Next Article in Journal
Discovery of a 4-Hydroxy-3′-Trifluoromethoxy-Substituted Resveratrol Derivative as an Anti-Aging Agent
Previous Article in Journal
Internal Vibrations of Pyridinium Cation in One-Dimensional Halide Perovskites and the Corresponding Halide Salts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of the Extraction Method on the Polyphenolic Profile and the Antioxidant Activity of Psidium guajava L. Leaf Extracts

by
Daniela Gutierrez Montiel
1,
Alma Lilian Guerrero Barrera
1,*,
Guillermo Cristian Guadalupe Martínez Ávila
2,*,
María Dolores Gonzalez Hernandez
2,
Norma Angelica Chavez Vela
3,
Francisco Javier Avelar Gonzalez
4 and
Flor Yazmin Ramírez Castillo
1
1
Laboratorio de Biología Celular y Tisular, Departamento de Morfología, Centro de Ciencias Básicas, Universidad Autónoma de Aguascalientes, Aguascalientes CP 20100, Mexico
2
Laboratorio de Química y Bioquímica, Facultad de Agronomía, Universidad Autónoma de Nuevo León, General Escobedo CP 66050, Mexico
3
Laboratorio de Biotecnología, Departamento Ingeniería Bioquímica, Centro de Ciencias Básicas, Universidad Autónoma de Aguascalientes, Aguascalientes CP 20100, Mexico
4
Laboratorio de Estudios Ambientales, Departamento de Fisiología y Farmacología, Centro de Ciencias Básicas, Universidad Autónoma de Aguascalientes, Aguascalientes CP 20100, Mexico
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(1), 85; https://doi.org/10.3390/molecules29010085
Submission received: 13 November 2023 / Revised: 9 December 2023 / Accepted: 19 December 2023 / Published: 22 December 2023
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
The leaves of Psidium guajava L. are an agro-industrial by-product with an outstanding content of polyphenolic compounds; however, there are many factors which can affect the phytochemical profile when valuing this type of plant material, such as temperatures and extraction times involving in the extraction methods applied. In this context, this study analyzed the impact of different extraction methods (Soxhlet, maceration and ultrasound-assisted extraction) on the phytochemical profile (FTIR and UPLC-MS) and the antioxidant activity (ABTS, FRAP and Folin–Ciocalteu) of guava leaf extracts. A yield of phenolic compounds per gram of guava leaf was obtained within the range of 16 to 45 mg/g; on the other hand, the I C 50 values determined with the ABTS assay ranged between 78 ± 4 to 152 ± 12 µg/mL. The methanolic extract obtained by Soxhlet was the one with the best reducing power, both in the FRAP assay and in the Folin–Ciocalteu assay. Finally, bioactive compounds such as quercetin, kaempferol and avicularin were identified in the guava leaf extract. It was concluded that the purification of polyphenolics compounds improves the antioxidant capacity, and that the extraction method greatly influences the phytochemical profile and activity of the extracts.

Graphical Abstract

1. Introduction

Recently, there has been a growing interest from the scientific community in phytochemicals, non-nutritive bioactive compounds of plant origin, which have beneficial properties for health, as well as great antioxidant power, which makes them potential nutraceuticals, products that are able to prevent diseases and thereby increasing its value for the society [1].
One of the most abundant groups of phytochemicals are polyphenolic compounds, which are secondary plant metabolites having highly varied structures but are characterized by the presence of aromatic rings and hydroxyl groups [2]. Polyphenolic compounds are recognized as antioxidants for their ability to donate hydrogen atoms and/or electrons to free radicals, breaking the oxidation chain [3]. This ability has become a topic of great importance considering that oxidative stress has been associated with many diseases, such as cancer, hypertension, diabetes mellitus, atherosclerosis and neurological disorders [3].
Oxidative stress occurs when there is no balance between pro-oxidant species and antioxidants [4]. It should be noted that oxidation is not a specific problem for humans, it also appears in food, e.g., the oxidation of lipids and proteins, which is a serious problem, as it can cause the destruction of essential nutrients, bad odors and even the generation of toxic compounds in food systems [5].
Downstream processing involves essential steps in discovering bioactive compounds from raw plant materials, including the extraction methods. Different factors such as solvent type, temperature, time and the solid–liquid ratio affect the extraction efficiency, so they must be carefully selected, considering the technique to be used [6]. The most popular techniques are maceration, infusion and continuous hot extraction, such as Soxhlet; however, alternative methods such as ultrasonic or microwave-assisted extraction and supercritical fluid extractions have become available, and have gained interest because they are faster and, above all, respectful of the environment by reducing solvent and energy consumption [7].
Each extraction technique has its own advantages and disadvantages, in the case of Soxhlet extraction, it allows to extract a large quantity and variety of compounds in a relatively short time; furthermore, the solvent can be reused and the filtration and/or centrifugation process for separation the plant material can be avoided. However, generally, high temperatures are used, and, consequently, the degradation of phytochemicals, including polyphenolic compounds, may occur [6]. Therefore, in this study, we evaluated other options which did not require high temperatures: maceration due to its simplicity and wide use [8] and ultrasound-assisted extraction that facilitates the breaking of plant cells, increasing the contact area between plant material and solvent, decreasing heat requirements [6]. It is important to highlight that the extraction of polyphenolic compounds is a challenge since they can be unstable and the biological activity can be damaged and/or lost by high temperatures and the presence of oxygen and light; therefore, the choice of the extraction technique and all its parameters is key to obtain good yields and at the same time maintain the integrity of the compounds [9].
In this sense, Psidium guajava L. is a native American shrub of great economic importance for its fruit, the guava, a berry with firm pulp and numerous seeds that is marketed worldwide. The harvest P. guajava generates a large number of leaves as by-products, which are rich in polyphenolic compounds, and it has also been reported that they have antimicrobial and antioxidant activity [10]. Unfortunately, about 1.3 billion tons of agricultural by-products, including leaves, seeds and skins are wasted each year [2] and are landfilled or incinerated even though they may contain large amounts of bioactive compounds. These practices only increase the environmental load and the total cost of production [11,12]. In the specific case of guava, approximately 80 kg of this by-product per metric ton of fresh fruit is produced during guava processing [13].
The valorization of guava leaves is an opportunity for producing countries such as Mexico, India and China, where large amounts of leaves are available [8]. However, for its use, it is necessary to consider that, depending on the geographical location, weather conditions, the presence of different pathogens, the variety, etc., changes can occur in the phytochemical composition of the plant organs [14,15], so it is extremely important to consider the parts of the plant material.
In the present study, the impact of different extraction methods on the antioxidant activity and the phytochemical profile of guava leaf extracts was evaluated, considering two conventional techniques and a green alternative. In addition, to our best knowledge, this is the first report in which the polyphenolic compounds of the guava leaf are purified with amberlite XAD-16, achieving improvements in their bioactivity.

2. Results

The leaves of P. guajava L. are recognized for their high and diverse content of polyphenolic compounds [16,17,18], as well as for their wide use in traditional medicine to treat conditions such as stomach pains, wounds, cavities and coughs [10]. In recent decades, evidence has demonstrated the beneficial properties of polyphenols, for example, their antioxidant, antimicrobial and anti-inflammatory activity [3,19]. The extraction of polyphenolic compounds from agro-industrial residues is a viable option to value this type of waste, minimizing environmental damage and generating value-added products, such as natural antioxidants in functional foods or for the development of dietetic supplements [19]. This section presents the tentative identification of the polyphenolic compounds present in guava leaf extracts as well as their antioxidant activity, discussing the differences observed according to the extraction method used, which consists of a given technique, temperature and time.

2.1. Purification of Polyphenolic Compounds

Table 1 shows the different extraction methods used and the yields of polyphenolic compounds obtained using them, which ranged from 16 to 45 mg·g−1. The highest values were observed with the extraction method of Soxhlet, using methanol as an extracting solvent (44 mg·g−1), and ultrasound-assisted extraction at 30 °C with distilled water (45 mg·g−1). On the other hand, all other extraction methods had an approximate yield of around 20 mg·g−1.
Multiple authors have reported different yields of total polyphenolic compounds in guava leaves. For example, Farag et al. [20] observed a total content of 59.267 ± 0.348 mg GAE (gallic acid equivalent) ·g−1; Sowmya et al. [21] reported a content of 41.33 ± 0.92 mg GAE·g−1 and 37.60 ± 0.26 mg GAE·g−1 in two guava varieties; and Laily et al. [22] published a content of 101.93 mg GAE·g−1. Although these concentrations may be higher than those obtained in this study, it is important to mention that, in the works mentioned, a purification process for polyphenolic compounds was not carried out, so the total content may be overestimated, especially considering that the Folin–Ciocalteu assay was used for the determination. Even if this assay is well established and widely used, it should be noted that since it is based on a redox reaction, compounds other than phenols, for example, reducing sugars and ascorbic acid, can also reduce the Folin–Ciocalteu reagent [23]. It is worth mentioning that the yield is not directly related to the phenolic composition of the samples or to their antioxidant activity; therefore, a greater number of polyphenolic compounds does not always mean better antioxidant activity [24].

2.2. Antioxidant Activity

The antioxidant activity was evaluated using three different assays: Folin–Ciocalteu and FRAP (ferric-reducing antioxidant power), both to analyze the reducing capacity of the samples, and ABTS (2,2′-Azino-bis (3-ethylbenzthiazoline-6-sulfonic acid)), to analyze the ability to inhibit the radical cation ABTS•+.
Preliminarily, the three tests (Folin–Ciocalteu, FRAP and ABTS•+) were carried out with both the crude and purified solutions of the guava leaf extracts at 0.250 mg·mL−1, and it was observed that in all the tests, the samples with purified polyphenolics had a higher antioxidant activity in comparison with the raw samples, as shown in Figure 1.
This behavior is especially interesting given that Folin–Ciocalteu and FRAP are not specific tests for polyphenols, since they can be reduced by other agents such as reducing sugars, amino acids and ascorbic acid [25], which could be present in raw samples [26]. For this reason, it is important to carry out a purification process, since it allows us to eliminate inert and undesirable components that can interfere with the study and/or that have limited antioxidant activity that represses the activity of polyphenols [27].
The determination of the I C 50 (the minimum extract concentration at which 50% of the free radicals are inhibited) values were carried out with the ABTS•+ assay. The results are within the range from 78 to 152 µg·mL−1, as shown in Table 2. It is worth mentioning that the lowest values of I C 50 indicate a greater capacity to inhibit free radicals of the samples, which in the present study were exhibited by the methanolic extracts obtained by maceration and ultrasound (87 to 78 µg·mL−1).
Other studies with guava leaves have reported lower I C 50 values: 24.37 ± 3.85 μg·mL−1 [28], 31.19 ± 5.01 to 72.31 ± 3.57 μg·mL−1 [29] and 3.23 ± 0.24 to 8.26 ± 1.06 μg·mL−1 [30]. This may be since the polyphenol content and, therefore, its antioxidant activity may be affected by different factors, including the extraction conditions (technique, temperature, solvent, time, etc.), climatic conditions and soil quality [14,15,31]. Even so, guava leaf collected in Aguascalientes, Mexico presented relevant antioxidant activity compared to other plant extracts characterized by exhibiting different functional properties such as avocado leaf extracts ( I C 50 = 269.56 ± 6.52 to 442.72 ± 9.62 μg·mL−1) [32], rosemary leaf extracts ( I C 50 = 70 ± 4.67 μg·mL−1) [33] and oriental ebony leaf extracts ( I C 50 = 108.7 μg·mL−1) [34].
Similarly, it has been observed that guava leaves have a concentration of polyphenolic compounds and an antioxidant capacity higher than other parts of the bush such as seeds, fruit and bark [30,35]; consequently, the present study shows that guava leaf extracts can be considered an alternative for obtaining bioactive compounds with multiple applications in industry.
The FRAP and Folin–Ciocalteu assays were performed with the purified guava leaf extracts at respective concentrations of I C 50 (Table 1). In both tests, it was observed that the extract with the highest reducing capacity was the one obtained using the Soxhlet extraction method with methanol as the extracting solvent, while the other extracts exhibited a similar reducing capacity, as can be seen in Figure 2 and Figure 3.
Temperature is an important factor for the extraction of phytochemicals [8]. Increases in temperature can result in improvements in extraction thanks to better diffusion and solubilization [36,37]; however, they could also generate the degradation of compounds and decrease antioxidant activity [38,39]. In our case, the extraction method with Soxhlet is the one that used the highest temperature (65 °C), and this favored the yield of polyphenolic compounds and antioxidant activity, specifically the reducing activity.
In recent years, the search for natural antioxidants has gained importance in view of the need to replace fossil-derived resources, as well as to avoid the use of synthetic antioxidants, which can be cytotoxic and carcinogenic [40,41]. The different guava leaf polyphenolic extracts studied exhibited good antioxidant activity, property useful in a wide variety of applications, such as in the food [42], cosmetic [43,44] and pharmaceutical industries [45], meaning they could be a low-cost alternative of natural origin. For example, the methanolic extract obtained by Soxhlet, on account of its significant reducing capacity, may be a good candidate for the green synthesis of nanoparticles [46,47], while the methanolic extracts obtained by maceration and ultrasound-assisted extraction may be good candidates as nutraceuticals or cosmetics given their ability to inhibit free radicals [48]. Obviously, previous studies on toxicity, stability and bioavailability as well as a solvent removal process are necessary before the extract can have an industrial application.

2.3. Characterization by FTIR

The dried purified methanolic extracts obtained by three different methods (Soxhlet, maceration at 25 °C and ultrasound-assisted extraction at 30 °C) were selected due to their notable results in antioxidant activity and yield in polyphenolic compounds for analysis by Fourier transform infrared spectroscopy (FTIR). It should be noted that, as the samples analyzed are purified extracts, the peaks of the spectra correspond to the different functional groups present in the polyphenolic compounds.
In Figure 4, from 3000 to 3600 c m 1 , the broad and strong band corresponds to the stretching vibrations of the O–H bond, which indicates the presence of functional groups such as hydroxyls. In addition, the presence of O–H groups is confirmed since there is vibration between 1600 and 1300 c m 1 , 1200 and 1000 c m 1 and 800 and 600 c m 1 . The presence of peaks from 3000 to 2900 c m 1 corresponds to the stretching of the C–H bond, which is characteristic of aliphatic functional groups. On the other hand, from 2200 to 1950 c m 1 we can observe two small vibrations caused by the resonance effect of aromatic compounds. The presence of the carbonyl functional group is confirmed since peaks are observed from 1850 to 1650 c m 1 corresponding to the stretching of the C–O bond. The intense and well-defined peak between 1650 and 1600 c m 1 indicates the presence of carboxylic acids, while the two peaks around 1615 and 1495 c m 1 indicate the presence of double bounds for the vibration of the C–C bond, confirming the presence of aromatic compounds. Finally, peaks from 1500 to 600 c m 1 correspond to the fingerprint area, which is specific and unique; the bands present from 1000 to 1300 c m 1 are due to vibrations of the C–O bonds found in esters, carboxyls, ethers and hydroxyls.
Other studies on the phytochemicals from guava leaf extracts show similar spectra. Lok et al. [49] analyzed guava leaf extracts obtained with three different solvents, namely distilled water, ethanol and n-hexane, and they also observed the stretching bands of the C–H and O-H bonds; on the other hand, Nagpal et al. [50] and Lahlou et al. [51] also reported the vibration of the functional groups C=O, O–H, C–O and C–H.

2.4. UPLC-MS

The purified methanolic extract obtained by Soxhlet was selected for analysis by UPLC-MS given its outstanding reducing activity and yield of polyphenolic compounds, in comparison with the other two techniques evaluated. The tentative identification of the phytochemicals is presented in Table 3.
In the extract, 13 compounds were identified, among these were quercetin, considered the most active and powerful antioxidant of guava leaves [52]; catechin; kaempferol; avicularin; and guavinoside B and C, all of which exhibit different beneficial activities for the human health; e.g., anti-inflammatory, antimicrobial and antitumor activity [53,54,55,56,57,58,59,60,61]. Therefore, guava leaves are a rich source of polyphenolic compounds and a potential nutraceutical. On the other hand, there were also compounds that could not be identified, giving rise to new investigations for the identification, characterization and in vitro evaluation of the activity of these phytochemicals.
For additional information, see the Supplementary Materials.
Table 3. Tentative identification of phenolic compounds in Psidium guajava L. leaf extracts.
Table 3. Tentative identification of phenolic compounds in Psidium guajava L. leaf extracts.
Extraction MethodNo.Tentative IdentityTr (min)m/z expm/z
Calculated
Molecular FormulaFragmentsReference
Soxhlet with methanol1Not identified0.749249.0305248.034588 C 19 H 5 O 113.0318, 181.0322, 207.0358
2Vescalagin7.71933.2043933.07178 C 41 H 26 O 26 466.1265, 179.0356, 289.1466[17]
3Catechin9.587289.147289.079587 C 15 H 14 O 6 179.0359, 207.0354, 287.1305, 245.1503[17]
4Not identified18.88603.1823602.181312 C 40 H 27 O 6 179.0356, 207.0351, 235.9931, 257.02
5Casuarinin/Casuarictin Isomer19.637935.2216935.08743 C 41 H 28 O 26 467.1334, 145.9832, 385.2003, 478.1260[17]
6Not identified33.071381.0783380.076847 C 24 H 13 O 5 379.0626, 301.1106, 299.0956
7Not identified33.475381.0775380.076847 C 24 H 13 O 5 299.0951, 301.110, 302.1134, 379.0617, 271.098
8Quercetin glucuronide34.788477.1651477.075289 C 21 H 18 O 13 463.1837, 299.0954, 301.1106[17]
9Reynoutrin 37.414433.1703433.08546 C 20 H 18 O 11 431.1533, 181.0318, 235.9926, 300.1021, 415.2863[17]
10Guajaverin 40.242433.1713433.08546 C 20 H 18 O 11 431.1543, 300.1037, 301.1096, 391.9761[17]
11Avicularin40.798433.1707433.08546 C 20 H 18 O 11 431.1534, 300.1033, 302.1121[17]
12Myrciaphenone B48.222481.1956481.106589 C 21 H 22 O 13 479.1792, 417.1714, 365.9648, 257.0243, 235.9927, 239.9696, 207.0348, 181.0316, 179.0352[17]
13Guavinoside C57.06585.1982585.096419 C 27 H 22 O 15 583.1833, 304.9899, 285.9811, 235.9928, 257.0249, 352.9331[17]
14Not identified63.473551.2103550.207527 C 34 H 31 O 7 541.1789, 343.1274, 328.1021,
15Guavinoside B64.887571.2532571.15354 C 28 H 28 O 13 569.2360, 481.2711, 257.0239[17]
16Not identified78.22711.5146710.401989 C 30 H 63 O 18 701.4839, 549.4456, 503.4370
17Not identified82.967695.5202694.51974 C 51 H 67 O 685.4873, 533.4503, 487.4390
18Luteolin 7-O-malonyl-glucoside86.2533.4523533.101504 C 24 H 22 O 14 487.4415, 488.445, 523.4203, 501.4174[62]
19Kaempferol 3-O-(6″-malonyl-glucoside)86.907533.4519533.101504 C 24 H 22 O 14 487.4413, 488.4441, 523.4199[62]
20Chrysoeriol 7-O-(6″-malonyl-glucoside)87.816547.4318547.117154 C 25 H 24 O 14 501.4212, 502.4242, 427.0584, 533.4493[62]

3. Materials and Methods

3.1. Plant Material

The collection of guava leaves was carried out manually and randomly from different specimens free of pesticides in Aguascalientes, Mexico, in November 2021. The plant sample was transferred in a botanical press to the laboratory where only green leaves, without damage from insects or pests, were selected. Subsequently, the leaves were thoroughly washed with distilled water to remove traces of dust and other contaminants and dried at 40 °C for 72 h [35,63]. Finally, the sample was pulverized with an electric processor and the obtained powder was stored at room temperature in an airtight container protected from light [64].

3.2. Extraction of Phytochemicals

Three different extraction techniques were tested: Soxhlet, maceration and ultrasound-assisted extraction (UAE). In all cases, a solid–liquid ratio of 1:20 was used (5 g of plant sample per 100 mL of solvent). It was decided that two different solvents should be used, namely methanol, since multiple articles [65,66,67] and previous work performed by our laboratory have demonstrated its better extractive power, and distilled water, as a green alternative. The continuous extraction by Soxhlet was carried out using 7 siphons [68], while the maceration lasted 8 days and 2 temperatures were evaluated: 25 °C and 37 °C [69,70,71]. The ultrasound-assisted extraction lasted 40 min and 2 temperatures were evaluated: 23 °C and 30 °C [72,73,74,75]. The extracts obtained via maceration and UAE were centrifuged (5000 rpm for 17 min) and filtered (0.2 µm) to eliminate guava leaf particles [17].

3.3. Solvent Elimination

The aqueous extracts were subjected to low temperatures (−48 °C) and a vacuum for 5 days in alyophilizer (FreeZone, Labconco, Kansas City, MO, USA) to remove distilled water [76]. On the other hand, the methanolic extracts were subjected to 50 °C in an oven to eliminate the solvent [32]. In all cases, a green to reddish brown powder was obtained, which was stored in an Eppendorf tube at room temperature protected from light until use.

3.4. Purification of Polyphenolic Compounds

Phenolic compounds from guava leaves were purified with the commercial adsorbent Amberlite XAD-16 (Sigma-Aldrich, Saint Louis, MO, USA). Briefly, the lyophilized aqueous extracts were solubilized in distilled water while the methanolic extracts were solubilized in 80% methanol and the alcohol was removed using a rotary evaporator to obtain the water solubilized extract. Subsequently, 20 mL of the extracts were added to a column packed with amberlite XAD-16 as a stationary phase and distilled water was added to eliminate sugars and other compounds present in the extract, and finally, the polyphenolic compounds were eluted with absolute ethanol. The solvent was removed in an oven at 50 °C for 24 h and the crystals obtained were kept protected from light at room temperature [32,77].
The yield of polyphenols per gram of plant material was determined as follows:
Y i e l d = M i l l i g r a m s   o f   p h e n o l i c   c o m p o u n d s   o b t a i n e d G r a m s   o f   p l a n t   m a t e r i a l   u s e d   f o r   e x t r a c t i o n

3.5. Antioxidant Capacity Tests

The antioxidant capacity of both the crude guava leaf extracts and their purified polyphenolic compounds was analyzed. Three different assays were carried out in microplates: Folin–Ciocalteu, ABTS and FRAP.
The reducing power was determined by adding 25 μL of Folin–Ciocalteu reagent and 25 μL of sodium carbonate (75 g/L) to 25 μL of properly diluted sample (1:4 v/v). The obtained mixture was homogenized and incubated at 40 °C for 30 min. After that, 200 μL of distilled H2O were added, and the absorbance at 750 nm was recorded [78]. Results were expressed as gallic acid equivalents in micrograms per milliliter (GAE μg/mL) using the calibration curve prepared with the same standard.
The ABTS•+ radical scavenging capacity assay was carried out according to the methodology proposed by Hernández et al. [77]. Briefly, a solution of ABTS (7 mM) and one of potassium persulfate (2.45 mM) were mixed (2:1) and were allowed to rest for 12 h at room temperature; then, it was adjusted with absolute ethanol until it reached an absorbance of 0.7 nm. Subsequently, five microliters of each test sample and calibration curve were pipetted in triplicate into the microplate, and 95 μL of the adjusted ABTS•+ solution was added. After 1 min the absorbance at 734 nm was measured. According to the Trolox calibration curve, results were expressed as the percentage inhibition of ABTS•+ radicals or as I C 50 (sample concentration needed to inhibit 50% of radicals).
The iron-reducing antioxidant power (FRAP) was determined by mixing 5 µL of the samples to be analyzed with 12 µL of phosphate buffer (pH 7) and 22 µL of 1% potassium ferrocyanide, and then incubated at 50 °C for 20 min. Then, 12 µL of 10% trichloroacetic acid, 45 µL of distilled water and 10 µL of ferric chloride were added to read the absorbance at 700 nm [79]. Results were reported as the μg gallic acid equivalent per milliliter (GAE μg/mL).
Initially, as a preliminary test, antioxidant capacity tests were carried out with solutions of the extracts (both crude and purified) at 0.250 mg·mL−1 to observe their behavior. Subsequently, the I C 50 of the purified extracts were determined with the ABTS•+ assay, and finally, the FRAP and Folin–Ciocalteu assays were performed with the purified extracts at their determined I C 50 .
Statistical analysis was performed using Excel, MiniTab and GraphPad Prism 8.0.1 software.

3.6. FTIR (Fourier Transform Infrared Spectroscopy)

The polyphenolic compounds of the methanolic extracts obtained via Soxhlet, maceration at 25 °C and ultrasound at 30 °C were analyzed using Fourier transform infrared spectroscopy (Agilent Technologies, Santa Clara, CA, USA, FTIR model Cary 630 coupled to a zinc selenide crystal (ZnSe) ATR). The obtained powder after the purification process was deposited on the surface of the reader and secured by means of the equipped press. The spectra were acquired at a range of 4000–600 cm−1 across 32 scans with a resolution of 2 cm−1 [80].

3.7. UPLC-MS

The analysis was carried out with an Acquity UPLC system (Waters, Milford, MA, USA) consisting of an auto-sampler, a binary pump equipped with a 10 µL Loop (partial Loop injection mode) and a BEH PHENYL column (2.1 mm × 100 mm, 1.7 µm; WATERS, Waxford, Ireland). The solvents used were (A) water + 0.1% (v/v) formic acid and (B) acetonitrile + 0.1% (v/v) formic acid at a constant flow rate of 0.3 m L · m i n 1 . The elution gradient (for 113 min) was 100% A, gradually decreasing until reaching 10% A and 90% B, to move from normal conditions (100% A) one minute later to re-equilibrate the column. MS detection was performed on a Q-ToF quadrupole orthogonal acceleration time-of-flight mass spectrometer (Q-TOF™, Waters, Milford, MA, USA) equipped with an electrospray ionization source (ESI). The sample acquisition mode was in negative ionic polarity, analysis mode in sensitivity and normal dynamic range, in a mass range of 50 to 1200 Da, sweep conditions of 0.5 s 1 , a centroid data format, a collision energy of 6 V and a cone voltage of 40 V.

4. Conclusions

The present study shows that guava leaves are a rich resource in polyphenolic compounds, the antioxidant activity of which could be of great interest to the pharmaceutical and food industry; thus, guava-producing countries should value this agro-industrial by-product.
About 13 compounds were identified in the analyzed guava leaf extract, including quercetin, catechin, hyperin and guajaverin. Of the three methods studied, Soxhlet is postulated as an advantageous option for the recovery of phenolic compounds, given that it does not require a significant amount of time, no processes are needed to separate the plant material from the solvent, the solvent can be reused and the extract obtained exhibits outstanding reducing activity and yield in phenolic compounds compared to maceration and ultrasound-assisted extraction. In addition, different commercially available systems allow the Soxhlet extraction method to be scaled and even automated for industrial applications. Furthermore, the purification of phenolic compounds generated improvements in the antioxidant activity in the three assays carried out (FRAP, Folin–Ciocalteu and ABTS•+) in our study, demonstrating that this process may be a way to increase the bioactivity of these phytochemicals.
It should be noted that, for its future industrial application, it will be necessary to face different challenges, especially to improve the long-term stability of polyphenolic compounds and increase their bioavailability; this can be achieved, for example, through encapsulation. In addition, more studies are needed on its toxicity, in vivo activity and safety.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29010085/s1.

Author Contributions

Conceptualization, D.G.M., G.C.G.M.Á. and A.L.G.B.; methodology, D.G.M., G.C.G.M.Á. and M.D.G.H.; writing—original draft preparation, D.G.M., G.C.G.M.Á. and A.L.G.B.; formal analysis, D.G.M., M.D.G.H. and G.C.G.M.Á.; investigation, D.G.M., F.Y.R.C., N.A.C.V. and F.J.A.G.; resources, D.G.M., A.L.G.B. and G.C.G.M.Á.; writing—review and editing, D.G.M., N.A.C.V., F.J.A.G. and F.Y.R.C.; supervision, A.L.G.B., G.C.G.M.Á., N.A.C.V. and F.J.A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by internal project PIB22-3 of the Universidad Autonoma de Aguascalientes. DGM was supported by CONACYT CVU 1091197.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

Gutiérrez-Montiel thanks to National Council for Humanities, Sciences and Technologies of Mexico for her postgraduate scholarship. All the authors thank to Universidad Autónoma de Aguascalientes and Universidad Autónoma de Nuevo León for all the facilities given for institutional cooperation.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Thakur, M.; Singh, K.; Khedkar, R. Phytochemicals. In Functional and Preservative Properties of Phytochemicals; Elsevier: Amsterdam, The Netherlands, 2020; pp. 341–361. [Google Scholar]
  2. Wang, Z.; Li, S.; Ge, S.; Lin, S. Review of Distribution, Extraction Methods, and Health Benefits of Bound Phenolics in Food Plants. J. Agric. Food Chem. 2020, 68, 3330–3343. [Google Scholar] [CrossRef] [PubMed]
  3. Albuquerque, B.R.; Heleno, S.A.; Oliveira, M.B.P.P.; Barros, L.; Ferreira, I.C.F.R. Phenolic Compounds: Current Industrial Applications, Limitations and Future Challenges. Food Funct. 2021, 12, 14–29. [Google Scholar] [CrossRef]
  4. Pisoschi, A.M.; Pop, A.; Iordache, F.; Stanca, L.; Predoi, G.; Serban, A.I. Oxidative Stress Mitigation by Antioxidants—An Overview on Their Chemistry and Influences on Health Status. Eur. J. Med. Chem. 2021, 209, 112891. [Google Scholar] [CrossRef] [PubMed]
  5. Choe, E.; Min, D.B. Mechanisms of Antioxidants in the Oxidation of Foods. Compr. Rev. Food Sci. Food Saf. 2009, 8, 345–358. [Google Scholar] [CrossRef]
  6. Abu Bakar, F.I.; Abu Bakar, M.F.; Abdullah, N.; Endrini, S.; Fatmawati, S. Optimization of Extraction Conditions of Phytochemical Compounds and Anti-Gout Activity of Euphorbia Hirta L. (Ara Tanah) Using Response Surface Methodology and Liquid Chromatography-Mass Spectrometry (LC-MS) Analysis. Evid. Based Complement. Altern. Med. 2020, 2020, 4501261. [Google Scholar] [CrossRef] [PubMed]
  7. Dhanani, T.; Shah, S.; Gajbhiye, N.A.; Kumar, S. Effect of Extraction Methods on Yield, Phytochemical Constituents and Antioxidant Activity of Withania Somnifera. Arab. J. Chem. 2017, 10, S1193–S1199. [Google Scholar] [CrossRef]
  8. Alara, O.R.; Abdurahman, N.H.; Ukaegbu, C.I. Extraction of Phenolic Compounds: A Review. Curr. Res. Food Sci. 2021, 4, 200–214. [Google Scholar] [CrossRef] [PubMed]
  9. Osorio-Tobón, J.F. Recent Advances and Comparisons of Conventional and Alternative Extraction Techniques of Phenolic Compounds. J. Food Sci. Technol. 2020, 57, 4299–4315. [Google Scholar] [CrossRef]
  10. Gutierrez-Montiel, D.; Guerrero-Barrera, A.L.; Chávez-Vela, N.A.; Avelar-Gonzalez, F.J.; Ornelas-García, I.G. Psidium Guajava L.: From Byproduct and Use in Traditional Mexican Medicine to Antimicrobial Agent. Front. Nutr. 2023, 10, 1108306. [Google Scholar] [CrossRef]
  11. Kuila, A.; Mukhopadhyay, M. Biorefinery Production Technologies for Chemicals and Energy; Kuila, A., Mukhopadhyay, M., Eds.; Wiley: Hoboken, NJ, USA, 2020; ISBN 9781119591429. [Google Scholar]
  12. Muzaffar, K.; Ahmad Sofi, S.; Mir, S.A. Handbook of Fruit Wastes and By-Products; CRC Press: Boca Raton, FL, USA, 2022; ISBN 9781003164463. [Google Scholar]
  13. Lim, S.Y.; Tham, P.Y.; Lim, H.Y.L.; Heng, W.S.; Chang, Y.P. Potential Functional Byproducts from Guava Purée Processing. J. Food Sci. 2018, 83, 1522–1532. [Google Scholar] [CrossRef]
  14. Hardege, J.D. Nereidid Polychaetes as Model Organisms for Marine Chemical Ecology. Hydrobiologia 1999, 402, 145–161. [Google Scholar] [CrossRef]
  15. Lavola, A.; Salonen, A.; Virjamo, V.; Julkunen-Tiitto, R. Phytochemical Variation in the Plant-Part Specific Phenols of Wild Crowberry (Empetrum Hermaphroditum Hagerup) Populations. Phytochem. Lett. 2017, 21, 11–20. [Google Scholar] [CrossRef]
  16. Lorena, C.; Ressaissi, A.; Serralheiro, M.L. Bioactives from Psidium Guajava Leaf Decoction: LC-HRMS-MS-Qtof Identification, Bioactivities and Bioavailability Evaluation. Food Chem. Adv. 2022, 1, 100003. [Google Scholar] [CrossRef]
  17. Díaz-de-Cerio, E.; Verardo, V.; Gómez-Caravaca, A.; Fernández-Gutiérrez, A.; Segura-Carretero, A. Exploratory Characterization of Phenolic Compounds with Demonstrated Anti-Diabetic Activity in Guava Leaves at Different Oxidation States. Int. J. Mol. Sci. 2016, 17, 699. [Google Scholar] [CrossRef] [PubMed]
  18. Purba, R.A.P.; Paengkoum, P. Farang (Psidium Guajava L.) Dried Leaf Extracts: Phytochemical Profiles, Antioxidant, Anti-Diabetic, and Anti-Hemolytic Properties for Ruminant Health and Production. Molecules 2022, 27, 8987. [Google Scholar] [CrossRef]
  19. Rosales, T.K.O.; Fabi, J.P. Valorization of Polyphenolic Compounds from Food Industry By-Products for Application in Polysaccharide-Based Nanoparticles. Front. Nutr. 2023, 10, 1144677. [Google Scholar] [CrossRef]
  20. Farag, R.S.; Abdel-Latif, M.S.; Abd El Baky, H.H.; Tawfeek, L.S. Phytochemical Screening and Antioxidant Activity of Some Medicinal Plants’ Crude Juices. Biotechnol. Rep. 2020, 28, e00536. [Google Scholar] [CrossRef]
  21. Sowmya, B.H.; Anandhi, D.U. Quantification of Total Phenolics, Flavonoids and Evaluation of in Vitro Free Radical Scavenging Activities in Psidium Guajava L. Indian. J. Pharm. Sci. 2020, 82, 578–585. [Google Scholar] [CrossRef]
  22. 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]
  23. Lawag, I.L.; Nolden, E.S.; Schaper, A.A.M.; Lim, L.Y.; Locher, C. A Modified Folin-Ciocalteu Assay for the Determination of Total Phenolics Content in Honey. Appl. Sci. 2023, 13, 2135. [Google Scholar] [CrossRef]
  24. Jiménez-Moreno, N.; Volpe, F.; Moler, J.A.; Esparza, I.; Ancín-Azpilicueta, C. Impact of Extraction Conditions on the Phenolic Composition and Antioxidant Capacity of Grape Stem Extracts. Antioxidants 2019, 8, 597. [Google Scholar] [CrossRef]
  25. Huang, D.; Ou, B.; Prior, R.L. The Chemistry behind Antioxidant Capacity Assays. J. Agric. Food Chem. 2005, 53, 1841–1856. [Google Scholar] [CrossRef]
  26. Kumar, M.; Tomar, M.; Amarowicz, R.; Saurabh, V.; Nair, M.S.; Maheshwari, C.; Sasi, M.; Prajapati, U.; Hasan, M.; Singh, S.; et al. Guava (Psidium Guajava L.) Leaves: Nutritional Composition, Phytochemical Profile, and Health-Promoting Bioactivities. Foods 2021, 10, 752. [Google Scholar] [CrossRef]
  27. Barbosa-Pereira, L.; Pocheville, A.; Angulo, I.; Paseiro-Losada, P.; Cruz, J.M. Fractionation and Purification of Bioactive Compounds Obtained from a Brewery Waste Stream. Biomed. Res. Int. 2013, 2013, 408491. [Google Scholar] [CrossRef]
  28. Le, T.B.; Le, T.D.; Nguyen, T.T.; Do, X.C.; Lam, C.D.; Bui, N.A.; Nguyen, H.T.L.; Tran, N.G.H.; Bui, T.B.H. Chemical Composition and Antioxidant Activity of Psidium Guajava L. Leaves. Can. Tho Univ. J. Sci. 2021, 13, 66–70. [Google Scholar] [CrossRef]
  29. Wang, L.; Wu, Y.; Bei, Q.; Shi, K.; Wu, Z. Fingerprint Profiles of Flavonoid Compounds from Different Psidium Guajava Leaves and Their Antioxidant Activities. J. Sep. Sci. 2017, 40, 3817–3829. [Google Scholar] [CrossRef]
  30. 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]
  31. Chirinos, R.; Rogez, H.; Campos, D.; Pedreschi, R.; Larondelle, Y. Optimization of Extraction Conditions of Antioxidant Phenolic Compounds from Mashua (Tropaeolum Tuberosum Ruíz & Pavón) Tubers. Sep. Purif. Technol. 2007, 55, 217–225. [Google Scholar] [CrossRef]
  32. Castro-López, C.; Bautista-Hernández, I.; González-Hernández, M.D.; Martínez-Ávila, G.C.G.; Rojas, R.; Gutiérrez-Díez, A.; Medina-Herrera, N.; Aguirre-Arzola, V.E. Polyphenolic Profile and Antioxidant Activity of Leaf Purified Hydroalcoholic Extracts from Seven Mexican Persea Americana Cultivars. Molecules 2019, 24, 173. [Google Scholar] [CrossRef]
  33. Saini, A.; Pandey, A.; Sharma, S.; Suradkar, U.S.; Ambedkar, Y.R.; Meena, P.; Raman, R.; Gurjar, A.S. Assessment of Antioxidant Activity of Rosemary (Rosmarinus Officinalis) Leaves Extract. J. Pharmacogn. Phytochem. 2020, 9, 14–17. [Google Scholar]
  34. Singh, G.; Passsari, A.K.; Leo, V.V.; Mishra, V.K.; Subbarayan, S.; Singh, B.P.; Kumar, B.; Kumar, S.; Gupta, V.K.; Lalhlenmawia, H.; et al. Evaluation of Phenolic Content Variability along with Antioxidant, Antimicrobial, and Cytotoxic Potential of Selected Traditional Medicinal Plants from India. Front. Plant Sci. 2016, 7, 407. [Google Scholar] [CrossRef] [PubMed]
  35. Shabbir, H.; Kausar, T.; Noreen, S.; Rehman, H.U.; Hussain, A.; Huang, Q.; Gani, A.; Su, S.; Nawaz, A. In Vivo Screening and Antidiabetic Potential of Polyphenol Extracts from Guava Pulp, Seeds and Leaves. Animals 2020, 10, 1714. [Google Scholar] [CrossRef] [PubMed]
  36. Dorta, E.; Lobo, M.G.; Gonzalez, M. Reutilization of Mango Byproducts: Study of the Effect of Extraction Solvent and Temperature on Their Antioxidant Properties. J. Food Sci. 2012, 77, C80–C88. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, Q.-W.; Lin, L.-G.; Ye, W.-C. Techniques for Extraction and Isolation of Natural Products: A Comprehensive Review. Chin. Med. 2018, 13, 20. [Google Scholar] [CrossRef] [PubMed]
  38. Suzery, M.; Nudin, B.; Nurwahyu Bima, D.; Cahyono, B. Effects of Temperature and Heating Time on Degradation and Antioxidant Activity of Anthocyanin from Roselle Petals (Hibiscus Sabdariffa L.). Int. J. Sci. Technol. Manag. 2020, 1, 238–288. [Google Scholar] [CrossRef]
  39. Ghafoor, K.; Ahmed, I.A.M.; Doğu, S.; Uslu, N.; Fadimu, G.J.; Al Juhaimi, F.; Babiker, E.E.; Özcan, M.M. The Effect of Heating Temperature on Total Phenolic Content, Antioxidant Activity, and Phenolic Compounds of Plum and Mahaleb Fruits. Int. J. Food Eng. 2019, 15, 20170302. [Google Scholar] [CrossRef]
  40. Liu, R.; Mabury, S.A. Synthetic Phenolic Antioxidants: A Review of Environmental Occurrence, Fate, Human Exposure, and Toxicity. Environ. Sci. Technol. 2020, 54, 11706–11719. [Google Scholar] [CrossRef]
  41. Rumpf, J.; Burger, R.; Schulze, M. Statistical Evaluation of DPPH, ABTS, FRAP, and Folin-Ciocalteu Assays to Assess the Antioxidant Capacity of Lignins. Int. J. Biol. Macromol. 2023, 233, 123470. [Google Scholar] [CrossRef]
  42. Tran, T.T.T.; Ton, N.M.N.; Nguyen, T.T.; Le, V.V.M.; Sajeev, D.; Schilling, M.W.; Dinh, T.T.N. Application of Natural Antioxidant Extract from Guava Leaves (Psidium Guajava L.) in Fresh Pork Sausage. Meat Sci. 2020, 165, 108106. [Google Scholar] [CrossRef]
  43. Chiari-Andréo, B.G.; Trovatti, E.; Marto, J.; Almeida-Cincotto, M.G.J.D.; Melero, A.; Corrêa, M.A.; Chiavacci, L.A.; Ribeiro, H.; Garrigues, T.; Isaac, V.L.B. Guava: Phytochemical Composition of a Potential Source of Antioxidants for Cosmetic and/or Dermatological Applications. Braz. J. Pharm. Sci. 2017, 53, e16141. [Google Scholar] [CrossRef]
  44. Cherubim, D.J.; Martins, C.V.; Fariña, L.; Lucca, R.A. Polyphenols as Natural Antioxidants in Cosmetics Applications. J. Cosmet. Dermatol. 2020, 19, 33–37. [Google Scholar] [CrossRef]
  45. Khan, M.K.; Paniwnyk, L.; Hassan, S. Polyphenols as Natural Antioxidants: Sources, Extraction and Applications in Food, Cosmetics and Drugs. In Plant Based “Green Chemistry 2.0”. Green Chemistry and Sustainable Technology; Springer: Singapore, 2019; pp. 197–235. [Google Scholar]
  46. Eze, F.N.; Ovatlarnporn, C.; Nalinbenjapun, S.; Sripetthong, S. Ultra-Fast Sustainable Synthesis, Optimization and Characterization of Guava Phenolic Extract Functionalized Nanosilver with Enhanced Biomimetic Attributes. Arab. J. Chem. 2022, 15, 104167. [Google Scholar] [CrossRef]
  47. Wang, L.; Wu, Y.; Xie, J.; Wu, S.; Wu, Z. Characterization, Antioxidant and Antimicrobial Activities of Green Synthesized Silver Nanoparticles from Psidium Guajava L. Leaf Aqueous Extracts. Mater. Sci. Eng. C 2018, 86, 1–8. [Google Scholar] [CrossRef]
  48. Lobo, V.; Patil, A.; Phatak, A.; Chandra, N. Free Radicals, Antioxidants and Functional Foods: Impact on Human Health. Pharmacogn. Rev. 2010, 4, 118. [Google Scholar] [CrossRef]
  49. Lok, B.; Sandai, D.; Baharetha, H.M.; Nazari, V.M.; Asif, M.; Tan, C.S.; Majid, A.A. Anticancer Effect of Psidium Guajava (Guava) Leaf Extracts against Colorectal Cancer through Inhibition of Angiogenesis. Asian Pac. J. Trop. Biomed. 2020, 10, 293. [Google Scholar] [CrossRef]
  50. Nagpal, T.; Alam, S.; Khare, S.K.; Satya, S.; Chaturvedi, S.; Sahu, J.K. Effect of Psidium Guajava Leaves Extracts on Thermo-Lipid Oxidation and Maillard Pathway Born Food Toxicant Acrylamide in Indian Staple Food. J. Food Sci. Technol. 2022, 59, 86–94. [Google Scholar] [CrossRef]
  51. Lahlou, J.; Amraoui, B.E.; El-Wahidi, M.; Bamhaoud, T. Chemical Composition, Antioxidant and Antimicrobial Activities of Moroccan Species of Psidium Guajava Extracts. Rocz. Panstw. Zakl. Hig. 2022, 73, 65–77. [Google Scholar] [CrossRef]
  52. Naseer, S.; Hussain, S.; Naeem, N.; Pervaiz, M.; Rahman, M. The Phytochemistry and Medicinal Value of Psidium Guajava (Guava). Clin. Phytoscience 2018, 4, 32. [Google Scholar] [CrossRef]
  53. Nguyen, T.L.A.; Bhattacharya, D. Antimicrobial Activity of Quercetin: An Approach to Its Mechanistic Principle. Molecules 2022, 27, 2494. [Google Scholar] [CrossRef]
  54. Tian, C.; Liu, X.; Chang, Y.; Wang, R.; Lv, T.; Cui, C.; Liu, M. Investigation of the Anti-Inflammatory and Antioxidant Activities of Luteolin, Kaempferol, Apigenin and Quercetin. South Afr. J. Bot. 2021, 137, 257–264. [Google Scholar] [CrossRef]
  55. Azeem, M.; Hanif, M.; Mahmood, K.; Ameer, N.; Chughtai, F.R.S.; Abid, U. An Insight into Anticancer, Antioxidant, Antimicrobial, Antidiabetic and Anti-Inflammatory Effects of Quercetin: A Review. Polym. Bull. 2023, 80, 241–262. [Google Scholar] [CrossRef]
  56. Fan, F.-Y.; Sang, L.-X.; Jiang, M. Catechins and Their Therapeutic Benefits to Inflammatory Bowel Disease. Molecules 2017, 22, 484. [Google Scholar] [CrossRef]
  57. Sharifi-Rad, J.; Quispe, C.; Castillo, C.M.S.; Caroca, R.; Lazo-Vélez, M.A.; Antonyak, H.; Polishchuk, A.; Lysiuk, R.; Oliinyk, P.; De Masi, L.; et al. Ellagic Acid: A Review on Its Natural Sources, Chemical Stability, and Therapeutic Potential. Oxid. Med. Cell Longev. 2022, 2022, 3848084. [Google Scholar] [CrossRef]
  58. Patel, D.K.; Patel, K. Therapeutic Importance of Avicularin for the Treatment of Liver Inflammatory Disorders. Metabolism 2021, 116, 154549. [Google Scholar] [CrossRef]
  59. Jamieson, S.; Wallace, C.E.; Das, N.; Bhattacharyya, P.; Bishayee, A. Guava (Psidium Guajava L.): A Glorious Plant with Cancer Preventive and Therapeutic Potential. Crit. Rev. Food Sci. Nutr. 2023, 63, 192–223. [Google Scholar] [CrossRef]
  60. Li, Y.; Xu, J.; Li, D.; Ma, H.; Mu, Y.; Huang, X.; Li, L. Guavinoside B from Psidium Guajava Alleviates Acetaminophen-Induced Liver Injury via Regulating the Nrf2 and JNK Signaling Pathways. Food Funct. 2020, 11, 8297–8308. [Google Scholar] [CrossRef]
  61. Jan, R.; Khan, M.; Asaf, S.; Lubna; Asif, S.; Kim, K.-M. Bioactivity and Therapeutic Potential of Kaempferol and Quercetin: New Insights for Plant and Human Health. Plants 2022, 11, 2623. [Google Scholar] [CrossRef]
  62. Rothwell, J.A.; Perez-Jimenez, J.; Neveu, V.; Medina-Remon, A.; M’Hiri, N.; Garcia-Lobato, P.; Manach, C.; Knox, C.; Eisner, R.; Wishart, D.S.; et al. Phenol-Explorer 3.0: A Major Update of the Phenol-Explorer Database to Incorporate Data on the Effects of Food Processing on Polyphenol Content. Database 2013, 2013, bat070. [Google Scholar] [CrossRef]
  63. Rajput, R.; Kumar, K. Protective Effect of Ethanolic Extract of Guava Leaves (Psidium Guajava L.) in Alloxan-Induced Diabetic Mice. Mater. Today Proc. 2021, 47, 437–439. [Google Scholar] [CrossRef]
  64. Azizan, N.A.; Wahab, N.Z.A.; Mohamad, N.A.; Shambely, A.S.; Othman, A.S. Antimicrobial Activity of Psidium Guajava Leaves Extract Against Foodborne Pathogens. Int. J. Psychosoc. Rehabil. 2020, 24, 318–326. [Google Scholar]
  65. Truong, D.-H.; Nguyen, D.H.; Ta, N.T.A.; Bui, A.V.; Do, T.H.; Nguyen, H.C. Evaluation of the Use of Different Solvents for Phytochemical Constituents, Antioxidants, and In Vitro Anti-Inflammatory Activities of Severinia Buxifolia. J. Food Qual. 2019, 2019, 8178294. [Google Scholar] [CrossRef]
  66. Nobossé, P.; Fombang, E.N.; Mbofung, C.M.F. Effects of Age and Extraction Solvent on Phytochemical Content and Antioxidant Activity of Fresh Moringa Oleifera L. Leaves. Food Sci. Nutr. 2018, 6, 2188–2198. [Google Scholar] [CrossRef]
  67. Nagaraja, S.K.; Nayaka, S.; Kumar, R.S. Phytochemical Analysis, GC–MS Profiling, and In Vitro Evaluation of Biological Applications of Different Solvent Extracts of Leonotis Nepetifolia (L.) R.Br. Flower Buds. Appl. Biochem. Biotechnol. 2023, 195, 1197–1215. [Google Scholar] [CrossRef]
  68. Aguilar-Morones, S.; Guerrero-Barrera, A.L.; Díaz-Villaseñor, J.M. Efecto Del Extracto de Hoja de Psidium Guajava L. En La Formación de Biopelículas de Escherichia Coli Uropatógena in-Vitro. Bachelor’s Thesis, Universidad Autonoma de Aguascalientes, Aguascalientes, Mexico, 2018. [Google Scholar]
  69. Tachakittirungrod, S.; Okonogi, S.; Chowwanapoonpohn, S. Study on Antioxidant Activity of Certain Plants in Thailand: Mechanism of Antioxidant Action of Guava Leaf Extract. Food Chem. 2007, 103, 381–388. [Google Scholar] [CrossRef]
  70. Palomares Salvador, I. Obtención de Fenoles Totales de Extractos Hidroalcohólicos en Diferentes Tiempos de Maceración de Hojas de Guayaba (Psidium Guajava L.) y de Aguacate (Persea Americana Mill.). 2019. Available online: https://repositorioslatinoamericanos.uchile.cl/handle/2250/2909123 (accessed on 18 May 2023).
  71. Coelho, J.M.P.; Johann, G.; da Silva, E.A.; Palú, F.; Vieira, M.G.A. Extraction of Natural Antioxidants from Strawberry Guava Leaf by Conventional and Non-Conventional Techniques. Chem. Eng. Commun. 2021, 208, 1131–1142. [Google Scholar] [CrossRef]
  72. Li, J.; Wu, C.; Li, F.; Yu, R.; Wu, X.; Shen, L.; Liu, Y.; Zeng, W. Optimization of Ultrasound-Assisted Water Extraction of Flavonoids from Psidium Guajava Leaves by Response Surface Analysis. Prep. Biochem. Biotechnol. 2019, 49, 21–29. [Google Scholar] [CrossRef]
  73. Amalia, A.; Suryono, S.; Endro Suseno, J.; Kurniawati, R. Ultrasound–Assisted Extraction Optimization of Phenolic Compounds from Psidium Guajava L. Using Artificial Neural Network-Genetic Algorithm. J. Phys. Conf. Ser. 2018, 1025, 012020. [Google Scholar] [CrossRef]
  74. Ratu Ayu, H.; Suryono, S.; Endro Suseno, J.; Kurniawati, R. Determination of the Ultrasound Power Effects on Flavonoid Compounds from Psidium Guajava L. Using ANFIS. J. Phys. Conf. Ser. 2018, 1025, 012024. [Google Scholar] [CrossRef]
  75. Setyaningsih, W.; Saputro, I.E.; Palma, M.; Barroso, C.G. Stability of 40 Phenolic Compounds during Ultrasound-Assisted Extractions (UAE). AIP Conf. Proc. 2016, 1755, 080009. [Google Scholar]
  76. Saraya, S.; Kanta, J.; Sarisuta, N.; Temsiririrkkul, R.; Suvathi, Y.; Samranri, K.; Chumnumwat, S. Development of Guava Extract Chewable Tablets for Anticariogenic Activity against Streptococcus Mutans. Mah Univ. J. Pharma Sci. 2008, 35, 18–23. [Google Scholar]
  77. Hernández, C.; Ascacio-Valdés, J.; De la Garza, H.; Wong-Paz, J.; Aguilar, C.N.; Martínez-Ávila, G.C.; Castro-López, C.; Aguilera-Carbó, A. Polyphenolic Content, in Vitro Antioxidant Activity and Chemical Composition of Extract from Nephelium Lappaceum L. (Mexican Rambutan) Husk. Asian Pac. J. Trop. Med. 2017, 10, 1201–1205. [Google Scholar] [CrossRef] [PubMed]
  78. Castro-López, C.; Sánchez-Alejo, E.J.; Saucedo-Pompa, S.; Rojas, R.; Aranda-Ruiz, J.; Martínez-Avila, G.C.G. Fluctuations in Phenolic Content, Ascorbic Acid and Total Carotenoids and Antioxidant Activity of Fruit Beverages during Storage. Heliyon 2016, 2, e00152. [Google Scholar] [CrossRef] [PubMed]
  79. Bautista-Hernández, I.; Aranda-Ledesma, N.E.; Rojas, R.; Tafolla-Arellano, J.C.; Martínez-Ávila, G.C.G. Antioxidant Activity of Polyphenolic Compounds Obtained from Euphorbia Antisyphilitica By-Products. Heliyon 2021, 7, e06734. [Google Scholar] [CrossRef] [PubMed]
  80. Aranda-Ledesma, N.E.; González-Hernández, M.D.; Rojas, R.; Paz-González, A.D.; Rivera, G.; Luna-Sosa, B.; Martínez-Ávila, G.C.G. Essential Oil and Polyphenolic Compounds of Flourensia Cernua Leaves: Chemical Profiling and Functional Properties. Agronomy 2022, 12, 2274. [Google Scholar] [CrossRef]
Figure 1. Mean and SEM of crude and purified samples in different antioxidant capacity assays.
Figure 1. Mean and SEM of crude and purified samples in different antioxidant capacity assays.
Molecules 29 00085 g001
Figure 2. Folin–Ciocalteu reagent reducing capacity according to the extraction method. SoxMeOH: Soxhlet, methanol, 65 °C, 5 h; Sox H 2 O : Soxhlet, distilled water, 100 °C, 5 h; M37: maceration, methanol, 37 °C, 192 h; M25: maceration, methanol, 25 °C, 192 h; UAEMeOH30: ultrasound, methanol, 30 °C, 0.66 h; UAEMeOH23: ultrasound, methanol, 23 °C, 0.66 h; UAE H 2 O 30: ultrasound, distilled water, 30 °C, 0.66 h; UAE H 2 O 23: ultrasound, distilled water, 23 °C, 0.66 h. Methods that do not share a letter are significantly different (p < 0.05).
Figure 2. Folin–Ciocalteu reagent reducing capacity according to the extraction method. SoxMeOH: Soxhlet, methanol, 65 °C, 5 h; Sox H 2 O : Soxhlet, distilled water, 100 °C, 5 h; M37: maceration, methanol, 37 °C, 192 h; M25: maceration, methanol, 25 °C, 192 h; UAEMeOH30: ultrasound, methanol, 30 °C, 0.66 h; UAEMeOH23: ultrasound, methanol, 23 °C, 0.66 h; UAE H 2 O 30: ultrasound, distilled water, 30 °C, 0.66 h; UAE H 2 O 23: ultrasound, distilled water, 23 °C, 0.66 h. Methods that do not share a letter are significantly different (p < 0.05).
Molecules 29 00085 g002
Figure 3. Iron-reducing capacity according to the extraction method. SoxMeOH: Soxhlet, methanol, 65 °C, 5 h.; Sox H 2 O : Soxhlet, distilled water, 100 °C, 5 h.; M37: maceration, methanol, 37 °C, 192 h; M25: maceration, methanol, 25 °C, 192 h; UAEMeOH30: ultrasound, methanol, 30 °C, 0.66 h; UAEMeOH23: ultrasound, methanol, 23 °C, 0.66 h; UAE H 2 O 30: ultrasound, distilled water, 30 °C, 0.66 h; UAE H 2 O 23: ultrasound, distilled water, 23 °C, 0.66 h. Methods that do not share a letter are significantly different (p < 0.05).
Figure 3. Iron-reducing capacity according to the extraction method. SoxMeOH: Soxhlet, methanol, 65 °C, 5 h.; Sox H 2 O : Soxhlet, distilled water, 100 °C, 5 h.; M37: maceration, methanol, 37 °C, 192 h; M25: maceration, methanol, 25 °C, 192 h; UAEMeOH30: ultrasound, methanol, 30 °C, 0.66 h; UAEMeOH23: ultrasound, methanol, 23 °C, 0.66 h; UAE H 2 O 30: ultrasound, distilled water, 30 °C, 0.66 h; UAE H 2 O 23: ultrasound, distilled water, 23 °C, 0.66 h. Methods that do not share a letter are significantly different (p < 0.05).
Molecules 29 00085 g003
Figure 4. FTIR analysis of the methanolic extracts of guava leaves. M825: maceration at 25 °C; SoxMeOH: Soxhlet at 65 °C; UAE30: ultrasound-assisted extraction at 30 °C.
Figure 4. FTIR analysis of the methanolic extracts of guava leaves. M825: maceration at 25 °C; SoxMeOH: Soxhlet at 65 °C; UAE30: ultrasound-assisted extraction at 30 °C.
Molecules 29 00085 g004
Table 1. Extraction yield of purified polyphenolic compounds.
Table 1. Extraction yield of purified polyphenolic compounds.
Extraction MethodExtraction Yield
TechniqueSolventTemperature (°C)Time (min)mg of Phenolic Compounds/g of Guava Leaf
SoxhletMethanol65544
SoxhletDistilled water100524
MacerationMethanol3719221
MacerationMethanol2519222
UltrasoundMethanol300.6618
UltrasoundMethanol230.6619
UltrasoundDistilled water300.6645
Table 2. I C 50 values according to the extraction method obtained via the ABTS•+ assay.
Table 2. I C 50 values according to the extraction method obtained via the ABTS•+ assay.
Extraction Method I C 50 (µg·mL−1)
Soxhlet, methanol, 65 °C, 5 h 119 ± 6   B
Soxhlet, distilled water, 100 °C, 5 h 137 ± 12   A B
Maceration, methanol, 37 °C, 192 h 89 ± 3   C
Maceration, methanol, 25 °C, 192 h 87 ± 6   C
Ultrasound, methanol, 30 °C, 0.66 h 78 ± 4   C
Ultrasound, methanol, 23 °C, 0.66 h 78 ± 4   C
Ultrasound, distilled water, 30 °C, 0.66 h 123 ± 4   B
Ultrasound, distilled water, 23 °C, 0.66 h 152 ± 12   A
Different letters mean significant (p < 0.05) differences between the extraction methods.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gutierrez Montiel, D.; Guerrero Barrera, A.L.; Martínez Ávila, G.C.G.; Gonzalez Hernandez, M.D.; Chavez Vela, N.A.; Avelar Gonzalez, F.J.; Ramírez Castillo, F.Y. Influence of the Extraction Method on the Polyphenolic Profile and the Antioxidant Activity of Psidium guajava L. Leaf Extracts. Molecules 2024, 29, 85. https://doi.org/10.3390/molecules29010085

AMA Style

Gutierrez Montiel D, Guerrero Barrera AL, Martínez Ávila GCG, Gonzalez Hernandez MD, Chavez Vela NA, Avelar Gonzalez FJ, Ramírez Castillo FY. Influence of the Extraction Method on the Polyphenolic Profile and the Antioxidant Activity of Psidium guajava L. Leaf Extracts. Molecules. 2024; 29(1):85. https://doi.org/10.3390/molecules29010085

Chicago/Turabian Style

Gutierrez Montiel, Daniela, Alma Lilian Guerrero Barrera, Guillermo Cristian Guadalupe Martínez Ávila, María Dolores Gonzalez Hernandez, Norma Angelica Chavez Vela, Francisco Javier Avelar Gonzalez, and Flor Yazmin Ramírez Castillo. 2024. "Influence of the Extraction Method on the Polyphenolic Profile and the Antioxidant Activity of Psidium guajava L. Leaf Extracts" Molecules 29, no. 1: 85. https://doi.org/10.3390/molecules29010085

Article Metrics

Back to TopTop