotoxic effect on MDA-MB-231 cells.[40]
TaiwanDecoction (30 min)-Human prostate carcinoma (DU-145)MTT, ELISA, gelatinolytic zymography, wound scratch, and chicken chorioallantoic membrane assaysAt 0.25 mg/mL: cell suppression (IC50 0.57 mg/mL). ↓ Expressions of VEGF (76.9%), IL-6 (98.8%) and IL-8 (98%), and MMP-2 (100%) and MMP-9 (100%). Suppressed the cell migration (30.9%) and the angiogenesis.[41]
TaiwanDecoction (1 h)TPC: 470.0 mg/g Individual compounds: gallic acid (348), catechin (102), epicatechin (60), rutin (100), quercetin (102), and rutin (100) in mg/gHuman prostate epithelial (PZ-HPV-7) and DU-145MTT assayAt 1 mg/mL: 100% suppression DU-145 cells. PZ-HPV-7 cells followed an auto-decaying process. Cell-killing rate coefficient (kapp) = 0.03 × 103 phenolic compounds cells/mg h.[42]
BrazilSoxhlet with DCM. Maceration with EtOHGuajadial, psidial A, and psiguajadial A and BLeukemia (K-562), MCF-7, ovarian cancer (NCI/ADR-RES), lung (NCI-H460), melanoma (UACC-62), prostate (PC-3), colon (HT-29), ovarian (OVCAR-3), and kidney (786-0)Protocol established by NCI (ELISA test)At 250 µg/mL: Anti-proliferative activity DCM > EtOH, inhibition growths: 26 (OVCAR-3)-65 (UACC-62) µg/mL due to the major compounds.[43]
JapanMaceration with sonication in MeOH:H2O 80% (v/v) (3 h) and isolation CPOPC-3 and MCF-7MTT, annexin V antibody, TUNEL, and western blot assaysAt 50 µg/mL: ↓ cell proliferation, ↑ early and late apoptotic effect, down-regulation of PI3K/AKT/mTOR/S6K1 pathway, up-regulation of MAPKs, JNK, ERKs, and p38 MAPK.[44]
ThailandHydrodistillation Human mouth epidermal carcinoma (KB) and murine leukemia (P388)MTT assayAt 0.15 mg/mL: KB: 75% cytotoxic effect, IC50 = 0.04 mg/mL; At 0.08 mg/mL: P388: 80% cytotoxic effect, IC50 = 0.05 mg/mL.[45]
JamaicaMaceration in hexane (4 days)-Leukemia (Kasumi-1)MTT assayIC50 = 200 µg/mL.[46]
JapanMaceration with sonication in MeOH:H2O 80% (v/v) (3 h). Fractionation with hexane60 compounds (in hexane fraction): β-eudesmol (11.98%), α-copaene (7.97%), phytol (7.95%), α-patchoulene (3.76%), and CPO (3.63%)Human prostate cancer (PC-3 and LNCaP)MTT, annexin V antibody, TUNEL, and western blot assaysAt 150 µg/mL: PC-3: ↑ apoptotic effect of the hexane fraction (15%), ↓ effect on early apoptotic cells, ↑ effect for late apoptosis, via the suppression of PI3K/AKT/mTOR/S6K1 and MAPK signalling cascades in both cell lines.[47]
β-Caryophyllene oxide (CPO); c-jun NH2-terminal kinases (JNK); dichloromethane (DCM); inhibitory concentration (IC50); mammalian target of rapamycin (mTOR); mitogen-activated protein kinases (MAPKs); phosphatidylinositol 3-kinase (PI3K); prostaglandin endoperoxide H synthase (PGHS); protein kinase B (AKT); ribosomal protein S6 kinase beta-1 (S6K1); signal-related kinases (ERKs); tetrazolium (MTT); total flavonoid content (TFC); total phenolic content (TPC); ↑ increases the affect; ↓ decreases the effect.
Table 3. In vitro assays against diseases of the blood and immune system.
Table 3. In vitro assays against diseases of the blood and immune system.
OriginExtraction MethodMajor ConstituentCellsAssayMain ResultsRef.
KoreaMaceration in MeOH:H2O 70% (v/v) (5 days)-LPS-stimulated RAW 264.7 (Mouse macrophage)Griess, MTT, ELISA kit, western blot, transient transfection, and luciferase assaysAt 125 µg/mL: no cytotoxic effect, ↑ 44–62% inhibition rates. ↓ LPS-induced NO and PEG2 ↓ iNOS and COX-2 (↓ I-κBα degradation, ↓ activation NF-κB).[48]
PalestineMaceration in DCM:MeOH 50% (v/v) (24 h)-L929sA fibroblastTransfection and luciferase assaysAt 62.5 µg/mL: ↓ expression of IL-6 and NF-κB luciferase reporter gene construct via the NF-κB transactivation level, since no ↓ inhibition of NF-κB/DNA binding.[40]
KoreaExtraction in MeOH:H2O 70% (v/v) (6 h)TPC: 426.84 mg (GAE)/gLPS-stimulated RAW 264.7 MTT, Griess, and ELISA test assaysAt 30 µg/mL: no cytotoxic effect. ↓ LPS-induced NO (52.58%) and the production of PGE2 (43.45).[49]
KoreaExtraction in EtOH:H2O 55% (v/v) (4.9 h, 47 °C)Gallic acid (0.2) and catechin (4.4) in mg/gLPS-stimulated RAW 264.7 MTT, Griess, ELISA test, RT-PCR, and total western blot assaysAt 50 µg/mL: no cytotoxic effect. ↓ LPS-induced NO (>65%) by ↓ iNOS, ↓ PGE2 (to basal level) via ↓ COX-2 mRNA. ↓ IL-6. ↓ iNOS and COX-2 due to the down-regulation of ERK1/2 pathway, because no effect was found to other proteins at the dose tested.[50]
IndiaMaceration in MeOH:H2O 90% (v/v) (x3)-LPS-stimulated in Labeo rohita head-kidney macrophagesMTT, Greiss, ELISA, RT-PCR, and western blot assaysAt 200 µg/mL, ↓ LPS-induced NO (75%) by ↓ iNOS-mRNA, ↓ PGE2 (45%) via ↓ production COX-2-mRNA, TNF-α, IL-1β, IL-10, and mRNA expression. Suppressed phosphorylation of MAPK (↓ I-κBα degradation ↓ activation NF-κB).[51]
KoreaSoxhlet with EtOH:H2O 55% (v/v) (4.9 h, 47 °C)Gallic acid (0.09) and catechin (0.72) in mg/gLPS-stimulated RAW 264.7MTT, Greiss and ELISA test assaysAt 30 µg/mL: no cytotoxic effect. ↓ LPS-induced NO (47.5%) and PGE2 (45.8).[52]
IndiaMaceration with agitation in MeOH and EtOH (24 h)-Human bloodHRBC membrane stabilization methodAt 200 µg/mL: ↑ 13.8–14.4% prevention of lysis of the membrane.[53]
IndonesiaMaceration with agitation in EtOH:H2O 96% (v/v) (6 h)TPC: 101.93 mg GAE/gHuman lymphocyteMTT assay0.5 μg/mL: Stimulation index 1.54%.[54]
Cyclooxygenase-2 (COX-2); dichloromethane (DCM); gallic acid equivalent (GAE); human red blood cell (HRBC); inducible nitric oxide synthase (iNOS); inhibitor of kappa B (I-κBα); interleukin-1β (I-1β); lipopolysaccharide (LPS); mitogen-activated protein kinases (MAPKs); nitric oxide (NO); prostaglandin E2 (PEG2); reverse transcription-polymerase chain reaction RT-PCR; tetrazolium (MTT); total phenolic content (TPC); transcriptional nuclear factor-κB (NF-κB); Tumor necrosis factor alpha (TNF-α); ↑ increases the affect; ↓ decreases the effect.
Table 4. Compounds in guava leaves with anti-diabetic properties in in vitro assays.
Table 4. Compounds in guava leaves with anti-diabetic properties in in vitro assays.
OriginCompoundAssayMain ResultsRef.
IndiaEthyl acetate fractionIn vitro glycation of BSA-fluorescence measurementIn vitro AGEs formation with IC50 of 38.95 ± 3.08 μg/mL.[55]
TaiwanGallic acid, catechin and quercetinIn vitro glycation of BSA-fluorescence measurement; Fructosamine assay and Girard-T assayAt 100 μg/mL: 80% inhibitory effects on the formation of α-dicarbonyl compounds at a concentration of 50 µg/mL, inhibitory effects on AGEs formation in BSA glycation systems.[56]
ChinaQuercetin, kaempferol, myricetinRat intestinal sucrase and maltase inhibitory activities; Porcine pancreatic α-amylase inhibitory activityAt 1.5 mg/mL: inhibitory activities with IC50 values of 3.5, 5.2 and 3.0 mM against sucrase, with IC50 values of 4.8, 5.6 and 4.1 mM against maltase and with IC50 values of 4.8, 5.3 and 4.3 mM against α-amylase, respectively. Synergistic effect against α-glucosidase.[57]
ChinaWater-soluble polysaccharides, including GP90 and P90α-Glucosidase inhibition assayα-Glucosidase inhibition activity with an EC50 of 2.27 µg/mL and 0.18 mg/mL.[59]
-Peltatoside, hyperoside, isoquercitrin, guaijaverin and flavonol-glycosidesSpectrophotometric assay; absorption assay into CaCo-2 cellsConcentration of the compounds (0.01 to 0.06 µmol/mL). Individual flavonol-glycosides inhibited DP-IV dose-dependently. The ethanolic guava leaves extract (380 µg/mL) showed a dose-dependent inhibition of DP-IV, with an IC50 of 380 μg/mL test assay solution; the highest uptake was from Guaijaverin.[60]
KoreaQuercetin and catechinFructose transport in CaCo-2 cell systemsAt 1 mg/mL: inhibition of fructose uptake (55%). At 30 µg/mL: quercetin contributed to both, GLUT2 and 5 transporters, and catechin to GLUT5-mediated fructose uptake inhibition.[61]
IndiaGuavanoic acidSpectrophotometric assayAt 27 μg/mL: remarkable PTP1B inhibitory activity (90%) and in vitro stability in various physiological medium including saline, histidine, cysteine, BSA, HSA and buffers (pH 5, 7 and 9). IC50 = 1.14 μg/mL.[64]
Indian-Hexane, methanol, ethanol and aqueous leaf extractsInhibitory glucose diffusionAt 50 g/L: the methanol extract was the most potent with the lowest mean glucose concentration of 201 ± 1.69 mg/dL at the end of 27 h (↓ 93% uptake).[63]
Japan70% Ethanol extractOil Red O Assay; Real-Time RTAt 100 μg/mL: inhibition of 3T3-L1 differentiation via down-regulation of adipogenic transcription factors and markers (mRNA levels of PPAR-γ, C/EBP-α, and aP2), and suppression of mitotic clonal expansion (at day 4 and 8).[65]
TaiwanAqueous extractGlucose uptake test; bicinchonic acid method; Western-blot analysisAt 400 µg/mL: ↑ IR (25.1%), p-IR (46.2%), p-IRS (51.2%), PI3K (32.2%), Akt (46.1%), p-Akt (36.3%), GLUT-2 (46.8%), and total glycogen synthase (45.5%).[66]
TaiwanVescalaginGlucose-uptake testAt 100 µg/mL: Enhancement of glucose uptake in TNF-α-induced insulin-resistant.[67]
Advanced glycation end products (AGEs); bovine serum albumin (BSA), dipeptidyl peptidase (DP); effective concentration (EC50); glucose transporter 2 and 5 (GLUT-2; GLUT-5); human serum albumin (HSA); inhibitory concentration (IC50); insulin receptor (IR); insulin receptor substrate (p-IRS (Tyr)); p85 regulatory subunit of phospho-inositide 3 kinase (PI3K (p85)); phosphorylation of the insulin receptor (p-IR (Tyr)); protein kinase B (p-Akt (Ser)); tumor necrosis factor (TNF); ↓ decreases the effect.
Table 5. In vitro assays against diseases related to the digestive system.
Table 5. In vitro assays against diseases related to the digestive system.
OriginExtraction MethodMicroorganism(s)/CellsAssayMain ResultsRef.
IndiaSoxhlet with MeOH (4.5 h)S. mutans strainsAgar well diffusion assay, effect on acid production, on sucrose-dependent adherence to smooth glass surfaces, and on sucrose-induced cellular aggregation, and MATH assaysMIC > 5 mg/mL (MeOH). MIC = 2–4 mg/mL (guaijaverin)
At sub-MIC (0.125–2 mg/mL): ↑ pH (5 to 6–7), hydrophobicity indexes (3.2–72%), ↓ sucrose-dependent adherence (34–84%) and aggregation.
[73]
MalaysiaDecoctionS. sanguinis and S. mutansNAM model systemAt 60.95 mg/mL: MIC = 7.62 (S. sanguinis) and 3.81(S. mutans.) mg/mL. MBC values = 15.24 and 30.48 mg/mL, respectively. At 0.5 mg/mL: ↓ adherence 57 and 60% (single-species) and 88–89% (dual-species).[74]
MalaysiaSonication with H2O (10 min)S. sanguinis, S. mitis, and Actinomyces spp.MATH assayAt 1 mg/mL: ↓ 54.1%, 49.9% and 40.6%, respectively, cell-surface hydrophobicity. At 20 mg/mL: was 64.7, 60.5, and 55.5%, respectively.[75]
MalaysiaDecoctionS. sanguinis, S. mitis, and Actinomyces spp.Bacterial growth and generation time rates determinationsAt 4 mg/mL: Time growth = 1.22 (S. sanguinis, Actinomyces spp) and 2.06 h (S. mitis) ↓ growth 42.6%, 51.2% and 55%.[76]
IndiaMaceration with stirring in EtOH (2 days)S. mutans, S. sanguinis, and S. salivariusAgar well diffusion assayAt 10 mg/mL: inhibition zones of 21.17, 18.58, and 23.00 mm, respectively.[77]
IndiaMaceration (2 days) and Soxhlet (6 h) with EtOH, H2O, and EtOH:H2O 50% (v/v)S. mutans and S. mitisAgar well diffusion assay, sucrose-dependent adherence and cellular co-aggregation activities, and biofilm formation sterile acrylic tooth determinationsAt 15 mg/mL: inhibition zone for H2O (11.8 mm) to EtOH:H2O (25 mm), both by Soxhlet. MIC = 1 mg/mL.
EtOH:H2O extract: at >0.05 mg/mL: ↓ adherence and co-aggregation, at MIC, ↓ the viable count of dental biofilm (3.50 log10 CFU/mL).
[78]
IndiaSoxhlet with EtOH:H2O 50% (v/v) (6 h)S. mutans and S. mitisMATH assayAt >1 mg/mL ↓ hydrophobicity (index < 40%).[79]
IndiaMaceration with stirring (2 days) and Soxhlet with EtOHS. mutans, S. sanguinis, and S. salivariusAgar well diffusion assayAt 10 mg/mL: ↑ inhibition zones for maceration extracts (19–23 mm).[80]
GhanaMaceration with agitation in EtOH:H2O 70% (v/v) (24 h)Aggregatibacter actinomycetemcomitans strainsAgar well diffusion assay, release of the cytosol enzyme lactate dehydrogenase, fluorescence assisted cell sorter, and ELISA assaysNo growth inhibitory effect, although neutralized the cell death and pro-inflammatory response, and restored the morphological alterations induced by the leukotoxin. These effects were due to the direct binding of guava compounds and the leukotoxin.[81]
IndiaMaceration in Ac, EtOH, chloroform, MeOH and H2O (15 days at 22 °C)Neisseria catarrhalis, S. mutans, S. salivarius, Streptococcus viridans, Bacillus megaterium, and P. aeruginosaAgar well diffusion assay↑ Inhibition zones in Ac (15–29 mm), except for N. catarrhalis (20 mm in MeOH).[82]
IndiaMaceration in MeOH (72 h). Fractionation with ethyl acetateS. aureus and S. mutansHRBC membrane stabilization method, disc and agar well diffusion assaysMeOH and ethyl acetate fraction ↑ protection (84–99%) to the inflammatory response. Inhibition zones (25–100 µg/mL) = 10.5 to 22 mm by both methods. MICs = 0.48 (ethyl acetate) and 0.62 (MeOH) mg/mL.[83]
TaiwanMaceration EtOH, Ac, H2O (room temperature and 60 °C) (24 h)Clone 9 rat liver cellsWST-1 and ALT assaysAt >500 μg/mL cytotoxic effect of EtOH and Ac and 600 μg/mL for H2O. At <200 μg/mL normal values were observed for H2O and Ac, and EtOH (<500 μg/mL).
At <100 μg/mL: Hepato-protective effect in EtOH and H2O (full range).
[85]
Alanin aminotransferase (ALT); colony forming unit (CFU); human red blood cell (HRBC); microbial adhesion to hydrocarbon test (MATH); minimum bactericidal concentration (MBC); minimum fungicidal concentration (MFC); minimum inhibitory concentration (MIC); nordini’s artificial mouth (NAM); Tetrazolium (WST-1); ↑ increases the affect; ↓ decreases the effect.
Table 6. In vivo anti-bacterial effect.
Table 6. In vivo anti-bacterial effect.
OriginExtraction MethodSubjectTreatmentMain ResultsRef.
ThailandMaceration in H2O, EtOH, and ether (24 h)Oreochromis niloticusAeromonas hydrophilaLD50 = 3.44 × 106 CFU/mL. ↓ Mortality of the subjects.[95]
China-Penaeus monodonYellow-head virus, white spot syndrome virus, and Vibrio harveyiSurvival rate = 80–95% (↑ Weight (2 to 6 g)). In serum (↑ feed): ↓ PO (7.50 U/mL) and SOD (178.33 U/mL), ↑ NOS (64.80 U/mL). In hepato-pancreas: ↑ SOD (57.32 U/mg), ACP (23.28 U/mg), AKP (19.35 U/mg), and LSZ (3459.946 U/mg).[96]
NigeriaMaceration with agitation in EtOH:H2O 80% (v/v) (24 h)Albino ratsT. b. bruceiAt 300 mg/kg: ↓ parasitemia; ↑ survival in 24 days.[97]
NigeriaMaceration with agitation in EtOH:H2O 80% (v/v) (24 h)Albino ratsT. b. bruceiAdministration 1–7 days. ↑ GSH: liver (5.4 to 8.1), kidney (3.3 to 6.0), and serum (0.8 to 2.4), restored in kidney and serum. In the brain, no effect was found. ↓ MDA: serum (13.9 to 5.9), brain (42.8 to 18.1), kidney (27.3 to 17.6), and liver (38.2 to 19.2).[98]
IndiaDecoction of the leaves (10 min)BALB/c micePlasmodium bergheiAt 350 and 1000 mg/kg ↓ parasitemia (73.7% and 85.8%); ↑ survival 15 and 18 days.[99]
IndiaExtraction in EtOH:H2O 50% (v/v)Swiss miceCitrobacter rodentiumAt 300 mg/kg: ↓ infection (day 4) of the treatment, and no infection at day 19 (control group at day 24).[100]
NigeriaHidrodistillation and fractionation with ethyl acetateISA brown male chicksE. coliAt 100 mg/kg: In 10 days ↓ signs of villous collapse (stunting, matting and fusion of villi), number of wet droppings (12-6); ↑ activity, weight gaining, and feed intake (from 27 to 45 g) in contrast to the infected ones (from 30 to 18 g); ↓ bacterial shedding load (from 60 to 45 CFU/mL).[101]
NigeriaDecoction of the leavesAdult miceV. choleraAt 250 mg/kg: Histopathological observations: mild degenerative, secretory, and inflammatory changes with goblet cells and with most of the exudate (neutrophils and lymphocytes).[102]
India-Adult male goatHaemonchus contortus90 Days feeding: ↑ Hb (7.2 to 8.6 g/dL), PCV (20.2 to 29.3%), total protein (4.8 to 6.3 g/dL), GLO (2.3 to 3.8 g/dL) (↑ control (2.8)), glucose (43.9 to 52.6 g/dL), and calcium (8.7 to 9.6 mg/dL); ↓ blood urea (47.9 to 29.8 mg/dL) (↓ control (41)). Phosphorus balance, serum albumin levels and serum enzyme activity did not show variation.[103]
Acid phosphatase (ACP); alkaline phosphatase (AKP); colony forming unit (CFU); globulin (GLO); glutathione (GSH); hemoglobin (Hb); lysozyme (LSZ); malondialdehyde (MDA); median lethal dose (LD50); nitric oxide synthase (NOS); packed cell volume (PCV); prophenoloxidase (PO); superoxide dismutase (SOD); ↑ increases the affect; ↓ decreases the effect.
Table 7. In vivo studies against diseases of the blood and immune system.
Table 7. In vivo studies against diseases of the blood and immune system.
OriginExtraction MethodSubjectTreatmentMain ResultsRef.
NigeriaMaceration with agitation in EtOH:H2O 80% (v/v) (24 h)Wistar ratsT. b. brucei/no infectedTreatment (1–7 days) at 150 mg/kg: ↑ Hb (6.5 to 10.7 g/dL), PCV (28.6 to 34.4%), RBCC (4.1 to 5.0 × 1012/L), MCV (53.6 to 64.3 fL), and MCHC (21.4 to 31.4 g/dL); ↓ WBC (23.2 to 19.4 × 109/dL) and neutrophil levels (28.9 to 27.3 × 103/mL).Compared to no infected subjects: similar values that obtained in treated-infected animals but with opposite conclusions.[105]
NigeriaExtraction in chloroform (24 h)MiceNo infectedTreatment (28 days) at 45.9 mg/mL: no differences in Hb (12 to 11 g/dL), PCV (37 to 35%), RBCC (6.1 to 5.1 × 106/L), and MCHC (33 to 32 g/dL), and neutrophil levels (13 to 12%); ↑ lymphocyte levels (85 to 92%) and MCV (61 to 69 fL).[106]
KoreaExtraction in EtOH:H2O 55% (v/v) (4.9 h, 47 °C)Sprague-Dawley rats and miceFreund’s complete adjuvant-induced hyperalgesia/LPS-induced endotoxic shockAt 400 mg/kg: PWL restored; ↑ 67% survival rate (72 h) by ↓ TNF-α (500 to 325 pg/mL) and IL-6 (80 to 58 ng/mL).[50]
BrazilTurbo-extraction in water and acetone: H2O 70% (v/v) (20 min)Swiss miceCarrageenan-induced peritonitis, acetic acid-induced abdominal writhing and hot plate testAt 50mg/kg: number of leukocyte migration into the peritoneal cavity H2O < H2O -acetone extract. No central analgesic activity. Peripheral analgesic activity: ↓ number writhing response (from 50 to 15 count).[19]
IndiaMaceration in EtOH (7 days)Wistar ratsAcetic acid-induced writhingAt 2 mg/kg ↓ 66% number writhing response (from 67 to 54 count). Comparable to diclofenac sodium (75%).[107]
IndiaDistillation with MeOH and H2OWistar ratsAcetic acid-induced writhing and hot plate testAt 10 and 30 mg/kg ↓ responses time (at 9.4 and 10.6 s) compared to the analgesic drug Pentazocine (14 s).[108]
Hemoglobin (Hb); interleukin-6 (IL-6); lipopolysaccharide (LPS); mean concentration hemoglobin count (MCHC); mean corpuscular volume (MCV); packed cell volume (PCV); paw withdrawal latency (PWL); red-blood cell counts (RBCC); tumor necrosis factor alpha (TNF-α); white-blood cell (WBC); ↑ increases the affect; ↓ decreases the effect.
Table 8. Endocrine and metabolic in vivo assays with guava leaves.
Table 8. Endocrine and metabolic in vivo assays with guava leaves.
OriginSubjectTreatmentMain ResultsRef.
NigeriaRabbitsHigh-cholesterol dietAt 250 mg/kg: ↓ TC (15%); ↑ HDL (69%); ↓ LDL (74%); ↓ hyperglycemia 43%.[109]
BrazilWistar ratsHigh-cholesterol dietAt 369.89 mg phenolic compound in the extract/g: ↓ TC (29–35%), TG (59–73%); ↑ HDL (46%); ↓ VLDL+LDL; ↓ enzyme activity (SOD (6.2 to 5.7 U/mg protein), GP (4.6 to 2.3 µmol/g protein).[110]
KoreaLeprdb/Leprdb juvenile and adult miceDiabetes spontaneous mutationAt 10 mg/kg: 87% inhibition PTP1B; ↓ glucose levels 31% and 42% respectively.[111]
IranWistar ratStreptozotocin-induced diabetesAt 1mg/L: ↓ Ca/Mg ratio (18 to 12), glucose level, TG (100 to 65 mg/dL), TC (68 to 48 mg/dL), ↑ HDL (18 to 40 mg/dL), ↓ LDL, and VLDL to normal levels; ↓ alteration in vascular reactivity (110 to 50 mmHg).[112]
TaiwanSprague-Dawley ratsLow-dose streptozotocin and nicotinamide-induced diabetesAt 400 mg/kg: ↓ blood glucose level (230 to 140 mg/dL); ↑ plasma insulin level and glucose utilization (normal levels); ↑ enzyme activity (hepatic hexokinase (8 to 11 U/mg protein), phosphofructokinase (18 to 25 U/mg protein) and glucose-6-phosphate dehydrogenase (11 to 25 U/mg protein).[113]
IndiaSprague-Dawley ratsStreptozotocin-induced diabetesAt 100 mg/kg: ↓ blood glucose level (4 to 1 mg/mL) and lipid peroxidation (2 to 1 mmol/100 g tissue); ↑ enzyme activity (CAT (6 to 10 × 103 U/mg protein), SOD (6 to 10 U/mg protein), GPx (0.4 to 0.6 U/mg protein), GRd (0.1 to 0.3 U/mg protein).[55]
NigeriaAlbino ratsAlloxan-induced diabetesAt 200 mg/kg: ↑ average weight (99 to 209g); ↓ blood glucose level (15 to 8 mmol/L); ↓ alanine aminotransferase activity (32 to 24 U/L).[114]
IndiaAlbino ratsAlloxan-induced diabetesAt 500 mg/kg: ↓ blood glucose level, TC (231 to 163 mg/dL), TG (133 to 69 mg/dL), LDL (186 to 126 mg/dL), VLDL (26 to 13 mg/dL); ↑ HDL (18 to 23 mg/dL).[115]
NigeriaWister rats-At 150 mg/kg: ↑ ALP (300, 175and 650 IU), AST (500, 400, 450 IU), ALT (1200, 1200, 1800 IU), ACP (750, 650, 900 IU) activity in the kidney, liver, and serum, respectively.[116]
NigeriaMice-At 49.3 mg/mL: ↑ AST (93 to 126 iµ/L), ALT (30 to 35 iµ/L), ALP (57 to 66 iµ/L), conjugate bilirubin (0.2 to 0.3 mg/dL) and creatinine (0.9 to 1.2 mg/dL).[106]
NigeriaAlbino rats-At 150 mg/kg: ↑ serum urea (2.9 to 6 mmol/L) and creatinine (2.7 to 4 mmol/L); ↓ concentration of serum Na+ (122 to 99 mmol/L).[117]
acid phosphatase (ACP); alanine aminotransferase (ALT); alkaline phosphatase (ALP); aspartate aminotransferase (AST); catalase (CAT); glutathione peroxidase (GPx); glutathione reductase (GRd); high-density lipoprotein (HDL) cholesterol; low-density lipoprotein (LDL) cholesterol; protein tyrosine phosphatase 1B (PTP1B); superoxide dismutase enzyme (SOD); total cholesterol (TC); triglycerides (TG); very low-density lipoprotein (VLDL) cholesterol; ↑ increases the affect; ↓ decreases the effect.
Table 9. In vivo assays for digestive system related diseases.
Table 9. In vivo assays for digestive system related diseases.
OriginExtraction MethodSubjectTreatmentMain ResultsRef.
IndiaExtraction with MeOHWistar ratsASP, PL, and EtOH-induced ulcersAt 200 mg/kg: PL-induced ulcers: ↓ 64% ulcer formation (ui = 2.1), ↓ GV (5 to 2 mL), acid secretion (88 to 64 mEq/L/100 g), ↑ pH (2 to 5).Comparable to omeprazole; ASP (↓ 70.5%, ui = 2.5) and EtOH (↓ 70.4%, ui = 8.7)-induced systems.[124]
NigeriaMaceration in H2O (24 h)Albino ratsEtOH-induced ulcersAt 1000 mg/kg: ↓ MNL (9.4 to 2) ui (4.7 to 1).[125]
NigeriaMaceration with agitation in MeOH (24 h)Wistar ratsEtOH-induced ulcersAt 1000 mg/kg: ↓ ui (17.7 to 6.3), ↑ protection (64.4%).[126]
IndiaMaceration in EtOH:H2O 90% (v/v) (72 h). Wistar ratsPL and EtOH-induced ulcersAt 200 mg/kg: PL-induced: ↓ ulcer formation (77 to 84%), ui (5 to 1.3), GV (1.4 to 0.5 mL/100g), and acid secretion (28 to 23 mEq/L); ↑ pH (2.0 to 3.4). EtOH-induced: ↓ (63% to 79%, ui = 1.6 to 5.6), and gastric lesions (5.6–1.9).[127]
South AfricaMaceration in H2O (48 h)Wistar rats and BALB/c miceCastor oil-induced diarrhea and castor oil-induced enteropoolingAt 400 mg/kg: ↑ 83.3% rat protection, ↓ 75% fluid accumulation in rats; ↓ 87.73% transit in rats and 77.2% in mice; ↓ 64.35% of contractions in mice.[128]
NigeriaSoxhlet with EtOH:H2O 70% (v/v)Wistar ratsCastor oil-induced diarrhea At 80 mg/kg: ↓ 53.03% transit in rats and ↓ 67.70% intestinal contractions.[129]
PakistanMaceration with EtOHBALB/c mice, rabbit jejunumCastor oil-induced diarrhea, K+-induced motilityAt 1 g/kg: ↑ 81.1% mice protection; Spasmolytic effect (0.3–1 mg/mL) ↓ spontaneous contractions EC50 = 0.66 mg/mL in rabbits.[130]
IndiaDecoction (1 h)Wistar ratsCCl4, PCM, and TAA-induced liver injuryAt 500 mg/kg: CCl4: ↓ ALT (384 to 17 U/L), AST (642 to 152 U/L), ALP (750 to 489 U/L), and bilirubin (1.6 to 0.3 mg/dL), ↓ control levels; PCM: ↓ ALT (384 to 87 U/L), AST (642 to 179 U/L), ALP (750 to 338 U/L), and bilirubin (1.6 to 0.6 mg/dL); TAA: ↓ ALT (337 to 32 U/L), AST (438 to 237 U/L), and ALP (770 to 479 U/L).[131]
IndiaSoxhlet with EtOHWistar ratsPCM-induced liver injuryAt 400 mg/kg: ↓ SGOT (475 to 370), SGPT (158 to 128), ALP (814 to 729), and bilirubin (0.7 to 0.6); ↑ total protein (5.15 to 5.6), albumin (2.6 to 3.1), and GLO (2.1 to 2.4). Histopathological observations: less diffuse granular degeneration and mild periportal lymphocytic infiltration.[132]
IndiaDecoction (1 h)Wistar ratsAcetaminophen-induced liver injuryAt 500 mg/kg: ↓ AST (121 to 77 IU/L), ALT (80 to 57 IU/L), ALP (115 to 67 IU/L), and total bilirubin (4 to 2 mg/dL). Restored: total protein (5 to 7 g/dL), LPO (7 to 2 nmol/mg protein), GPx (13 to 19 µmol/mg protein), GSH (15 to 23 µmol/mg protein), CAT (14 to 24 µmol/mg protein), and SOD (48 to 63 µmol/mg protein). Histopathological observations: normal lobular structure.[133]
EgyptMaceration with agitation in EtOH:H2O 70% (v/v) (24 h)Albino ratsCCl4-induced liver injuryAt 500 mg/kg: ↓ ALT (94 to 55 U/mL), AST (199 to 82 U/mL), GGT (71 to 23 U/mL), lysosomal enzymes (50%), and LPO (7 to 3 nmol/mg protein); ↑ SOD (15 to 39 U/mg protein), CAT (5 to 15 µg/mg protein), GSH (6 to 8 µg/mg protein), GST (13 to 25 mM/min/mg protein), total protein (48 to 58 g/L), albumin (29 to 38 g/L), GLO (19 to 21 g/L).[134]
EgyptDecoction (1 h)Wistar ratsPCM-induced liver injury↓ AST (342 to 156 U/L), ALT (359 to 80 U/L), ALP (288 to 263 U/L), LDH (207 to 143 U/L), GGT (11 to 7 U/L), and total bilirubin (0.3 to 0.2 mg/dL). Restored SOD (13 to 24 U/g) and CAT (5 to 17 U/g).[135]
Alkaline phosphatase (ALP); Alanine aminotransferase (ALT); aspirin (ASP); aspartate aminotransferase(AST); catalase (CAT); carbon tetrachloride (CCl4); ethanol (EtOH); gamma glutamyl transferase (GGT); gastric volume (GV); globulin (GLO); glutathione (GSH); glutathione peroxidase (GPx); glutathione S-transferase (GST); lactate dehydrogenase (LDH); lipid peroxidation (LPO); mean number lesions (MNL); paracetamol (PCM); pyloric ligation (PL), Serum glutamic oxaloacetic transaminase (SGOT); Serum glutamic pyruvic transaminase (SGPT), superoxide dismutase (SOD); thioacetamide (TAA); ulcer index (ui); ↑ increases the affect; ↓ decreases the effect.

Share and Cite

MDPI and ACS Style

Díaz-de-Cerio, E.; Verardo, V.; Gómez-Caravaca, A.M.; Fernández-Gutiérrez, A.; Segura-Carretero, A. Health Effects of Psidium guajava L. Leaves: An Overview of the Last Decade. Int. J. Mol. Sci. 2017, 18, 897. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18040897

AMA Style

Díaz-de-Cerio E, Verardo V, Gómez-Caravaca AM, Fernández-Gutiérrez A, Segura-Carretero A. Health Effects of Psidium guajava L. Leaves: An Overview of the Last Decade. International Journal of Molecular Sciences. 2017; 18(4):897. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18040897

Chicago/Turabian Style

Díaz-de-Cerio, Elixabet, Vito Verardo, Ana María Gómez-Caravaca, Alberto Fernández-Gutiérrez, and Antonio Segura-Carretero. 2017. "Health Effects of Psidium guajava L. Leaves: An Overview of the Last Decade" International Journal of Molecular Sciences 18, no. 4: 897. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18040897

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