Next Article in Journal
Multicenter Surveillance of Antimicrobial Resistance among Gram-Negative Bacteria Isolated from Bloodstream Infections in Ghana
Next Article in Special Issue
Anti-Staphylococcal Activities of Rosmarinus officinalis and Myrtus communis Essential Oils through ROS-Mediated Oxidative Stress
Previous Article in Journal
Implementation of MRSA Nasal Swabs as an Antimicrobial Stewardship Intervention to Decrease Anti-MRSA Therapy in COVID-19 Infection
Previous Article in Special Issue
A Comparative Study on Chemical Compositions and Biological Activities of Four Amazonian Ecuador Essential Oils: Curcuma longa L. (Zingiberaceae), Cymbopogon citratus (DC.) Stapf, (Poaceae), Ocimum campechianum Mill. (Lamiaceae), and Zingiber officinale Roscoe (Zingiberaceae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

trans-Cinnamaldehyde as a Novel Candidate to Overcome Bacterial Resistance: An Overview of In Vitro Studies

Department of Physiology and Pharmacology “V. Erspamer”, Sapienza University of Rome, P.le Aldo Moro 5, 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
Antibiotics 2023, 12(2), 254; https://doi.org/10.3390/antibiotics12020254
Submission received: 17 December 2022 / Revised: 22 January 2023 / Accepted: 25 January 2023 / Published: 27 January 2023

Abstract

:
The increasing of drug-resistant bacteria and the scanty availability of novel effective antibacterial agents represent alarming problems of the modern society, which stimulated researchers to investigate novel strategies to replace or assist synthetic antibiotics. A great deal of attention has been devoted over the years to essential oils that contain mixtures of volatile compounds and have been traditionally exploited as antimicrobial remedies. Among the essential oil phytochemicals, remarkable antimicrobial and antibiotic-potentiating activities have been highlighted for cinnamaldehyde, an α,β-unsaturated aldehyde, particularly abundant in the essential oils of Cinnamomum spp., and widely used as a food additive in industrial products. In line with this evidence, in the present study, an overview of the available literature has been carried out in order to define the bacterial sensitizing profile of cinnamaldehyde. In vitro studies displayed the ability of the substance to resensitize microbial strains to drugs and increase the efficacy of different antibiotics, especially cefotaxime, ciprofloxacin, and gentamicin; however, in vivo, and clinical trials are lacking. Based on the collected findings, cinnamaldehyde appears to be of interest as an adjuvant agent to overcome superbug infections and antibiotic resistance; however, future more in-dept studies and clinical investigations should be encouraged to clarify its efficacy and the mechanisms involved.

1. Introduction

The discovery of antibiotics is considered one of the most important achievements in the history of medicine since their use has significantly reduced morbidity and mortality associated with bacterial infections [1]. However, their inappropriate use and abuse have led to the emergence of antibiotic resistance at an alarming rate, which has resulted in drug treatment failure and the development of recurrent infections [2]. This phenomenon has been favored by incorrect prescriptions and a lack of adherence to therapies [3,4]. Approximately 700,000 people die every year due to infections caused by multidrug-resistant bacteria (MDR), and this number is expected to exceed 10 million deaths by 2050 [2].
An irresponsible use of antimicrobial agents has also been highlighted in veterinary and agricultural fields. In fact, large volumes of antibiotics, often unnecessary, are administered to food-producing animals, endangering human health due to the possible presence of drug residues in food and the selection of resistant bacteria [3].
Resistant bacteria, also known as superbugs, have limited treatment options, thus representing a serious threat to public health, and increasing the risk of death, especially in critically ill patients, immunocompromised subjects, and in the hospital setting [3,5]. The most severe chronic infections are frequently caused by six pathogenic bacteria, known by their acronym ESKAPE, which means Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa and Enterobacteria sp. [6]. Indeed, hospital infections caused by P. aeruginosa and A. baumanii are resistant to almost all antibiotics; additionally, the extended-spectrum of β-lactamases of the Enterobacteriaceae has limited the efficacy of the latest generations of penicillin and cephalosporins [7]. The loss of drug efficacy along with the emergence of novel superbugs increased the need for innovative therapies; particularly, substances able to increase the susceptibility of bacteria to drugs, thus acting as bacterial sensitizers, have been approached as promising strategies to overcome antibiotic resistance and achieve the expected antibacterial efficacy [8,9].
Many natural products, both phytocomplexes and pure compounds, have been studied as possible antibacterial and sensitizing agents [10]. Among them, essential oils, which are mixtures of naturally occurring volatile compounds with a characteristic smell and flavor, attracted a great deal of attention [11]. Terpenes represent the most abundant compounds of essential oils, with lower amounts of aromatic and aliphatic substances (e.g., aldehydes, phenols, alcohols, and heterocycles) [12].
Essential oils are known to possess a broad spectrum of bioactivities, including antimicrobial, anti-inflammatory, antioxidant, genoprotective, and antiproliferative [13,14,15,16,17,18,19,20,21]; The antimicrobial properties of essential oils have been known since antiquity and represent the most exploited up until now. They may act as both bacteriostatic and bactericide agents, being able to inhibit bacterial growth, thus blocking the bacteria’s reproductive ability, and to kill bacterial cells [22,23,24]. Usually, these effects are explained based on the lipophilicity of the essential oil constituents, especially monoterpenes, which can cross the bacterial wall and alter the cell permeability [11,24]. Moreover, they can alter the conformation of different fatty acids, polysaccharides, and phospholipid layers, causing disintegration of the bacterial cell wall [11,24]. These events can be reflected in membrane potential changes, disruption of transporters, and intracellular content leakage, which eventually lead to cell lysis and death [11]. The complex composition of essential oils also allows for hypothesizing the involvement of additional antimicrobial mechanisms, including the inhibition of bacterial enzymes and the interference with systems involved in energy production and the synthesis of structural components [24].
Among the essential oil compounds, a great interest has been devoted to cinnamaldehyde, also known as cinnamic aldehyde or 3-phenyl-2-propenal (Figure 1), an α,β-unsaturated aldehyde, belonging to the class of phenylpropanoids. It is widely used as a food additive in industrial products, such as drinks, candies, ice cream, chewing gum, and condiments [25], and it is rated safe (GRAS) by the United States Food and Drug Administration (FDA) and by the Flavor and Extract Manufacturer’s Association (FEMA) [5]. The FDA and the Council of Europe have recommended a daily intake of 1.25 mg/kg [25].
Cinnamaldehyde occurs naturally as a trans stereoisomer, namely (2E)-3-phenylprop-2-enal or trans-cinnamaldehyde, which is especially abundant in the essential oils from Cinnamomum spp. (Fam. Lauraceae), where it contributes to the typical aroma [26,27]. However, minor amounts (≤0.9%) of (2Z)-cinnamaldehyde in the essential oils of Cinnamomum spp. from Madagascar have been reported as well [28].
The bark of Cinnamomum cassia Nees (or Chinese cinnamon) and Cinnamomum verum J. Presl (or true cinnamon), which achieve about 85% and 90% content of trans-cinnamaldehyde, are considered its major natural sources [29], although other varieties (Table 1) can produce high amounts of the substance [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49].
The biosynthesis of cinnamaldehyde arises from the deamination of l-phenylalanine into a cinnamic acid by a phenylalanine-ammonia lyase, followed by the conversion into a cinnamoyl-CoA, mediated by a 4-coumarate-CoA ligase, which in turn is reduced to cinnamaldehyde by a cinnamoyl-CoA reductase [50].
Cinnamoyl moiety is a characteristic scaffold of cinnamaldehyde and its derivatives: it is considered as a Michael acceptor due to the presence of a α,α-unsaturated carbonyl pharmacophore, which can react with different electrophilic structures (e.g., enzymes, receptors), leading to several pharmacological effects [28]. Indeed, the substance has been found endowed with remarkable bioactivities in preclinical models (Figure 2), including antioxidant, anti-inflammatory, antimutagenic, antiproliferative, and neuroprotective ones [51,52]; moreover, its chemopreventive power has been reported [53]. Semisynthetic derivatives (e.g., α-hexylcinnamaldehyde) of cinnamaldehyde have also been studied to exploit the pharmacological properties of the lead compound and achieve improvements in its chemical stability [54,55,56].
Among the bioactivities of cinnamaldehyde, a remarkable broad spectrum of antibacterial and antifungal properties has been highlighted: the substance was especially effective against Escherichia coli, Staphylococcus aureus, Salmonella spp., and Bacillus spp. strains, acting through bactericidal mechanisms [28].
The antimicrobial capacity seems to arise from the ability of the substance to interact with the bacterial wall and disrupt its integrity; indeed, the aldehydic group can be easily absorbed by the hydrophilic group of the bacterial surfaces, then it can pass through the cell wall and start a process of inhibition and sterilization by destroying the polysaccharide structure, leading to leakage of ions, proteins, and nucleic acids [25,28]. Other mechanisms, such as the inhibition of biofilm formation and ATP production, along with the interference with the quorum sensing systems, have been reported too [28]. A special attention has also been devoted in the years to the antibiotic-potentiating properties of cinnamaldehyde, especially in superbug strains. In line with this evidence, in the present study, an overview of the available literature has been carried out in order to define the bacterial sensitizing profile of cinnamaldehyde and to highlight a future interest in this natural substance as a novel strategy to overcome antibiotic resistance.

2. Methodology

The existing literature in PubMed and Scopus databases was searched in November 2022 to select journal articles over a 20-year period (2002-today) focused on the antibacterial combination of cinnamaldehyde and antibiotics in resistant bacteria; combinations with antifungal agents have also been considered. English was chosen as the preferred language. The keywords “trans-cinnamaldehyde”, “cinnamaldehyde” “antibiotic”, “synergism”, and “combination”, and their combinations through the Boolean logical operator “AND” have been used. As a research strategy, the PRISMA methodology was applied to select eligible papers for the study [57]. Notably, the studies focused on herbal extracts or essential oils containing cinnamaldehyde, but not on the pure compound, along with studies in which the purity of the substance was low (<90%) or not specified, and studies assessing other substances, diverse bioactivities, or lacking data, were excluded.

3. Results

A total of 276 studies focused on the ability of cinnamaldehyde to potentiate the effect of antimicrobial drugs when used in combination (Figure 3). Among them, 24 records were removed as publications other than journal articles, while 129 were replicates in searched databases.
The screened 123 papers were further selected; out of which 25 studies focused on essential oils or herbal extracts containing cinnamaldehyde, and the other 14 on other substances, and were excluded. Moreover, out of 83 eligible papers, 4 reports were not included since they focused on other bioactivities; similarly, 61 records evaluating the antimicrobial properties of cinnamaldehyde alone, but not in combination with antimicrobial drugs, were removed too. Furthermore, 8 studies were not included for lacking data and another one since purity was not specified. At the end of the literature analysis, a total of 10 studies were considered eligible since they met the inclusion criteria.
Based on the selected studies, cinnamaldehyde has been found to be able to potentiate the antimicrobial properties of different drugs, although with specific potency and efficacy with respect to the drug and bacterial (or fungal) strain [58,59,60,61,62,63,64,65,66,67]. Usually, it produces synergistic or additive effects and allows for a significant reduction in the MIC (minimal inhibitory concentration) value of the combined drug, thus suggesting promising bacterial sensitizing properties. It is noteworthy that some of the susceptible bacteria [60,62,63,64,65] belonged to the ESKAPE group (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, Enterobacter sp.), known to be responsible for resistant infections.
To quantify the type of interaction (synergism or antagonism) between cinnamaldehyde and antibiotic drugs, the fractional inhibitory concentration index (FICI), which represents the sum of FIC concentrations of each component in the mixture, has been conserved [68]. Considering a given combination of two drugs A and B, FICA is the MIC of drug A in the presence of compound B divided by the MIC of drug A alone (FICA = [MICA(B)/MICA]), and vice versa for FICB (FICB = [MICB(A)/MICB]) [68]. The FICI value is the sum of FICA and FICB and reveals the degree of drug interaction: a lower than 0.5 FICI value indicates a synergistic interaction, values between 0.5 and 1 denote additive effects, while FICI values higher than 1 and 4 correspond to null and antagonistic effects, respectively [68].
The ability of cinnamaldehyde to affect drug efficacy in different bacterial strains (Gram-positive and Gram-negative) and fungi has been described and detailed in Table 2. For each microorganism, the strain, antibiotic drug, MIC value (expressed as µg/mL) of the antibiotic drug alone and in combination with cinnamaldehyde, the cinnamaldehyde concentration in combination (expressed as µg/mL), and the FICI value have been displayed.

3.1. Potentiating Effects of Cinnamaldehyde in Gram-Positive Bacteria

The potentiating effects of cinnamaldehyde were evaluated in different Gram-positive bacteria, including Listeria monocytogenes, Staphylococcus aureus and its methicillin-resistant strains, namely MRSA (methicillin-resistant S. aureus), and Streptococcus pyogenes (Table 2).
L. monocytogenes is a ubiquitous bacterium, implicated within the past decade in several outbreaks of foodborne disease [69]. It causes invasive syndromes, and case fatalities can be around 30% in specific high-risk population groups, such as the elderly, immunocompromised individuals, fetuses, and newborns [70]. Moreover, it may acquire antibiotic resistance genes from the plasmids and conjugative transposons of other organisms [71]. Only a few studies have evaluated the ability of cinnamaldehyde to synergize antibiotics in L. monocytogenes. Alves et al. [58] highlighted that the substance produced synergistic effects with nisin (0.50 FICI), a bacteriocin produced by Lactococcus lactis strains, reducing the MIC value by 4 folds.
S. aureus is a Gram-positive opportunistic pathogen that is responsible for many nosocomial and community-acquired infections. The attachment to medical implants and host tissue, and the establishment of a mature biofilm, all play an important role in the persistence of chronic infections [72,73]. Clinical use of methicillin led to the development of methicillin-resistant S. aureus (MRSA) strains [74], which increased the need for new therapeutic strategies to sensitize these strains to the antibiotic treatment.
Cinnamaldehyde has been assessed against S. aureus in association with conventional antibiotics and other antibacterial substances, such as nisin. In particular, two studies have highlighted the ability of cinnamaldehyde to significantly synergize nisin with 0.26 to 0.50 FICI values [58,59]. Remarkable synergistic effects were reported in combination with ampicillin, piperacillin, and bacitracin (0.24–0.37 FICI), with antibiotic MIC values reduced by about 8 folds [60]. The substance was also found to greatly synergize amikacin, amoxicillin, and gentamicin (0.19–0.50 FICI) in MRSA strains [61]; moreover, it lowered by about 2-fold the MIC value of ampicillin and ceftazidime (1.00 FICI), although without exhibiting synergistic effects [61]. Both additive and synergistic interactions were recorded in combination with cefoxitin, oxacillin, and vancomycin [61].
Table 2. Effect of cinnamaldehyde in combination with antimicrobial drugs in bacterial strains.
Table 2. Effect of cinnamaldehyde in combination with antimicrobial drugs in bacterial strains.
BacteriaStrainAntibioticMIC [µg/mL]Antibiotic and trans-
Cinnamaldehyde Combination
FICI/Type of InteractionRef.
Cinnamaldehyde Concentration [µg/mL]MIC [µg/mL] (RR)
Gram-positive
Listeria monocytogenesATCC 15313Nisin12516.2562.5 (4)0.50/Synergism[58]
Staphylococcus aureusJL10001Nisin16502 (8)0.32/Synergism[59]
JL10002, JL10006, JL10008, JL100111662.51 (16)0.31/Synergism
JL1000, JL10005, JL10009, JL10013321252 (16)0.31/Synergism
JL1000416125 2 (8)0.37/Synergism
JL10007 JL100121662.5 2 (8)0.37/Synergism
JL100103262.5 4 (8)0.37/Synergism
ATCC 292133250 2 (16)0.26/Synergism
ATCC 2592311025 27.5 (4)0.50/Synergism[58]
bla ZAmpicillin3241.34 (8)0.25/Synergism[60]
Bacitracin3241.34 (8)0.24/Synergism
Piperacillin128 0.37/Synergism
Methicillin-resistant Staphylococcus aureus (MRSA)ATCC 33571Amikacin31.231.25 7.8 (4)0.38/Synergism[61]
Dps-131.231.25 3.9 (8)0.25/Synergism
Dps-362.531.25 3.9 (16)0.19/Synergism
ATCC 33571Amoxicillin62.5125 7.8 (8)0.63/Additive effect
Dps-112562.5 31.25 (4)0.5/Synergism
Dps-312531.25 15.6 (8)0.25/Synergism
ATCC 33571Ampicillin62.5125 31.25 (2)1.00/Additive effect
Dps-131.3125 7.8 (4)0.75/Additive effect
Dps-362.5125 15.6 (4)0.75/Additive effect
ATCC 33571Cefoxitin31.2125 7.8 (4)0.75/Additive effect
Dps-162.5125 7.8 (8)0.62/Additive effect
Dps-325031.25 31.25 (4)0.50/Synergism
ATCC 33571Ceftazidime125125 62.5 (2)1.00/Additive effect
Dps-1125125 62.5 (2)1.00/Additive effect
Dps-3250125 62.5 (4)0.75/Additive effect
ATCC 33571Gentamicin3.9125 0.97 (4)0.75/Additive effect
Dps-112531.25 31.25 (4)0.37/Synergism
Dps-325062.5 62.5 (4)0.50/Synergism
ATCC 33571Oxacillin62.5125 15.6 (4)0.75/Additive effect
Dps-1500125 250 (2)1.00/Additive effect
Dps-350031.25 62.5 (8)0.25/Synergism
ATCC 33571Vancomycin25031.25 31.25 (8)0.25/Synergism
Dps-1250125 125 (2)1.00/Additive effect
Dps-3500125 250 (2)1.00/Additive effect
Streptococcus pyogeneserm BErythromycin>51241.6>256 (8)1.00/Additive effect[60]
Nitrofurantoin0.13/Synergism
Gram-negative
Escherichia coli28 clinically isolated strainsCefotaxime5120.221 (512)0.07–0.30/75% synergism [62]
Ciprofloxacin5120.118 (64)0.07–0.50/39.6% synergism
ATCC 11775Erythromycin16100 4 (4)0.50/Synergism[63]
ATCC 2373932- -0.30/Synergism
8WT64100 16 (4)0.50/Synergism
02:062716100 4 (4)0.50/Synergism
ATCC 23739Tetracycline32--0.30/Synergism
ATCC 23739Novobiocin128--0.20/Synergism
8WT6432 32 (2)1.00/Additive effect
02:062712810032 (4)0.50/Synergism
ATCC 11775Bacitracin>512-->1.00/Lacking effect
ATCC 23739>512-->1.00/Lacking effect
8WT>512-->1.00/Lacking effect
02:0627>512-->1.00/Lacking effect
N00 666Ampicillin>512 0.37/Synergism[60]
Bacitracin>512165.2>64 (8)0.63/Additive effect
Erythromycin51241.364 (8)0.24/Synergism
Novobiocin6441.38 (8)0.24/Synergism
Piperacillin>51241.3>64 (8)0.24/Synergism
Tetracycline128 0.37/Synergism
Klebsiella sp.33 clinically isolated strainsCefotaxime5120.050.5 (1024)0.10–0.50/42.4% synergism[62]
Ciprofloxacin5120.032 (256)0.07–0.50/60.6% synergism
Pseudomonas aeruginosaPAO1Carbenicillin128396.5 64 (2)0.75/Additive effect[64]
Colistin7.86396.5 1.96 (4)0.50/Synergism
Erythromycin256396.5 128 (2)0.75/Additive effect
Tobramycin1443.8396.5 721.9 (2)0.75/Additive effect
Gentamicin4.07.50.25 (16)0.37/Synergism[65]
Salmonella typhimuriumSGI 1Ampicillin>51241.3 >64 (8)0.25/Synergism[60]
Bacitracin>51241.3>64 (8)0.24/Synergism
Erythromycin102441.3 128 (8)0.24/Synergism
Novobiocin25641.3 32 (8)0.24/Synergism
Piperacillin>512165.2 >64 (8)0.63/Additive effect
Tetracycline64 0.37/Synergism
Fungi
Aspergillus fumigatusMTCC 2550Fluconazole200525 (8)0.19/Synergism[66]
Malassezia pachydermatis30 isolated strainsClotrimazole0.03–64 (GM 4.5)1.25–40 (GM 3.15)0.063–8 (GM 0.52)0.064–2.125 (GM: 0.52)/40% synergism
60% null effect
[67]
Fluconazole1–64
(GM 9.4)
1.25–40
(GM 6.64)
0.25–16 GM 0.7 (4)0.066–12 (GM 0.73)/
26.6% synergism
70% antagonism
Ketoconazole0.015–4 (GM 0.08)1.25–160 (GM 5.48)0.016–0.062 (GM 0.02)0.093–6.006 (GM 1.55)/
23.3% synergism
30% null effect
46,6% antagonism
Itraconazole0.0039–1 (GM 0.02)1.25–160 (GM 4.66)0.016–0.125 (GM 0.02)0.007–16.52 (GM: 0.85)/
30.0% synergism
56.6% null effect
13.3% antagonism
Miconazole0.03–64 (GM 8.96)1.25–40 (GM 2.17)0.016–8 (GM 0.72)0.039–2.003 (GM: 0.31)/
66.6% synergism
33.3% null
Nystatin4–64 (GM 41.96)1.25–20 (GM 2.22)0.25–64 (GM 29.2)0.062–1.25 (GM 0.31)/
70% synergism
30% null effect
Terbinafine0.03–64 (GM 2.57)1.25–40 (GM 8.31)0.125–8 (GM 0.29)0.046–4.5 (GM: 0.97)/
16.6% synergism
70% null effect
13.3% antagonism
Trichophyton rubrumIO A-9Fluconazole2001.2525 (8)0.16/Synergism[66]
At last, possible potentiating effects of cinnamaldehyde were evaluated in Streptococcus pyogenes, which is an exclusive human Gram-positive bacterial pathogen, characterized by high virulence and mortality risk [75]. Palaniappan et al. [60] highlighted synergistic effects of cinnamaldehyde with nitrofurantoin (0.13 FICI value) in Streptococcus pyogenes, while an additive effect was observed in combination with ampicillin (1.00 FICI).

3.2. Potentiating Effects of Cinnamaldehyde in Gram-Negative Bacteria

The substance was assessed in combination with different antibiotics in many Gram-negative strains, including Escherichia coli, Klebsiella spp., Pseudomonas aeruginosa, and Salmonella typhimurium (Table 2). E. coli and Klebsiella spp., belonging to the Enterobacteriaceae, are usually part of the intestinal flora but can also contribute to a wide range of both community- and hospital-acquired infections [76]. Klebsiella spp. are also responsible for opportunistic nosocomial infections, with a high incidence of resistant strains [77,78]. β-Lactam antibiotics are usually administered to treat their infections, although the resistance to these drugs causes serious pharmacological and medical issues [76]. E. coli belongs to the resident flora in the lower intestinal tract of warm-blooded animals, such as humans, but can also be found as an environmental contaminant as a consequence of the release of feces or wastewater effluent [77].
Dhara et al. [62] showed that cinnamaldehyde synergized ciprofloxacin (0.07–0.50 FICI) and cefotaxime (0.10–0.50 FICI) in Klebsiella spp. in 60.6% and 42.4% of cases, respectively; the MIC values of ciprofloxacin and cefotaxime were lowered by 256 and 1024 folds, respectively. Furthermore, cinnamaldehyde exhibited synergistic effects (≤0.5 FICI) in combination with erythromycin, tetracycline, cefotaxime, ciprofloxacin, ampicillin, and piperacillin in E. coli, although with weak or null effects in combination with novobiocin and bacitracin [60,63].
As for Pseudomonas aeruginosa, a common Gram-negative environmental organism that can cause severe infections in humans owing to its natural resistance to antibiotics and the ability to form biofilms [79], Topa et al. [64] demonstrated that cinnamaldehyde produced synergistic effects with colistin (0.50 FICI) and additive effects with carbenicillin, tobramycin, and erythromycin (0.75 FICI). Recently, Chada et al. [65] highlighted a synergist interaction of cinnamaldehyde with gentamicin in P. aeruginosa (0.375 FICI), with a 4-fold lowering of the antibiotic MIC. Moreover, the substance exhibited a quorum quenching (QQ) potential, being able to attenuate the quorum sensing (QS) circuits, particularly by downregulating QS genes and abrogating the biosynthesis of key factors involved in bacterial virulence and biofilm formation [65]. The antivirulence properties of cinnamaldehyde in combination with gentamicin were also confirmed in a Caenorhabditis elegans model infected with a P. aeruginosa infection [65]. These findings highlight an interest in cinnamaldehyde as a possible anti-quorum sensing agent to be exploited in combination with antibiotics in the battle against P. aeruginosa and deserve further in vivo studies for confirmation.
At last, the possible synergistic potential of cinnamaldehyde has been evaluated in S. typhimurium in combination with different antibiotics [60]. This bacterium primarily affects the intestinal lumen and often causes diarrhea in infants and young children, leading to food poisoning. Furthermore, the development of drug resistance by S. typhimurium strains led to serious complications in clinical patients [80]. Palaniappan et al. [60] showed remarkable synergistic effects of cinnamaldehyde in combination with ampicillin, tetracycline, erythromycin, bacitracin, and novobiocin (0.24–0.37 FICI) in S. typhimurium, reducing the MIC values of all the tested antibiotics by about 8 folds.

3.3. Potentiating Effects of Cinnamaldehyde in Fungi

Cinnamaldehyde has also been assayed as a possible strategy to counteract fungi infections, and some studies highlighted its ability to potentiate the effects of some antifungal drugs (Table 2): particularly, it partly synergized azole drugs in Aspergillus fumigatus, Trichophyton rubrum, and Malassezia pachydermatis fungi, being especially effective in combination with fluconazole (<0.2 FICI) [66,67].

4. Discussion

The increasing prevalence of drug-resistant bacteria and the lack of effective antibiotics have highly alarmed the scientific community, leading researchers to investigate natural substances as novel strategies to both directly affect bacterial infections and synergize synthetic antibiotics. Among natural compounds, cinnamaldehyde attracted special attention owing to its antibacterial properties and the ability to resensitize microbial strains to drugs [60], thus suggesting a possible interest in the battle against antibiotic resistance.
In this study, we selected ten in vitro studies, which are not available in vivo or in clinical trials, using the following criteria: >90% purity of cinnamaldehyde and combination of this substance with antimicrobial agents to counteract resistant bacteria. The purity of cinnamaldehyde is a key issue, since the presence of impurities in minor compounds can affect the activity of the tested substance, leading to unreliable results.
Based on the selected studies, the most efficient synergism was found when cinnamaldehyde (0.03–0.05 µg/mL) was assessed in combination with cefotaxime or ciprofloxacin in 33 clinical isolates of Klebsiella sp; in fact, the MIC values were lowered by 1024 and 256 folds, respectively [62]. Similar results were obtained in 28 clinical isolates of Escherichia coli, where cinnamaldehyde (0.11–0.22 µg/mL) lowered the MIC value of cefotaxime by 512 folds, and that of ciprofloxacin by 64 folds [62]. Interesting synergistic effects of cinnamaldehyde were also highlighted in combination with tetracycline in Escherichia coli, where a MIC reduction of 4- to 8-fold was registered; similar potentiating effects were produced in combination with erythromycin, novobiocin, ampicillin, and piperacillin [62,63].
Cinnamaldehyde also produced synergistic effects in combination with colistin and gentamicin in Pseudomonas aeruginosa, reducing the MIC values by 4- and 16-fold, respectively [64,65], and in MRSA strains in combination with amikacin (16-fold reduction of the antibiotic MIC), gentamicin, and vancomycin, followed by oxacillin and amoxicillin; the substance was found effective at concentrations from 31.25 to 62.5 µg/mL, corresponding to 1/8 and 1/4 of the MIC value [61]. Similarly, a notable antibacterial activity of the combination of cinnamaldehyde and nisin (i.e., 25 to 125 µg/mL cinnamaldehyde and 1/8 of the antibiotic MIC) was reported in S. aureus [58,59]. The substance (16.25–41.3 µg/mL) also potentiated the antibiotic effects of ampicillin, tetracycline, erythromycin, bacitracin, and novobiocin in Salmonella typhimurium and those of nisin (62.5 µg/mL) in Listeria monocytogenes ATCC 15313, reducing the MIC value by 4- to 8-fold [58,60].
Schlemmer et al. [67] demonstrated a partial synergism between cinnamaldehyde and fluconazole, ketoconazole, itraconazole, clotrimazole, miconazole, terbinafine, and nystatin against Malassezia pachydermatis. Additionally, potentiating effects towards fluconazole (25 µg/mL corresponding to 1/8 of MIC) were reported in Aspergillus fumigatus and Trichophyton rubrum [66]. Additive effects were achieved when cinnamaldehyde (at a halved MIC value) was administered in combination with ampicillin or cefotaxime in Staphylococcus aureus [61], piperacillin in Salmonella typhimurium [60], erythromycin in Streptococcus pyogenes [60], and bacitracin against Escherichia coli [60,61]. Null or antagonistic effects of cinnamaldehyde with some antimicrobial agents were reported as well [63,67].
In this respect, Tetard et al. [81] showed that cinnamaldehyde (>256 µg/mL) triggers an upregulation of the efflux pumps of the resistance-nodulation-cell division (RND) family in P. aeruginosa, especially of the multidrug efflux system MexAB-OprM, which can lead to increased drug extrusion and lowered antibiotic efficacy. This effect was found to be transient and persistent until the compound is degraded into cinnamic alcohol, which lacks the ability to induce the efflux pumps [81]. Moreover, the authors highlighted that the resistance induced by cinnamaldehyde in P. aeruginosa was modest and gained after several days of exposure at concentrations higher than 900 µg/mL [82]. Furthermore, the mutation mechanisms and the clinical impact remain to be clarified. It is important to outline that the concentrations of cinnamaldehyde inducing bacterial sensitization in P. aeruginosa [64] were at least 3- to 120-fold lower than those responsible for the resistance, suggesting that opposite effects can occur depending on the concentrations of the substance; more in-depth studies could clarify this issue.
In regard to the mechanisms accounting for the bacterial sensitizing properties of cinnamaldehyde, the substance has been shown to affect multiple targets, including the bacterial wall, biofilm, quorum sensing system, cell metabolism, and factors involved in cell survival (Figure 4), which in turn can contribute to the potentiation of the antibiotic efficacy and the overcoming of resistance.
As also reported for other essential oil compounds [21,22], Shi et al. [59] hypothesized that the synergistic effects of cinnamaldehyde in combination with nisin in S. aureus ATCC 29213 could arise from its ability to damage the bacterial wall and alter its permeability, thus affecting the antibiotic absorption and impairing the bacterial cell homeostasis, leading to autolysis and cell death. Indeed, an 54.5% membrane damage was induced by the combination of cinnamaldehyde and nisin with respect to the drug alone (28% damage) [59]. Similarly, Chadha et al. [65] hypothesized that cinnamaldehyde can cause membrane permeabilization and disruption along with oxidative damage, thus facilitating the penetration of the antibiotic gentamicin into cell and making the bacterial cell more susceptible to its antimicrobial activity.
Wang et al. [61] highlighted that the substance was able to destroy the bacterial wall and biofilm of MRSA and to downregulate the transcription and translation of the antibiotic resistance gene mecA; these effects could explain the synergism with non-beta-lactam antibiotics. Dhara et al. [62] also reported alterations in the cell surface morphology, shrinkage of the cell surface, and cytoplasm lowering in Gram-negative bacteria, i.e., E. coli and K. pneumoniae, after treatment with cinnamaldehyde, likely as a consequence of permeability and osmotic changes induced by the substance. Moreover, deep pores, disruption of the cytoplasmic membrane, and decomposition of inner organelles on cell surfaces were revealed after treatment with the combination of cinnamaldehyde and cefotaxime/ciprofloxacin [62].
Gram-negative bacteria carry an outer membrane characterized by an asymmetric hydrophobic bilayer composed of phospholipids and lipopolysaccharides (LPS), the latter playing a crucial role in the bacteria’s protection [83]. Most antibiotics are absorbed through the outer membrane to reach their targets: hydrophobic drugs are able to pass the membrane by diffusion mechanisms, while hydrophilic ones, like β-lactams, exploit the bacterial porins to be transferred into cells [84]. Any alteration in the outer membrane of Gram-negative bacteria, including changes in the hydrophobic properties and porin mutations, can lower the antibiotic permeability, thus leading to bacterial resistance [84]. In this respect, the results obtained by Dhara et al. [62] strengthen the hypothesis that an impairment in the bacterial wall by cinnamaldehyde is a key mechanism of its bacterial sensitizing activity.
Some studies also highlighted that cinnamaldehyde significantly inhibited the biofilm formation and the expression of the biofilm regulatory gene hld in methicillin-resistant Staphylococcus aureus [61]. Moreover, the combined treatments of cinnamaldehyde with colistin and tobramycin potentiated the drug’s ability to inhibit biofilm formation, leading to a complete inhibition of the process [64].
Biofilm is composed by a complex community of microbes that can adhere to a surface or form aggregates, enclosed in an extracellular polysaccharide matrix [85,86]. It enhances the bacterial resistance to hostile environmental conditions, allows the cellular exchange of plasmids encoding for antibiotic resistance, and impairs the activation of the immune system response, thus favoring the bacterial invasion [86]. It is also responsible for the development of persistent infections [87]. The biofilm inhibition by cinnamaldehyde can arise from different mechanisms, among which is a block of the quorum sensing system (QS), as recently highlighted by Chadha et al. [65] in combination with gentamicin in P. aeruginosa.
QS represents a cell-to-cell communication mechanism that occurs extensively in both Gram-positive and Gram-negative bacteria [88]. It consists of enzymes, receptors, and factors that regulate various bacterial functions, including biofilm production, sporulation, motility, and virulence [88,89]. The QS signal molecules are characterized by a low molecular weight and can be classified into different classes, including acyl homoserine lactones (AHLs), furanosyl borate diesters (AI2), cis-unsaturated fatty acids (DSF family signals), and peptides [89]. In P. aeruginosa, the QS system harbors two complete AHL circuits, namely LasI/LasR and RhlI/RhlR, with LasI/R being hierarchically positioned upstream of the RhlI/R circuit [88].
Chadha et al. [65] reported that cinnamaldehyde was able to affect the QS system in P. aeruginosa PAO1 by downregulating the QS and virulence genes (e.g., las, rhl, rhlAB, aprA, toxA, plcH) and abrogating the biosynthesis of AHL (acyl-homoserine lactones) molecules, involved in the QS processes. Similarly, Topa et al. [74] showed that cinnamaldehyde inhibited the expression of the LasB, RhlA, and PqsA QS systems in P. aeruginosa. Cinnamaldehyde exhibited a quorum quenching (QQ) potential, being able to affect the QS system at subinhibitory concentrations [65]. As also confirmed by molecular docking studies, the effect can be attributed to the ability of the substance to easily gain access to the active site of the QS receptors of P. aeruginosa, owing to its relatively small size; furthermore, being structurally similar to the AHL molecules, 3-oxo-C12-HSL and C4-HSL, it can strongly interact with the QS receptors, thus attenuating the QS circuits, inhibiting the biofilm formation, and lowering the bacterial virulence and motility [65]. Particularly, it has been hypothesized that cinnamaldehyde may abrogate the twitching motility in P. aeruginosa by inhibiting the mechanotactic functions of type IV pilus and the swimming and swarming motilities because of its anti-QS properties [65]. Furthermore, an inhibition of the EPS (extracellular polymeric substance) production by cinnamaldehyde, especially in relation to the alginate and rhamnolipid components, seems to directly modulate the pseudomonal biofilm formation and demonstrates the anti-fouling properties of the natural substance against P. aeruginosa [65].
Other mechanisms have also been proposed to explain the synergistic effects of cinnamaldehyde in combination with antibiotics. Particularly, Thirapanmethee et al. [90] showed that the substance blocked the polymerization, assembly, and bundling of the bacterial protein FtsZ in Acinetobacter baumanni, involved in the control of cell division [90,91]. Furthermore, some studies highlighted an ATP depletion by cinnamaldehyde [92,93], which could be reflected in an impairment of the bacterial function and survival.

5. Conclusions

Altogether, the collected evidence suggests a possible interest in cinnamaldehyde as an adjuvant strategy to synergize or support the effects of synthetic antibiotics against bacteria, especially against resistant strains and superbugs. However, as a small and heterogeneous group of in vitro studies, more in-depth mechanistic evidence and clinical investigations should be encouraged to clarify the promises and challenges of cinnamaldehyde in antibiotic resistance.

Author Contributions

Conceptualization, F.U. and A.D.S.; methodology, A.D.S.; software, F.U. and A.D.S.; visualization, F.U. and A.D.S.; supervision, A.D.S.; writing—original draft preparation, F.U. and A.D.S.; writing—review and editing, F.U. and A.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The Authors thank Annabella Vitalone and Silvia Di Giacomo for the precious advice and “Enrico and Enrica Sovena” Foundation (Italy) for the support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghosh, D.; Veeraraghavan, B.; Elangovan, R.; Vivekanandan, P. Antibiotic resistance and epigenetics: More to it than meets the eye. Antimicrob. Agents Chemother. 2020, 64, e02225-19. [Google Scholar] [CrossRef] [PubMed]
  2. Simonsen, G.S.; Tapsall, J.W.; Allegranzi, B.; Talbot, E.A.; Lazzari, S. The antimicrobial resistance containment and surveillance approach--a public health tool. Bull. World Health Organ. 2004, 82, 928–934. [Google Scholar] [PubMed]
  3. Dadgostar, P. Antimicrobial resistance: Implications and costs. Infect. Drug Resist. 2019, 12, 3903–3910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bowler, P.; Murphy, C.; Wolcott, R. Biofilm exacerbates antibiotic resistance: Is this a current oversight in antimicrobial stewardship? Antimicrob. Resist. Infect. Control 2020, 9, 162. [Google Scholar] [CrossRef]
  5. Uruén, C.; Chopo-Escuin, G.; Tommassen, J.; Mainar-Jaime, R.C.; Arenas, J. Biofilms as promoters of bacterial antibiotic resistance and tolerance. Antibiotics 2020, 10, 3. [Google Scholar] [CrossRef]
  6. Parrino, B.; Carbone, D.; Cirrincione, G.; Diana, P.; Cascioferro, S. Inhibitors of antibiotic resistance mechanisms: Clinical applications and future perspectives. Future Med. Chem. 2020, 12, 357–359. [Google Scholar] [CrossRef] [Green Version]
  7. Levy, S.B.; Marshall, B. Antibacterial resistance worldwide: Causes, challenges and responses. Nat. Med. 2004, 10, S122–S129. [Google Scholar] [CrossRef]
  8. Khameneh, B.; Iranshahy, M.; Soheili, V.; Bazzaz, B.S.F. Review on plant antimicrobials: A mechanistic viewpoint. Antimicrob. Resist. Infect. Control 2019, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, C.; Chowdhury, M.A.K.; Huo, Y.; Gong, J. Phytogenic compounds as alternatives to in-feed antibiotics: Potentials and challenges in application. Pathogens 2015, 4, 137–156. [Google Scholar] [CrossRef] [Green Version]
  10. Stan, D.; Enciu, A.M.; Mateescu, A.L.; Ion, A.C.; Brezeanu, A.C.; Stan, D.; Tanase, C. Natural compounds with antimicrobial and antiviral effect and nanocarriers used for their transportation. Front. Pharmacol. 2021, 12, 723233. [Google Scholar] [CrossRef]
  11. Brochot, A.; Guilbot, A.; Haddioui, L.; Roques, C. Antibacterial, antifungal, and antiviral effects of three essential oil blends. Microbiologyopen 2017, 6, e00459. [Google Scholar] [CrossRef] [PubMed]
  12. Falleh, H.; Ben Jemaa, M.; Saada, M.; Ksouri, R. Essential oils: A promising eco-friendly food preservative. Food Chem. 2020, 330, 127268. [Google Scholar] [CrossRef] [PubMed]
  13. Zengin, G.; Menghini, L.; Di Sotto, A.; Mancinelli, R.; Sisto, F.; Carradori, S.; Cesa, S.; Fraschetti, C.; Filippi, A.; Angiolella, L.; et al. Chromatographic analyses, in vitro biological activities, and cytotoxicity of Cannabis sativa L. essential oil: A multidisciplinary study. Molecules 2018, 23, 3266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Garzoli, S.; Pirolli, A.; Vavala, E.; Di Sotto, A.; Sartorelli, G.; Božović, M.; Angiolella, L.; Mazzanti, G.; Pepi, F.; Ragno, R. Multidisciplinary approach to determine the optimal time and period for extracting the essential oil from Mentha suaveolens Ehrh. Molecules 2015, 20, 9640–9655. [Google Scholar] [CrossRef] [Green Version]
  15. Menghini, L.; Ferrante, C.; Carradori, S.; D’Antonio, M.; Orlando, G.; Cairone, F.; Cesa, S.; Filippi, A.; Fraschetti, C.; Zengin, G.; et al. Chemical and bioinformatics analyses of the anti-leishmanial and anti-oxidant activities of hemp essential oil. Biomolecules 2021, 11, 272. [Google Scholar] [CrossRef]
  16. Di Sotto, A.; Gullì, M.; Acquaviva, A.; Tacchini, M.; Di Simone, S.C.; Chiavaroli, A.; Recinella, L.; Leone, S.; Brunetti, L.; Orlando, G.; et al. Phytochemical and pharmacological profiles of the essential oil from the inflorescences of the Cannabis sativa L. Ind. Crops Prod. 2022, 183, 114980. [Google Scholar] [CrossRef]
  17. Zhao, Q.; Zhu, L.; Wang, S.; Gao, Y.; Jin, F. Molecular mechanism of the anti-inflammatory effects of plant essential oils: A systematic review. J. Ethnopharmacol. 2023, 301, 115829. [Google Scholar] [CrossRef]
  18. Sharma, M.; Grewal, K.; Jandrotia, R.; Batish, D.R.; Singh, H.P.; Kohli, R.K. Essential oils as anticancer agents: Potential role in malignancies, drug delivery mechanisms, and immune system enhancement. Biomed. Pharmacother. 2022, 146, 112514. [Google Scholar] [CrossRef]
  19. Shala, A.; Singh, S.; Hameed, S.; Khurana, S.M.P. Essential oils as alternative promising anti-candidal agents: Progress and prospects. Curr. Pharm. Des. 2022, 28, 58–70. [Google Scholar] [CrossRef]
  20. Chaudhari, A.K.; Singh, V.K.; Kedia, A.; Das, S.; Dubey, N.K. Essential oils and their bioactive compounds as eco-friendly novel green pesticides for management of storage insect pests: Prospects and retrospects. Environ. Sci. Pollut. Res. Int. 2021, 28, 18918–18940. [Google Scholar] [CrossRef]
  21. Sharifi-Rad, J.; Sureda, A.; Tenore, G.C.; Daglia, M.; Sharifi-Rad, M.; Valussi, M.; Tundis, R.; Sharifi-Rad, M.; Loizzo, M.R.; Ademiluyi, A.O.; et al. Biological activities of essential oils: From plant chemoecology to traditional healing systems. Molecules 2017, 22, 70. [Google Scholar] [CrossRef] [PubMed]
  22. Pandey, A.K.; Kumar, P.; Singh, P.; Tripathi, N.N.; Bajpai, V.K. Essential oils: Sources of antimicrobials and food preservatives. Front. Microbiol. 2017, 7, 2161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Nazzaro, F.; Fratianni, F.; De Martino, L.; Coppola, R.; De Feo, V. Effect of essential oils on pathogenic bacteria. Pharmaceuticals 2013, 6, 1451–1474. [Google Scholar] [CrossRef] [PubMed]
  24. Khorshidian, N.; Yousefi, M.; Khanniri, E.; Mortazavian, A.M. Potential application of essential oils as antimicrobial preservatives in cheese. Innov. Food Sci. Emerg. Technol. 2018, 45, 62–72. [Google Scholar] [CrossRef]
  25. Qu, S.; Yang, K.; Chen, L.; Liu, M.; Geng, Q.; He, X.; Li, Y.; Liu, Y.; Tian, J. Cinnamaldehyde, a promising natural preservative against Aspergillus flavus. Front. Microbiol. 2019, 10, 2895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Yanakiev, S. Effects of Cinnamon (Cinnamomum spp.) in dentistry: A review. Molecules 2020, 25, 4184. [Google Scholar] [CrossRef]
  27. Didehdar, M.; Chegini, Z.; Tabaeian, S.P.; Razavi, S.; Shariati, A. Cinnamomum: The new therapeutic agents for inhibition of bacterial and fungal biofilm-associated infection. Front. Cell. Infect. Microbiol. 2022, 12, 930624. [Google Scholar] [CrossRef]
  28. Chalchat, J.C.; Valade, I. Chemical composition of leaf oils of Cinnamomum from Madagascar: C. zeylanicum Blume, C. camphora L., C. fragrans Baillon and C. angustifolium. J. Essent. Oil Res. 2000, 12, 537–540. [Google Scholar] [CrossRef]
  29. Doyle, A.A.; Stephens, J.C. A review of cinnamaldehyde and its derivatives as antibacterial agents. Fitoterapia 2019, 139, 104405. [Google Scholar] [CrossRef]
  30. Khallouki, F.; Hmamouchi, M.; Younos, C.; Soulimani, R.; Bessiere, J.M.; Essassi, E.M. Antibacterial and molluscicidal activities of the essential oil of Chrysanthemum viscidehirtum. Fitoterapia 2000, 71, 544–546. [Google Scholar] [CrossRef]
  31. Ali, N.A.M.; Mohtar, M.; Shaari, K.; Rahmanii, M.; Ali, A.M.; Jantan, I.B. Chemical composition and antimicrobial activities of the essential oils of Cinnamomum aureofulvum Gamb. J. Essent. Oil Res. 2002, 14, 135–138. [Google Scholar] [CrossRef]
  32. Salleh, W.M.N.H.; Ahmad, F.; Yen, K.H.; Zulkifli, R.M. Essential oil compositions of Malaysian Lauraceae: A mini review. Pharm. Sci. 2016, 22, 60–67. [Google Scholar] [CrossRef] [Green Version]
  33. Al-Dhubiab, B.E. Pharmaceutical applications and phytochemical profile of Cinnamomum burmannii. Pharmacogn. Rev. 2012, 6, 125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ooi, L.S.; Li, Y.; Kam, S.L.; Wang, H.; Wong, E.Y.; Ooi, V.E. Antimicrobial activities of cinnamon oil and cinnamaldehyde from the Chinese medicinal herb Cinnamomum cassia Blume. Am. J. Chin. Med. 2006, 34, 511–522. [Google Scholar] [CrossRef] [PubMed]
  35. Firmino, D.F.; Cavalcante, T.T.A.; Gomes, G.A.; Firmino, N.C.S.; Rosa, L.D.; de Carvalho, M.G.; Catunda Jr, F.E.A. Antibacterial and antibiofilm activities of Cinnamomum sp. essential oil and cinnamaldehyde: Antimicrobial activities. Sci. World J. 2018, 2018, 7405736. [Google Scholar] [CrossRef] [Green Version]
  36. Dai, D.N.; Lam, N.T.; Chuong, N.T.; Ngan, T.Q.; Truong, N.C.; Ogunwande, I.A. Essential oils of Cinnamomum curvifolium (Lour.) Nees and Cinnamomum mairei H. Lev. Am. J. Essent. Oil. Nat. Prod. 2019, 7, 11–14. [Google Scholar]
  37. Son, L.C.; Dai, D.N.; Thai, T.H.; Huyen, D.D.; Thang, T.D.; Ogunwande, I.A. The leaf essential oils of four Vietnamese species of Cinnamomum (Lauraceae). J. Essent. Oil Res. 2013, 25, 267–271. [Google Scholar] [CrossRef]
  38. Li, Y.; Tan, B.; Cen, Z.; Fu, Y.; Zhu, X.; He, H.; Kong, D.; Wu, H. The variation in essential oils composition, phenolic acids and flavonoids is correlated with changes in antioxidant activity during Cinnamomum loureirii bark growth. Arab. J. Chem. 2021, 14, 103249. [Google Scholar] [CrossRef]
  39. Li, R.; Wang, Y.; Jiang, Z.T.; Jiang, S. Chemical composition of the essential oils of Cinnamomum loureirii Nees from China obtained by hydrodistillation and microwave-assisted hydrodistillation. J. Essent. Oil Res. 2010, 22, 129–131. [Google Scholar] [CrossRef]
  40. Cheng, S.S.; Liu, J.Y.; Tsai, K.H.; Chen, W.J.; Chang, S.T. Chemical composition and mosquito larvicidal activity of essential oils from leaves of different Cinnamomum osmophloeum provenances. J. Agric. Food Chem. 2004, 52, 4395–4400. [Google Scholar] [CrossRef]
  41. Islam, R.; Khan, R.I.; Al-Reza, S.M.; Jeong, Y.T.; Song, C.H.; Khalequzzaman, M. Chemical composition and insecticidal properties of Cinnamomum aromaticum (Nees) essential oil against the stored product beetle Callosobruchus maculatus (F.). J. Sci. Food Agric. 2009, 89, 1241–1246. [Google Scholar] [CrossRef]
  42. Abdelwahab, S.I.; Zaman, F.Q.; Mariod, A.A.; Yaacob, M.; Ahmed Abdelmageed, A.H.; Khamis, S. Chemical composition, antioxidant and antibacterial properties of the essential oils of Etlingera elatior and Cinnamomum pubescens Kochummen. J. Sci. Food Agric. 2010, 90, 2682–2688. [Google Scholar] [CrossRef] [PubMed]
  43. Haider, S.Z.; Lohani, H.; Bhandari, U.; Naik, G.G.; Chauhan, N.K. Nutritional Value and Volatile Composition of Leaf and Bark of Cinnamomum tamala from Uttarakhand (India). J. Essent. Oil-Bear. Plants 2018, 21, 732–740. [Google Scholar] [CrossRef]
  44. Sharma, V.; Rao, L.J.M. An overview on chemical composition, bioactivity and processing of leaves of Cinnamomum tamala. Crit. Rev. Food Sci. Nutr. 2014, 54, 433–448. [Google Scholar] [CrossRef]
  45. Farias, A.P.P.; Monteiro, O.S.; da Silva, J.K.R.; Figueiredo, P.L.B.; Rodrigues, A.A.C.; Monteiro, I.N.; Maia, J.G.S. Chemical composition and biological activities of two chemotype-oils from Cinnamomum verum J. Presl growing in North Brazil. J. Food Sci. Technol. 2020, 57, 3176–3183. [Google Scholar] [CrossRef] [PubMed]
  46. Parihar, A.K.S.; Kulshrestha, M.; Sahu, U.; Karbhal, K.S.; Inchulkar, S.R.; Chauhan, N.S. Quality control of Dalchini (Cinnamomum zeylanicum): A review. Adv. Tradit. Med. 2021. [Google Scholar] [CrossRef]
  47. Gursale, A.; Dighe, V.; Parekh, G. Simultaneous quantitative determination of cinnamaldehyde and methyl eugenol from stem bark of Cinnamomum zeylanicum Blume using RP-HPLC. J. Chromatogr. Sci. 2010, 48, 59–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Javidnia, K.; Miri, R.; Soltani, M.; Khosravi, A.R. Constituents of the essential oil of Marrubium astracanicum Jacq. from Iran. J. Essent. Oil Res. 2007, 19, 559–561. [Google Scholar] [CrossRef]
  49. Vernin, G.; Vernin, C.; Pieribattesti, J.C.; Roque, C. Analysis of the volatile compounds of Psidium cattleianum sabine fruit from Reunion Island. J. Essent. Oil Res. 1998, 10, 353–362. [Google Scholar] [CrossRef]
  50. Bang, H.B.; Lee, Y.H.; Kim, S.C.; Sung, C.K.; Jeong, K.J. Metabolic engineering of Escherichia coli for the production of cinnamaldehyde. Microb. Cell Fact. 2016, 15, 16. [Google Scholar] [CrossRef] [Green Version]
  51. Zhao, J.; Zhang, X.; Dong, L.; Wen, Y.; Zheng, X.; Zhang, C.; Chen, R.; Zhang, Y.; Li, Y.; He, T.; et al. Cinnamaldehyde inhibits inflammation and brain damage in a mouse model of permanent cerebral ischaemia. Br. J. Pharmacol. 2015, 172, 5009–5023. [Google Scholar] [CrossRef] [PubMed]
  52. Hajinejad, M.; Ghaddaripouri, M.; Dabzadeh, M.; Forouzanfar, F.; Sahab-Negah, S. Natural cinnamaldehyde and its derivatives ameliorate neuroinflammatory pathways in neurodegenerative diseases. BioMed Res. Int. 2020, 2020, 1034325. [Google Scholar] [CrossRef] [PubMed]
  53. Hong, S.H.; Ismail, I.A.; Kang, S.M.; Han, D.C.; Kwon, B.M. Cinnamaldehydes in cancer chemotherapy. Phytother. Res. 2016, 30, 754–767. [Google Scholar] [CrossRef]
  54. Di Giacomo, S.; Mazzanti, G.; Sarpietro, M.G.; Di Sotto, A. α-Hexylcinnamaldehyde inhibits the genotoxicity of environmental pollutants in the bacterial reverse mutation assay. J. Nat. Prod. 2014, 77, 2664–2670. [Google Scholar] [CrossRef] [PubMed]
  55. Sarpietro, M.G.; Di Sotto, A.; Accolla, M.L.; Castelli, F. Interaction of α-Hexylcinnamaldehyde with a biomembrane model: A possible MDR reversal mechanism. J. Nat. Prod. 2015, 78, 1154–1159. [Google Scholar] [CrossRef]
  56. Di Giacomo, S.; Di Sotto, A.; El-Readi, M.Z.; Mazzanti, G.; Wink, M. α-Hexylcinnamaldehyde synergistically increases doxorubicin cytotoxicity towards human cancer cell lines. Anticancer Res. 2016, 36, 3347–3351. [Google Scholar] [PubMed]
  57. Welch, V.; Petticrew, M.; Petkovic, J.; Moher, D.; Waters, E.; White, H.; Tugwell, P. Extending the PRISMA statement to equity-focused systematic reviews (PRISMA-E 2012): Explanation and elaboration. Int. J. Equity Health 2015, 14, 92. [Google Scholar] [CrossRef] [Green Version]
  58. Alves, F.C.B.; Barbosa, L.N.; Andrade, B.F.M.T.; Albano, M.; Furtado, F.B.; Marques Pereira, A.F.; Rall, V.L.M.; Fernandes Júnior, A. Short communication: Inhibitory activities of the antibiotic nisin combined with phenolic compounds against Staphylococcus aureus and Listeria monocytogenes in cow milk. J. Dairy Sci. 2016, 99, 1831–1836. [Google Scholar] [CrossRef] [Green Version]
  59. Shi, C.; Zhang, X.; Zhao, X.; Meng, R.; Liu, Z.; Chen, X.; Guo, N. Synergistic interactions of nisin in combination with cinnamaldehyde against Staphylococcus aureus in pasteurized milk. Food Control 2017, 71, 10–16. [Google Scholar] [CrossRef]
  60. Palaniappan, K.; Holley, R.A. Use of natural antimicrobials to increase antibiotic susceptibility of drug resistant bacteria. Int. J. Food Microbiol. 2010, 140, 164–168. [Google Scholar] [CrossRef]
  61. Wang, S.; Kang, O.H.; Kwon, D.Y. Trans-Cinnamaldehyde exhibits synergy with conventional antibiotic against methicillin-resistant Staphylococcus aureus. Int. J. Mol. Sci. 2021, 22, 2752. [Google Scholar] [CrossRef]
  62. Dhara, L.; Tripathi, A. Cinnamaldehyde: A compound with antimicrobial and synergistic activity against ESBL-producing quinolone-resistant pathogenic Enterobacteriaceae. Eur. J. Clin. Microbiol. Infect. Dis. 2020, 39, 65–73. [Google Scholar] [CrossRef] [PubMed]
  63. Visvalingam, J.; Palaniappan, K.; Holley, R.A. In vitro enhancement of antibiotic susceptibility of drug resistant Escherichia coli by cinnamaldehyde. Food Control 2017, 79, 288–291. [Google Scholar] [CrossRef]
  64. Topa, S.H.; Palombo, E.A.; Kingshott, P.; Blackall, L.L. Activity of cinnamaldehyde on quorum sensing and biofilm susceptibility to antibiotics in Pseudomonas aeruginosa. Microorganisms 2020, 8, 455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Chadha, J.; Ravi; Singh, J.; Chhibber, S.; Harjai, K. Gentamicin augments the quorum quenching potential of cinnamaldehyde in vitro and protects Caenorhabditis elegans from Pseudomonas aeruginosa infection. Front. Cell. Infect. Microbiol. 2022, 12, 899566. [Google Scholar] [CrossRef] [PubMed]
  66. Sajjad, M.; Khan, A.; Ahmad, I. Antifungal activity of essential oils and their synergy with fluconazole against drug-resistant strains of Aspergillus fumigatus and Trichophyton rubrum. Appl. Microbiol. Biotechnol. 2011, 90, 1083–1094. [Google Scholar]
  67. Schlemmer, K.B.; Jesus, F.P.K.; Tondolo, J.S.M.; Weiblen, C.; Azevedo, M.I.; Machado, V.S.; Botton, S.A.; Alves, S.H.; Santurio, J.M. In vitro activity of carvacrol, cinnamaldehyde and thymol combined with antifungals against Malassezia pachydermatis. J. Mycol. Med. 2019, 29, 375–377. [Google Scholar] [CrossRef]
  68. Gómara, M.; Ramón-García, S. The FICI paradigm: Correcting flaws in antimicrobial in vitro synergy screens at their inception. Biochem. Pharmacol. 2019, 163, 299–307. [Google Scholar] [CrossRef]
  69. Farber, J.M.; Peterkin, P.I. Listeria monocytogenes, a food-borne pathogen. Microbiol. Rev. 1991, 55, 476–511. [Google Scholar] [CrossRef]
  70. Lomonaco, S.; Nucera, D.; Filipello, V. The evolution and epidemiology of Listeria monocytogenes in Europe and the United States. Infect. Genet. Evol. 2015, 35, 172–183. [Google Scholar] [CrossRef]
  71. Baquero, F.; Lanza, V.F.; Duval, M.; Coque, T.M. Ecogenetics of antibiotic resistance in Listeria monocytogenes. Mol. Microbiol. 2020, 113, 570–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Lister, J.L.; Horswill, A.R. Staphylococcus aureus biofilms: Recent developments in biofilm dispersal. Front. Cell. Infect. Microbiol. 2014, 4, 178. [Google Scholar] [CrossRef] [PubMed]
  73. Jenul, C.; Horswill, A.R. Regulation of Staphylococcus aureus virulence. Microbiol Spectr. 2019, 7, 1–21. [Google Scholar] [CrossRef]
  74. Lakhundi, S.; Zhang, K. Methicillin-Resistant Staphylococcus aureus: Molecular characterization, evolution, and epidemiology. Clin. Microbiol. Rev. 2018, 31, e00020-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Fiedler, T.; Köller, T.; Kreikemeyer, B. Streptococcus pyogenes biofilms formation, biology, and clinical relevance. Front. Cell. Infect. Microbiol. 2015, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  76. Rood, I.G.H.; Li, Q. Review: Molecular detection of extended spectrum-β-lactamase- and carbapenemase-producing Enterobacteriaceae in a clinical setting. Diagn. Microbiol. Infect. Dis. 2017, 89, 245–250. [Google Scholar] [CrossRef]
  77. Jang, J.; Hur, H.G.; Sadowsky, M.J.; Byappanahalli, M.N.; Yan, T.; Ishii, S. Environmental Escherichia coli: Ecology and public health implications-a review. J. Appl. Microbiol. 2017, 123, 570–581. [Google Scholar] [CrossRef] [Green Version]
  78. Herridge, W.P.; Shibu, P.; O’Shea, J.; Brook, T.C.; Hoyles, L. Bacteriophages of Klebsiella spp., their diversity and potential therapeutic uses. J. Med. Microbiol. 2020, 69, 176–194. [Google Scholar]
  79. Mielko, K.A.; Jabłoński, S.J.; Milczewska, J.; Sands, D.; Łukaszewicz, M.; Młynarz, P. Metabolomic studies of Pseudomonas aeruginosa. World J. Microbiol. Biotechnol. 2019, 35, 178. [Google Scholar] [CrossRef] [Green Version]
  80. Yang, C.; Li, H.; Zhang, T.; Chu, Y.; Zuo, J.; Chen, D. Study on antibiotic susceptibility of Salmonella typhimurium L forms to the third and fourth generation cephalosporins. Sci Rep. 2020, 10, 3042. [Google Scholar] [CrossRef] [Green Version]
  81. Tetard, A.; Zedet, A.; Girard, C.; Plésiat, P.; Llanesa, C. Cinnamaldehyde induces expression of efflux pumps and multidrug resistance in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2019, 63, e01081-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Tetard, A.; Gaillot, S.; Dubois, E.; Aarras, S.; Valot, B.; Phan, G.; Plésiat, P.; Llanes, C. Exposure of Pseudomonas aeruginosa to Cinnamaldehyde Selects Multidrug Resistant Mutants. Antibiotics 2022, 11, 1790. [Google Scholar] [CrossRef] [PubMed]
  83. Delcour, A.H. Outer membrane permeability and antibiotic resistance. Biochim. Biophys. Acta. 2009, 1794, 808–816. [Google Scholar] [CrossRef] [PubMed]
  84. Breijyeh, Z.; Jubeh, B.; Karaman, R. Resistance of Gram-negative bacteria to current antibacterial agents and approaches to resolve it. Molecules 2020, 25, 1340. [Google Scholar] [CrossRef] [Green Version]
  85. Rabin, N.; Zheng, Y.; Opoku-Temeng, C.; Du, Y.; Bonsu, E.; Sintim, O.H. Biofilm formation mechanisms and targets for developing antibiofilm agents. Future Med. Chem. 2015, 7, 493–512. [Google Scholar] [CrossRef]
  86. Roy, R.; Tiwari, M.; Donelli, G.; Tiwari, V. Strategies for combating bacterial biofilms: A focus on anti-biofilm agents and their mechanisms of action. Virulence 2018, 9, 522–554. [Google Scholar] [CrossRef] [Green Version]
  87. Epstein, A.K.; Wong, T.S.; Belisle, R.A.; Boggs, E.M.; Aizenberg, J. Liquid-infused structured surfaces with exceptional anti-biofouling performance. Proc. Natl. Acad. Sci. USA 2012, 109, 13182–13187. [Google Scholar] [CrossRef] [Green Version]
  88. Solano, C.; Echeverz, M.; Lasa, I. Biofilm dispersion and quorum sensing. Curr. Opin. Microbiol. 2014, 18, 96–104. [Google Scholar] [CrossRef] [Green Version]
  89. Yang, D.; Wang, H.; Yuan, H.; Li, S. Quantitative structure activity relationship of cinnamaldehyde compounds against wood decaying fungi. Molecules 2016, 21, 1563. [Google Scholar] [CrossRef] [Green Version]
  90. Thirapanmethee, K.; Kanathum, P.; Khuntayaporn, P.; Huayhongthong, S.; Surassmo, S.; Chomnawang, M.T. Cinnamaldehyde: A plant-derived antimicrobial for overcoming multidrug-resistant Acinetobacter baumannii infection. Eur. J. Integr. Med. 2021, 48, 101376. [Google Scholar] [CrossRef]
  91. Vasconcelos, N.G.; Croda, J.; Simionatto, S. Antibacterial mechanisms of cinnamon and its constituents: A review. Microb. Pathog. 2018, 120, 198–203. [Google Scholar] [CrossRef] [PubMed]
  92. Oussalah, M.; Caillet, S.; Lacroix, M. Mechanism of action of Spanish oregano, Chinese cinnamon, and savory essential oils against cell membranes and walls of Escherichia coli O157:H7 and Listeria monocytogenes. J. Food Prot. 2006, 69, 1046–1055. [Google Scholar] [CrossRef] [PubMed]
  93. Nowotarska, S.W.; Nowotarski, K.; Grant, I.R.; Elliott, C.T.; Friedman, M.; Situ, C. Mechanisms of antimicrobial action of cinnamon and oregano oils, cinnamaldehyde, carvacrol, 2,5-dihydroxybenzaldehyde, and 2-hydroxy-5-methoxybenzaldehyde against Mycobacterium avium subsp. paratuberculosis (Map). Foods 2017, 6, 72. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of cinnamaldehyde.
Figure 1. Chemical structure of cinnamaldehyde.
Antibiotics 12 00254 g001
Figure 2. Pharmacological properties of cinnamaldehyde.
Figure 2. Pharmacological properties of cinnamaldehyde.
Antibiotics 12 00254 g002
Figure 3. Study selection by PRISMA flow diagram about the ability of cinnamaldehyde to synergize antimicrobial drugs against superbugs.
Figure 3. Study selection by PRISMA flow diagram about the ability of cinnamaldehyde to synergize antimicrobial drugs against superbugs.
Antibiotics 12 00254 g003
Figure 4. Possible mechanisms underlying the bacterial sensitizing properties of trans-cinnamaldehyde.
Figure 4. Possible mechanisms underlying the bacterial sensitizing properties of trans-cinnamaldehyde.
Antibiotics 12 00254 g004
Table 1. Natural occurrence of trans-cinnamaldehyde in plant essential oils.
Table 1. Natural occurrence of trans-cinnamaldehyde in plant essential oils.
Plant Species/FamilyPlant Parttrans-Cinnamaldehyde
(%)
Ref.
Chrysanthemum viscidehirtum Schott Tell/LauraceaeLeaf2.1[30]
Aerial parts0.7
Cinnamomum angustifolium Lukman/LauraceaeLeaf and bark0.2[28]
Cinnamomum aureofulvum Gamble/LauraceaeBark46.6[31,32]
Cinnamomum burmannii Nees & T. Nees/LauraceaeLeaf45–62[33]
Bark17–32
Cinnamomum cassia Nees/LauraceaeBark85[28,34,35]
Cinnamomum curvifolium Nees/LauraceaeLeaf8.9[36]
Steam bark1.2
Cinnamomum durifolium Kosterm/LamiaceaeAerial parts0.6[37]
Cinnamomum loureirii Nees/LauraceaeBark50.2–92.9[38,39]
Cinnamomum mairei H. Léveillé/LauraceaeLeaf1.9[36]
Steam bark6.5
Cinnamomum osmophloeum Kaneh/LauraceaeLeaf79.8[40,41]
Cinnamomum pubescens Kochummen/LauraceaeLeaf56.1[42]
Cinnamomum sericans Hance/LauraceaeLeaf0.6[37]
Cinnamomum tamala Nees Eberm/LauraceaeLeaf
Bark
68.7–79.4
64.8
[43,44]
[43]
Cinnamomum verum J. Presl/LauraceaeLeaf0.6[28,45]
Bark89.3
Cinnamomum zeylanicum Blume/LauraceaeBark44.2–68.7[32,35,46,47]
Leaf1–5[46]
Marrubium astracanicum Jacq./LauraceaeAerial parts2.2[48]
Psidium cattleianum Sabine/LamiaceaeAerial parts2.2[49]
Fruit0.6
Teucrium persicum Boiss/MyrtaceaeAerial parts0.4[48]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Usai, F.; Di Sotto, A. trans-Cinnamaldehyde as a Novel Candidate to Overcome Bacterial Resistance: An Overview of In Vitro Studies. Antibiotics 2023, 12, 254. https://doi.org/10.3390/antibiotics12020254

AMA Style

Usai F, Di Sotto A. trans-Cinnamaldehyde as a Novel Candidate to Overcome Bacterial Resistance: An Overview of In Vitro Studies. Antibiotics. 2023; 12(2):254. https://doi.org/10.3390/antibiotics12020254

Chicago/Turabian Style

Usai, Federica, and Antonella Di Sotto. 2023. "trans-Cinnamaldehyde as a Novel Candidate to Overcome Bacterial Resistance: An Overview of In Vitro Studies" Antibiotics 12, no. 2: 254. https://doi.org/10.3390/antibiotics12020254

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop