Research Article | Open Access
Abdulhakim Bawadekji1 , Mohd. Imran2 and Mohammad Akram Randahawa3
1Department of Biological Sciences, College of Science, Northern Border University, P.O. Box 1321, Arar 91431, Saudi Arabia.
2Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Northern Border University, P.O. Box 840, Rafha 91911, Saudi Arabia.
3Department of Pharmacology, Hazrat Bari Sarkar Medical & Dental College, Tarlai 45550, Islamabad, Pakistan.
J. Pure Appl. Microbiol., 2019, 13 (4): 2189-2194 | Article Number: 5888
https://doi.org/10.22207/JPAM.13.4.31 | © The Author(s). 2019
Received: 28/09/2019 | Accepted: 29/10/2019 | Published: 11/12/2019
Abstract

The purpose of this work was to perform the phytochemical and antimicrobial activity evaluation of the water-immiscible solvent extracts of Olive tree leaves (Olea europaea L., Family: Oleaceae). Seven sample extracts of the leaves of O. europaea were obtained using dichloromethane (DCM), dichloroethane (DCE) along with their mixtures with chloroform (CH), and ethyl acetate (EA). The phytochemical studies were carried out using the standard procedures. Serial plate dilution technique was used to perform the antimicrobial activity of the extracts. The phytochemical tests revealed the presence of steroids, terpenoids, saponins and flavonoids in all extracts of olive leaves. The 1:1 mixture of ethyl acetate and dichloromethane (EA:DCM) exhibited equivalent MIC (20 µgml-1, 100%) values concerning Staphylococcus aureus, Escherichia coli, and Pseudomonas aeruginosa with respect to ofloxacin (MIC = 20 µgml-1, 100%). Similarly, the 1:1 mixture of ethyl acetate and dichloroethane (EA:DCE) exhibited equivalent MIC (20 µgml-1, 100%) values concerning E. coli and P. aeruginosa with respect to ofloxacin (MIC = 20 µgml-1, 100%). The 1:1 mixture of EA:DCM and EA:DCE showed moderate antifungal activity (MIC = 20 µgml-1, 75%) with respect to fluconazole against Aspergillus niger and Candida albicans. The water immiscible solvent extracts showed more potent antimicrobial activity than the water extract of the olive leaves, which might be because of the presence of the lipophilic compounds in the water immiscible solvent extracts of the olive leaves. It is also expected that by increasing the concentration of ethyl acetate in EA:DCM & EA:DCE solvent system may further provide better MIC values.

Keywords

Olive leaves, Water immiscible solvent, Extract, Antimicrobial.

Introduction

Antimicrobial resistance (AMR) has initiated the emergence of multidrug-resistant (MDR) pathogens and has also made management of many bacterial infections problematic1. It is informed that the infections triggered by antibiotic resistant bacteria execute about twenty thousand patients yearly in the USA and also primes to economic loss2. The circumstanaces of AMR infections are growing in Saudi Arabia due to inappropriate usage of antibiotics as well as owing to the socioeconomic and demographic physiognomies of different population residing in Saudi Arabia3,4. Additional factor causative to the expansion of AMR is letdown to ascertain newer antimicrobial agents5,6. At present, synthetic antimicrobials are extensively used for prophylactic or to cure many infections. The haphazard use of synthetic antimicrobials poses a severe threat to humankind5,6, because multidrug resistance is developing among the disease triggering microorganisms. Consequently, researchers are motivated to develop antimicrobial phytomedicines, which are non-toxic and can support to overcome the appearance of multidrug resistance concern5,6.

Olive tree (Olea europaea L.) relates to the Oleaceae family, which contains twenty-four genus and nine hundred species7. Many parts like leaves, fruits, and stems of this plant have been investigated for its biological effects7-12, for example, antioxidant activity, anti-inflammatory activity, antithrombotic activity, antihypertensive activity, cardioprotective activity, and hypoglycemic activity. The various extracts of the olive tree leaves have also been investigated for their antimicrobial activity13-20. According to the literature, mainly the extracts of the water-miscible solvents, have been investigated for the antimicrobial activity7-20. Because of the above facts, it has been decided to perform the phytochemical and antimicrobial activity evaluation of the water-immiscible solvent extracts of Olive tree leaves (Olea europaea L., Family: Oleaceae) that can lead to the identification of the possible new class of the phytoconstituents as antimicrobial agents.

Materials and Methods

Collection of the Plant Material
The leaves of the Olive tree (Olea europaea L., 2 kg, semi-dried) were acquired from irrigated plantation on 1st of April 2019 at Arar city, in northern border region, Saudi Arabia. The olive leaves were validated by Directorate of Environment, Water and Agriculture at Northern Border region at Arar city. Specimen were kept the facility of College of Pharmay at Rafha city with the reference N. Oli 19. The leaves were cleaned and dried in air for seven days at room temperature (25°C-30°C). The completely dried material was ground into a powder using a grinder. The coarse powder was sifted and kept in polythene bags for the experiment/extraction purpose.

Preparation of the Extracts
The powdered leaves (50 g) were taken in a 1000 ml flask. Dichloromethane (DCM, 500 ml) was added to the flask. The flask was plugged, and the combination was stirred at 25°C for about 30 minutes. The flask was kept for three days at (25°C-30°C) with infrequent shaking. The content was filtered with Whatman filter paper, and the obtained filtrate was concentrated in a rotary evaporator to get the semisolid residue. On the other hand, and in the similar manner, the extracts of dichloroethane (DCE), and the 1:1 mixtures of chloroform:dichloroethane (CH:DCE), chloroform:dichloromethane (CH:DCM), dichloroethane: dichloromethane (DCE:DCM), ethyl acetate:dichloroethane (EA:DCE), and ethyl acetate: dichloromethane (EA:DCM) were obtained.

Phytochemical Studies
The Mayer’s test & Wagner’s test for alkaloids; FeCl3 test and Lead acetate test for tannins; Baljet test and Kellar Killani test for cardenolides; Liebermann-Burchard test for steroids; Salkowski’s test for terpenoids; Foam test for saponins; Borntrager’s test for Anthraquinone; and Flavonoid test for Flavonoids were carried out using the prepared extracts of Olive leaves as per the standard procedures21,22.

Antimicrobial Screening
Serial plate dilution procedure was employed to perform the antimicrobial activity of the extracts23-26. The minimum inhibitory concentrations (MICs) of the prepared extracts, fluconazole, and ofloxacin were defined against six microorganisms. Different dilutions (10-200 µgml-1) of fluconazole, ofloxacin, and the extracts were prepared in sterilised dimethylformamide (DMF) solvent. The sterilised DMF was also used as control. The nutrient agar media was used for antibacterial activity assessment while Sabouraod dextrose media was used for the antifungal activity assessment. Three bacteries and two fungi; Escherichia coli Migula 1895 Castellani and Chalmers 1919, Pseudomonas aeruginosa (Schröter, 1872) Migula 1900, Staphylococcus aureus Rosenbach 1884, Aspergillus niger Van Tieghem and Candida albicans (C.–P. Robin) Berkhout (1923) were tested for antimicrobial activity. All strains were purchased from the College of Pharmacy, Rafha, Saudi Arabia.

Statistical Analysis
The data (N = 3, Mean±Standard Error Mean) was analysed by SPSS-software, in which p < 0.05 specified the significant results.

RESULTS AND DISCUSSION

Seven sample extracts of the leaves of O. europaea were obtained using dichloromethane (DCM), dichloroethane (DCE) along with their mixtures with chloroform (CH), and ethyl acetate (EA) (Table 1).

Table (1):
Physical data of the extracts of O. europaea leaves.

Extract
Percentage Yield
State
Colour
DCM
12
Semisolid
Green
DCE
14
Semisolid
Green
DCM:DCE (1:1)
12
Semisolid
Green
CH:DCM (1:1)
12
Semisolid
Green
CH:DCE (1:1)
15
Semisolid
Green
EA:DCM (1:1)
14
Semisolid
Light Green
EA:DCE (1:1)
15
Semisolid
Light Green

The greenish semisolid extracts were obtained in the yield of 12% to 15%, wherein the highest yield (15%) was obtained for the 1:1 mixture of EA and DCE, and the 1:1 mixture of CH and DCE. The lowest yield (12%) was obtained for the DCM, 1:1 mixture of DCM and DCE, and 1:1 mixture of CH and DCM. It is suggested that this yield variation was because of the solvent/solvent system polarity. Standard chemical tests21,22 were used to determine the presence of alkaloids, tannins, cardenolides, steroids, terpenoids, saponins, anthraquinones, and flavonoids (Table 2).

Table (2):
Phytochemical assessment data of the extracts of O. europaea leaves.

Phytochemical
DCM
DCE
DCM:DCE (1:1)
CH:DCM (1:1)
CH:DCE (1:1)
EA:DCM (1:1)
EA:DEC (1:1)
Alkaloids
Anthraquinone
Cardenolides
Flavonoids
+
+
+
+
+
+
+
Saponins
+
+
+
+
+
+
+
Steroids
+
+
+
+
+
+
+
Tannins
Terpenoids
+
+
+
+
+
+
+

(+) = Present & (-) = Absent

These tests discovered the presence of steroids, terpenoids, saponins and flavonoids in all extracts of olive leaves, wherein these tests revealed the absence of alkaloids, tannins, cardenolides, and tannins, cardenolides, and anthraquinones. These findings were in concurrence with the previously published report on the phytochemistry of the leaves of O. europaea27, wherein it was shown that steroids, terpenoids, saponins and flavonoids are among the main chemical constituents of the leaves of O. europaea.

Serial plate dilution technique23-26 was employed to perform the antimicrobial activity of the olive leaves extracts against two fungi and four bacteria fluconazole and ofloxacin as standard drugs, respectively. The MIC values of the extracts, ofloxacin, and fluconazole concerning the tested microorganisms are presented in Table 3.

Table (3):
Antimicrobial assessment data of the extracts of O. europaea leaves.

Solvent Extract MIC (µgml-1)(% MIC concerning the standard)
Escherichia coli
(Gram -ve)
Klebsiella pneumonia
(Gram -ve)
Pseudomonas aeruginosa
(Gram -ve)
Staphylococcus aureus
(Gram +ve)
Aspergillus niger Candida albicans
DCM 25
(80%)
30
(66.66%)
30
(66.66%)
40
(50%)
40
(37.5%)
40
(37.5%)
DCE 25
(80%)
30
(66.66%)
30
(66.66%)
25
(80%)
30
(50%)
30
(50%)
DCM:DCE (1:1) 30
(66.66%)
30
(66.66%)
25
(80%)
30
(66.66%)
30
(50%)
30
(50%)
CH:DCM (1:1) 30
(66.66%)
35
(57.14%)
30
(66.66%)
40
(50%)
50
(30%)
40
(37.5%)
CH:DCE (1:1) 25
(80%)
25
(80%)
25
(80%)
25
(80%)
25
(60.0%)
25
(60.0%)
EA:DCM (1:1) 20
(100.0%)
25
(80.0%)
20
(100.0%)
20
(100.0%)
20
(75.0%)
20
(75.0%)
EA:DCE (1:1) 20
(100%)
25
(80%)
20
(100%)
25
(80%)
20
(75%)
20
(75%)
Fluconazole 15
(100%)
15
(100%)
Ofloxacin 20
(100%)
20
(100%)
20
(100%)
20
(100%)

*p<0.05 & N = 3.

For comparison, the MIC values of standard drugs (fluconazole & ofloxacin) are considered 100%. All the values exhibited statistically significant (p < 0.5) results. It is obvious from Table 3 data that the 1:1 mixture of EA:DCM exhibited equivalent MIC (20 µgml-1, 100%) values concerning S. aureus, E. Coli, and P. aeruginosa with respect to ofloxacin (MIC = 20 µgml-1, 100%). Similarly, the 1:1 mixture of EA:DCE exhibited equivalent MIC (20 µgml-1, 100 %) values concerning E. coli and P. aeruginosa with respect to ofloxacin (MIC = 20 µgml-1, 100 %). The 1:1 mixture of EA:DCM and EA:DCE also showed most promising MIC values ranging from 75% to 80% with respect to standard drugs, ofloxacin and fluconazole, against K. pneumoniae, C. albicans, and A. niger. All other extracts revealed antimicrobial activity ranging from 50% to 80% with respect to standard drugs against the tested microorganisms. As per published reports28, the conventional antibiotics have MIC range from 15 µgml-1 to 105 µgml-1. Accordingly, it can be assumed that our extracts are potent antimicrobials. However, this generalized assumption may not be applicable to our extracts because our extracts are not single compounds like conventional antibiotics, rather they may contain a mixture of many compounds. It has been suggested that the water-immiscible solvent systems (EA:DCM & EA:DCE) having more polarity had better antimicrobial activity in terms of the MIC values. On the other hand, the water-immiscible solvent systems having lesser polarity had mild to moderate antimicrobial activity in terms of the MIC values. It is expected that the antimicrobial activity of the water-immiscible solvent extracts is attributed to the occurrence of steroids, terpenoids, saponins and flavonoids in the solvent extracts. Secondly, more amount of steroids, terpenoids, saponins and flavonoids might be existing in the higher polarity solvent system of the present study (EA:DCM & EA:DCE), which attributed to the equivalent MIC values against ofloxacin, and fluconazole27. Thirdly, the water-immiscible solvent systems are generally more lipophilic. The lipophilic constituents can easily interrupt with the lipophilic cell membrane of the microbes. This may be the reason for better antimicrobial activity of the EA:DCM & EA:DCE solvent system20,25. The results also suggested that our extracts have lesser MIC value than the water extract of the olive leaves13. The water extracts of the olive leaves13 exhibited MIC values of 0.13 mgml-1 (P. aeruginosa), 0.3 mgml-1 (K. pneumonia and E. coli), and 0.6 mgml-1 (S. aureus), whereas our extracts showed their MIC values in the range of 20 µgml-1 to 40 µgml-1. This shows that our extracts are more potent antimicrobials as compared to the water extract of the olive leaves. The possible reason for this phenomenon is that the bacterial cell wall mainly consists of lipid. Our water immiscible extracts contain lipid soluble compounds, which can easily pass through the lipid membrane of the microbes. The water extract mainly contains water soluble compounds, which may not be able to cross the lipid membrane of the microbes20,25.

CONCLUSION

The water immiscible solvent extracts showed more potent antimicrobial activity than the water extract of the olive leaves, which might be because of the presence of the lipophilic compounds in the water immiscible solvent extracts of the olive leaves. The 1:1 mixture of EA:DCM & EA:DCE exhibited the most promising and in some cases, equivalent MIC values with respect to ofloxacin and fluconazole. The promising MIC values of these two solvent system extracts might be due to the presence of more amount of steroids, terpenoids, saponins and flavonoids. It is also expected that by increasing the concentration of ethyl acetate in EA:DCM & EA:DCE solvent system may further provide better MIC values.

Declarations

ACKNOWLEDGMENTS
The authors gratefully acknowledge the approval and the support of this research study by the grant no. 7451-MED-2017-1-8-F from the Deanship of Scientific Research at Northern Border University, Arar, Saudi Arabia.

CONFLICT OF INTEREST
The authors declare that there is no conflict of interest.

AUTHORS’ CONTRIBUTION
All authors have made substantial, direct and intellectual contribution to the work and approved it for publication.

FUNDING
Grant no. 7451-MED-2017-1-8-F from the Deanship of Scientific Research at Northern Border University, Arar, Saudi Arabia.

ETHICS STATEMENT
This article does not contain any studies with human participants or animals performed by any of the authors.

AVAILABILITY OF DATA
All datasets generated or analyzed during this study are included in the manuscript and/or the Supplementary Files.

References
  1. Zowawi HM. Antimicrobial resistance in Saudi Arabia. Saudi Med. J., 2016; 37(9): 935-940.
    Crossref
  2. Harbarth S, Balkhy HH, Goossens H, Jarlier V, Kluytmans J, Laxminarayan R, Saam M, Belkum AV, Pittet D. Antimicrobial resistance: one world, one fight!. Antimicrob Resist Infect Control, 2015; 4: 49.
    Crossref
  3. Emeka PM, Al-Omar M, Khan TM. Public attitude and justification to purchase antibiotics in the Eastern region Al Ahsa of Saudi Arabia. Saudi Pharm. J., 2014; 22(6): 550-554.
    Crossref
  4. El Zowalaty ME, Belkina T, Bahashwan SA, El Zowalaty AE, Tebbens JD, Abdel-Salam HA, Khalil AI, Daghriry SI, Gahtani MA, Madkhaly FM, Nohi NI, Khodari RH, Sharahili RM, Dagreery KA, Khormi M, Habibah SA, Medrba BA, Gahtani AA, Hifthi RY, Zaid JM, Amshan AW, Alneami AA, Noreddin A, Vlcek J. Knowledge, awareness, and attitudes toward antibiotic use and antimicrobial resistance among Saudi population. Int. J. Clin. Pharm., 2016; 1-8.
    Crossref
  5. Jindal AK, Pandya K, Khan ID. Antimicrobial resistance: A public health challenge. Med. J. Armed Forces India, 2015; 71(2): 178-181.
    Crossref
  6. Brandt C, Makarewicz O, Fischer T, Stein C, Pfeifer Y, Werner G, Pletz MW. The bigger picture: the history of antibiotics and antimicrobial resistance displayed by scientometric data. Int. J. Antimicrob. Agents, 2014; 44(5): 424-430.
    Crossref
  7. Abunab H, Dator WL, Hawamdeh S, Effect of olive leaf extract on glucose levels in diabetes-induced rats: A systematic review and meta-analysis, J. Diabetes, 2017; 9(10), 947-957.
    Crossref
  8. Hashmi MA, Khan A, Hanif M, Farooq U, Perveen S. Traditional Uses, Phytochemistry, and Pharmacology of Olea europaea (Olive). Evid Based Complement Alternat Med., 2015; 2015: 541591.
    Crossref
  9. Boss A, Bishop KS, Marlow G, Barnett MP, Ferguson LR, Evidence to Support the Anti-Cancer Effect of Olive Leaf Extract and Future Directions, Nutrients, 2016; 8(8). pii: E513.
    Crossref
  10. Vogel P, Kasper Machado I, Garavaglia J2, Zani VT, de Souza D, Morelo Dal Bosco S, Polyphenols benefits of olive leaf (Olea europaea L.) to human health, Nutr. Hosp., 2014, 31(3): 1427-33.
  11. Abaza L, Taamalli A, Nsir H, Zarrouk M. Olive tree (Olea europeae L.) Leaves: Importance and Advances in the Analysis of Phenolic Compounds. Antioxidants (Basel), 2015, 4(4): 682-698.
    Crossref
  12. El SN, Karakaya S. Olive tree (Olea europaea) leaves: potential beneficial effects on human health. Nutr. Rev., 2009, 67(11): 632-638.
    Crossref
  13. Markin D, Duek L, Berdicevsky I. In vitro antimicrobial activity of olive leaves. Mycoses, 2003; 46(3-4): 132-136.
    Crossref
  14. Motamedifar M, Nekooeian A.A, Moatari A. The effect of hydroalcoholic extract of olive leaves against herpes simplex virus type 1. Iran J. Med. Sci., 2007; 32(4): 222-226.
  15. Pereira AP, Ferreira IC, Marcelino F, Valentao P, Andrade PB, Seabra R, Estevinho L, Bento A, Pereira JA. Phenolic compounds and antimicrobial activity of olive (Olea europaea L. Cv. Cobrancosa) leaves. Molecules, 2007; 12(5): 1153-1162.
    Crossref
  16. Aliabadi, M.A., R.K. Darsanaki, M.L. Rokhi, M. Nourbakhsh and G. Raeisi. Antimicrobial activity of olive leaf aqueous extract. Annals of Biol. Res., 2012; 3(8): 4189-4191.
  17. Keskin D, Ceyhan N, Ugur A, Durgan Dbeys A. Antimicrobial activity and chemical constitutions of west anatolian olive (Olea europaea L.) leaves. Journal of Food, Agriculture & Environment, 2012; 10(2): 99-102.
  18. AbbasvaliM, Esmaeili Koutamehr M, Moshtaghi H, Eskandari M. Effect of olive leaf extracts on Salmonella typhimurium. Online Journal of Veterinary Research, 2015; 19(9): 632-641.
  19. Salma Nassrullah Malik. Antibacterial Activity of Olive (Olea europaea) Leaves and Arugula (Eruca sativa) Seeds Extract. International Journal of Pharmacognosy and Phytochemical Research, 2015; 7(2): 307-310.
  20. Gumgumjee N, Hajar A. Antimicrobial Activities and phytochemical properties of Saudi Olea europaea subsp. cuspidata. Life Sci. J., 2014; 11(6s): 232-237.
  21. Asghari, G., Palizban, A., Bakhshaei, B. Quantitative analysis of the nutritional components in leaves and seeds of the Persian Moringa peregrina (Forssk.) Fiori. Pharmacog Res., 2015; 7(3): 242-248.
    Crossref
  22. Trease, G., Evans, W. 1972. Pharmacognosy, Univ. Press, Aberdeen, Great Britain. Robbers, J., M. Speedie, and V. Tyler. 1996. Pharmacognosy and pharmacobiotechnology. Williams and Wilkins, Baltimore.
  23. Barry, A.L. Procedures and theoretical considerations for testing antimicrobial agents in agar media. In: Lorian V, Ed. Antibiotics in Laboratory Medicine, 3rd edn. Baltimore, Williams & Wilkins, 1991, pp 1-16. 24.
  24. Varma, R.S., Khan, Z.K., Singh, A.P. Eds. Antifungal agents: past, present, future prospects. Lucknow: National Academy of Chemistry and Biology, India, 1998, pp 55- 128.
  25. Imran, M., Abida, Khan, S.A. Synthesis and antimicrobial activity of some 2-amino-4-(7-substituted/unsubstitutedcoumarin-3-yl)-6-(chlorosubstitutedphenyl)pyrimidines. Trop. J. Pharm. Res., 2015; 14(7): 1265-1272.
    Crossref
  26. Imran, M., Abida, Alsalman, A.J. Synthesis and evaluation of antimicrobial activity of some 2-morpholinomethylamino-4-(7-unsubstituted/substitutedcoumarin-3-yl)-6-chlorosubstituted phenyl pyrimidines. Trop. J. Pharm. Res., 2016; 15(2): 393-404.
    Crossref
  27. Salma Nassrullah Malik, Antibacterial Activity of Olive (Olea europaea) Leaves and Arugula (Eruca sativa) Seeds Extract, International Journal of Pharmacognosy and Phytochemical Research, 2015; 7(2): 307-310.
  28. Cermak P, Olsovska J, Mikyska A, Dusek M, Kadleckova Z, Vanicek J, Nyc O, Sigler K, Bostikova V, Bostik P. Strong antimicrobial activity of xanthohumol and other derivatives from hops (Humulus lupulus L.) on gut anaerobic bacteria. APMIS Journal of Pathology, Microbiology and Immunology, 2017; 125: 1033–1038.
    Crossref

Article Metrics

Article View: 2044

Share This Article

© The Author(s) 2019. Open Access. This article is distributed under the terms of the Creative Commons Attribution 4.0 International License which permits unrestricted use, sharing, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.