Skip to main content

Antibiotic resistance modifying ability of phytoextracts in anthrax biological agent Bacillus anthracis and emerging superbugs: a review of synergistic mechanisms

Abstract

Background and objectives

The chemotherapeutic management of infections has become challenging due to the global emergence of antibiotic resistant pathogenic bacteria. The recent expansion of studies on plant-derived natural products has lead to the discovery of a plethora of phytochemicals with the potential to combat bacterial drug resistance via various mechanisms of action. This review paper summarizes the primary antibiotic resistance mechanisms of bacteria and also discusses the antibiotic-potentiating ability of phytoextracts and various classes of isolated phytochemicals in reversing antibiotic resistance in anthrax agent Bacillus anthracis and emerging superbug bacteria.

Methods

Growth inhibitory indices and fractional inhibitory concentration index were applied to evaluate the in vitro synergistic activity of phytoextract-antibiotic combinations in general.

Findings

A number of studies have indicated that plant-derived natural compounds are capable of significantly reducing the minimum inhibitory concentration of standard antibiotics by altering drug-resistance mechanisms of B. anthracis and other superbug infection causing bacteria. Phytochemical compounds allicin, oleanolic acid, epigallocatechin gallate and curcumin and Jatropha curcas extracts were exceptional synergistic potentiators of various standard antibiotics.

Conclusion

Considering these facts, phytochemicals represents a valuable and novel source of bioactive compounds with potent antibiotic synergism to modulate bacterial drug-resistance.

Background

The emergence of multidrug-resistant bacteria has become a threat to global public health and invoked problems, resulting in inadequate treatment for infectious diseases. The chemotherapeutic management of bacterial infections has become more challenging in recent years due to the development of antimicrobial resistance in pathogenic bacteria and their certain populations evolving into formidable super drug-resistant strains known as superbugs that are capable of causing serious illnesses [1, 2]. The bacterium Bacillus anthracis was one of the organisms of interest in the recent past due to its ability to cause the life-threatening illness anthrax. It can be difficult to treat anthrax if progressed to advanced stages, due to the virulence nature of the pathogen and therefore, research has been undertaken to discover novel therapies to treat the disease more safely and effectively [3]. Antibiotics have the ability to restrict the growth and replication of bacteria by inhibiting bacterial cellular components associated with the synthesis of cell walls, proteins and nucleic acids, along with the suspension of folate metabolism and depolarization of the cell membrane [4, 5]. The successful management of bacterial infections has been achieved in the past due to the development of antibiotic agents. It seems, however, that the golden era of these synthetically produced antibiotics has come to a near end due to their irrational use. Although antibiotics have broad-spectrum activities, bacteria have evolved to combat the action of these agents through various resistance mechanisms, such as the production of antibiotic molecule inactivating enzymes, the modification and mutation of antibiotic binding sites, suspension of bacterial cell membrane porin activity associated and the expression of efflux pumps [6, 7]. Another ingenious strategy of overcoming antimicrobials is the deployment of autoinducer molecules by certain bacteria to mediate quorum sensing. Quorum sensing (QS) is a bacterial cell-to-cell biochemical communication process that involves the activation of specific signals to coordinate pathogenic behaviors and assist bacteria to acclimatize against unfavorable conditions imposed by the proximal environment that they exist. Signal molecules responsible for mediating bacterial quorum sensing include autoinducing peptides, autoinducer-2 and acyl-homoserine lactone. Quorum sensing in bacteria can facilitate biofilm formation which agitates the penetration of antibiotic molecules and therefore, it is a major contributor towards antibiotic resistance [8]. Despite the complex nature of issues related to antibiotic resistance, no individual nation has independently succeeded in addressing this major public health problem up to date.

The rise of antimicrobial resistance has mediated interest in research focusing on the significance of medicinal plants and their phytochemical compositions. Long before the invention of modern antibiotics, folklore medicine was able to use the therapeutic efficacy of these medicinal herbs and integrate their potentials in the treatment of infectious diseases. So far, the World Health Organization (WHO) reports that about 80% of the populations residing in Asia and Africa rely on traditional medicine for their primary health care needs [9]. These traditional therapeutic methods have been considered as possibly the safest alternative sources of antimicrobial agents available, by which the involvement of medicinal herbs in treating infectious diseases has paved the way for the development of modern medicine [10]. Plants are reservoirs of chemical agents with therapeutic properties beneficial to mankind [11]. Bioactive compounds naturally extracted from whole plants or from different parts of plants, like leaves, bark, stem, roots, fruits, fruit rind, seeds and flowers, can serve as novel sources for the management of infectious diseases caused by pathogenic microorganisms as an alternative to synthetic drugs [12, 13].

Certain phytochemical compounds are capable to interact synergistically with antibiotics already available, which can be potentially an effective way to combat the phenomenon of resistance. There is evidence that combinations of natural compounds from plants can facilitate or improve the interaction of antibiotics with their target in the pathogen and thus reduce the emergence of resistance through mechanisms of resistance modification [1]. This combined therapeutic strategy can also reduce drug/dose-related side effects to the consumer since lower concentrations of both agents can be used. Therefore, the objectives of this review are to provide an update of the literature review on the synergism between antibiotics and plant extracts, presentation of experimental data on antibiotic-potentiating mechanisms of plant-derived compounds and scientific evidence that support the successive pre-clinical application of synergistic effects of combined plant compounds to serve as a starting point for the discovery of novel antibacterial agents that are capable of neutralizing infections and reversing antibiotic resistance in anthrax causative organism B. anthracis and the Centers for Disease Control and Prevention (CDC) classified emerging bacterial superbugs.

Antibacterial drug discovery and development

Long before the twentieth century, the management and treatment of infectious diseases were based mainly on folk medicine. There is evidence that complex mixtures with antibacterial properties have been applied among ancient populations for over two thousand years [14,15,16]. Post-mortem analysis has revealed the presence of traces of tetracycline like compound that has been incorporated into the dentals of early Sudanese populations that lived around A.D 350–500. The presence of this compound in their corpses has lead to the impression that these populations may have used it as a medicine or included in their diet [17, 18]. A similar finding was reported in ancient populations living in the Dakhleh Oasis, in Egypt, around the time of the late Roman Empire [19]. There is a popular anecdote that showed how to use the red soil in the Hashemite Kingdom of Jordan as a source of antimicrobials to treat skin infections. Interestingly, the bacterium known as Actinomycetes, which is generally found in such soils, produces modern antibiotics, such as actinomycin, streptomycin, erythromycin, nystatin, amphotericin and vancomycin [20, 21]. Traditional Chinese medicine consists of a large summary of medicinal herbs used for millennia in the treatment of many infections caused by bacteria [22, 23]. Their application of active compounds from ancient medicinal herbs has enriched the arsenal of many antibacterial agents used in modern medicine [24]. The modern era of antibacterial agents began with the discovery of penicillin extracted from a mould specimen known as Penicillium notatum by Sir Alexander Fleming in 1928. Penicillin caught the attention of many ancient scientists, as the compound was able to stop the growth of a wide range of bacteria [25,26,27]. During the time of its discovery, penicillin became the most popular therapeutic agent due to the wide application and the magnitude of its therapeutic outcomes. The technologies used and developed to produce penicillin became the basis for the production of all subsequent antibiotics currently in use [28, 29]. Most antibiotics currently in existence, such as cephalosporins, penicillins, macrolides, tetracyclines, vancomycin, teicoplanin, daptomycin and rifamycin have been synthetically derived from natural products [30, 31]. According to the World Health Organization, more than 11% of modern drugs are derivatives from plants [32]. Advanced technologies such as high throughput screening, combinational chemistry and genomic applications have been implemented to invent new antibacterial molecules to reverse antibiotic resistance [33]. An investigation conducted by Kim Lewis in 2001 lead to the key discovery of synergistic compounds of plant origin. His finding elucidated that a compound known as 5'-methoxyhydnocarpin-D isolated from the extracts of Berberis fremontii was able to potentiate the antibacterial action of berberine, inhibiting the activity of multidrug efflux pumps in Gram-positive and Gram-negative bacteria [34,35,36]. Novel approaches have been taken to combat antibiotic resistance such as the development of therapeutics based on anti-QS agents or bacterial quorum quenchers from natural products and bacterial vaccines. Unlike conventional treatments involving antibiotics, these novel therapies can be more potent and robust in combating advanced conditions of antibiotic resistance and bacterial virulence [37]. The implementation of bacterial vaccines is well evident in controlling diseases like tetanus, diphtheria, cholera, bacterial meningitis, typhoid fever and even anthrax, where measures have been taken to neutralize the outbreak of the disease in Swedish nature reserves in 2011 by vaccinating the resident animals against anthrax when treatment with penicillin was ineffective. Given this ideal, it has been predicted that such approaches will be nigh-impervious to resistance in bacterial populations and more robustly prevent the spread of infection [38, 39].

Anthrax and the biological agent Bacillus anthracis

Anthrax is a serious enzootic infectious disease transmitted from infected livestock animals to humans. The biological agent responsible for causing anthrax is a Gram-positive endospore forming bacilliform bacterium known as Bacillus anthracis. Anthrax can be primarily acquired through direct contact and the consumption of contaminated meat. The most common forms of the disease under natural settings are cutaneous and gastrointestinal anthrax. Other but rare means of acquiring the disease include the inhalation of bacterial endospores that can result in pulmonary anthrax. The endospores remain dormant until being inhaled by a host and internalized, where they mature into toxin producing virulent bacterial cells in thoracic lymph nodes that cause acute and severe infection. The tactical delivery of concentrated endospores obtained from wild type B. anthracis is a strategy used in biological warfare and bio-terrorism. The disease is endemic to agricultural regions of southwestern and central Asia, Central and South America, southern sub-Saharan Africa, the Caribbean and Eastern Europe [40]. The largest agricultural outbreak of anthrax was reported from Zimbabwe, with cases of infection exceeding 10,000 between 1979 and 1985. It has been reported that nearly all cases of the infection occurred in Zimbabwe outbreak were cutaneous anthrax [41]. In the year 1979, about 79 cases related to inhalational anthrax were reported from Sverdlovsk region of Russia, in which 68 of those cases became ultimately fatal. The Sverdlovsk incident was the largest outbreak of human anthrax ever documented in history and is believed to have been caused due to an accident occurred in a Soviet military affiliated Microbiology facility that lead to the release of aerosolized anthrax spores [42]. The most recent incident related to anthrax bioterrorism was reported in 2001, in which concentrated spores of highly virulent B. anthracis were delivered using postal letters that resulted in about five fatalities among the 22 infected [43]. The disease incidence was significantly reduced during the twentieth century. Hence, anthrax continued to represent globally outside the United States, with an occurrence of approximately 2000 cases annually by the end of the twentieth century. A majority of these worldwide cases were associated with cutaneous anthrax [42]. The standard therapy for anthrax include with antibiotics like penicillin G procaine, doxycycline or ciprofloxacin being first-line treatment for the infection [44]. Newer anthrax therapeutic agents like monoclonal antibody based Anthim (obiltoxaximab) and the anthrax immune globulin based Anthrasil were also deployed solo or in combination with antibiotics to control the infection more effectively [41, 45]. In 2015, the Food and Drug Administration (FDA) approved BioThrax which is an immunologically active B. anthracis antigen vaccine against anthrax to prevent the disease. It is the only anthrax vaccine that has received FDA approval up to date [46]. The treatment of acute anthrax can be difficult due to the virulence properties of B. anthracis, in which the bacterium and its endospore are both encapsulated with a protective polysaccharide coating that allows immune evasion from phagocytes like macrophages. Bacterial exotoxins secreted by B. anthracis are known as edema toxin and lethal toxin which cause diarrhea and flu-like symptoms. The entry of these exotoxins into host cells and initial pathogenesis is facilitated by a major virulence factor present in B. anthracis known as the protective antigen. The polysaccharide capsule and other associated virulence factors are expressed by bacterial DNA plasmids, pXO1 and pXO2 present in B. anthracis. Although the bacterium can be eradicated by antibiotic agents, the toxins produced by the bacterium remain nonresponsive to antibiotic therapy. Hence, the CDC recommends the employment of a combined course of rapid antibiotic therapy involving two or three antibiotics along with anthrax anti-toxin therapy in order to prevent the accumulation of exotoxins in the body [47,48,49]. Antibiotic resistance in B. anthracis has been documented. One study showed that 11.5% out of 96 isolates of B. anthracis recovered from France between the time period of 1994 and 2000 indicated resistance amoxicillin and penicillin G. The same investigation revealed that all 96 isolates of B. anthracis tested were resistant to cotrimoxazole [50]. Although drug resistance mechanisms of B. anthracis has not yet been fully exploited, a study conducted by Price et al. showed that efflux-pump encoding bacterial genes gyrA, gyrB and parC can mediate cross-resistance to fluoroquinolone antibiotics like ciprofloxacin in B. anthracis [51]. Another study stated that B. anthracis consist of genes bla1 and bla2 that are capable of expressing beta-lactamases against β-lactam antibiotics [52]. Despite the treatment of anthrax infection with currently available antibiotics, the introduction of safer and more efficient chemotherapeutic options are required. Studies have demonstrated anti-B. anthracis activity of novel compounds extracted from medicinal plants and therefore, new insights involving the efficiency of plant-derived compounds and antibiotic combinations exhibiting anti-anthrax potential are needed to be addressed.

Superbug bacteria

The continuous or inappropriate use of antibiotics has resulted in the development of extensive drug resistance in bacteria. Overtime, these organisms will progressively advance and evolve into superbug bacteria as an adaptive response to selective antibiotic pressure [53, 54]. The WHO has defined these bacteria as “Superbugs” due to the infections caused by these organisms are no longer treatable with existing antibiotic agents [55]. The CDC has categorized these organisms as urgent, serious, concerning and watch-list pathogenic threats. Antibiotic resistance is highly prevalent among Gram-positive bacteria, in which some of the well known examples of superbugs include methicillin-resistant Staphylococcus aureus (MRSA) and vancomycin-resistant enterococci (VRE). The United States national antimicrobial surveillance data has indicated the emergence more serious superbug infections associated with extended-spectrum beta-lactamases (ESBL) or carbapenemase producing organisms like Klebsiella pneumoniae and Eschericia coli, MDR pathogens like Pseudomonas aeruginosa, Vibrio cholera and Clostridium difficile, Salmonella spp. and drug-resistant Mycobacterium tuberculosis. However, some of the most worrisome threats up to date are associated with emerging superbugs like carbapenem-resistant Acinetobacter spp., particularly Acinetobacter baumannii [56].

Methicillin-resistant Staphylococcus aureus

Methicillin-resistant Staphylococcus aureus (MRSA) are Gram-positive bacteria arranged in round-shaped clusters. They are among the most successive pathogens in the modern time period. S. aureus can exist as part of the human flora and can cause opportunistic infections. Being genetically diverse, these organisms can evolve into epidemic strains like MRSA [57, 58]. These MRSA are formidable clinical threats and are considered as archetypal hospital superbugs that are responsible for causing serious bloodborne infections like sepsis and infective endocarditis [59, 60]. According to CDC, more than 120,000 cases of hospitalizations reported in the United States from 1999–2000 were due to S. aureus associated infections. Among these cases MRSA accounts for 43.2% of infections in those hospitalized [61]. Antibiotic agents vancomycin or daptomycin are usually the first-line treatment for MRSA bacteraemia and related infections. However, in case of severe infection, combined therapy with flucloxacillin is given for more effective treatment. The primary reason for methicillin and other β-lactam-resistance in S. aureus is due to the expression of a foreign PBP known as PBP2a by the mecA gene present in them. PBP2a variant binds with β-lactam antibiotics with reduced avidity, which mediates resistance to this class of antibiotics. Lower affinity of PBP2a to β-lactam agents allows MRSA to replicate due to peptidoglycan synthesis taking place in the presence of β-lactams antibiotics that are capable of inactivating transpeptidase activity of PBPs. PBP2a is composed of a non-penicillin-binding protein and transpeptidase domain. Mutations in S. aureus associated genes like mprF, yycH, and dltA are also known for conferring cross resistance to daptomycin. S. aureus are also capable of acquiring genes that encode for antibiotic resistance from their predecessors. MRSA consist of a wide range of dynamic virulence factors that include immune-evasive bacterial surface factors (e.g. protein A and capsule) and tissue invasion promoting enzymes like hyaluronidase and toxins (e.g. leukocidins and haemolysins) for mediating pathogenesis. Two additional virulence factors known as Panton–Valentine leukocidin and arginine-catabolic mobile element have been discovered in a MRSA isolate called USA300, which facilitated the rapid spread of the strain by improving its adaptability to the pH of the human skin [62, 63]. Research has been undergoing in order to introduce newer and more effective therapies for MRSA infections such as the development of vaccines and potent natural products.

Vancomycin-resistant enterococci

Vancomycin-resistant enterococci (VRE) are round-shaped Gram-positive bacteria that can cause serious MDR infections and persistent colonization in humans. Enterococci are opportunistic inhabitants that exist in the environment with an exceptional ability to adapt and evolve to transmit antibiotic-resistant determinants [64]. VRE can cause life-threatening infections in humans such as bloodstream infections like sepsis, endocarditis and pyelonephritis, in which most of these are nosocomial [65]. Tan et al. reported that a survey conducted from 2014 to 2016 showed that 523 out of 5,357 patients from health-care facilities in Singapore suffered from infections caused by VRE. An outbreak of VRE related infections has also been reported in 1997 from acute-care hospitals in the United Sates [64]. Linezolid is usually recommended as the first-line treatment for VRE and whereas, daptomycin, tigecycline and quinupristin-dalfopristin combined therapy are considered as last resort antibiotic agents for the management of enterococci related infections. The primary mechanism of drug resistance in enterococci involves the alteration of pathways associated with peptidoglycan synthesis, which specifically substitute D-Alanine-D-Alanine (D-Ala-D-Ala), to either D-Alanine-D-Serine (D-Ala-D-Ser) or D-Alanine-D-Lactate (D-Ala-D-Lac). These termini in VRE cell walls bind poorly with glycopeptide antibiotics like vancomycin [66]. A total of eight van gene clusters, vanA, vanB, vanD, vanE, vanG, vanL, vanM, and vanN are responsible for expressing elements required for antibiotic resistance in enterococci with vanA and vanB being the most abundant [67]. VanA is responsible for mediating the primary mechanism for antibiotic resistance in enterococci [68]. In addition to drug-resistance mechanisms, these enterococci consist of virulence factors like DNAse, caseinase and gelatinase to promote pathogenesis [69].

Extended-spectrum beta-lactamases and carbapenemase producing Klebsiella pneumoniae

Extended-spectrum β-lactamases (ESBL) and carbapenemase producing Klebsiella pneumoniae are rod-shaped Gram-negative bacteria that cause high morbidity and mortality among hospitalized patients under intensive-care and neonatal intensive-care. These organisms are capable of producing carbapenemase against carbapenems and a rapidly evolving class of β-lactamase enzymes known as extended-spectrum β-lactamase, which have the ability to hydrolyze the β-lactam ring of a range of third/fourth-generation cephalosporin antibiotics and render them ineffective. EBSL and carbapenemase mediated drug resistance to numerous antibiotics make it challenging to treat infections caused by these organisms. Klebsiella spp. are ubiquitous in nature that belong to the family of bacteria called Enterobacteriaceae. These organisms exist in the natural environment and are a part of the human flora. K. pneumoniae are opportunistic pathogens, which have the ability to colonize the respiratory tract, gastrointestinal tract, genitourinary tract and eyes of those vulnerable [70]. Appropriate first-line treatment for EBSL producing K. pneumoniae include with antibiotics like amoxicillin/clavulanic acid, ceftriaxone, ciprofloxacin or cotrimoxazole) [71]. In case of carbapenemase producing K. pneumoniae, the treatment options include high-dose or combined antibiotic therapy with meropenem, tigecycline and/or colistin, gentamicin or fosfomycin depending on susceptibility [72]. Müller-Schulte et al. stated that 94% of infections reported from the University Teaching Hospital in Bouaké, West Africa from 2016–2017 were caused by EBSL producing K. pneumonia [73]. A survey conducted from 2014–2015 in long-term acute care hospitals based in the United States indicated that nearly 25% of infections caused in hospitalized patients were due to carbapenemase producing K. pneumonia [74]. Genes responsible for coding EBSL mediated antibiotic resistance in K. pneumoniae include blaSHV, CTX-M and TEM [75]. Navon-Venezia et al. indicated that plasmid genes blaVIM-1, blaOXA-48, blaVIM and blaNDM-1 are responsible for carbapenemase mediated drug resistance in K. pneumoniae [76]. Additionally, K. pneumoniae consist of virulence factors for facilitation pathogenesis such as capsular (K) antigen for evading phagocytosis, O antigen for invasion of host cells and siderophores like enterobactin and aerobactin for iron acquisition [77].

Extended-spectrum beta-lactamases and carbapenemase producing Eschericia coli

ESBL and carbapenemase producing Eschericia coli are Gram-negative rod-shaped bacteria. These organisms are responsible for causing serious community and hospital-acquired infections worldwide, especially in places where inadequate hygienic practices are common and poor sanitation. E. coli is well known for causing gastroenteritis and infections associated with the urinary tract [78, 79]. In case of EBSL producing E. coli, the organism is responsible for causing bacteraemia in more than 5000 cases of hospitalized patients in the United Kingdom [80]. A prevalence survey conducted in a Spanish University Hospital indicated that 7.69% and 1.83% of admitted patients out of 10,643 suffered from infections associated with ESBL producing E. coli and carbapenemase producing E. coli respectively [81]. Antibiotics like colistin and carbapenems are usually the first-line treatment for ESBL producing E. coli [82]. Fritzenwanker et al. suggest that ertapenem infusion with meropenem or doripenem combine antibiotic therapy is given infections caused by carbapenemase producing E. coli [83]. A study conducted by Overdevest et al. showed that ESBL producing E. coli harbored plasmid genes like blaCTX-M-1 and blaTEM-52 [84]. Shin et al. detected blaNDM–5 gene in high level carbapenemase-resistant E. coli [85]. Aside from antibiotic resistance, E. coli has a number of virulence factors for mediating pathogenesis like heat-labile toxin, heat-stable toxin, enterohaemolysin, shiga-like toxin, enteroaggregative heat-stable, enterotoxin, haemolysin, cytotoxic necrotizing factor, uropathogenic specific protein and invasin for host cell invasion and K1-capsule and intimin for immune evasion and cellular attachment [86, 87].

Multidrug-resistant Pseudomonas aeruginosa

Pseudomonas aeruginosa are Gram-negative bacteria arranged in rods or bacilli. These organisms can be found in the environment (e.g. soil and water) and are known for causing blood borne infections and pneumonia in humans under opportunistic conditions [88]. P. aeruginosa are also associated with hospital-acquired infections, in which MDR P. aeruginosa is responsible for causing 32,600 infections among patients who were hospitalized and 2700 fatalities in the United States in 2017 [89]. Monotherapy and combined therapy with antibiotic agents like ceftolozane-tazobactam, ceftazidime-avibactam, cefiderocol and imipenem-relebactam/cilastatin are used for the treatment of infections caused by MDR strains of P. aeruginosa [90]. The most common mechanism of antibiotic resistance in P. aeruginosa is associated with the overproduction of drug efflux pump systems like MexAB-OprM, MexEF-OprN, MexXY-OprM and MexCD-OprJ induced by mex gene mutations. These multi-drug efflux pumps function as antibiotic molecule extruders. Apart from efflux pumps, these organisms also consist of genes like AmpC that code for the production of β-lactamases and OprD for encoding alterations in type II topoisomerases (DNA gyrase) to mediate resistance against fluoroquinolone and carbapenem antibiotics [91]. P. aeruginosa also expresses a number of virulence factors such as protease A, exotoxins, phospholipase C and cytotoxins for host cell invasion and pyoverdine and QS system regulatory proteins essential for the formation of biofilms, that plays a vital role in host immune evasion and antibiotic resistance [92].

Multidrug-resistant Vibrio cholera

Vibrio cholera are comma-shape Gram-negative bacteria well known for causing the severe water-borne acute, diarrheal illness cholera. These organisms have become a major threat to public health, particularly in the developing world [93]. According to recent reports, the V. cholerae O1 El Tor variant is responsible for causing cholera outbreaks worldwide [94]. The transmission of V. cholerae generally occurs via the faecal-oral route by ingesting contaminated water and food [95]. Typically, doxycycline is used as the first-line treatment for cholera infection caused by V. cholera [96]. According to CDC, V. cholerae colonizes the small intestine to cause cholera and an estimation of 2.9 million cases and 95,000 fatalities occur annually worldwide [97]. WHO estimates that 1.3 to 4 million cases of pathogenic infections and 21,000–143,000 fatalities reported across the globe are due to cholera [98]. It has been reported that V. cholerae have caused seven pandemics related to cholera in different countries [99]. MDR V. cholerae has a number of antibiotic resistance mechanisms such as active antibiotic molecule efflux, reduced cell wall permeability to antibiotics, alteration of binding targets sites for antibiotics via undergoing post-transcriptional or translational modifications (e.g. mutations in topoisomerase and DNA gyrase) and hydrolysis or chemo modification of antibiotic agents. These resistance mechanisms are expressed by genes like blaNDM-1, blaDHA-1, carR, ant 3’, tet(M), tetD, foIP, qacEΔ1, mph2, mel, armA, rmtB, rmtC, rmtF, aphA1, arr2, bcr, mphRK, mrx, blaP, vigA, blaCTX-M, sh ble, floR, cat, aacA, aphD, tetG, aac-Ib, qnrVC3, ereA2, bla, strA, strB, sul2, mdtH, rpsl, dfrA, dhfrII, aad3’ and mph in MDR V. cholerae [100]. Major virulence factors necessary for mediating pathogenesis and host cell invasion in V. cholerae include the cholera, toxin-coregulated pilus and O antigen [101].

Multidrug-resistant Clostridioides difficile

Clostridioides difficile are Gram-positive spore-forming bacteria that are ranged in rods or bacilli. These organisms are responsible for causing colitis and nosocomial diarrhea. C. difficile are opportunistic pathogens, in which they colonize the small intestine when the gut microbiota is disrupted as a result of antibiotic misuse [102]. Metronidazole is the first-line treatment mild to moderate infection, whereas advanced forms of infection are treated with vancomycin and fidaxomicin monotherapy or combined therapy. C. difficile infection is more prevalent among the elderly who have prescribed antibiotics for other conditions [103, 104]. CDC reports that 223,900 patients admitted to hospitals and 12,800 fatalities in the United States in 2017 were associated with C. difficile infections [105]. Data obtained from the French National Uniform Hospital in 2016 indicated that 3.6 cases per 10,000 acute care patient days account for infections caused by C. difficile [106]. MDR strains of C. difficile consist of genes like gyrA for mediating moxifloxacin and rpoB for mediating resistance against rifampicin. Moreover, C. difficile consist of genes associated with tetracycline resistance like tetM and genes that code for aminoglycoside-modifying enzymes like aac(6′)-aph(2″) and aadE. C. difficile also expresses genes like mef(A)-msr(D), ermG, and vat for coding resistance to antibiotics lincosamide, streptogramins and macrolide. A mutation associated with Cys721Ser PBP in C. difficile has been speculated to contribute resistance towards imipenem [107]. The bacterial exotoxins TcdA, TcdB and binary toxin are the common virulence factors associated with C. difficile for host invasion and promoting pathogenesis [108].

Drug-resistant Mycobacterium tuberculosis

Mycobacterium tuberculosis are acid-fast bacilli that has been categorized under serious threats by the CDC. These organisms are well known for causing the highly infectious lung associated disease named tuberculosis (TB) [109]. The transmission of M. tuberculosis occurs via droplet nuclei from an infected person [110]. TB caused by MDR, XDR or pandrug-resistant (PDR) strains of M. tuberculosis poses a serious threat to the public health worldwide, in which the disease has claimed the lives of 1.3 million and about 8.6 million cases of the infection has been reported in 2012 [111]. The WHO in 2016 estimated that there were 600,000 cases of TB and 240,000 fatalities attributed to the disease have been reported [112]. FDA approved first-line treatment anti-TB agents include isoniazid, rifampin, ethambutol, pyrazinamide [113]. In case of MDR or XDR TB, anti-TB drugs like pretomanid in combination with linezolid and bedaquiline are given [114]. Bacillus Calmette–Guérin immunotherapy is generally used to prevent TB, which uses a live-attenuated vaccine derived from Mycobacterium bovis to immunize against the TB infection [115]. M. tuberculosis consists of several genes capable of mediating antibiotic resistance via drug target modulation. These include katG, inhA, ndh and ahpC targeted against isoniazid, rpoB against rifampicin, pncA and rspA targeted to counter pyrazinamide, embCAB and embR that modulates binding sites of ethambutol, rpsL, rrs and gidB targeted against streptomycin, rrs and eis modulates binding sites of amikacin/kanamycin, ethA, inhA, ethR, ndh and mshA to counter ethionamide and gyrA and gyrB modifies DNA gyrase against fluoroquinolones [116]. The major virulence factors for promoting pathogenesis in M. tuberculosis include phthiocerol dimycocerosate for host cell invasion and phenolic glycolipids involved in the evasion of host immune responses and inducing macrophage toxicity [117].

Multidrug-resistant Salmonella spp.

Salmonella spp. are Gram-negative zoonotic disease causing enteric bacteria. Over 2600 serotypes of Salmonella have been identified, which are responsible for causing gastrointestinal diseases such as food poisoning. Depending on the nature of symptoms, Salmonella infections can be classified as non-typhoidal, paratyphoidal and typhoidal, in which both paratyphoidal and typhoidal Salmonella causes high fever (typhoid fever). Most cases of foodborne infections have been found to be associated with Salmonella enterica serovar Enteritidis [118,119,120]. The transmission of Salmonella occurs via the fecal–oral route by ingesting contaminated food [121]. Generally, the first-line treatment for Salmonella infections include with antibiotics like fluoroquinolones for adults and azithromycin for children pediatric patients. Alternatively, ceftriaxone can also be used as a first-line antibiotic therapy for Salmonella [122]. The FDA has approved Salmonella vaccines such as Vi bacterial polysaccharide (Vi antigen) under the brand name Typhim Vi and Vivotif that uses a live-attenuated ty21a strain via oral administration for the immunization against typhoid fever [123]. According to CDC, about 1.35 million community infections, 26,500 cases of hospitalizations, and 420 fatalities in the United States annually are associated with Salmonella [124]. Zhang et al. stated that about 70–80% of outbreaks of foodborne illnesses in China are caused by Salmonella bacteria [125]. A global study conducted by WHO indicated that 21,650,974 cases of Salmonella infections caused typhoid fever, which resulted in death among 216,510 of the infected and 5,412,744 cases suffered from paratyphoid fever [126]. Zhang et al. revealed that mutations in AcrAB gene mediate antibiotic resistance in Salmonella by overexpressing AcrAB-TolC bacterial efflux pumps. The findings of the same study showed that genetic mutations in GyrA and GyrB can alter DNA Gyrase in Salmonella, which are known for facilitating the development of resisatnce against ciprofloxacin. Another study revealed that aac(6′)-I gene is frequently associated with aminoglycoside resistance. Moreover, the β-lactamase producing blaCMY-2 gene and tetR** gene that encodes for tetracycline resistance were abundantly present in Salmonella [127]. There are a number of associated with Salmonella such as the Vi capsular antigen, somatic O antigen, H antigen (flagella), fimbriae and type III secretion systems that include Salmonella pathogenicity island 1 (SPI-1), SPI-2, which promote host cell invasion and pathogenesis [128, 129]. The cytolethal distending toxin in Salmonella has been found to cause typhoid fever among the infected. Other salmonellosis mediating toxins like pertussis-like toxin A and pertussis-like toxin B were also have been found to be expressed by these organisms [130].

Acinetobacter baumannii

Acinetobacter baumannii is accountable for causing most community and hospital-acquired infections. Overtime, these organisms can evolve into XDR or PDR superbugs as a result of continuous selective pressure and rendering the majority or all existing antibiotics ineffective. The CDC has alerted and listed A. baumannii as an organism that needs to be considered as an urgent threat [56, 131, 132]. A. baumannii is a coccobacillus Gram-negative bacterium that is known for colonizing the gastrointestinal, respiratory tract and the oral cavity of humans. It is also recognized as a formidable opportunistic pathogen that causes many forms of severe recalcitrant infections. A. baumannii infections are resulted frequently due to wound contamination [133]. However, it is a clinically dominant bacterial species that has a pronounced tendency to cause healthcare-associated nosocomial infections [134, 135]. A. baumannii has been listed under ESKAPE pathogens and is the most aggravating member of the Acb-complex (A. calcoaceticus, A. baumannii and Acinetobacter genomic species 13TU) that have been found to show high resistance to antibiotic agents which increases the risk of mortality among hospitalized patients under intensive care [133, 134, 136,137,138,139,140]. Although infections caused by this bacterium were able to keep under control in early 1970s, A. baumannii lately re-emerged as MDR and XDR strains with marked resistance to most antibiotics like gentamicin, nalidixic acid, minocycline, carbenicillin, and ampicillin. The bacterium exhibited resistance to a majority of antibiotics during early 1990s and by late 1990s, the only treatment of choice was carbapenems in combination with rifampicin [141, 142]. Presently, infections caused by MDR and XDR strains of A. baumannii were being treated with antibiotics like polymyxin B, colistin and tigecycline. However, more new strains of A. baumannii have been frequently reported that can exhibit resistance to the aforementioned antibiotics [143, 144]. The extensive resistance to antibiotics in A. baumannii is primarily due to the prevalence of adaptive multidrug efflux pumps like adeA, adeB, adeC, adeDE, adeABC, adeFGH, adeXYZ and adeIJK in the bacterium [145]. A. baumannii harbors a multitude of virulence factors for facilitating host cell invasion and pathogenesis such as the biofilm promoter outer membrane protein A, surface antigen 1, lipid A, phospholipase, secretion systems (type 1, type 2, type 4, type 5, type 6 and type V), siderophores for iron acquisition, binding domains for the acquisition of zinc and magnesium and drug resistance promoting QS systems [146]. These rapid mutations in A. baumannii make it one of the most challenging biological factors to human health and public health-care systems. The recent emergence of pandrug-resistant superbugs like A. baumannii has indicated an urgent necessity for the discovery of novel antibacterial agents and chemotherapeutic strategies [137, 139].

Plant-derived compounds as antibacterial agents

Plants are natural factories capable of producing a series of different phytochemical compounds. These compounds were produced in response to adverse biotic and abiotic environmental conditions. Phytoconstituents have a major impact on other plants, animals and microorganisms in their immediate environment that surrounds them [147]. Plant-derived constituents are biologically active organic compounds and are generally defined as secondary metabolites. These secondary metabolites are structurally diverse compounds that are classified into three primary groups as phenolic compounds (phenolic acids, simple phenols, flavonoids, quinones, coumarins and tannins), alkaloids and terpenes (Fig. 1). These compounds can be isolated from crude extracts and essential oils of plants. Complex mixtures of phytochemicals are represented in crude extracts that contain primary and secondary metabolites of different classes, chemical and biosynthetic. These compounds share some of the common mutual characteristics, such as volatility and/or polarity. Since antiquity, extracts obtained from medicinal plants have been known to have broad-spectrum antimicrobial activities and have been frequently studied and reviewed. Their profound antibacterial activity is generally recognized as a safe substance and, with the minimal risk of developing bacterial resistance, has qualified them as suitable sources for the development of new antibacterial agents [148, 149].

Fig. 1
figure 1

Chemical structures of plant compounds generally isolated from plants

Mechanism of action of plant-derived antibacterial compounds

The efficiency of antibacterial compounds derived from plants depends on several factors, such as features of test microorganism (type, species and strain), botanical source and composition of the bioactive phytochemical compounds, as well as the stage of development, time of harvesting of the plant material and most importantly, the method of plant extraction. Due to the complex nature of the compounds present in crude extracts of plants, they can exhibit multiple mechanisms of action on bacteria. These include the suspension of bacterial growth, function or viability, targeting bacteria virulence factors, potentiating the effectiveness of antibiotics as agents that modify bacterial resistance. Similar to antibiotic agents, these phytochemicals can inhibit the growth and replication of bacteria, disrupting the structure and function of the bacterial cell membrane [150], interrupting the synthesis of nucleic acids such as DNA or RNA [151], disrupting the intermediary metabolism [152], and the coagulatory induction of bacterial cytoplasmic constituents [153].

Several studies have been conducted to understand and illustrate the antibacterial action of phenolic compounds such as flavonoids, coumarins and tannins. Flavonoids are a diverse group of polyphenolic compounds that have the ability to inhibit the activities of DNA gyrase and DNA topoisomerase, energy metabolism mediated by NADH-cytochrome C reductase or inhibition of ATP synthase and the interruption of components involved in the synthesis of the cell wall and cell membrane [154]. Possible targets of quinones include peptidoglycan from the bacterial cell wall and enzymes associated with the cell membrane [151]. It is known that tannins cause the destabilization of the cell membrane and alterations in metabolic pathways and inactivation of membrane-bound proteins [155]. Phytochemicals like coumarins mediate the delay in bacterial cell respiration. Terpenes disrupt the bacterial cell membrane due to their lipophilic nature. Alkaloids are some of the most widely studied plant-derived compounds which can intercalate with bacterial DNA and enzymes associated with nucleic acids as esterase, DNA or RNA polymerases [156]. Examples of antibacterial mechanisms of action of plant secondary metabolites against CDC classified bacterial superbugs and anthrax biological agent B. anthracis are elucidated in Table 1.

Table 1 Representative studies on phytochemical constituents generally used for antibacterial activity against B. anthracis and superbug bacteria

Methods for evaluating the antibiotic synergistic activity of plant-derived compounds

Growth inhibitory indices

The agar diffusion assay based synergistic activity of antibiotic-plant extract combinations can be evaluated using the application of the growth inhibitory indices (GIIs), calculated according to the formula below:

$$\rm{GIIs }= \frac{\rm{Inhibiton \, zone \, diameter \, of \, antibiotic }+\rm{ plant \, extract \, combined \, disk}}{\rm{Inhibition \, zone \, diameter \, of \, plant \, extract \, and \, antibiotic \, in \, individual \, action}}$$

The GIIs value > 1 will be considered as synergistic, 1 as additive, and < 1 as antagonistic [157].

Fractional inhibitory concentration index

The fractional inhibitory concentration index (FICI) is used for the evaluation of synergism between two antimicrobial compounds in micro/macrobroth dilution assays. The FICI of the antibiotic-plant extract combination agent can be estimated using the standard formula shown below:

$${\rm{FIC}}_{\rm{Plant \, extract}/\rm{compound }}= \frac{\rm{MIC \, of \, plant \, extract}/\rm{compound \, in \, combination \, with \, antibiotic}}{\rm{MIC \, of \, plant \, extract}/\rm{compound}}$$
$${\rm{FIC}}_{\rm{Antibiotic }}= \frac{\rm{MIC \, of \, plant \, extract}/\rm{compound \, in \, combination \, with \, antibiotic }}{\rm{MIC \, of \, antibiotic}}$$
$${\rm{FIC \, index}}_{\rm{Plant \, and \, antibiotic \, combination}}= {\rm{FIC}}_{\rm{Plant \, extract}/\rm{compound }}+ {\rm{FIC}}_{\rm{Antibiotic}}$$

FIC index ≤ 0.5 will be considered as synergistic, > 0.5 but < 1 as partially synergistic, additive when = 1, indifferent when > 1 but < 4 and ≥ 4 as antagonistic [158].

Synergistic interaction between phytoextracts and antibiotics

Synergism between plant-derived compounds and existing antibiotic agents is an effective and an efficient way to manage the development of bacterial multi-drug resistance [159, 160]. Several studies have shown the significance of this type of synergistic interaction in the discovery of novel antibacterial agents. Phytochemicals are cable of interacting with synthetic antibiotic agents. This phytochemical and antibiotic interaction has been classified as antagonistic, additive or synergistic. The term antagonistic is given when a plant-derived compound reduces the effectiveness of an antibiotic agent against a certain type of bacteria, whereas the terms additive and synergistic are assigned to compounds that can enhance the antibacterial activity of the antibiotic [161]. An additive effect is usually considered as the baseline effect for detecting synergy in antimicrobial assays, in which such effect can be theoretically expected from a combination of multiple antimicrobial agents when the synergistic effect is absent. Synergistic effect can be defined as a combined effect that is significantly greater than the additive effect. A plant extract/compound fused with an antibiotic agent can be considered as a synergistic product when their combined action is superior to that of their individual antibacterial activity [162]. The distinctive action of phytochemical-antibiotic synergism is the ability to overcome antimicrobial resistance. Besides reducing antibiotic resistance, another advantage of this type of synergism is that it can reduce the minimum inhibitory concentration of an antibiotic agent, which also lower the dose needed for its effect to take place and mitigation of possible adverse effects [9, 163].

A recent study indicated that epigallocatechin gallate isolated from Camellia sinensis, which is a variant of green tea was able to potentiate the antibacterial action of sulfamylon (mafenide acetate) against a clinical isolate of A. baumannii [164]. Another study showed that phytoextracts obtained from plants like Alstonia scholaris, Adenium obesum, Cerbera odollam, Cerbera manghus, Nerium oleander, Holarrhena antidysenterica, Plumeria obtuse, Wrightia pubescens, Thevetia peruviana, Punica granatum, Terminalia bellirica, Quisqualis indica, Terminalia sp. and Terminalia chebula were able to synergistically potentiate the activity of seven antibiotic agents like cephazolin, rifampicin, meropenem, gentamicin, erythromycin, streptomycin, fusidic acid and novobiocin against A. baumannii ATCC 19606 [165,166,167,168]. Knezevic et al. reported that plant extracts obtained from Eucalyptus camaldulensis showed ciprofloxacin, polymyxin B and gentamicin potentiating effect when used against different strains of A. baumannii [169]. Isolated phytochemical compounds like tannic acid, catechol, cinnamic acid, ellagic acid, ferulic acid, gallic acid and syringic acid have exhibited novobiocin potentiating ability against A. baumannii JVC1053 [170]. A study indicates that extracts of Levisticum officinale was able to synergistically enhance the antibacterial action of cipofloxacin against MDR A. baumannii NCTC 13305 [171]. An investigation conducted by Mandal et al. showed that ethanolic extracts of Ocimum sanctum potentiated the antibacterial action of trimethoprim and chloramphenicol against S. typhi with highest GIIs ranging from 1.2 to 1.3 [157]. Plant extracts obtained from Peganum harmala L, Cassis italic Mill, Carthamus tinctorius have exhibited the antibiotic potentiating ability of ampicillin, cefotaxime, vancomycin, chloramphenicol, tetracyclines against A. baumannii with a FICI of 0.5 [172, 173]. Antibiotics like cefotaxime, tetracyclines, vancomycin, ampicillin and chloramphenicol showed enhanced antibacterial activity against clinical isolates of A. baumannii when used in conjunction with phytoextracts of Terminalia chebula and Senna italica Mill [167, 174]. Liu et al. revealed that isolated compounds present in Pithecellobium clypearia mediated synergistic antibacterial activity against A. baumannii when used in combination with imipenem, cefoperazone, ceftazidime, levofloxacin, amikacin, tetracycline and polymyxin B sulfate [175]. A study indicates that the isolated phytochemical compound known as berberine exhibited antibiotic potentiating ability of ciprofloxacin and Imipenem against A. baumannii [176]. Allicin (Fig. 2) isolated from Allium sativum (garlic) indicated FICIs of 0.5 and 0.38 when used in combination with cefazolin and oxacillin respectively against S. aureus and FICIs of 0.25 for both antibiotics when used in combination against S. epidermidis [158]. Ekambaram et al. showed that rosmarinic acid (Fig. 2) was able to synergistically potentiate the antibacterial activity of vancomycin, amoxicillin and ofloxacin when used in combination against S. aureus and MRSA with FICIs of 0.5 for each combination [177]. One study indicated that oleanolic acid (Fig. 2) was able to enhance the antibacterial action of kanamycin and gentamicin against A. baumannii. A FICI of 0.313 for gentamicin and 0.375 for kanamycin was indicated when combined with oleanolic acid [178]. Another study showed the antibiotic potentiating ability of Zingiber cassumunar extracts when used in combination with a range of broad-spectrum antibiotic agents like trimethoprim-sulfamethoxazole, amoxicillin, amoxicillin-clavulanic acid, piperacillin-tazobactam, ceftazidime, cefepime, ceftriaxone, imipenem, cefotaxime, meropenem, ertapenem, tetracycline, gentamicin, amikacin, gentamicin, doxycycline, ciprofloxacin and levofloxacin against an XDR strain of A. baumannii [179]. Isolated compounds from plants taken for the investigation of antibiotic synergism include bisbenzylisoquinoline, tetrandrine, carvacrol, curcumin, murucoidin, catechol, cinnamic acid, ferulic acid, gallic acid, syringic acid, berberine, methyl gallate, ethyl gallate, pyrogallol, myricetin-3-O-α-L-rhamnopyranoside, quercetin-3-o-α-l-rhamnopyranoside, 5,3',4',5'-tetrahydroxy-flavan-7-gallate, ellagic acid, (E)-3,2',4'-trihydroxy-3'-methoxychalcone, (2S)-5,7,2'-trihydroxyflavonone, 7-methyljuglone, isoimperatorin and tannic acid (Table 2 and Figs. 1, 2). A study indicate that crude seed and seed oil extracts of Jatropha curcas showed synergistic activity against both MDR and clinical strains of E. coli, P. monteilii, P. aeruginosa, MRSA, E. faecalis, MDR A. baumannii and P. chlororaphis when used in combination with cephazolin, rifampicin, fusidic acid ciprofloxacin, cefotaxime, rifampicin, moxifloxacin and ofloxacin. The seed oil extract of J. curcas combined with cefotaxime showed the best synergism in the study with an FICI of 0.005 against the clinical isolate of P. aeruginosa [180]. An investigation revealed that theaflavin (Fig. 2), an antioxidant polyphenol compound present in black tea was able to potentiate the antibacterial action of ampicillin with a FICI of 0.35 when tested against S. maltophilia [181]. Seed and root extracts obtained from Peganum harmala L. synergistically potentiated the antibacterial action of novobiocin and carbenicillin against B. anthracis isolated from a clinical specimen [182]. Moreover, a study conducted by Kouitcheu et al. showed that extracts obtained from Cylicodiscus gabunensis, Picralima nitida, Cassia arereh and Trichilia emetic interacted synergistically with gentamicin, erythromycin and kanamycin against V. cholera. These extracts were able to reduce the MICs of the selected antibiotic by 2 to 16 fold for V. cholera [183]. The hydro-alcholic extracts of Oliveria decumbens showed a synergistic effect when used in combination with tetracycline, oxacillin and doxycycline against Brucella melitensis. According to the investigation, O. decumbens-doxycycline combination indicated the highest anti-Brucella synergistic activity with a zone of inhibition that is 9 mm larger than the antibiotic alone [184]. Tsevelmaa et al. investigated the synergistic antibiotic potentiating activity of Caryopteris mongolica Bunge root extract against B. melitensis under in vitro conditions. The study showed that C. mongolica significantly reduced the neutrophil phagocytic activity of B. melitensis infected female BALB/c mice when used in combination with doxycycline compared to the antibiotic alone [185]. A study conducted by El-Tawab showed that plant extracts obtained from Camellia sinensis, Thymus vulgaris, Zingiber officinale, Curcuma longa and Pelargonium graveolens potentiated the antibacterial action of amoxicillin, gentamicin, doxycyciline, difloxacin against Listeria monocytogenes. P. graveolens-difloxacin combination indicated the highest anti-Listeria synergistic activity in the study, which induced an inhibition zone diameter of 8 mm larger than difloxacin alone [186].

Fig. 2
figure 2figure 2figure 2

Chemical structures of isolated plant compounds studied for antibiotic synergistic activity

Table 2 List of phytoextracts used in combination with antibiotics against B. anthracis and superbug bacteria for synergistic activity

Mechanism of synergistic activity of phytoextracts and antibiotic combined agents

Plant-derived compounds and agents combined with antibiotics have broad antibacterial activity against many types of bacteria. Several studies have indicated various antibacterial mechanisms of these combined compounds that highlight their ability to reverse antibiotic resistance. These mechanisms include the modification of active sites in the bacterial cell wall and the plasma membrane to increase the permeability of the antibiotic molecule, inhibition of extracellular enzymes that catalyze the modification or degradation of antibiotics, inactivation of efflux pumps to facilitate the intracellular accumulation of antibiotic molecules and disruption of quorum sensing signal molecules and their corresponding receptors (Fig. 3) [163]. Examples of phytoextract-antibiotic combinations and their synergistic effects/mechanisms against Gram-positive bacteria and A. baumannii are elucidated in Table 2.

Fig. 3
figure 3

Mechanisms of action of phytochemical synergists in reversing antibiotic resistance in bacteria. Plant-derived compounds are capable of modifying the course of drug resistance in bacteria by interacting with antibiotic binding target sites/receptors, inhibition of antibiotic degrading enzymes, increasing cell membrane permeability to antibiotic molecules and disruption of drug extruding efflux pumps like MFS: Major Facilitator Super-family, SMR: Small Multidrug Resistance super-family, MATE: Multidrug and Aoxic Compound Extrusion super-family, RND: Resistance-nodulation-cell Division super-family, ABC: ATP-binding Cassette super-family. ATP: Adenosine triphosphate, ADP: Adenosine diphosphate, H + : Hydrogen, EGCG: Epigallocatechin gallate, PBP2a: Penicillin-binding protein 2a

Plant-derived synergists as inhibitors of antibiotic binding site modification

Bacteria are capable of modifying the antibiotic binding target sites known as receptors (e.g. penicillin-binding proteins) to mediate antibiotic resistance. These alterations will no longer permit the binding of the antibiotic molecule to its specific receptor and permeate into the bacterial cell, rendering the antibiotic ineffective [187, 188]. Examples of this type of plant-derived synergistic compounds include corilagin, tellimagrandin I, pinoresinol, tiliroside, coronarin D, totatrol, baicalin, momorcharaside B and magnatriol B (Fig. 2) [1, 189,190,191,192]. Corilagin is a type of tannins isolated from Arctostaphylos uva-ursi, which indicates a MIC of 128 μg/mL against MRSA. However, the MIC dropped 2000 fold when used in combination with Oxacillin and β-Lactam antibiotics. Corilagin indicated strong synergism with an FICI of 0.5 with bactericidal action against MRSA [193]. Tellimagrandin I is another tannin compound that indicates a FICI of 0.39 for MRSA when used in combination with β-Lactam antibiotics. The combination of antibiotics and Tellimagrandin I had a MIC reduction of 128–512 fold when compared to the isolated phytochemical compound [194]. Phenolic compounds extracted from Verbena officinalis, Magnolia officinalis, Daphne genkwa and Momordica charantia such as pinoresinol, tiliroside, momorcharaside B, magnatriol B indicated a FICI of 0.375 for MRSA when used in combination with oxacillin. Bacteriological studies have indicated that these phytochemicals are capable of inhibiting PBP2a or PBP4 in MRSA [195]. Another study revealed that proanthocyanidin (Fig. 2) isolated from Vaccinium macrocarpon Aiton was able to synergistically interact with levofloxacin against H. pylori. Morphological investigations of the study revealed the reduction of PBP2a synthesis in H. pylori by proanthocyanidin [196]. A novel study revealed that garlic extracts obtained from the plant species Allium sativum L., which predominantly composed of phthalic acid (Fig. 2) and conceivably allicin showed synergistic antibiotic potentiating activity when used in combination with tetracycline, penicillin, rifampicin against the potential anthrax causing bio-agent B. anthracis strain Sterne 34F2. The indicative FICIs ranged from 0.5 to 0.8 for the selected plant-antibiotic combinations and microscopic analysis in the study detected garlic extract induced morphological disruptions on the cell wall of B. anthracis [49]. Isolated compounds like cinnamic acid, ferulic acid and p-coumaric acid are capable of inhibiting the synthesis of S. aureus cell membrane when combined with amikacin [197].

Plant-derived synergists as inhibitors of antibiotic degrading/modifying enzymes

Certain bacteria can produce extracellular enzymes like β-lactamases and transacetylase that can chemically alter or even degrade antibiotic molecules. These enzymes can effectively retard the action of the antibiotic and render the antibiotic agent ineffective against the bacterium [33, 198,199,200]. However, studies have shown naturally occurring plant-based compounds that can synergistically interact with antibiotics to overcome these bacterial defenses. Examples of this type of phytochemical synergists include baicalin, rugosin B, 5-O-Methylglovanon and epigallocatechin gallate (Fig. 2) [1]. Baicalin extracted from Scutellaria amoena is one of the generally studied examples of plant-derived compounds contributing to this type of synergism, which was able to inhibit the activity of β-lactamases in MRSA and facilitated the antibacterial action of β-Lactam antibiotics [1, 201]. Epigallocatechin gallate is a polyphenolic compound that belongs to a class of catechin. An investigation revealed that epigallocatechin gallate (Fig. 2) isolated from tea extracts was able to reduce the MIC to 4 mg/L of ampicillin/sulbactam when used in combination with the antibiotic. The compound indicated a good FIC between 0.19 and 0.56 for MRSA and another study revealed that the compound had a FICI between 0.126 and 0.625 for 28 strains of MRSA. The study showed that epigallocatechin gallate can reduce the activity of penicillinase and β-lactamases in MRSA [202]. 5-O-methylglovanon isolated from Glycosmis plants is an isoprenyl flavonoid compound with broad-spectrum antibacterial activity. The compound can lower the production of β-lactamases to facilitate the action of ampicillin in S. epidermidis and S. aureus [203].

Plant-derived synergists as inhibitors of active bacterial efflux pumps

Efflux pumps are among the most common bacterial defenses that lead to antibiotic resistance. These bacterial structures have the ability to extrude antibiotic molecules at a faster rate than the antibiotic can diffuse in the bacterial cell [204]. Efflux pumps are structurally present in Gram-positive and Gram-negative bacteria [205, 206]. There are several genes involved in the expression of these efflux pumps in Gram-positive and Gram-negative bacteria. Examples of such classes of genes include Tet, Acr, Ydh, Mex, Bla, Mdtef, and Nor [207]. These efflux pumps can be classified into five groups, depending on their capacity and drug extrusion mechanisms, such as MFS, SMR, MATE, RND and ABC (Fig. 3) [208]. Several studies have identified a number of plant-derived compounds that can counter the effects of these efflux pumps. Examples of such phytochemical synergists that modulated antibiotic resistance against Gram-positive bacteria include carnosic acid, carnosol, baicalin, erybraedin-a, sophoraflavanone-G, 2,6-dimethyl-4-phenyl-pyridine-3,5-dicarboxylic acid diethyl ester, myricetin, tiliroside, carnosic acid, carnosol, piperine, indoline, indirubin, capsaicin, kaempferol-3-o-α-l-(2,4-bis-E-pcoumaroyl) rhamnoside, reserpine, epicatechin gallate, 5'-methoxyhydnocarpin-D, pheophorbide A, isoflavonoids and tannic acid (Fig. 2) [1, 209,210,211]. Plant-derived synergists capable of modulating drug resistance facilitated by efflux pumps in A. baumannii include conessine and epigallocatechin gallate (Fig. 2).

The indoline compound indirubin isolated from Wrightia tinctoria indicated a high FICI of 0.45 for S. aureus SA199B when used in combination with ciprofloxacin. The compound was able to inhibit the NorA gene expressed efflux pump of the bacterium [212]. Capsaicin extracted from chili peppers (Capsicum) indicated similar synergistic action for S. aureus SA199 and S. aureus SA199B targeting their efflux pumps when used in combination with ciprofloxacin. Nevertheless, the compound reduced the MIC of ciprofloxacin by 2 to fourfold [213]. Carnosic acid and carnosol are terpenes isolated from Rosmarinus officinalis indicated a MIC of 64 μg/mL and 16 μg/mL respectively, against MDR S. aureus. However, the MIC decreased 32 fold for carnosol and 16 fold for carnosic acid when used in combination with erythromycin at a lower concentration of 10 μg/mL. It was found that their synergistic action was targeted at the S. aureus NorA efflux pumps [165]. Baicalin isolated from Thymus vulgaris L and Scutellaria baicalensis indicated synergistic action targeting NorA and TetK efflux pumps expressed in MRSA when used in combination with β-lactam antibiotics and tetracycline [201]. The flavonoid compound kaempferol rhamnoside demonstrated the ability to inhibit the activity of NorA efflux pumps in S. aureus when used in combination with ciprofloxacin and synergistically reduce the MIC by eightfold compared to the compound alone against the bacterium [214]. Epicatechin gallate (Fig. 2) is a type of catechin isolated from green tea extracts were able to interact synergistically with tetracycline to inactivate TetK and TetB efflux pumps expressing Staphylococcus spp. [215]. Another study indicated a fourfold reduction in MIC and efflux pump inhibitory activity in norfloxacin-resistant S. aureus when epicatechin gallate was used in combination with norfloxacin against the bacterium [216]. Epigallocatechin gallate isolated from Camellia sinensis potentiated the antibacterial activity of tetracycline against S. aureus by inhibiting the activity of TetK efflux pumps [215, 217]. The alkaloid compound reserpine was able to reduce the MIC of moxifloxacin, ciprofloxacin and sparfloxacin fourfold against S. aureus. Bacteriological studies indicated that reserpine was able to synergistically inhibit the activity of multi-drug efflux pumps expressed by the NorA gene in S. aureus [218]. Moreover, another study showed that reserpine was able to inactivate efflux pumps present in S. aureus, S. pneumonia and B. subtilis when used in combination with norfloxacin, tetracycline and ciprofloxacin [219, 220]. Recent investigations revealed that 5'-methoxyhydnocarpin-D and pheophorbide A (Fig. 2) isolated from Berberis fremontii, tiliroside isolated from Herissantia tiubae and piperine purified from Piper nigrum extracts were able to potentiate the antibacterial action of amikicine, ampicillin, tetracycline, lomefloxacin, norfloxacin and ofloxacin by inhibiting the activity of NorA multi-drug efflux pumps in S. aureus [1, 35, 221]. Phytoextracts obtained from Punica granatum (pomegranate) was able to potentiate the action of gentamicin, chloramphenicol, tetracycline, ampicillin, and oxacillin by inhibiting the activity of NorA efflux pumps in MRSA [31]. Another study indicated that conessine isolated from Holarrhena antidysenterica was able to synergistically potentiate the antibacterial action of novobiocin and rifampicin by inhibiting the activity of multidrug efflux pumps expressed by the AdeIJK gene in XDR A. baumannii [207]. A recent study revealed that curcumin (Fig. 2) purified and isolated from Curcuma longa Linn. was able to potentiate the antibacterial action of polymyxins when used in combination against MDR strains of A. baumannii. Curcumin-polymyxin B combinations indicated remarkably high FICIs of 0.156, 0.375, 0.068 for AB12, AB14, AB16 and NCTC 19606 strains of A. baumannii respectively. Bacteriological studies in the investigation elucidated that curcumin was able to reverse polymyxin resistance in MDR A. baumannii by modulating the activity of EtBr and EmrAB efflux pumps [222,223,224]. Furthermore, purified plant compounds like tannic acid and ellagic acid enhanced the antibacterial action of novobiocin, coumermycin, chlorobiocin, rifampicin and fusidic acid by reducing the MIC of each antibiotic by 2–fourfold. Bacteriological investigations in the study indicated that tannic acid and ellagic acid disrupted the activity of multidrug efflux pumps present in A. baumannii [225]. One study indicated that phytoextracts obtained from plants like Erythrina variegata, Jatropha elliptica, Cytisus striatus and Persea lingue also synergistically potentiated the action of antibiotics by inactivating efflux pumps present in drug-resistant Gram-positive bacteria like MRSA and VRE [1, 226,227,228]. Artemisinins (Fig. 2), an AcrAB‑TolC gene associated bacterial efflux pump disruptor isolated from Artemisia annua was able to potentiate the antibacterial action of penicillin G, cefazolin, ampicillin, cefoperazone and cefuroxime when used in combination against E. coli, which indicated FICIs of < 0.5 for each antibiotic [229]. Isolated compounds like Cathinone and Theobromine (Fig. 2) worked synergistically with ciprofloxacin and tetracycline when used in combination against S. typhimurium and K. pneumoniae. Both phytochemical compounds lowered the MICs of ciprofloxacin and tetracycline by 2–fourfold via the inhibition of efflux pumps expressed by AcrAB‑TolC gene [230]. Phenolic compounds like p-Coumaric acid, caffeic acid, vanillic acid, sinapic acid, gallic acid and taxifolin (Fig. 2) reduced the MICs of ciprofloxacin by 32 fold and erythromycin by 16 fold when used in combination against C. jejuni. These phenolic compounds disrupted the activity of efflux pumps in C. jejuni expressed by CmeABC gene [231]. Purified phytochemical compounds like ferulenol extracted from Ferula communis and plumbagin (Fig. 2) isolated from Plumbago zeylanica synergistically lowered the MIC of isoniazid against Mycobacterium spp. via the inhibition of efflux pumps expressed by NorA gene [232]. Furthermore, a study conducted by Sharma et al. revealed that piperine isolated from Piper nigrum L. (black pepper) reduced the MIC of rifampicin against the tuberculosis causing M. tuberculosis H37Rv strain by 4–eightfold. Bacteriological studies indicated that piperine was able to synergistically inhibit the activity of multi-drug efflux pumps expressed by the Rv1258c gene in M. tuberculosis [233]. Catechin (Fig. 2) compounds, epigallocatechin and epicatechin gallates were the first herbal drugs to receive FDA approval in 2006. The leaf extract of Camellia sinensis consists of about 85% to 95% catechins and the essence of the plant was used in the topical management and treatment of genital warts. [234]. Curcumin isolated from Curcuma longa Linn. is another FDA approved plant based natural product that has proven benefits in clinical trials, and capable of potentiating synthetic antibiotics when used in combination with polymyxin B against MRSA and A. baumannii [222,223,224]. Cranberry juice extract of Vaccinium macrocarpon Aiton which is abundant of proanthocyanidin also received FDA approval for treating uropathogenic E. coli [235, 236]. A derivative phytochemical compound known as salicylate (Fig. 2) mediated synergistic action against the pulmonary pathogen Burkholderia cenocepacia when used in combination with trimethoprim, ciprofloxacin and chloramphenicol. The study indicated that salicylate reduced the MIC of the selected antibiotics by tenfold and the combinations potentially inhibited the activity of efflux-pumps in B. cenocepacia [237].

Plant-derived synergists as biofilm formation/quorum sensing antagonists

Plant-derived synergists are also capable of negating the process of quorum sensing in bacteria. A novel study conducted by Christensen et al. indicated that horseradish juice and curcumin supplemented with furanone C-30 was able to induce significant synergistic quorum quenching activity when combined with tobramycin against P. aeruginosa PAO1 in female BALB/c mice [238]. Bacteriological studies indicated that curcumin and phytochemicals from horseradish juice reduced of the secretion of autoinducer molecules like C-4 and C12- homoserine lactones from the bacterium. A similar study detected In vitro synergistic anti-quorum sensing activity of curcumin with gentamicin and azithromycin combinations against P. aeruginosa. The study also concluded that plant-antibiotic combinations were able to reduce the activity of N-acyl-homoserine lactone autoinducer signaling molecules and down regulate virulent genes like rhlA, lasB and rhl associated with quorum sensing in P. aeruginosa [239, 240]. Furthermore, synergistic anti-biofilm and anti-quorum sensing activities were detected when phytochemical compounds baicalin, hamamelitannin and cinnamaldehyde (Fig. 2) were combined with vancomycin, clindamycin and tobramycin against clinical isolates of B. cenocepacia, S. aureus and E. coli in both in vitro assays and in greater wax moth (Galleria mellonella), Caenorhabditis elegans nematode and female BALB/c mice models used in in vivo assessments [241]. Moreover, naturally synthesized phytochemical compounds like furanones, particularly the halogenated variant known as (5Z)-4-bromo-5-(bromomethylene)-3-butyl-2(5H)-furanone (Fig. 2) was able to attenuate the QS system of B. anthracis [319, 320].

Concluding remarks and future perspectives

The problem of antibiotic resistance is growing rapidly, and the prospects for the application of antibiotic agents in the future have reached uncertainty. Despite the mass production of antibiotics by the pharmaceutical industries in recent decades, bacteria have shown greater resistance to these antibiotics. Plants are remarkable and phenomenal sources of new bioactive compounds with broad-spectrum antibacterial properties. These compounds can assign direct action or interact synergistically with antibiotics to work against bacteria. The following review summarizes the findings of recent investigations based on phytoextracts in combination with existing antibiotics in the context of their drug resistance modulating potential against the anthrax causative organism Bacillus anthracis and MDR and XDR strains of emerging bacterial superbugs. Phytochemical-antibiotic combinations have shown promising results as agents with different mechanisms for modifying and reversing antibiotic resistance. For instance, phytochemicals such as epigallocatechin gallate can interact synergistically with different classes of antibiotics. Depending on the bacterium, this compound can mediate synergism and increase the potency of antibiotics, deactivating β-lactamases and multidrug efflux pumps. Pre-clinical studies have shown that these synergistic compounds can significantly reduce the MIC of bacteria when used in combination with antibiotics. The motivation in antimicrobial synergy research leads to the discovery and production of new antimicrobial agents. However, the underlying action mechanisms of synergistic compounds have not yet been fully explored. A profile that indicates a complete understanding of the pharmacokinetics and pharmacodynamics of the combination agents are required to qualify as a standardized and effective antimicrobial drug. Furthermore, in vivo and nano-medicine drug delivery studies based on combined synergists of plant compound-antibiotics can be deployed for better understand the toxicological responses and bioavailability of the combined agents, to determine their true relevance and safety in the treatment of bacterial infections in humans. Advanced techniques such as isobolograms and phytochemical paradigms can be used to analyze and utilize regions of synergistic interaction between mixtures of antibacterial drugs. At present, the availability of experimental data based on antibiotic-potentiating mechanisms of plant synergists against Bacillus anthracis and antibiotic resistance modulating effects of plant based QS antagonists are limited and therefore, broadening of these studies are imperative. Furthermore, it is necessary to exploit drug resistance modulating potentials of novel combinative products focusing on plant-derived antibodies and antibiotics against bacterial superbugs and B. anthracis. The efficiency of plant-antibiotic synergists and their drug resistance modulating mechanisms are needed to be investigated on recently CDC listed superbugs like Bordetella pertussis and Mycoplasma genitalium and other infectious disease causing pathogens like Rickettsia rickettsii, Neisseria spp., Yersinia pestis and Francisella tularensis.

Availability of data and materials

The following review was based on data extracted from published research article available in all relevant databases with no limitation up to 10th January 2021.

Abbreviations

β:

Beta

MIC:

Minimum inhibitory concentration

MRSA:

Methicillin-resistant Staphylococcus aureus

MSSA:

Methicillin-sensitive Staphylococcus aureus

VRE:

Vancomycin-resistant Enterococcus faecium

PBP:

Penicilin-binding protein

S. aureus :

Staphylococcus aureus

S. pneumonia :

Streptococcus pneumoniae

B. subtilis :

Bacillus subtilis

S. epidermidis :

Staphylococcus epidermidis

FICI:

Fractional Inhibitory Concentration Index

μg/mL:

Microgram per milliliters

μg/μL:

Microgram per microliters

L. monocytogenes :

Listeria monocytogenes

E. faecium :

Enterococcus faecium

O. oeni :

Oenococcusoeni

L.hilgardii:

Lactobacillus hilgardii

S. agalactiae :

Streptococcus agalactiae

S. pyogenes :

Streptococcus pyogenes

B. cereus :

Bacillus cereus

B. anthracis :

Bacillus anthracis

A. baumannii :

Acinetobacter baumannii

P. aeruginosa :

Pseudomonas aeruginosa

E. coli :

Escherichia coli

S. typhimurium :

Salmonella typhimurium

K. pneumoniae :

Klebsiella pneumoniae

C. jejuni :

Campylobacter jejuni

S. pneumonia :

Streptococcus pneumoniae

S. maltophilia :

Stenotrophomonas maltophilia

P. monteilii :

Pseudomonas monteilii

P. chlororaphis :

Pseudomonas chlororaphis

C. difficile :

Clostridium difficile

H. pylori :

Helicobacter pylori

M. chelonei :

Mycobacterium chelonae

M. smegmatis :

Mycobacterium smegmatis

M. intracellulare :

Mycobacterium intracellulare

M. xenopei :

Mycobacterium xenopei

M. tuberculosis :

Mycobacterium tuberculosis

A. xylosoxidans :

Achromobacter xylosoxidans

P. mirabilis :

Proteus mirabilis

P. vulgaris :

Proteus vulgaris

S. sonnei :

Shigella sonnei

E. cloacae :

Enterobacter cloacae

S. faecalis :

Streptococcus faecalis

L. ivanovii :

Listeria ivanovii

V. vulnificus :

Vibrio vulnificus

V. fluvialis :

Vibrio fluvialis

V. cholerae :

Vibrio cholerae

B. cenocepacia :

Burkholderia cenocepacia

B. melitensis :

Brucella melitensis

MFS:

Major facilitator super-family

SMR:

Small multidrug resistance super-family

MATE:

Multidrug and aoxic compound extrusion super-family

RND:

Resistance-nodulation-cell division super-family

ABC:

ATP-binding cassette super-family

mg/mL:

Milligram per milliliter

DNA:

Deoxyribonucleic acid

RNA:

Ribonucleic acid

ATP:

Adenosine triphosphate

sp.:

Species (single)

spp.:

Species (multiple)

ESBL:

Extended-spectrum beta-lactamases

ESKAPE:

Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, And Enterobacter species

NADH:

Nicotinamide adenine dinucleotide

MIC:

Minimum inhibitory concentration

ADP:

Adenosine diphosphate

H + :

Hydrogen

EGCG:

Epigallocatechin gallate

PBP2a:

Penicillin-binding protein 2a

WHO:

World Health Organization

CDC:

Centers for Disease Control and Prevention

FDA:

Food and Drug Administration

ATCC:

American Type Culture Collection

NCTC:

National Collection of Type Cultures

SPI:

Salmonella pathogenicity island

MDR:

Multidrug-resistant

XDR:

Extensively drug-resistant

PDR:

Pandrug-resistant

QS:

Quorum sensing

References

  1. Ayaz M, Ullah F, Sadiq A, Ullah F, Ovais M, Ahmed J, Devkota H. Synergistic interactions of phytochemicals with antimicrobial agents: potential strategy to counteract drug resistance. Chem Biol Interact. 2019;308:294–303.

    CAS  PubMed  Google Scholar 

  2. Davies J, Davies D. Origins and evolution of antibiotic resistance. Microbiol Mol Biol Rev. 2010;74:417–33.

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Sweeney DA, Hicks CW, Cui X, Li Y, Eichacker PQ. Anthrax infection. Am J Respir Crit Care Med. 2011;184:1333–41.

    PubMed  PubMed Central  Google Scholar 

  4. Antimicrobial Mechanisms of Escherichia coli. https://www.intechopen.com/books/-i-escherichia-coli-i-recent-advances-on-physiology-pathogenesis-and-biotechnological-applications/antimicrobial-mechanisms-of-i-escherichia-coli-i-. Accessed 12 July 2017.

  5. Reygaert WC. An overview of the antimicrobial resistance mechanisms of bacteria. AIMS Microbiol. 2018;4(3):482–501.

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Liu Y, Ding S, Shen J, Zhu K. Nonribosomal antibacterial peptides that target multidrug-resistant bacteria. Nat Prod Rep. 2019;36(4):573–92.

    CAS  PubMed  Google Scholar 

  7. Neyfakh A, Bidnenko V, Chen L. Efflux-mediated multidrug resistance in Bacillus subtilis: similarities and dissimilarities with the mammalian system. Proc Natl Acad Sci. 1991;88(11):4781–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Jiang Q, Chen J, Yang C, Yin Y, Yao K. Quorum sensing: a prospective therapeutic target for bacterial diseases. Biomed Res Int. 2019;2019:1–15.

    Google Scholar 

  9. Traditional medicine: growing needs and potential. Geneva: World Health Organization; 2002. WHO policy perspectives on medicines.

  10. Kumar HNK, Chandana E, Preethi SD. In vitro antimicrobial activity and phytochemical screening of Aloe vera Linn. Int J Curr Pharm Res. 2012;4(3):45–7.

    Google Scholar 

  11. Sandigawad AM. In vitro evaluation of antibacterial activity of bark and flower extracts of Pimenta officinalis Lindi. Adv Biores. 2010;1(2):61–8.

    Google Scholar 

  12. Chanda S, Dudhatra S, Kaneria M. Antioxidative and antibacterial effects of seeds and fruit rind of nutraceutical plants belonging to the Fabaceae family. Food Funct. 2010;1(3):308.

    CAS  PubMed  Google Scholar 

  13. Gupta R, Nand P, Drabu S. Insignificant anti-acne activity of Azadirachta indica leaves and bark. J Pharm Negative Results. 2012;3(1):29.

    Google Scholar 

  14. Lindblad W. Review paper: considerations for determining if a natural product is an effective wound-healing agent. Int J Low Extrem Wounds. 2008;7(2):75–81.

    PubMed  Google Scholar 

  15. Guay I, Boulanger S, Isabelle C, Brouillette E, Chagnon F, Bouarab K, Marsault E, Malouin F. Tomatidine and analog FC04–100 possess bactericidal activities against Listeria, Bacillus and Staphylococcus spp. BMC Pharm Toxicol. 2018. https://doi.org/10.1186/s40360-018-0197-2.

    Article  Google Scholar 

  16. Nickell LG. Antimicrobial activity of vascular plants. Econ Bot. 1959;13(4):281–318.

    CAS  Google Scholar 

  17. Bassett E, Keith M, Armelagos G, Martin D, Villanueva A. Tetracycline-labeled human bone from ancient Sudanese Nubia (AD 350). Science. 1980;209(4464):1532–4.

    CAS  PubMed  Google Scholar 

  18. Nelson M, Dinardo A, Hochberg J, Armelagos G. Brief communication: mass spectroscopic characterization of tetracycline in the skeletal remains of an ancient population from Sudanese Nubia 350–550 CE. Am J Phys Anthropol. 2010;143(1):151–4.

    PubMed  Google Scholar 

  19. Cook M, Molto E, Anderson C. Fluorochrome labelling in Roman period skeletons from Dakhleh oasis, Egypt. Am J Phys Anthropol. 1989;80(2):137–43.

    CAS  PubMed  Google Scholar 

  20. Falkinham J, Wall T, Tanner J, Tawaha K, Alali F, Li C, Oberlies N. Proliferation of antibiotic-producing bacteria and concomitant antibiotic production as the basis for the antibiotic activity of Jordan’s Red Soils. Appl Environ Microbiol. 2009;75(9):2735–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Clardy J. Using genomics to deliver natural products from symbiotic bacteria. Genome Biol. 2005;6(9):232.

    PubMed  PubMed Central  Google Scholar 

  22. Aminov R. A brief history of the antibiotic era: lessons learned and challenges for the future. Front Microbiol. 2010. https://doi.org/10.3389/fmicb.2010.00134.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Forrest RD. Early history of wound treatment. J R Soc Med. 1982;75(3):198–205.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Wong R, Hägg U, Samaranayake L, Yuen M, Seneviratne C, Kao R. Antimicrobial activity of Chinese medicine herbs against common bacteria in oral biofilm. A pilot study. Int J Oral Maxillofac Surg. 2010;39(6):599–605.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Ligon BL. Penicillin: its discovery and early development. Semin Pediatr Infect Dis. 2004;15(1):52–7.

    PubMed  Google Scholar 

  26. Monciardini P, Iorio M, Maffioli S, Sosio M, Donadio S. Discovering new bioactive molecules from microbial sources. Microb Biotechnol. 2014;7(3):209–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Jeśman C, Młudzik A, Cybulska M. History of antibiotics and sulphonamides discoveries. Polski Merkuriusz Lekarski Organ Polskiego Towarzystwa Lekarskiego. 2011;30(179):20–322.

    Google Scholar 

  28. Kardos N, Demain A. Penicillin: the medicine with the greatest impact on therapeutic outcomes. Appl Microbiol Biotechnol. 2011;92(4):677–87.

    CAS  PubMed  Google Scholar 

  29. da Cunha R, Fonesca LP, Calado CRC. Antibiotic discovery: where have we come from, where do we go? Antibiotics. 2019;8(2):45.

    Google Scholar 

  30. Min BR, Pinchak WE, Merkel R, Walker S, Tomita G, Anderson RC. Comparative antimicrobial activity of tannin extract from perennial plants on mastitis pathogens. Sci Res Essay. 2008;3(2):66–73.

    Google Scholar 

  31. Braga L, Leite A, Xavier K, Takahashi J, Bemquerer M, Chartone-Souza E, Nascimento A. Synergic interaction between pomegranate extract and antibiotics against Staphylococcus aureus. Can J Microbiol. 2005;51(7):541–7.

    CAS  PubMed  Google Scholar 

  32. Rates SMK. Plants as source of drugs. Toxicon. 2001;39(5):603–13.

    CAS  PubMed  Google Scholar 

  33. Livermore D, Blaser M, Carrs O, Cassell G, Fishman N, Guidos R, Levy S, Powers J, Norrby R, Tillotson G, Davies R, Projan S, Dawson M, Monnet D, Keogh-Brown M, Hand K, Garner S, Findlay D, Morel C, Wise R, Bax R, Burke F, Chopra I, Czaplewski L, Finch R, Livermore D, Piddock L, White T. Discovery research: the scientific challenge of finding new antibiotics. J Antimicrob Chemother. 2011;66(9):1941–4.

    CAS  PubMed  Google Scholar 

  34. Bush K. Antibacterial drug discovery in the 21st century. Clin Microbiol Infect. 2004;10(Suppl 4):10–7.

    PubMed  Google Scholar 

  35. Lewis K. In search of natural substrates and inhibitors of MDR pumps. J Mol Microbiol Biotechnol. 2001;3(2):247–54.

    CAS  PubMed  Google Scholar 

  36. Stermitz FR, Lorenz P, Tawara JN, Zenewicz LA, Lewis K. Synergy in a medicinal plant: antimicrobial action of berberine potentiated by 5’-methoxyhydnocarpin, a multidrug pump inhibitor. Proc Natl Acad Sci. 2000;97(4):1433–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Fleitas Martínez O, Cardoso M, Ribeiro S, Franco O. Recent advances in anti-virulence therapeutic strategies with a focus on dismantling bacterial membrane microdomains, toxin neutralization, quorum-sensing interference and biofilm inhibition. Front Cell Infect Microbiol. 2019;9:74. https://doi.org/10.3389/fcimb.2019.00074.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Rodrigues C, Plotkin S. Impact of vaccines; health, economic and social perspectives. Front Microbiol. 2020;11:1526. https://doi.org/10.3389/fmicb.2020.01526.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Ågren J, Finn M, Bengtsson B, Segerman B. Microevolution during an anthrax outbreak leading to clonal heterogeneity and penicillin resistance. PLoS ONE. 2014;9: e89112.

    PubMed  PubMed Central  Google Scholar 

  40. What is anthrax? | CDC https://www.cdc.gov/anthrax/basics/index.html.

  41. Anthrax—NORD (National Organization for Rare Disorders). https://rarediseases.org/rare-diseases/anthrax/.

  42. Kamal S, Rashid A, Bakar M, Ahad M. Anthrax: an update. Asian Pac J Trop Biomed. 2011;1:496–501.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. The threat of an anthrax attack | CDC. https://www.cdc.gov/anthrax/bioterrorism/threat.html.

  44. Bell DM, Kozarsky PE, Stephens DS. Clinical issues in the prophylaxis, diagnosis, and treatment of anthrax. Emerg Infect Dis. 2002;8(2):222–5. https://doi.org/10.3201/eid0802.010521.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Products approved for anthrax. https://www.fda.gov/drugs/bioterrorism-and-drug-preparedness/products-approved-anthrax.

  46. Beasley DWC, Brasel TL, Comer JE. First vaccine approval under the FDA Animal Rule. NPJ Vaccines. 2016;1:16013. https://doi.org/10.1038/npjvaccines.2016.13.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Lee C, Wright M, Greene A, Aldosary H, Cock I. Preliminary evaluations of the antibacterial activity of Tasmannia lanceolata against Bacillus anthracis: natural resource probing to prevent anthrax. Pharmacogn Commun. 2019;9(4):124–9.

    CAS  Google Scholar 

  48. CDC—Anthrax Sterne, General Information—NCZVED. https://www.cdc.gov/nczved/divisions/dfbmd/diseases/anthrax_sterne/#:~:text=The%20Sterne%20strain%20(34F2)%20of%20Bacillus%20anthracis%20was%20discovered%20in,or%20animals%20have%20been%20reduced.

  49. Kaur R, Tiwari A, Manish M, Maurya I, Bhatnagar R, Singh S. Common garlic (Allium sativum L.) has potent anti-Bacillus anthracis activity. J Ethnopharmacol. 2021;264: 113230.

    CAS  PubMed  Google Scholar 

  50. Cavallo J, Ramisse F, Girardet M, Vaissaire J, Mock M, Hernandez E. Antibiotic susceptibilities of 96 isolates of Bacillus anthracis isolated in France between 1994 and 2000. Antimicrob Agents Chemother. 2002;46(7):2307–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Price L, Vogler A, Pearson T, Busch J, Schupp J, Keim P. In vitro selection and characterization of Bacillus anthracis mutants with high-level resistance to ciprofloxacin. Antimicrob Agents Chemother. 2003;47(7):2362–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Chen Y, Tenover F, Koehler T. β-Lactamase gene expression in a penicillin-resistant Bacillus anthracis strain. Antimicrob Agents Chemother. 2004;48(12):4873–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Kaur S, Hariharan S, Dharmaraj S. Superbugs: the powerful warriors. Int J Pharm Sci Res. 2020;11:1506–26.

    CAS  Google Scholar 

  54. Rajendran R. Superbug infection. J Drug Metab Toxicol. 2018. https://doi.org/10.4172/2157-7609.1000238.

    Article  Google Scholar 

  55. Antimicrobial resistance. https://www.who.int/news-room/fact-sheets/detail/antimicrobial-resistance.

  56. Antibiotic-resistant germs: new threats. https://www.cdc.gov/drugresistance/biggest-threats.html#extend.

  57. Lee A, de Lencastre H, Garau J, Kluytmans J, Malhotra-Kumar S, Peschel A, Harbarth S. Methicillin-resistant Staphylococcus aureus. Nat Rev Dis Primers. 2018;4:18033.

    PubMed  Google Scholar 

  58. Silva-Santana G, Cabral-Oliveira G, Oliveira D, Nogueira B, Pereira-Ribeiro P, Mattos-Guaraldi A. Staphylococcus aureus biofilms: an opportunistic pathogen with multidrug resistance. Rev Med Microbiol. 2020;32:12–21.

    Google Scholar 

  59. Sizar O, Unakal C. Gram positive bacteria. Treasure Island (FL): StatPearls; 2019.

    Google Scholar 

  60. Woodford N, Livermore DM. Infections caused by Gram-positive bacteria: a review of the global challenge. J Infect. 2009;59(Suppl 1):S4–16. https://doi.org/10.1016/S0163-4453(09)60003-7.

    Article  PubMed  Google Scholar 

  61. Boucher H, Corey G. Epidemiology of methicillin-resistant Staphylococcus aureus. Clin Infect Dis. 2008;46:S344–9.

    PubMed  Google Scholar 

  62. Turner N, Sharma-Kuinkel B, Maskarinec S, Eichenberger E, Shah P, Carugati M, Holland T, Fowler V. Methicillin-resistant Staphylococcus aureus: an overview of basic and clinical research. Nat Rev Microbiol. 2019;17:203–18.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Fishovitz J, Hermoso J, Chang M, Mobashery S. Penicillin-binding protein 2a of methicillin-resistant Staphylococcus aureus. IUBMB Life. 2014;66:572–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Tan D, Htun H, Koh J, Kanagasabai K, Lim J, Hon P, Ang B, Chow A. Comparative epidemiology of vancomycin-resistant enterococci colonization in an acute-care hospital and its affiliated intermediate- and long-term care facilities in Singapore. Antimicrob Agents Chemother. 2018. https://doi.org/10.1128/AAC.01507-18.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Ahmed M, Baptiste K. Vancomycin-resistant enterococci: a review of antimicrobial resistance mechanisms and perspectives of human and animal health. Microb Drug Resist. 2018;24(5):590–606.

    CAS  PubMed  Google Scholar 

  66. O’Driscoll T, Crank CW. Vancomycin-resistant enterococcal infections: epidemiology, clinical manifestations, and optimal management. Infect Drug Resist. 2015;8:217–30.

    PubMed  PubMed Central  Google Scholar 

  67. Hashimoto Y, Kurushima J, Nomura T, Tanimoto K, Tamai K, Yanagisawa H, Shirabe K, Ike Y, Tomita H. Dissemination and genetic analysis of the stealthy vanB gene clusters of Enterococcus faecium clinical isolates in Japan. BMC Microbiol. 2018. https://doi.org/10.1186/s12866-018-1342-1.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Miller W, Munita J, Arias C. Mechanisms of antibiotic resistance in enterococci. Expert Rev Anti Infect Therapy. 2014;12:1221–36.

    CAS  Google Scholar 

  69. Biswas P, Dey S, Sen A, Adhikari L. Virulence markers of vancomycin resistant enterococci isolated from infected and colonized patients. J Glob Infect Dis. 2014;6(4):157.

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Gupta A, Ampofo K, Rubenstein D, Saiman L. Extended spectrum β lactamase-producing Klebsiella pneumoniae infections: a review of the literature. J Perinatol. 2003;23(6):439–43.

    PubMed  Google Scholar 

  71. Rammaert B, Goyet S, Beauté J, Hem S, Te V, Try P, Mayaud C, Borand L, Buchy P, Guillard B, Vong S. Klebsiella pneumoniae related community-acquired acute lower respiratory infections in Cambodia: clinical characteristics and treatment. BMC Infect Dis. 2012. https://doi.org/10.1186/1471-2334-12-3.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Bassetti M, Peghin M. How to manage KPC infections. Ther Adv Infect Dis. 2020;7:1–12.

    CAS  Google Scholar 

  73. Müller-Schulte E, Tuo M, Akoua-Koffi C, Schaumburg F, Becker S. High prevalence of ESBL-producing Klebsiella pneumoniae in clinical samples from central Côte d’Ivoire. Int J Infect Dis. 2020;91:207–9.

    PubMed  Google Scholar 

  74. Han JH, Goldstein EJ, Wise J, Bilker WB, Tolomeo P, Lautenbach E. Epidemiology of carbapenem-resistant Klebsiella pneumoniae in a network of long-term acute care hospitals. Clin Infect Dis. 2017;64(7):839–44.

    PubMed  Google Scholar 

  75. Ghafourian S, Sekawi Z, Neela V, Khosravi A, Rahbar M, Sadeghifard N. Incidence of extended-spectrum beta-lactamase-producing Klebsiella pneumoniae in patients with urinary tract infection. Sao Paulo Med J. 2012;130(1):37–43.

    PubMed  Google Scholar 

  76. Navon-Venezia S, Kondratyeva K, Carattoli A. Klebsiella pneumoniae: a major worldwide source and shuttle for antibiotic resistance. FEMS Microbiol Rev. 2017;41(3):252–75.

    CAS  PubMed  Google Scholar 

  77. Arato V, Raso M, Gasperini G, Berlanda Scorza F, Micoli F. Prophylaxis and treatment against Klebsiella pneumoniae: current insights on this emerging anti-microbial resistant global threat. Int J Mol Sci. 2021;22(8):4042.

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Furchtgott L, Wingreen N, Huang K. Mechanisms for maintaining cell shape in rod-shaped Gram-negative bacteria. Mol Microbiol. 2011;81(2):340–53.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Mahmud Z, Kabir M, Ali S, Moniruzzaman M, Imran K, Nafiz T, Islam M, Hussain A, Hakim S, Worth M, Ahmed D, Johnston D, Ahmed N. Extended-spectrum beta-lactamase-producing Escherichia coli in drinking water samples from a forcibly displaced, densely populated community setting in Bangladesh. Front Public Health. 2020. https://doi.org/10.3389/fpubh.2020.00228.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Day M, Hopkins K, Wareham D, Toleman M, Elviss N, Randall L, Teale C, Cleary P, Wiuff C, Doumith M, Ellington M, Woodford N, Livermore D. Extended-spectrum β-lactamase-producing Escherichia coli in human-derived and foodchain-derived samples from England, Wales, and Scotland: an epidemiological surveillance and typing study. Lancet Infect Dis. 2019;19(12):1325–35.

    CAS  PubMed  Google Scholar 

  81. Díaz-Agero Pérez C, López-Fresneña N, Rincon Carlavilla A, Hernandez Garcia M, Ruiz-Garbajosa P, Aranaz-Andrés J, Maechler F, Gastmeier P, Bonten M, Canton R. Local prevalence of extended-spectrum beta-lactamase (ESBL) producing Enterobacteriaceae intestinal carriers at admission and co-expression of ESBL and OXA-48 carbapenemase in Klebsiella pneumoniae: a prevalence survey in a Spanish University Hospital. BMJ Open. 2019;9(3): e024879.

    PubMed  PubMed Central  Google Scholar 

  82. Tulara N. Nitrofurantoin and fosfomycin for extended spectrum beta-lactamases producing Escherichia coli and Klebsiella pneumoniae. J Glob Infect Dis. 2018;10(1):19.

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Fritzenwanker M, Imirzalioglu C, Herold S, Wagenlehner FM, Zimmer KP, Chakraborty T. Treatment options for carbapenem- resistant gram-negative infections. Dtsch Arztebl Int. 2018;115(20–21):345–52.

    PubMed  PubMed Central  Google Scholar 

  84. Overdevest I. Extended-spectrum B-lactamase genes of Escherichia coli in chicken meat and humans, the Netherlands. Emerg Infect Dis. 2011;17:1216–22.

    PubMed  PubMed Central  Google Scholar 

  85. Shin H, Kim Y, Han D, Hur H. Emergence of high level carbapenem and extensively drug resistant Escherichia coli ST746 producing NDM-5 in influent of wastewater treatment plant, Seoul, South Korea. Front Microbiol. 2021. https://doi.org/10.3389/fmicb.2021.645411.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Shabana II. Detection of the enteroaggregative heat-stable enterotoxin 1 nucleotide sequences among diarrhegenic Escherichia coli. Austin J Clin Immunol. 2014;1(2):1009.

    Google Scholar 

  87. Sarowska J, Futoma-Koloch B, Jama-Kmiecik A, Frej-Madrzak M, Ksiazczyk M, Bugla-Ploskonska G, Choroszy-Krol I. Virulence factors, prevalence and potential transmission of extraintestinal pathogenic Escherichia coli isolated from different sources: recent reports. Gut Pathogens. 2019. https://doi.org/10.1186/s13099-019-0290-0.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Diggle SP, Whiteley M. Microbe profile: Pseudomonas aeruginosa: opportunistic pathogen and lab rat. Microbiology (Reading). 2020;166(1):30–3.

    CAS  Google Scholar 

  89. Pseudomonas aeruginosa infection. https://www.cdc.gov/hai/organisms/pseudomonas.html.

  90. Daikos GL, da Cunha CA, Rossolini GM, Stone GG, Baillon-Plot N, Tawadrous M, Irani P. Review of ceftazidime-avibactam for the treatment of infections caused by Pseudomonas aeruginosa. Antibiotics. 2021;10:1126.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Pang Z, Raudonis R, Glick BR, Lin T-J, Cheng Z. Antibiotic resistance in pseudomonas aeruginosa: mechanisms and alternative therapeutic strategies. Biotechnol Adv. 2019;37:177–92.

    CAS  PubMed  Google Scholar 

  92. Alonso B, Fernández-Barat L, Di Domenico EG, Marín M, Cercenado E, Merino I, de Pablos M, Muñoz P, Guembe M. Characterization of the virulence of Pseudomonas aeruginosa strains causing ventilator-associated pneumonia. BMC Infect Dis. 2020;20:909.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Thomas S, Lekshmi N, Joseph I, Ramamurthy T. Changing facades of Vibrio cholerae: an enigma in the epidemiology of cholera. Indian J Med Res. 2018;147(2):133.

    PubMed  PubMed Central  Google Scholar 

  94. D’Mello-Guyett L, Gallandat K, Van den Bergh R, Taylor D, Bulit G, Legros D, Maes P, Checchi F, Cumming O. Prevention and control of cholera with household and community water, sanitation and hygiene (WASH) interventions: a scoping review of current international guidelines. PLoS ONE. 2020;15(1): e0226549.

    PubMed  PubMed Central  Google Scholar 

  95. Marin M, Thompson C, Freitas F, Fonseca E, Aboderin A, Zailani S, Quartey N, Okeke I, Vicente A. Cholera outbreaks in Nigeria are associated with multidrug resistant atypical El Tor and non-O1/non-O139 Vibrio cholerae. PLoS Negl Trop Dis. 2013;7(2): e2049.

    PubMed  PubMed Central  Google Scholar 

  96. Williams PCM, Berkley JA. Guidelines for the management of paediatric cholera infection: a systematic review of the evidence. Paediatr Int Child Health. 2018;38(sup1):S16–31.

    PubMed  PubMed Central  Google Scholar 

  97. Sources of Infection and Risk Factors | Cholera | CDC. https://www.cdc.gov/cholera/infection-sources.html.

  98. Cholera. https://www.who.int/news-room/fact-sheets/detail/cholera.

  99. Hu D, Liu B, Feng L, Ding P, Guo X, Wang M, Cao B, Reeves P, Wang L. Origins of the current seventh cholera pandemic. Proc Natl Acad Sci. 2016;113(48):E7730–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Das B, Verma J, Kumar P, Ghosh A, Ramamurthy T. Antibiotic resistance in Vibrio cholerae: understanding the ecology of resistance genes and mechanisms. Vaccine. 2020;38:A83–92.

    CAS  PubMed  Google Scholar 

  101. Nelson E, Harris J, Glenn Morris J, Calderwood S, Camilli A. Cholera transmission: the host, pathogen and bacteriophage dynamic. Nat Rev Microbiol. 2009;7(10):693–702.

    CAS  PubMed  Google Scholar 

  102. Rineh A, Kelso M, Vatansever F, Tegos G, Hamblin M. Clostridium difficile infection: molecular pathogenesis and novel therapeutics. Expert Rev Anti Infect Therapy. 2014;12:131–50.

    CAS  Google Scholar 

  103. D’Ostroph AR, So TY. Treatment of pediatric Clostridium difficile infection: a review on treatment efficacy and economic value. Infect Drug Resist. 2017;10:365–75.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Asempa TE, Nicolau DP. Clostridium difficile infection in the elderly: an update on management. Clin Interv Aging. 2017;12:1799–809.

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Clostridioides difficile Infection | HAI | CDC. https://www.cdc.gov/hai/organisms/cdiff/cdiff_infect.html.

  106. Colomb-Cotinat M, Assouvie L, Durand J, Daniau C, Leon L, Maugat S, Soing-Altrach S, Gateau C, Couturier J, Arnaud I, Astagneau P, Berger-Carbonne A, Barbut F. Epidemiology of Clostridioides difficile infections, France, 2010 to 2017. Eurosurveillance. 2019;24(35):1800638.

    PubMed Central  Google Scholar 

  107. Isidro J, Menezes J, Serrano M, Borges V, Paixão P, Mimoso M, Martins F, Toscano C, Santos A, Henriques A, Oleastro M. Genomic study of a Clostridium difficile multidrug resistant outbreak-related clone reveals novel determinants of resistance. Front Microbiol. 2018;9:2994.

    PubMed  PubMed Central  Google Scholar 

  108. Awad MM, Johanesen PA, Carter GP, Rose E, Lyras D. Clostridium difficile virulence factors: insights into an anaerobic spore-forming pathogen. Gut Microbes. 2014;5(5):579–93.

    PubMed  PubMed Central  Google Scholar 

  109. Nour-Neamatollahi A, Siadat S, Yari S, Tasbiti A, Ebrahimzadeh N, Vaziri F, Fateh A, Ghazanfari M, Abdolrahimi F, Pourazar S, Bahrmand A. A new diagnostic tool for rapid and accurate detection of Mycobacterium tuberculosis. Saudi J Biol Sci. 2018;25(3):418–25.

    CAS  PubMed  Google Scholar 

  110. Nardell EA. Transmission and institutional infection control of tuberculosis. Cold Spring Harb Perspect Med. 2015;6(2): a018192.

    PubMed  Google Scholar 

  111. Palomino JC, Martin A. Drug resistance mechanisms in Mycobacterium tuberculosis. Antibiotics (Basel). 2014;3(3):317–40.

    Google Scholar 

  112. Wallis R, Cohen T, Menzies N, Churchyard G. Pan-tuberculosis regimens: an argument for. Lancet Respir Med. 2018;6(4):239–40.

    PubMed  Google Scholar 

  113. Treatment of TuberculosisAmerican Thoracic Society, CDC, and Infectious Diseases Society of America. https://www.cdc.gov/mmwr/preview/mmwrhtml/rr5211a1.htm.

  114. FDA approves new drug for treatment-resistant forms of tuberculosis that affects the lungs. https://www.fda.gov/news-events/press-announcements/fda-approves-new-drug-treatment-resistant-forms-tuberculosis-affects-lungs.

  115. O’Connor E, Teh J, Kamat A, Lawrentschuk N. Bacillus Calmette Guérin (BCG) vaccination use in the fight against COVID-19—what’s old is new again? Future Oncol. 2020;16(19):1323–5.

    CAS  PubMed  Google Scholar 

  116. Smith T, Wolff KA, Nguyen L. Molecular biology of drug resistance in Mycobacterium tuberculosis. Curr Top Microbiol Immunol. 2013;374:53–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Cambier C, Takaki K, Larson R, Hernandez R, Tobin D, Urdahl K, Cosma C, Ramakrishnan L. Mycobacteria manipulate macrophage recruitment through coordinated use of membrane lipids. Nature. 2013;505(7482):218–22.

    PubMed  PubMed Central  Google Scholar 

  118. Eng S, Pusparajah P, Ab Mutalib N, Ser H, Chan K, Lee L. Salmonella: a review on pathogenesis, epidemiology and antibiotic resistance. Front Life Sci. 2015;8(3):284–93.

    CAS  Google Scholar 

  119. Lane CR, LeBaigue S, Esan OB, Awofisyo AA, Adams NL, Fisher I, et al. Salmonella enterica Serovar Enteritidis, England and Wales, 1945–2011. Emerg Infect Dis. 2014;20(7):1097–104.

    PubMed  PubMed Central  Google Scholar 

  120. Shah DH, Casavant C, Hawley Q, Addwebi T, Call DR, Guard J. Salmonella Enteritidis strains from poultry exhibit differential responses to acid stress, oxidative stress, and survival in the egg albumen. Foodborne Pathog Dis. 2012;9(3):258–64.

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Salmonella (non-typhoidal). https://www.who.int/news-room/fact-sheets/detail/salmonella-(non-typhoidal).

  122. Advice to Clinicians | Multistate outbreak of Salmonella infections linked to raw chicken products | October 2018 | Salmonella | CDC . https://www.cdc.gov/salmonella/infantis-10-18/advice.html.

  123. Updated recommendations for the use of typhoid vaccine—Advisory Committee on Immunization Practices, United States, 2015. https://www.cdc.gov/mmwr/preview/mmwrhtml/mm6411a4.htm.

  124. Salmonella Homepage | CDC. https://www.cdc.gov/salmonella/index.html.

  125. Zhang L, Fu Y, Xiong Z, Ma Y, Wei Y, Qu X, Zhang H, Zhang J, Liao M. Highly prevalent multidrug-resistant salmonella from chicken and pork meat at retail markets in Guangdong, China. Front Microbiol. 2018. https://doi.org/10.3389/fmicb.2018.02104.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Crump JA, Luby SP, Mintz ED. The global burden of typhoid fever. Bull World Health Organ. 2004;82(5):346–53.

    PubMed  PubMed Central  Google Scholar 

  127. Zhang C, Ren S, Chang M, Chen P, Ding H, Jiang H. Resistance mechanisms and fitness of Salmonella Typhimurium and Salmonella Enteritidis mutants evolved under selection with ciprofloxacin in vitro. Sci Rep. 2017. https://doi.org/10.1038/s41598-017-09151-y.

    Article  PubMed  PubMed Central  Google Scholar 

  128. de Jong H, Parry C, van der Poll T, Wiersinga W. Host–pathogen interaction in invasive salmonellosis. PLoS Pathog. 2012;8(10): e1002933.

    PubMed  PubMed Central  Google Scholar 

  129. Jajere SM. A review of Salmonella enterica with particular focus on the pathogenicity and virulence factors, host specificity and antimicrobial resistance including multidrug resistance. Vet World. 2019;12(4):504–21.

    PubMed  PubMed Central  Google Scholar 

  130. Rodriguez-Rivera L, Bowen B, den Bakker H, Duhamel G, Wiedmann M. Characterization of the cytolethal distending toxin (typhoid toxin) in non-typhoidal Salmonella serovars. Gut Pathogens. 2015. https://doi.org/10.1186/s13099-015-0065-1.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Betts J, Hornsey M, Wareham D, La Ragione R. In vitro and in vivo activity of theaflavin-epicatechin combinations versus multidrug-resistant Acinetobacter baumannii. Infect Dis Therapy. 2017;6(3):435–42.

    Google Scholar 

  132. Magiorakos A, Srinivasan A, Carey R, Carmeli Y, Falagas M, Giske C, Harbarth S, Hindler J, Kahlmeter G, Olsson-Liljequist B, Paterson D, Rice L, Stelling J, Struelens M, Vatopoulos A, Weber J, Monnet D. Multidrug-resistant, extensively drug-resistant and pandrug-resistant bacteria: an international expert proposal for interim standard definitions for acquired resistance. Clin Microbiol Infect. 2012;18(3):268–81.

    CAS  PubMed  Google Scholar 

  133. Sabo VA, Svircev E, Mimica-Dukic N, Orcic D, Narancic J, Knezevic P. Anti-Acinetobacter baumannii activity of Rumex crispus L. and Rumex sanguineus L. extracts. Asian Pac J Trop Biomed. 2020;10(4):172.

    CAS  Google Scholar 

  134. Almasaudi SB. Acinetobacter spp. as nosocomial pathogens: epidemiology and resistance features. Saudi J Biol Sci. 2018;25(3):586–96.

    PubMed  Google Scholar 

  135. Xie R, Zhang XD, Zhao Q, Peng B, Zheng J. Analysis of global prevalence of antibiotic resistance in Acinetobacter baumannii infections disclosed a faster increase in OECD countries. Emerg Microbes Infect. 2018;7(1):31.

    PubMed  PubMed Central  Google Scholar 

  136. O’Shea MK. Acinetobacter in modern warfare. Int J Antimicrob Agents. 2012;39(5):363–75.

    CAS  PubMed  Google Scholar 

  137. Gerner-Smidt P. Ribotyping of the Acinetobacter calcoaceticus-Acinetobacter baumannii complex. J Clin Microbiol. 1992;30(10):2680–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Paduch R, Kandefer-Szerszeń M, Trytek M, Fiedurek J. Terpenes: substances useful in human healthcare. Arch Immunol Ther Exp. 2007;55(5):315–27.

    CAS  Google Scholar 

  139. Rice LB. Federal funding for the study of antimicrobial resistance in nosocomial pathogens: no ESKAPE. J Infect Dis. 2008;197(8):1079–81.

    PubMed  Google Scholar 

  140. Tiwari V, Roy R, Tiwari M. Antimicrobial active herbal compounds against Acinetobacter baumannii and other pathogens. Front Microbiol. 2015. https://doi.org/10.3389/fmicb.2015.00618.

    Article  PubMed  PubMed Central  Google Scholar 

  141. Zarrilli R, Crispino M, Bagattini M, Barretta E, Di Popolo A, Triassi M, Villari P. Molecular epidemiology of sequential outbreaks of Acintobacter baumannii in an intensive care unit shows the emergence of carbapenem resistance. J Clin Microbiol. 2004;4:946–53.

    Google Scholar 

  142. Seward RJ, Lambert T, Towner KJ. Molecular epidemiology of aminoglycoside resistance in Acinetobacter spp. J Med Microbiol. 1998;47:455–62.

    CAS  PubMed  Google Scholar 

  143. Ellis D, Cohen B, Liu J, Larson E. Risk factors for hospital-acquired antimicrobial-resistant infection caused by Acinetobacter baumannii. Antimicrob Resist Infect Control. 2015. https://doi.org/10.1186/s13756-015-0083-2.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Manchanda V, Sinha S, Singh N. Multidrug resistant Acinetobacter. J Glob Infect Dis. 2010;2(3):291.

    PubMed  PubMed Central  Google Scholar 

  145. Xu C, Bilya S, Xu W. adeABC efflux gene in Acinetobacter baumannii. New Microbes New Infect. 2019;30: 100549.

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Morris FC, Dexter C, Kostoulias X, Uddin MI, Peleg AY. The mechanisms of disease caused by Acinetobacter baumannii. Front Microbiol. 2019;10:1601.

    PubMed  PubMed Central  Google Scholar 

  147. Isler B, Doi Y, Bonomo RA, Paterson DL. New treatment options against carbapenem-resistant Acinetobacter baumannii infections. Antimicrob Agents Chemother. 2018, 63(1):e01110–e01118.

  148. Hartmann T. The lost origin of chemical ecology in the late 19th century. Proc Natl Acad Sci. 2008;105(12):4541–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Graziani R, Tosi G, Denti R. In vitro antimicrobial activity of SILVA FEED ENC® on bacterial strains of poultry origin. In: EPC 2006—12th European Poultry Conference. Beekbergen: World's Poultry Science Association; 2006:328.

  150. Radulovic N, Blagojevic P, Stojanovic-Radic Z, Stojanovic N. Antimicrobial plant metabolites: structural diversity and mechanism of action. Curr Med Chem. 2013;20(7):932–52.

    CAS  PubMed  Google Scholar 

  151. Cowan M. Plant products as antimicrobial agents. Clin Microbiol Rev. 1999;12(4):564–82.

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Daglia M. Polyphenols as antimicrobial agents. Curr Opin Biotechnol. 2012;23(2):174–81.

    CAS  PubMed  Google Scholar 

  153. Coppo E, Marchese A. Antibacterial activity of polyphenols. Curr Pharm Biotechnol. 2014;15(4):380–90.

    CAS  PubMed  Google Scholar 

  154. Cushnie TPT, Lamb AJ. Recent advances in understanding the antibacterial properties of flavonoids. Int J Antimicrob Agents. 2011;38:99–107.

    CAS  PubMed  Google Scholar 

  155. Buzzini P, Arapitsas P, Goretti M, Branda E, Turchetti B, Pinelli P, Ieri F, Romani A. Antimicrobial and antiviral activity of hydrolysable tannins. Mini-Rev Med Chem. 2008;8(12):1179–87.

    CAS  PubMed  Google Scholar 

  156. Kovacevic N. Basis of pharmacognosy. 3rd ed. Belgrade: Serbian School Book; 2004.

    Google Scholar 

  157. Mandal S, Mandal M, Pal N. Enhancing chloramphenicol and trimethoprim in vitro activity by Ocimum sanctum Linn. (Lamiaceae) leaf extract against Salmonella enterica serovar Typhi. Asian Pac J Trop Med. 2012;5(3):220–4.

    PubMed  Google Scholar 

  158. Cai Y, Wang R, Pei F, Liang B. Antibacterial activity of allicin alone and in combination with β-lactams against Staphylococcus spp. and Pseudomonas aeruginosa. J Antibiot. 2007;60(5):335–8.

    CAS  Google Scholar 

  159. Cock I, Cheesman M, Ilanko A, Blonk B. Developing new antimicrobial therapies: are synergistic combinations of plant extracts/compounds with conventional antibiotics the solution? Pharmacogn Rev. 2017;11(22):57.

    PubMed  PubMed Central  Google Scholar 

  160. Wagner H, Ulrich-Merzenich G. Synergy research: approaching a new generation of phytopharmaceuticals. Phytomedicine. 2009;16(2–3):97–110.

    CAS  PubMed  Google Scholar 

  161. Moussaoui F, Alaoui T. Evaluation of antibacterial activity and synergistic effect between antibiotic and the essential oils of some medicinal plants. Asian Pac J Trop Biomed. 2016;6(1):32–7.

    CAS  Google Scholar 

  162. Roell K, Reif D, Motsinger-Reif A. An introduction to terminology and methodology of chemical synergy—perspectives from across disciplines. Front Pharmacol. 2017. https://doi.org/10.3389/fphar.2017.00158.

    Article  PubMed  PubMed Central  Google Scholar 

  163. Stefanović O. Synergistic activity of antibiotics and bioactive plant extracts: a study against Gram-Positive and Gram-Negative Bacteria. Bacterial Pathogenesis and Antibacterial Control. https://doi.org/10.5772/intechopen.7202

  164. Osterburg A, Gardner J, Hyon SH, Neely A, Babcock G. Highly antibiotic-resistant Acinetobacter baumannii clinical isolates are killed by the green tea polyphenol (–)-epigallocatechin-3-gallate (EGCG). Clin Microbiol Infect. 2009;15(4):341–6.

    CAS  PubMed  Google Scholar 

  165. Oluwatuyi M, Kaatz G, Gibbons S. Antibacterial and resistance modifying activity of. Phytochemistry. 2004;65(24):3249–54.

    CAS  PubMed  Google Scholar 

  166. Chusri S, Siriyong T, Na-Phatthalung P, Voravuthikunchai S. Synergistic effects of ethnomedicinal plants of Apocynaceae family and antibiotics against clinical isolates of Acinetobacter baumannii. Asian Pac J Trop Med. 2014;7(6):456–61.

    PubMed  Google Scholar 

  167. Rabadia A, Kamat S, Kamat D. Study of synergistic action of cefotaxime and Terminalia chebula on Acinetobacter baumannii using checkerboard assay. Int J Pharm Pharm Sci. 2013;5(3):830–2.

    Google Scholar 

  168. Phatthalung P, Chusri S, Voravuthikunchai S. Thai ethnomedicinal plants as resistant modifying agents for combating Acinetobacter baumannii infections. BMC Complement Altern Med. 2012;12:56.

    PubMed  PubMed Central  Google Scholar 

  169. Knezevic P, Aleksic V, Simin N, Svircev E, Petrovic A, Mimica-Dukic N. Antimicrobial activity of Eucalyptus camaldulensis essential oils and their interactions with conventional antimicrobial agents against multi-drug resistant Acinetobacter baumannii. J Ethnopharmacol. 2016;178:125–36.

    CAS  PubMed  Google Scholar 

  170. Betts JW, Wareham DW. In vitro activity of curcumin in combination with epigallocatechin gallate (EGCG) versus multidrug-resistant Acinetobacter baumannii. BMC Microbiol. 2014;14(1):172.

    PubMed  PubMed Central  Google Scholar 

  171. Ebrahimi A, Eshraghi A, Mahzoonieh M, Lotfalian S. Antibacterial and antibiotic-potentiation activities of Levisticum officinale L. extracts on pathogenic bacteria. Int J Infect. 2016;4(2): e38768.

    Google Scholar 

  172. Abdel MES, Ahmed AE, Vajid NV, Mohanad A, Nasir AI. Evaluation of antimicrobial activity of Senna (Senna italica Mill) and its synergistic effect with antibiotic drugs. Eur J Biotechnol Biosci. 2017;14(3):69–73.

    Google Scholar 

  173. Abdel MES, Sherif MS, Ahmed AE, Mohanad A, Vajid NV. Evaluation of antimicrobial activity of safflower (Carthamus tinctorius) and its synergistic effect with antibiotic. EC Microbiol. 2018;14(3):160–6.

    Google Scholar 

  174. Abdel MES, Sherif MS, Ahmed AE, Vajid NV, Mohanad A, Nasir AI. Evaluation of antimicrobial and synergistic effects of selected medicinal plants of Hail area with antibiotics. Biosci Biotech Res Comm. 2017;10(1):44–50.

    Google Scholar 

  175. Liu C, Huang H, Zhou Q, Liu B, Wang Y, Li P, Liao K, Su W. Antibacterial and antibiotic synergistic activities of the extract from Pithecellobium clypearia against clinically important multidrug-resistant gram-negative bacteria. Eur J Integr Med. 2019;32: 100999.

    Google Scholar 

  176. Mahmoudi H, Zare Fahim N, Alikhani M, Shokoohizadeh L. Investigation of antimicrobial effect of berberine on ciprofloxacin and imipenem resistance Acinetobacter baumannii isolated from Hamadan Hospitals. Iran J Med Microbiol. 2020;14(1):44–54.

    Google Scholar 

  177. Ekambaram S, Perumal S, Balakrishnan A, Marappan N, Gajendran S, Viswanathan V. Antibacterial synergy between Rosmarinic acid and antibiotics against Methicillin resistant Staphylococcus aureus. J Intercult Ethnopharmacol. 2016;5(4):358.

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Shin B, Park W. Synergistic effect of oleanolic acid on aminoglycoside antibiotics against Acinetobacter baumannii. PLoS ONE. 2015;10(9): e0137751.

    PubMed  PubMed Central  Google Scholar 

  179. Boonyanugomol W, Kraisriwattana K, Rukseree K, Boonsam K, Narachai P. In vitro synergistic antibacterial activity of the essential oil from Zingiber cassumunar Roxb against extensively drug-resistant Acinetobacter baumannii strains. J Infect Public Health. 2017;10(5):586–92.

    PubMed  Google Scholar 

  180. Haq A, Siddiqi M, Batool S, Islam A, Khan A, Khan D, Khan S, Khan H, Shah A, Hasan F, Ahmed S, Badshah M. Comprehensive investigation on the synergistic antibacterial activities of Jatropha curcas pressed cake and seed oil in combination with antibiotics. AMB Express. 2019. https://doi.org/10.1186/s13568-019-0793-6.

    Article  PubMed  PubMed Central  Google Scholar 

  181. Betts JW, Murphy C, Kelly SM, Haswell SJ. Synergistic antibacterial effects of theaflavin in combination with ampicillin against hospital isolates of Stenotrophomonas Maltophilia. J Microbiol Biotechnol Food Sci. 2012;2(3):1068–78.

    Google Scholar 

  182. Darabpour E, Poshtkouhian Bavi A, Motamedi H, Seyyed Nejad SM. Antibacterial activity of different parts of Peganum harmala L. growing in Iran against multi-drug resistant bacteria. EXCLI J. 2011;10:252–63.

    PubMed  PubMed Central  Google Scholar 

  183. Kouitcheu LB, Emmanuel T, Kouam J, Zra T, Louis O. Synergetic effects of plant extracts and antibiotics on Vibrio cholerae O1 strains isolated from clinical specimens. Int J Biol. 2013. https://doi.org/10.5539/ijb.v5n3p64.

    Article  Google Scholar 

  184. Motamedi H, Darabpour E, Gholipour M, Seyyed Nejad SM. In vitro assay for the anti-Brucella activity of medicinal plants against tetracycline-resistant Brucella melitensis. J Zhejiang Univ Sci B. 2010;11(7):506–11. https://doi.org/10.1631/jzus.B0900365.

    Article  PubMed  PubMed Central  Google Scholar 

  185. Tsevelmaa N, Narangerel B, Odgerel O, Dariimaa D, Batkhuu J. Anti-Brucella activity of Caryopteris mongolica Bunge root extract against Brucella melitensis infection in mice. BMC Complementary Altern Med. 2018. https://doi.org/10.1186/s12906-018-2220-y.

    Article  Google Scholar 

  186. El- Tawab AAA, Ammar AA, Hamouda AM, Elhawary SAM. Interaction of some plant extracts with some antibiotics against Listeria monocytogenes from rabbits. Benha Vet Med J. 2018;35(2):120–33.

    Google Scholar 

  187. Munita JM, Arias CA. Mechanisms of antibiotic resistance. Microbiol Spectr. 2016. https://doi.org/10.1128/microbiolspec.VMBF-0016-2015.

    Article  PubMed  Google Scholar 

  188. Leonard F, Markey B. Meticillin-resistant Staphylococcus aureus in animals: a review. Vet J. 2008;175(1):27–36.

    CAS  PubMed  Google Scholar 

  189. Shiota S. Restoration of effectiveness of β-lactams on methicillin-resistant Staphylococcus aureus by tellimagrandin I from rose red. FEMS Microbiol Lett. 2000;185(2):135–8.

    CAS  PubMed  Google Scholar 

  190. Schroeder FC, del Campo ML, Grant JB, Weibel DB, Smedley SR, Bolton KL, Meinwald J, Eisner T. Pinoresinol: a lignol of plant origin serving for defense in a caterpillar. Proc Natl Acad Sci. 2006;103(42):15497–501.

    CAS  PubMed  PubMed Central  Google Scholar 

  191. Reuk-ngam N, Chimnoi N, Khunnawutmanotham N, Techasakul S. Antimicrobial activity of coronarin D and its synergistic potential with antibiotics. BioMed Res Int. 2014. https://doi.org/10.1155/2014/581985.

    Article  PubMed  PubMed Central  Google Scholar 

  192. Bigelow R, Cardelli J. The green tea catechins, (−)-Epigallocatechin-3-gallate (EGCG) and (−)-Epicatechin-3-gallate (ECG), inhibit HGF/Met signaling in immortalized and tumorigenic breast epithelial cells. Oncogene. 2006;25(13):1922–30.

    CAS  PubMed  Google Scholar 

  193. Shimizu M, Shiota S, Mizushima T, Ito H, Hatano T, Yoshida T, Tsuchiya T. Marked potentiation of activity of β-Lactams against methicillin-resistant Staphylococcus aureus by Corilagin. Antimicrob Agents Chemother. 2001;45(11):3198–201.

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Shiota S, Shimizu M, Mizusima T, Ito H, Hatano T, Yoshida T, Tsuchiya T. Restoration of effectiveness of Î2-lactams on methicillin-resistant Staphylococcus aureus by tellimagrandin I from rose red. FEMS Microbiol Lett. 2000;185(2):135–8.

    CAS  PubMed  Google Scholar 

  195. Kuok C, Hoi S, Hoi C, Chan C, Fong I, Ngok C, Meng L, Fong P. Synergistic antibacterial effects of herbal extracts and antibiotics on methicillin-resistant Staphylococcus aureus: a computational and experimental study. Exp Biol Med. 2017;242(7):731–43.

    CAS  Google Scholar 

  196. Vattem D, Lin Y, Ghaedian R, Shetty K. Cranberry synergies for dietary management of Helicobacter pylori infections. Process Biochem. 2005;40(5):1583–92.

    CAS  Google Scholar 

  197. Hemaiswarya S, Doble M. Synergistic interaction of phenylpropanoids with antibiotics against bacteria. J Med Microbiol. 2010;59(12):1469–76.

    CAS  PubMed  Google Scholar 

  198. Kamijo M, Kanazawa T, Funaki M, Nishizawa M, Yamagishi T. Effects of Rosa rugosa petals on intestinal bacteria. Biosci Biotechnol Biochem. 2008;72:773–7.

    CAS  PubMed  Google Scholar 

  199. Ingram J, Hassan H. The resistance of Pseudomonas aeruginosa to chloramphenicol. Can J Microbiol. 1975;21(8):1185–91.

    CAS  PubMed  Google Scholar 

  200. Livermore D. beta-Lactamases in laboratory and clinical resistance. Clin Microbiol Rev. 1995;8(4):557–84.

    CAS  PubMed  PubMed Central  Google Scholar 

  201. Liu I, Durham D, Richards R. Baicalin synergy with β-Lactam antibiotics against methicillin-resistant Staphylococcus aureus and other β-Lactam-resistant strains of S. aureus. J Pharm Pharmacol. 2000;52(3):361–6.

    CAS  PubMed  Google Scholar 

  202. Hu ZQ, Zhao WH, Hara Y, Shimamura T. Epigallocatechin gallate synergy with ampicillin/sulbactam against 28 clinical isolates of methicillin-resistant Staphylococcus aureus. J Antimicrob Chemother. 2001;48(3):361–4.

    CAS  PubMed  Google Scholar 

  203. Obiang-Obounou B, Jang Y. Enriching modern pharmacotherapy through synergy assessment for the combination of natural products and synthetic drugs. Arch Pharmacal Res. 2011;34(10):1579–81.

    CAS  Google Scholar 

  204. Worthington R, Melander C. Combination approaches to combat multidrug-resistant bacteria. Trends Biotechnol. 2013;31(3):177–84.

    CAS  PubMed  PubMed Central  Google Scholar 

  205. Putman M, van Veen H, Konings W. Molecular properties of bacterial multidrug transporters. Microbiol Mol Biol Rev. 2000;64(4):672–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  206. Langton K, Henderson P, Herbert R. Antibiotic resistance: multidrug efflux proteins, a common transport mechanism? Nat Prod Rep. 2004;22(4):439.

    Google Scholar 

  207. Pathania R, Sharma A, Gupta V. Efflux pump inhibitors for bacterial pathogens: from bench to bedside. Indian J Med Res. 2019;149(2):129.

    PubMed  PubMed Central  Google Scholar 

  208. Kumar S, He G, Kakarla P, Shrestha U, Ranjana K, Ranaweera I, Mark Willmon T, Barr S, Hernandez A, Varela M. Bacterial multidrug efflux pumps of the major facilitator superfamily as targets for modulation. Infect Disord Drug Targets. 2016;16(1):28–43.

    CAS  PubMed  Google Scholar 

  209. Prasad N, Muthusamy G, Shanmugam M, Ambudkar S. South Asian medicinal compounds as modulators of resistance to chemotherapy and radiotherapy. Cancers. 2016;8(3):32.

    CAS  PubMed Central  Google Scholar 

  210. Wang S, Duan M, Guan K, Zhou X, Zheng M, Shi X, Ye M, Guan W, Kuver A, Huang M, Liu Y, Dai K, Li X. Developmental neurotoxicity of reserpine exposure in zebrafish larvae (Danio rerio). Comp Biochem Physiol C Toxicol Pharmacol. 2019;223:115–23.

    CAS  PubMed  Google Scholar 

  211. Wu Y. Naturally derived anti-hepatitis B virus agents and their mechanism of action. World J Gastroenterol. 2016;22(1):188.

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Ponnusamy K, Ramasamy M, Savarimuthu I, Paulraj M. Indirubin potentiates ciprofloxacin activity in the NorA efflux pump of Staphylococcus aureus. Scand J Infect Dis. 2010;42(6–7):500–5.

    CAS  PubMed  Google Scholar 

  213. Kalia N, Mahajan P, Mehra R, Nargotra A, Sharma J, Koul S, Khan I. Capsaicin, a novel inhibitor of the NorA efflux pump, reduces the intracellular invasion of Staphylococcus aureus. J Antimicrob Chemother. 2012;67(10):2401–8.

    CAS  PubMed  Google Scholar 

  214. Holler J, Christensen S, Slotved H, Rasmussen H, Guzman A, Olsen C, Petersen B, Molgaard P. Novel inhibitory activity of the Staphylococcus aureus NorA efflux pump by a kaempferol rhamnoside isolated from Persea lingue Nees. J Antimicrob Chemother. 2012;67(5):1138–44.

    CAS  PubMed  Google Scholar 

  215. Sudano Roccaro A, Blanco AR, Giuliano F, Rusciano D, Enea V. Epigallocatechin-gallate enhances the activity of tetracycline in staphylococci by inhibiting its efflux from bacterial cells. Antimicrob Agents Chemother. 2004;48(6):1968–73.

    PubMed  Google Scholar 

  216. Gibbons S, Moser E, Kaatz G. Catechin gallates inhibit multidrug resistance (MDR) in Staphylococcus aureus. Planta Med. 2004;70(12):1240–2.

    CAS  PubMed  Google Scholar 

  217. Hu Z, Zhao W, Asano N, Yoda Y, Hara Y, Shimamura T. Epigallocatechin gallate synergistically enhances the activity of carbapenems against methicillin-resistant Staphylococcus aureus. Antimicrob Agents Chemother. 2002;46(2):558–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  218. Schmitz F, Fluit AC, Luckefahr M, Engler B, Hofmann B, Verhoef J, Heinz H, Hadding U, Jones M. The effect of reserpine, an inhibitor of multidrug efflux pumps, on the in-vitro activities of ciprofloxacin, sparfloxacin and moxifloxacin against clinical isolates of Staphylococcus aureus. J Antimicrob Chemother. 1998;42(6):807–10.

    CAS  PubMed  Google Scholar 

  219. Smith E, Williamson E, Zloh M, Gibbons S. Isopimaric acid from Pinus nigra shows activity against multidrug-resistant and EMRSA strains of Staphylococcus aureus. Phytother Res. 2005;19(6):538–42.

    CAS  PubMed  Google Scholar 

  220. Stavri M, Piddock L, Gibbons S. Bacterial efflux pump inhibitors from natural sources. J Antimicrob Chemother. 2006;59(6):1247–60.

    PubMed  Google Scholar 

  221. Falcão-Silva V, Silva D, Souza M, Siqueira-Junior J. Modulation of drug resistance in Staphylococcus aureus by a kaempferol glycoside from Herissantia tiubae (malvaceae). Phytother Res. 2009;23(10):1367–70.

    PubMed  Google Scholar 

  222. Betts J, Sharili A, La Ragione R, Wareham D. In vitro antibacterial activity of curcumin-polymyxin B combinations against multidrug-resistant bacteria associated with traumatic wound infections. J Nat Prod. 2016;79(6):1702–6.

    CAS  PubMed  Google Scholar 

  223. Dai C, Wang Y, Sharma G, Shen J, Velkov T, Xiao X. Polymyxins-curcumin combination antimicrobial therapy: safety implications and efficacy for infection treatment. Antioxidants. 2020;9(6):506.

    CAS  PubMed Central  Google Scholar 

  224. Kaur A, Sharma P, Capalash N. Curcumin alleviates persistence of Acinetobacter baumannii against colistin. Sci Rep. 2018. https://doi.org/10.1038/s41598-018-29291-z.

    Article  PubMed  PubMed Central  Google Scholar 

  225. Chusri S, Villanueva I, Voravuthikunchai SP, Davies J. Enhancing antibiotic activity: a strategy to control Acinetobacter infections. J Antimicrob Chemother. 2009;64(6):1203–11.

    CAS  PubMed  Google Scholar 

  226. Sato M, Tanaka H, Oh-Uchi T, Fukai T, Etoh H, Yamaguchi R. Antibacterial activity of phytochemicals isolated from Erythrina zeyheri against vancomycin-resistant enterococci and their combinations with vancomycin. Phytother Res. 2004;18(11):906–10.

    CAS  PubMed  Google Scholar 

  227. Marquez B, Neuville L, Moreau N, Genet J, dos Santos A, Caño de Andrade M, Goulart Sant’Ana A. Multidrug resistance reversal agent from Jatropha elliptica. Phytochemistry. 2005;66(15):1804–11.

    CAS  PubMed  Google Scholar 

  228. Abreu A, Coqueiro A, Sultan A, Lemmens N, Kim H, Verpoorte R, van Wamel W, Simões M, Choi Y. Looking to nature for a new concept in antimicrobial treatments: isoflavonoids from Cytisus striatus as antibiotic adjuvants against MRSA. Sci Rep. 2017;7(1):3777.

    PubMed  PubMed Central  Google Scholar 

  229. Li B, Yao Q, Pan X, Wang N, Zhang R, Li J, Ding G, Liu X, Wu C, Ran D, Zheng J, Zhou H. Artesunate enhances the antibacterial effect of β-lactam antibiotics against Escherichia coli by increasing antibiotic accumulation via inhibition of the multidrug efflux pump system AcrAB-TolC. J Antimicrob Chemother. 2011;66(4):769–77.

    CAS  PubMed  Google Scholar 

  230. Piddock L, Garvey M, Rahman M, Gibbons S. Natural and synthetic compounds such as trimethoprim behave as inhibitors of efflux in Gram-negative bacteria. J Antimicrob Chemother. 2010;65:1215–23.

    CAS  PubMed  Google Scholar 

  231. Oh E, Jeon B. Synergistic anti-Campylobacter jejuni activity of fluoroquinolone and macrolide antibiotics with phenolic compounds. Front Microbiol. 2015. https://doi.org/10.3389/fmicb.2015.01129.

    Article  PubMed  PubMed Central  Google Scholar 

  232. Mossa J, El-Feraly F, Muhammad I. Antimycobacterial constituents from Juniperus procera, Ferula communis and Plumbago zeylanica and theirin vitro synergistic activity with isonicotinic acid hydrazide. Phytother Res. 2004;18(11):934–7.

    CAS  PubMed  Google Scholar 

  233. Sharma S, Kumar M, Sharma S, Nargotra A, Koul S, Khan I. Piperine as an inhibitor of Rv1258c, a putative multidrug efflux pump of Mycobacterium tuberculosis. J Antimicrob Chemother. 2010;65(8):1694–701.

    CAS  PubMed  Google Scholar 

  234. Balaji G. Sinecatechins: a better prospect for treating anogenital warts. Indian J Sex Transm Dis AIDS. 2014;35(1):75–6.

    PubMed  PubMed Central  Google Scholar 

  235. Das S. Natural therapeutics for urinary tract infections—a review. Future J Pharm Sci. 2020. https://doi.org/10.1186/s43094-020-00086-2.

    Article  Google Scholar 

  236. FDA announces health claim for certain cranberry products and UTI. https://www.fda.gov/food/cfsan-constituent-updates/fda-announces-qualified-health-claim-certain-cranberry-products-and-urinary-tract-infections.

  237. Nair BM, Cheung KJ Jr, Griffith A, Burns JL. Salicylate induces an antibiotic efflux pump in Burkholderia cepacia complex genomovar III (B. cenocepacia). J Clin Invest. 2004;113(3):464–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  238. Christensen L, van Gennip M, Jakobsen T, Alhede M, Hougen H, Hoiby N, Bjarnsholt T, Givskov M. Synergistic antibacterial efficacy of early combination treatment with tobramycin and quorum-sensing inhibitors against Pseudomonas aeruginosa in an intraperitoneal foreign-body infection mouse model. J Antimicrob Chemother. 2012;67(5):1198–206.

    CAS  PubMed  Google Scholar 

  239. Bahari S, Zeighami H, Mirshahabi H, Roudashti S, Haghi F. Inhibition of Pseudomonas aeruginosa quorum sensing by subinhibitory concentrations of curcumin with gentamicin and azithromycin. J Glob Antimicrob Resist. 2017;10:21–8.

    PubMed  Google Scholar 

  240. Furiga A, Lajoie B, El Hage S, Baziard G, Roques C. Impairment of Pseudomonas aeruginosa biofilm resistance to antibiotics by combining the drugs with a new quorum-sensing inhibitor. Antimicrob Agents Chemother. 2015;60(3):1676–86.

    PubMed  Google Scholar 

  241. Brackman G, Cos P, Maes L, Nelis H, Coenye T. Quorum sensing inhibitors increase the susceptibility of bacterial biofilms to antibiotics in vitro and in vivo. Antimicrob Agents Chemother. 2011;55(6):2655–61.

    CAS  PubMed  PubMed Central  Google Scholar 

  242. Rafiq Z, Narasimhan S, Haridoss M, Vennila R, Vaidyanathan R. Punica granatum rind extract: antibiotic potentiator and efflux pump inhibitor of multidrug resistant Klebsiella pneumoniae clinical isolates. Asian J Pharm Clin Res. 2017;10(3):191.

    Google Scholar 

  243. Mahomoodally M, Dilmohamed S. Antibacterial and antibiotic potentiating activity of Vangueria madagascariensis leaves and ripe fruit pericarp against human pathogenic clinical bacterial isolates. J Tradit Complement Med. 2016;6(4):399–403.

    PubMed  Google Scholar 

  244. Wu S, Yang Z, Liu F, Peng W, Qu S, Li Q, Song X, Zhu K, Shen J. Antibacterial effect and mode of action of flavonoids from licorice against methicillin-resistant Staphylococcus aureus. Front Microbiol. 2019. https://doi.org/10.3389/fmicb.2019.02489.

    Article  PubMed  PubMed Central  Google Scholar 

  245. Ali Raza Naqvi S, Nadeem S, Komal S, Ali Asad Naqvi S, Samee Mubarik M, Yaqub Qureshi S, Ahmad S, Abbas A, Zahid M, Khan N, Shujat Raza S, Aslam N. Antioxidants: natural antibiotics. Antioxidants. 2019. https://doi.org/10.5772/intechopen.84864.

    Article  Google Scholar 

  246. Miyasaki Y, Rabenstein J, Rhea J, Crouch M, Mocek U, Kittell P, Morgan M, Nichols W, Van Benschoten M, Hardy W, Liu G. Isolation and characterization of antimicrobial compounds in plant extracts against multidrug-resistant Acinetobacter baumannii. PLoS ONE. 2013;8: e61594.

    CAS  PubMed  PubMed Central  Google Scholar 

  247. Chan BC, Lau CB, Jolivalt C, Lui S, Ganem-Elbaz C, Paris J, Litaudon M, Fung K, Leung P, Ip M. Chinese medicinal herbs against antibiotic-resistant bacterial pathogens. Sci Microbial Pathogens Commun Curr Res Technol Adv. 2011;2:773–81.

    Google Scholar 

  248. Betts JW, Kelly SM, Haswell SJ. Antibacterial effects of theaflavin and synergy with epicatechin against clinical isolates of Acinetobacter baumannii and Stenotrophomonas maltophilia. Int J Antimicrob Agents. 2011;38(5):421–5.

    CAS  PubMed  Google Scholar 

  249. Lamontagne BM, Isabelle C, Guay I, Brouillette E, Langlois JP, Jacques PÉ, Rodrigue S, Brzezinski R, Beauregard PB, Bouarab K, Boyapelly K, Boudreault PL, Marsault É, Malouin F. Tomatidine is a lead antibiotic molecule that targets Staphylococcus aureus ATP synthase subunit C. Antimicrob Agents Chemother. 2018;62(6):e02197-e2217. https://doi.org/10.1128/AAC.02197-17.

    Article  Google Scholar 

  250. Reiter J, Levina N, van der Linden M, Gruhlke M, Martin C, Slusarenko A. Diallylthiosulfinate (Allicin), a volatile antimicrobial from garlic (Allium sativum), kills human lung pathogenic bacteria, including MDR strains, as a vapor. Molecules. 2017;22(10):1711.

    PubMed Central  Google Scholar 

  251. Park H, ChoI K, Shin I. Antimicrobial activity of isothiocyanates (ITCs) extracted from horseradish (Armoracia rusticana) root against oral microorganisms. Biocontrol Sci. 2013;18(3):163–8.

    CAS  PubMed  Google Scholar 

  252. Campos F, Couto J, Figueiredo A, Tóth I, Rangel A, Hogg T. Cell membrane damage induced by phenolic acids on wine lactic acid bacteria. Int J Food Microbiol. 2009;135(2):144–51.

    CAS  PubMed  Google Scholar 

  253. Lawrence R, Tripathi P, Jeyakumar E. Isolation, purification and evaluation of antibacterial agents from aloe vera. Braz J Microbiol. 2009;40(4):906–15.

    CAS  PubMed  PubMed Central  Google Scholar 

  254. Wu Y, Bai J, Zhong K, Huang Y, Qi H, Jiang Y, Gao H. Antibacterial activity and membrane-disruptive mechanism of 3-p-trans-coumaroyl-2-hydroxyquinic acid, a novel phenolic compound from pine needles of Cedrus deodara, against Staphylococcus aureus. Molecules. 2016;21(8):1084.

    PubMed Central  Google Scholar 

  255. Pelletier RP. Effect of plant-derived molecules on Acinetobacter baumannii biofilm on abiotic surfaces. Honors Scholar Theses. University of Connecticut, Storrs; 2012.

  256. Zhao W, Hu Z, Hara Y, Shimamura T. Inhibition of penicillinase by epigallocatechin gallate resulting in restoration of antibacterial activity of penicillin against penicillinase-producing Staphylococcus aureus. Antimicrob Agents Chemother. 2002;46(7):2266–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  257. Tintino S, Morais-Tintino C, Campina F, Costa M, Menezes I, de Matos Y, Calixto-Júnior J, Pereira P, Siqueira-Junior J, Leal-Balbino T, Coutinho H, Balbino V. Tannic acid affects the phenotype of Staphylococcus aureus resistant to tetracycline and erythromycin by inhibition of efflux pumps. Bioorg Chem. 2017;74:197–200.

    CAS  PubMed  Google Scholar 

  258. Togashi N, Hamashima H, Shiraishi A, Inoue Y, Takano A. Antibacterial activities against Staphylococcus aureus of terpene alcohols with aliphatic carbon chains. J Essent Oil Res. 2010;22(3):263–9.

    CAS  Google Scholar 

  259. Jacobo-Salcedo Mdel R, Gonzalez-Espindola LA, Alonso-Castro AJ, Gonzalez-Martinez Mdel R, Domínguez F, Garcia-Carranca A. Antimicrobial activity and cytotoxic effects of Magnolia dealbata and its active compounds. Nat Prod Commun. 2011;6(8):1121–4.

    PubMed  Google Scholar 

  260. Zhang L, Xu S, Liang W, Mei J, Di Y, Lan H, Yang Y, Wang W, Luo Y, Wang H. Antibacterial activity and mode of action of Mentha arvensis ethanol extract against multidrug-resistant Acinetobacter baumannii. Trop J Pharm Res. 2015;14(11):2099–106.

    Google Scholar 

  261. Mun S, Joung D, Kim S, Park S, Seo Y, Gong R, Choi J, Shin D, Rho J, Kang O, Kwon D. The mechanism of antimicrobial activity of Sophoraflavanone B against methicillin-resistant Staphylococcus aureus. Foodborne Pathog Dis. 2014;11(3):234–9.

    CAS  PubMed  Google Scholar 

  262. Jeong K, Lee J, Kang D, Lee J, Shin S, Kim Y. Screening of flavonoids as candidate antibiotics against Enterococcus faecalis. J Nat Prod. 2009;72(4):719–24.

    CAS  PubMed  Google Scholar 

  263. Park B, Kim J, Kim M, Lee S, Takeoka G, Oh K, Kim J. Curcuma longa L. constituents inhibit sortase A and Staphylococcus aureus cell adhesion to fibronectin. J Agric Food Chem. 2005;53(23):9005–9.

    CAS  PubMed  Google Scholar 

  264. Mun S, Joung D, Kim Y, Kang O, Kim S, Seo Y, Kim Y, Lee D, Shin D, Kweon K, Kwon D. Synergistic antibacterial effect of curcumin against methicillin-resistant Staphylococcus aureus. Phytomedicine. 2013;20(8–9):714–8.

    CAS  PubMed  Google Scholar 

  265. Basile A, Sorbo S, Spadaro V, Bruno M, Maggio A, Faraone N, Rosselli S. Antimicrobial and antioxidant activities of coumarins from the roots of Ferulago campestris (Apiaceae). Molecules. 2019;14(3):939–52.

    Google Scholar 

  266. Tan N, Bilgin M, Tan E, Miski M. Antibacterial activities of pyrenylated coumarins from the roots of Prangos hulusii. Molecules. 2017;22(7):1098.

    PubMed Central  Google Scholar 

  267. El-Seedi H. Antimicrobial Arylcoumarins from Asphodelus microcarpus. J Nat Prod. 2007;70(1):118–20.

    CAS  PubMed  Google Scholar 

  268. Söderberg T, Johansson A. Gref R: toxic effects of some conifer resin acids and tea tree oil on human epithelial and fibroblast cells. Toxicology. 1996;107(2):99–109.

    PubMed  Google Scholar 

  269. Decarvalho C, Dafonseca M. Carvone: why and how should one bother to produce this terpene. Food Chem. 2006;95(3):413–22.

    CAS  Google Scholar 

  270. Althunibat O, Qaralleh H, Abboud M, Khleifat K, Majali IS, Susanti D, Dalaeen S, Al-Dalin A, Rayyan WA. Effect of thymol and carvacrol, the major components of Thymus capitatus on the growth of Pseudomonas aeruginosa. J Pure Appl Microbiol. 2016;10(1):367–74.

    CAS  Google Scholar 

  271. Yadav M, Chae S, Im G, Chung J, Song J. Eugenol: a phyto-compound effective against methicillin-resistant and methicillin-sensitive Staphylococcus aureus clinical strain biofilms. PLoS ONE. 2015;10(3): e0119564.

    PubMed  PubMed Central  Google Scholar 

  272. Girard M, Bee G. Invited review: tannins as a potential alternative to antibiotics to prevent coliform diarrhea in weaned pigs. Animal. 2019;14(1):95–107.

    PubMed  Google Scholar 

  273. Al-Zoreky N. Antimicrobial activity of pomegranate (Punica granatum L.) fruit peels. Int J Food Microbiol. 2009;134(3):244–8.

    CAS  PubMed  Google Scholar 

  274. Kapoor G, Saigal S, Elongavan A. Action and resistance mechanisms of antibiotics: a guide for clinicians. J Anaesthesiol Clin Pharmacol. 2017;33(3):300–5. https://doi.org/10.4103/joacp.JOACP_349_15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Min BR, Pinchak WE, Anderson RC, Callaway TR. Effect of tannins on the in vitro growth of Escherichia coli O157: H7 and in vivo growth of generic Escherichia coli excreted from steers. J Food Prot. 2007;70:543–50.

    CAS  PubMed  Google Scholar 

  276. Taguri T, Tanaka T, Kouno I. Antimicrobial activity of 10 different plant polyphenols against bacteria causing food-borne disease. Biol Pharm Bull. 2004;27:1965–9.

    CAS  PubMed  Google Scholar 

  277. Yao K, He Q, Ying JD, Shi B. The potential of wattle tannin extracts for fine use. Nat Prod Res. 2006;20:271–8.

    CAS  PubMed  Google Scholar 

  278. Becker H, Scher J, Speakman J, Zapp J. Bioactivity guided isolation of antimicrobial compounds from Lythrum salicaria. Fitoterapia. 2005;76(6):580–4.

    CAS  PubMed  Google Scholar 

  279. Guclu E, Genc H, Zengin M, Karabay O. Antibacterial activity of Lythrum salicaria against multidrug-resistant Acinetobacter baumannii and Pseudomonas aeruginosa. Annu Res Rev Biol. 2014;4(7):1099–105.

    Google Scholar 

  280. Pereda-Miranda R, Kaatz G, Gibbons S. Polyacylated oligosaccharides from medicinal Mexican Morning Glory Species as antibacterials and inhibitors of multidrug resistance in Staphylococcus aureus. J Nat Prod. 2006;69(3):406–9.

    CAS  PubMed  Google Scholar 

  281. Masoud A, Gouda A. Effect of some natural plant extracts against gram negative bacteria in Njran Area, Saudi Arabia. Egypt Acad J Biol Sci G Microbiol. 2012;4(1):85–92.

    Google Scholar 

  282. Nand P, Drabu S, Rk G. Insignificant anti-acne activity of Azadirachta indica leaves and bark. J Pharm Negat Results. 2012;3(1):29–33.

    Google Scholar 

  283. Zuo G, Li Y, Wang T, Han J, Wang G, Zhang Y, Pan W. Synergistic antibacterial and antibiotic effects of bisbenzylisoquinoline alkaloids on clinical isolates of methicillin-resistant Staphylococcus aureus (MRSA). Molecules. 2011;16(12):9819–26.

    CAS  PubMed  PubMed Central  Google Scholar 

  284. Sakagami Y, Mimura M, Kajimura K, Yokoyama H, Iinuma M, Tanaka T, Ohyama M. Anti-MRSA activity of sophoraflavanone G and synergism with other antibacterial agents. Lett Appl Microbiol. 1998;27(2):98–100.

    CAS  PubMed  Google Scholar 

  285. Magi G, Marini E, Facinelli B. Antimicrobial activity of essential oils and carvacrol, and synergy of carvacrol and erythromycin, against clinical, erythromycin-resistant Group A Streptococci. Front Microbiol. 2015;6:165. https://doi.org/10.3389/fmicb.2015.00165.

    Article  PubMed  PubMed Central  Google Scholar 

  286. Tan N, Yazıcı-Tütüniş S, Bilgin M, Tan E, Miski M. Antibacterial activities of pyrenylated coumarins from the roots of Prangos hulusii. Molecules. 2017;22(7):1098.

    PubMed Central  Google Scholar 

  287. Kali A, Bhuvaneshwar D, Charles P, Seetha K. Antibacterial synergy of curcumin with antibiotics against biofilm producing clinical bacterial isolates. J Basic Clin Pharm. 2016;7(3):93.

    CAS  PubMed  PubMed Central  Google Scholar 

  288. Teow S, Ali S. Synergistic antibacterial activity of Curcumin with antibiotics against Staphylococcus aureus. Pak J Pharm Sci. 2015;28(6):2109–14.

    CAS  PubMed  Google Scholar 

  289. Zhao W, Hu Z, Okubo S, Hara Y, Shimamura T. Mechanism of synergy between epigallocatechin gallate and β-Lactams against methicillin-resistant Staphylococcus aureus. Antimicrob Agents Chemother. 2001;45(6):1737–42.

    CAS  PubMed  PubMed Central  Google Scholar 

  290. Lin R, Chin Y, Lee M. Antimicrobial activity of antibiotics in combination with natural flavonoids against clinical extended-spectrum β-lactamase (ESBL)-producing Klebsiella pneumoniae. Phytother Res. 2005;19(7):612–7.

    CAS  PubMed  Google Scholar 

  291. Pao S, Paulsen I, Saier M. Major facilitator superfamily. Microbiol Mol Biol Rev. 1998;62(1):1–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  292. Stermitz F, Tawara-Matsuda J, Lorenz P, Mueller P, Zenewicz L, Lewis K. 5’-Methoxyhydnocarpin-D and Pheophorbide A: Berberis species components that potentiate berberine growth inhibition of resistant Staphylococcus aureus. J Nat Prod. 2000;63(8):1146–9.

    CAS  PubMed  Google Scholar 

  293. Takahashi O, Cai Z, Toda M, Hara Y, Shimamura T. Appearance of antibacterial activity of oxacillin against methicillin resistant Staphylococcus aureus (MRSA) in the presence of catechin. J Jpn Assoc Infect Dis. 1995;69(10):1126–34.

    CAS  Google Scholar 

  294. Chérigo L, Pereda-Miranda R, Fragoso-Serrano M, Jacobo-Herrera N, Kaatz G, Gibbons S. Inhibitors of bacterial multidrug efflux pumps from the resin glycosides of Ipomoea murucoides. J Nat Prod. 2008;71:1037–45.

    PubMed  Google Scholar 

  295. Seukep A, Kuete V, Nahar L, Sarker S, Guo M. Plant-derived secondary metabolites as the main source of efflux pump inhibitors and methods for identification. J Pharm Anal. 2020;10(4):277–90.

    PubMed  Google Scholar 

  296. Bhardwaj M, Singh B, Sinha D, Kumar V, Prasanna V, Varan S, Nirupama K, Archana S. Potential of herbal drug and antibiotic combination therapy: a new approach to treat multidrug resistant bacteria. Pharm Anal Acta. 2016. https://doi.org/10.4172/2153-2435.1000523.

    Article  Google Scholar 

  297. Musumeci R, Speciale A, Costanzo R, Annino A, Ragusa S, Rapisarda A, Pappalardo M, Iauk L. Berberis aetnensis C. Presl. extracts: antimicrobial properties and interaction with ciprofloxacin. Int J Antimicrob Agents. 2003;22(1):48–53.

    CAS  PubMed  Google Scholar 

  298. Khameneh B, Iranshahy M, Soheili V, Fazly BB. Review on plant antimicrobials: a mechanistic viewpoint. Antimicrob Resist Infect Control. 2019. https://doi.org/10.1186/s13756-019-0559-6.

    Article  PubMed  PubMed Central  Google Scholar 

  299. Zuo G, Li Y, Han J, Wang G, Zhang Y, Bian Z. Antibacterial and synergy of berberines with antibacterial agents against clinical multi-drug resistant isolates of methicillin-resistant Staphylococcus aureus (MRSA). Molecules. 2012;17(9):10322–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  300. Fadli M, Saad A, Sayadi S, Chevalier J, Mezrioui N, Pagès J, Hassani L. Antibacterial activity of Thymus maroccanus and Thymus broussonetii essential oils against nosocomial infection—bacteria and their synergistic potential with antibiotics. Phytomedicine. 2012;19(5):464–71.

    CAS  PubMed  Google Scholar 

  301. Mahboubi M, Ghazian BF. Antistaphylococcal activity of Zataria multiflora essential oil and its synergy with vancomycin. Phytomedicine. 2010;17(7):548–50.

    CAS  PubMed  Google Scholar 

  302. Kazmi M, Malik A, Hameed S, Akhtar N, Noor Ali S. An anthraquinone derivative from Cassia italica. Phytochemistry. 1994;36(6):761–3.

    CAS  Google Scholar 

  303. Chapelsky S, Batty S, Frost M, Mogridge J. Inhibition of anthrax lethal toxin-induced cytolysis of RAW264.7 cells by celastrol. PLoS ONE. 2008;3(1): e1421.

    PubMed  PubMed Central  Google Scholar 

  304. Slater L, Hett E, Mark K, Chumbler N, Patel D, Lacy D, Collier R, Hung D. Identification of novel host-targeted compounds that protect from anthrax lethal toxin-induced cell death. ACS Chem Biol. 2013;8(4):812–22.

    CAS  PubMed  PubMed Central  Google Scholar 

  305. Siriyong T, Srimanote P, Chusri S, Yingyongnarongkul B, Suaisom C, Tipmanee V, Voravuthikunchai S. Conessine as a novel inhibitor of multidrug efflux pump systems in Pseudomonas aeruginosa. BMC Complementary Altern Med. 2017. https://doi.org/10.1186/s12906-017-1913-y.

    Article  Google Scholar 

  306. Wultańska D, Piotrowski M, Pituch H. The effect of berberine chloride and/or its combination with vancomycin on the growth, biofilm formation, and motility of Clostridioides difficile. Eur J Clin Microbiol Infect Dis. 2020;39(7):1391–9.

    PubMed  PubMed Central  Google Scholar 

  307. Haroun MF, Al-Kayali RS. Synergistic effect of Thymbra spicata L. extracts with antibiotics against multidrug-resistant Staphylococcus aureus and Klebsiella pneumoniae strains. Iran J Basic Med Sci. 2016;19(11):1193–200.

    PubMed  PubMed Central  Google Scholar 

  308. Maisuria V, Okshevsky M, Déziel E, Tufenkji N. Proanthocyanidin interferes with intrinsic antibiotic resistance mechanisms of gram-negative bacteria. Adv Sci. 2019;6(15):1802333.

    Google Scholar 

  309. Ge F, Zeng F, Liu S, Guo N, Ye H, Song Y, Fan J, Wu X, Wang X, Deng X, Jin Q, Yu L. In vitro synergistic interactions of oleanolic acid in combination with isoniazid, rifampicin or ethambutol against Mycobacterium tuberculosis. J Med Microbiol. 2010;59(5):567–72.

    CAS  PubMed  Google Scholar 

  310. Mativandlela S, Muthivhi T, Kikuchi H, Oshima Y, Hamilton C, Hussein A, van der Walt M, Houghton P, Lall N. Antimycobacterial flavonoids from the leaf extract of Galenia africana. J Nat Prod. 2009;72(12):2169–71.

    CAS  PubMed  Google Scholar 

  311. Bapela N, Lall N, Fourie P, Franzblau S, Van Rensburg C. Activity of 7-methyljuglone in combination with antituberculous drugs against Mycobacterium tuberculosis. Phytomedicine. 2006;13(9–10):630–5.

    CAS  PubMed  Google Scholar 

  312. Guo N, Wu J, Fan J, Yuan P, Shi Q, Jin K, Cheng W, Zhao X, Zhang Y, Li W, Tang X, Yu L. In vitro activity of isoimperatorin, alone and in combination, against Mycobacterium tuberculosis. Lett Appl Microbiol. 2013;58(4):344–9.

    PubMed  Google Scholar 

  313. Labuschagné A, Hussein A, Rodríguez B, Lall N. Synergistic antimycobacterial actions of Knowltonia vesicatoria (L.f) Sims. Evid Based Complementary Altern Med. 2012;2012:1–9.

    Google Scholar 

  314. Dawoud MEA, Mawgoud YA, Dawoud TMG. Synergistic interactions between plant extracts, some antibiotics and/or their impact upon antibiotic-resistant bacterial isolates. Afr J Biotech. 2013;12(24):3835–46.

    Google Scholar 

  315. El-Sakhawy M, Al-Zaban M, Alharbi M, Abdelazim N, Zain M. Synergistic effect of medicinal plants and antibiotics on pathogenic bacteria. Int J Biol Pharm Allied Sci. 2015;4(9):5792–800.

    CAS  Google Scholar 

  316. Mantzourani I, Bontsidis C, Plessas S, Alexopoulos A, Theodoridou E, Tsigalou C, Voidarou C, Douganiotis G, Kazakos S, Stavropoulou E, Bezirtzoglou E. Comparative susceptibility study against pathogens using fermented cranberry juice and antibiotics. Front Microbiol. 2019. https://doi.org/10.3389/fmicb.2019.01294.

    Article  PubMed  PubMed Central  Google Scholar 

  317. Akinyele T, Igbinosa E, Akinpelu D, Okoh A. In vitro assessment of the synergism between extracts of Cocos nucifera husk and some standard antibiotics. Asian Pac J Trop Biomed. 2017;7(4):306–13.

    Google Scholar 

  318. Klessig D, Tian M, Choi H. Multiple targets of salicylic acid and its derivatives in plants and animals. Front Immunol. 2016. https://doi.org/10.3389/fimmu.2016.00206.

    Article  PubMed  PubMed Central  Google Scholar 

  319. Jones M, Jani R, Ren D, Wood T, Blaser M. Inhibition of Bacillus anthracis growth and virulence-gene expression by inhibitors of quorum-sensing. J Infect Dis. 2005;191(11):1881–8.

    CAS  PubMed  Google Scholar 

  320. Nazzaro F, Fratianni F, Coppola R. Quorum sensing and phytochemicals. Int J Mol Sci. 2013;14(6):12607–19.

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This paper has no authors to acknowledge.

Funding

This research did not receive any specific grant from a funding body.

Author information

Authors and Affiliations

Authors

Contributions

MKD contributed in giving the original idea, concept, also collaborated in designing, and writing, data interpretation and drafting of the article. TJK and JA contributed in editing the final version of the draft to be published. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Mackingsley Kushan Dassanayake.

Ethics declarations

Ethics approval and consent to participate

The study did not require the consent of the participants.

Consent for publication

The study did not require any consent for publication.

Competing interests

The authors have no competing interests to declare.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dassanayake, M.K., Khoo, TJ. & An, J. Antibiotic resistance modifying ability of phytoextracts in anthrax biological agent Bacillus anthracis and emerging superbugs: a review of synergistic mechanisms. Ann Clin Microbiol Antimicrob 20, 79 (2021). https://doi.org/10.1186/s12941-021-00485-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12941-021-00485-0

Keywords