Next Article in Journal
The Potential Protective Role of Naringenin against Dasatinib-Induced Hepatotoxicity
Next Article in Special Issue
The Safety of Bacteriophages in Treatment of Diseases Caused by Multidrug-Resistant Bacteria
Previous Article in Journal
Current Management and Future Perspectives in the Treatment of Lp(a) with a Focus on the Prevention of Cardiovascular Diseases
Previous Article in Special Issue
Mechanisms of Linezolid Resistance in Mycobacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CRISPR-Based Gene Editing in Acinetobacter baumannii to Combat Antimicrobial Resistance

by
Muhammad Junaid
1,2,
Krit Thirapanmethee
1,2,
Piyatip Khuntayaporn
1,2 and
Mullika Traidej Chomnawang
1,2,*
1
Department of Microbiology, Faculty of Pharmacy, Mahidol University, Bangkok 10400, Thailand
2
Antimicrobial Resistance Interdisciplinary Group (AmRIG), Faculty of Pharmacy, Mahidol University, Bangkok 10400, Thailand
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(7), 920; https://doi.org/10.3390/ph16070920
Submission received: 25 May 2023 / Revised: 19 June 2023 / Accepted: 20 June 2023 / Published: 23 June 2023
(This article belongs to the Special Issue Multidrug Resistance in Bacteria and New Therapeutic Options)

Abstract

:
Antimicrobial resistance (AMR) poses a significant threat to the health, social, environment, and economic sectors on a global scale and requires serious attention to addressing this issue. Acinetobacter baumannii was given top priority among infectious bacteria because of its extensive resistance to nearly all antibiotic classes and treatment options. Carbapenem-resistant A. baumannii is classified as one of the critical-priority pathogens on the World Health Organization (WHO) priority list of antibiotic-resistant bacteria for effective drug development. Although available genetic manipulation approaches are successful in A. baumannii laboratory strains, they are limited when employed on newly acquired clinical strains since such strains have higher levels of AMR than those used to select them for genetic manipulation. Recently, the CRISPR-Cas (Clustered regularly interspaced short palindromic repeats/CRISPR-associated protein) system has emerged as one of the most effective, efficient, and precise methods of genome editing and offers target-specific gene editing of AMR genes in a specific bacterial strain. CRISPR-based genome editing has been successfully applied in various bacterial strains to combat AMR; however, this strategy has not yet been extensively explored in A. baumannii. This review provides detailed insight into the progress, current scenario, and future potential of CRISPR-Cas usage for AMR-related gene manipulation in A. baumannii.

1. Introduction

The behavioral change in bacterial pathogenicity leading to AMR is one of the major constraints hindering several public health policies globally. It has been estimated that there might be a global economic loss of USD 100 trillion by the year 2050 if this threatening issue of antibiotic resistance continues [1]. Although there are several reasons for antibiotic resistance, the major one is due to a bottleneck in innovative research with a focus on the exploitation of the diversity of antibiotics against Gram-negative bacteria, while the existing antibiotics are in clinical trials mostly under phase II or phase III. This lack of research is an alarming situation for research scientists along with economic loss. The lives of millions of people are projected to be vulnerable to death by 2050 due to the emerging situation of AMR [2]. Other serious health issues related to AMR include the outbreak of different infectious diseases, the risk of common infectious diseases such as immunosuppression, intubation, catheterization, and other such procedures related to antibiotics [3]. In view of the above situation, it is very important to revolutionize therapeutic strategies to prevent AMR bacterial infections [4,5]. It is critical to focus on the most commonly occurring pathogenic bacteria, which have acquired AMR for most antibiotics.
Acinetobacter baumannii is among the most widely distributed bacteria that can adapt to various environments. However, this pathogenic bacterium has been found to be resistant to all classes of antibiotics, and the WHO has emphasized the necessity of classification and research on this particular bacterium due to its high environmental adaptability [6]. Additionally, A. baumannii is highly resistant to hot and humid ultraviolet rays and chemical disinfectants and can survive on dry-surface objects for more than 25 days. A. baumannii can be traced to healthcare providers and to various dry surfaces, which can eventually be enablers of drug resistance [7].
Investigating different mechanisms for understanding drug resistance should be a high priority for research scientists. However, genetic manipulation is considered one of the robust approaches for studying such mechanisms in A. baumannii [8]. In addition, genetic manipulation of A. baumannii is applicable to laboratory strains ATCC17978 and ATCC19606, while it becomes less efficient for strains isolated from hospitals and patients because of their higher genetic diversity and increased AMR [9]. Various therapeutic tools have been developed recently, including peptides, bacteriophage therapies, antibodies, bacteriocins, and antibacterial or anti-virulence substances that are based on nucleic acids [10]. Along with the above genetic tools, CRISPR-Cas based gene editing offers an exciting opportunity for specific manipulation of the targeted genes responsible for antibiotic resistance in a specific bacterial strain. Several studies have been reported on the utilization of the CRISPR-Cas system for understanding the mechanism of AMR. This review will therefore focus on the previous studies in which CRISPR-Cas-based gene editing has played a vital role in overcoming the pressing issue of AMR and different bacterial strains including A. baumannii.

2. Clinical Significance of A. baumannii

A. baumannii is an opportunistic pathogen responsible for many diseases in humans, including pneumonia, skin infections, wound-borne infections, urinary tract infections, soft tissue infections, meningitis, and bacteremia. Among these infections, bacteremia and pneumonia are the most commonly reported infections, which have significant rates of morbidity and mortality [11]. One trait of some Acinetobacter strains that facilitates transmission through fomite contamination in hospitals is their ability to endure environmental desiccation for weeks. A. baumannii is responsible for nearly 80% of ICU-acquired pneumonia in many regions including Asia, the Middle East, and Latin America [12]. Globally, pneumonia is responsible for nearly 64% of mortalities in tropical regions and most frequently affects those with diabetes, lung disorders, and smoking or alcohol addiction [13]. Yet, it is still uncertain whether the main cause of infection is host factors or bacterial virulence factors [11]. In Europe and the U.S., ICU-acquired infections implicated by A. baumannii range from 2–10% and 2.5%, respectively, with a more than 50% resistance to various antibiotics [14]. The ranges for ICU-acquired infections by A. baumannii are 2.1% for skin/soft tissue infections, 1.6% for urinary tract infections, ~33% for wound infections, and 34.1% for bacteremia, with a 10–47% mortality rate [15]. The death toll due to A. baumannii-acquired bacteremia has been reported at 37–52%. Although the infestation rate for meningitis is not significantly high, it has a greater motility rate of nearly 70% [16]. It has been observed that in many Asian countries, 51% and 82% of nosocomial infections are caused by drug-resistant Acinetobacter isolates, respectively, where more than 80% infection rates were detected in India, Malaysia, and Thailand and ~59% in China [17]. The overall death rate by A. baumannii infections ranges from 30–43.4% in Thailand [18]. The following factors primarily contribute to acquiring A. baumannii infections: prolonged admission in hospitals and especially in intensive care units (ICUs), aging, multimorbidity, weak immunity, antibiotic usage history, injuries and burns, surgery, prematurity in newborns, use of contaminated equipment, mechanical ventilation, and permanent usage of catheters [13,19]. Moreover, natural or manmade disasters including wars, earthquakes, and tsunamis also contribute to A. baumannii-acquired infections, especially skin and soft tissue infections [20,21].

3. Antimicrobial Resistance Mechanisms in A. baumannii

Bacteria have evolved to develop several mechanisms to neutralize the effect of antibiotics due to the introduction of a large number of new antibiotics and their consequent consumption around the world [22]. Antibiotic resistance mechanisms in bacteria are broadly divided into three categories, namely intrinsic, acquired, and adaptive resistance [23]. The bacterial genome is the sole determinant of intrinsic resistance, which is typically acquired through drug inactivation, decreased membrane permeability for the medication, or enhanced efflux of the antibiotic, ultimately restricting access to the target [24]. Moraxella catarrhalis, Salmonella typhimurium, Klebsiella pneumoniae, Pseudomonas aeruginosa PAO1, and Enterobacter cloacae ATCC 13047 are examples of bacteria showing intrinsic resistance [25,26].
Genetic mutations or post-translational modifications help bacteria to acquire resistance. They enable bacteria to become resistant to a certain type of antibiotic to which they were previously susceptible. Genetic manipulation processes (transformation, conjugation, or transduction) can lead to the acquisition of certain genes that can develop resistant phenotypes in bacteria against antibiotics by mutation and selection [27]. Strains of Escherichia coli, P. aeruginosa, A. baumannii, K. pneumoniae, and Vibrio cholerae are common examples of bacteria with acquired resistance [28,29,30,31]. Bacteria can temporarily avoid antibiotic effects under adaptive resistance, as this resistance mechanism is produced under the effect of exogenous stimuli. The action of environmental stimuli results in transient genetic effects that hydrolyze or modify the antibiotic and ultimately result in the inactivation of its activity [28]. Inactivation or removal of these exogenous environmental signals reverses the resistance mechanism. Various environmental factors include changes in pH, organic compounds (carbon and polyamines), ineffective antibiotic doses, and anaerobiosis [32]. This type of resistance is commonly presented by Salmonella enterica, S. enteritidis, E. coli, and P. aeruginosa [33,34].
In A. baumannii, many strains present resistance to most of the existing antibiotics (Table 1). The genetic plasticity of A. baumannii enables it to produce higher genetic mutations and genetic rearrangements, along with the flexibility to integrate external elements into its genome through mobile genetic elements. In particular, insertion sequences are viewed as one of the fundamental mechanisms influencing how bacterial genomes and, ultimately, evolution are shaped. To resist different kinds of antibiotics, A. baumannii can use a variety of resistance mechanisms [35,36,37]. Yet, resistance to a specific antibiotic family can be produced due to the combination of various distinct resistance mechanisms. Moreover, A. baumannii has the ability to produce biofilms, which enables it to survive longer on medical equipment such as ventilators in ICUs. Although the connection between the development of biofilms and antibiotic resistance is not yet clear [38,39], the most common mechanisms in A. baumannii conferring resistance to multiple antibiotic families are plasmid conjugation, transposon acquisition, or integron mobilization. The functional insertion sequences are critical in enhancing AMR and gene plasticity in A. baumannii [40,41].
A. baumannii has been classified into multidrug-resistant (MDR), extensively drug- resistant (XDR), and pan drug-resistant (PDR) phenotypes depending on its ability to respond to various antibiotics [37]. The isolate is non-susceptible to at least one agent in at least three antimicrobial categories is referred to be MDR. A. baumannii isolates that pose non-susceptible to at least one agent in all but two or fewer antimicrobial categories are classified as XDR phenotypes, whereas the PDR phenotype is an isolate with non-susceptibility to all agents in all antimicrobial categories. Antimicrobial categories for Acinetobacter spp. include antipseudomonal carbapenems, penicillins with β-lactamase inhibitors, antipseudomonal penicillins with β-lactamase inhibitors, antipseudomonal fluoroquinolones, extended-spectrum cephalosporins, aminoglycosides, folate pathway inhibitors, polymyxins, and tetracyclines [42]. The use of carbapenems to treat MDR A. baumannii is no longer effective [43] and has been replaced with polymyxins to treat MDR A. baumannii infections; however, these drugs can have nephrotoxicity and neurotoxicity [44]. Aminoglycoside resistance genes, β-lactamases, and methyltransferases have contributed to the development of MDR phenotypes in A. baumannii [45]. Various mechanisms are adapted by A. baumannii to confer resistance against antibiotics. Efflux pumps or reduced permeability in bacteria hinder the access of antibiotics to the target site in the cell. In some cases, bacteria can inactivate the antibiotic using enzymes that can hydrolyze and modify the antibiotic’s structure. Moreover, genetic mutations or modifications can help bacteria modify the specific target sites for antibiotics and thus attain resistance [46,47,48]. These mechanisms enable A. baumannii to resist various antibiotic families, including β-lactams, aminoglycosides, tetracyclines, erythromycin, macrolides, polymyxins, chloramphenicol, fluoroquinolones, and trimethoprim [37,47,48].
A. baumannii strains present resistance to various antibiotics [49,50,51,52,53]. Among these, β-lactamases are the most common resistance mechanisms that are divided into four classes (A–D). Classes A, C, and D β-lactamases are the active-site serine β-lactamases, while class B has zinc or any other heavy metal-dependent or metallo-β-lactamases (MBLs) in the catalytic site. The β-lactams family includes penicillin, carbapenems, cephalosporins, cephamycins, and monobactams [54]. The strains having class A β-lactamase enzymes show resistance to all penicillin and cephalosporins but are less effective against cephamycin and carbapenems. This class is the most common source of β-lactam resistance. The genes associated with class A β-lactamase resistance in A. baumannii include blaPER-(1, 2, and 7), blaSHV-(5, 12, and33), blaGES-(11 and 14), blaTEM-(1 and 92), blaCARB-10, blaCTX-M-(2 and 15), blaSCO-1, and blaVEB-1 and also contain the Klebsiella pneumoniae carbapenemase (KPC) enzymes, including KPC-(2, 3, and 5) [35,37,55,56,57,58]. Class B or MBLs show resistance to almost all β-lactam antibiotics, including carbapenem, but cannot hydrolyze monobactams [59,60]. Globally, a variety of MBLs have been identified in A. baumannii [47,61]; however, detection of MBLs by conventional methods is not very effective, so there is a need to apply more molecular strategies, including next-generation sequencing (NGS), to detect MBLs [62,63,64].
All A. baumannii strains contain chromosomally encoded non-inducible cephalosporinases, which form class C β-lactamases. Class C is also recognized as Acinetobacter-derived cephalosporinase (ADC) induced by the blaADC gene (formerly the blaAmpC gene) [65,66] and present resistance to penicillin, cefotenan, cephamycins, cefoxitin, and cephalosporins [48,59]. Class D β-lactamases can hydrolyze carbapenems and hence are known as carbapenem-hydrolyzing class D β-lactamase (CHLD) or oxacillinases (OXA). In A. baumannii, these lactamases can deactivate all β-lactams and provide resistance against carbapenem. The overexpression of OXA genes (chromosomal or plasmid encoded) enables A. baumannii to pose resistance against carbapenems [67]. Recent studies have shown the presence of various blaOXA enzymes (OXA-23, OXA-24, OXA-40, OXA-51, OXA-58, OXA-143, and OXA-235) in A. baumannii strains [68,69,70,71,72,73,74].
Enzymatic activity weakening the binding capacity of antibiotics, leading to changes in ribosomal target sites, efflux pumps, or permeability, provides resistance against aminoglycosides such as tobramycin, amikacin, and gentamicin in A. baumannii [75]. The aminoglycoside-resistant genes are specifically found in transposons, plasmids, chromosomal genomic islands, integrative conjugative elements, and chromosomes [76]. In A. baumannii strains, the role of efflux pumps involved in resistance against tetracyclines and tigecycline has been demonstrated by different studies [77,78,79,80,81]. Similarly, mutations in the gyrA gene are the main cause of resistance to fluroquinolones in A. baumannii; however, efflux pumps are also responsible for resistance to other groups of fluoroquinolones, including norfloxacin and ciprofloxacin [82]. A. baumannii has shown 50–73% resistance to fluoroquinolones, and in some regions, the resistance is up to 75–98% [83,84]. In addition, A. baumannii showed resistance to different macrolides, including erythromycin, oleandomycin, azithromycin, and clarithromycin [85]. It is resistant to polymyxins, especially exhibiting colistin resistance, which has significantly increased over time [86].
Table 1. Antimicrobial resistance mechanisms in A. baumannii against various antimicrobial categories.
Table 1. Antimicrobial resistance mechanisms in A. baumannii against various antimicrobial categories.
Antimicrobial CategoriesResistance
Mechanism
Class/Family/ActivityEnzymes/Genes/ProteinsReferences
β-lactamsβ-lactamasesClass AExtended-Spectrum β-lactamases
blaCARB-(4, 10)[76,87]
blaCTX-M-(2, 15, 43, 55, 115)[88,89,90]
blaPER-(1, 2, 3, 7)[91,92,93]
blaSHV-(5, 12, 33)[57,94,95]
blaVEB-(1, 3, 7)[96,97,98]
blaTEM-(1, 92, 116)[99,100,101]
Narrow-Spectrum β-lactamases
blaSCO-1[102]
Carbapenem-Hydrolyzing β-lactamases
blaGES-(1, 5, 11, 12, 14, 15)[103,104,105]
blaKPC-(2, 3, 5, 10)[55,56,106]
Class BblaFIM-1[107]
blaGIM-1[108]
blaIMP-(1, 2, 4, 5, 6, 8, 10, 11, 14, 16, 19, 24)[103,109,110]
blaNDM-(1, 2, 3)[111,112,113]
blaSIM-1[114]
blaSPM-1[115]
blaVIM-(1, 2, 3, 4, 6, 11)[116,117]
Class CblaAmpC-(69, 70, 71)[118,119,120]
blaADC-(11, 25, 30, 56, 76, 152, 196, 222)[121]
Class D (Oxacillinases or OXA family)blaOXA-(21, 37, 128) (Narrow spectrum)[122]
blaOXA-23 group, including blaOXA-(27, 49, 73, 102, 103, 105, 133, 134, 146, 165, 171,225, 239)[104,123,124,125]
blaOXA-24 group, including blaOXA-(25, 26, 27, 40, 72, 139, 160, 207, 40/24)[90,126,127]
blaOXA-48 group, including blaOXA-(48b, 162, 163, 181, 199, 204, 232, 247)[128,129]
blaOXA-51 group, including blaOXA-(64, 65, 66, 71, 75, 80, 82, 84, 86, 95, 98, 100, 104, 106, 113, 115)[69,130,131]
blaOXA-58 group, including blaOXA-(58, 96, 97, 164)[9,41,132]
blaOXA-143 group, including blaOXA-(143, 182, 231)[133,134,135]
blaOXA-235 group, including blaOXA-(235, 255)[136,137]
AminoglycosidesOveractive efflux pumpsResistance nodulation division (RND)AdeABC, AdeFGH, AdeIJK, AdeR, AdeS[47,69,138]
Reduced membrane permeability OmpA, Omp25, Omp33, OprB, OprC, OprD, OmpW, CarO[139,140]
Genetic mutationsPenicillin-binding protein (PBP)PBP2, PBP3, PBP6b, ftsI[47,141]
Overactive efflux pumpsRNDAdeABC, AmvA, AdeE, AdeR[139,142,143]
Genetic mutations16sRNA methylase genesarmA, rmt-(A, B, B1, C, D, E)[139,141,143]
Enzymatic inactivationAminoglycoside modifying enzymes (AME)AAC, APH, ANT[144,145,146,147]
TetracyclinesRibosomal protectionDissociation of tetracycline from ribosomeTet-M, Tet-O[148]
Overactive efflux pumpsRND and Tet pumpTet-(A, B, C, D, G, H, M, X), AdeABC, AdeIJK[47,80,144]
PolymyxinsGenetic mutationsLipid A, biotinMCR-(1, 4, 4.3), PmrCAB, Lps-(B, D), Lpx-(A, C, D), pldA, PheS, ZndP[47,149,150,151,152]
MacrolidesOveractive efflux pumpsSmall multidrug resistance (SMR) pumpAbeS[153]
FluoroquinolonesOveractive efflux pumpsRND and multidrug and toxic compound extrusionAdeABC, AbeM[154,155]
Genetic mutationsDNA gyrase, quinolone resistance pentapeptide repeat proteinGyrA, ParC, AAC, Qnr-(A, B, B19, S)[83,84]
Retrieved and modified from [37,156].

4. Latest Strategies to Combat Antimicrobial Resistance in Bacteria

Recently, various strategies and genetic tools have been developed against AMR bacteria. These tools are utilized for the genetic screening and manipulation of bacterial genomes for AMR. The use of antibiotic markers (non-clinical and non-antibiotic) [152,157], antimicrobial peptides [158], transposon mutagenesis and screening (for high-throughput genetic screening) [159,160,161], anti-virulence compounds [162], suicide plasmids and linear DNA fragments (for gene deletion) [152,163,164], homologous recombination and complementation [165,166], phage therapy [166,167,168], nanoparticles [169,170,171], enzymes [54,172], drug repurposing [173,174,175], and vaccines [176,177,178] are common approaches to overcome AMR in bacteria.
RNA-based strategies such as RNA silencing and interference, antisense oligonucleotides, and steric-blocking oligonucleotides are also proven to be effective against AMR bacteria. Resistance genes can be eliminated by enzymatically targeting the mRNA with these oligonucleotides [179]. Translation can be ceased in bacteria by RNA silencing, which is a built-in process in many bacteria. The cis- and trans-regions bind to the complementary regulatory regions present on a single mRNA strand (also known as antisense sequences) to halt the translation of certain genes [180]. Furthermore, this technique utilizes antisense RNA sequences to monitor resistant genes and mutations by creating antisense oligonucleotides that are continually redesigned to ensure that resistance is not encountered [181,182]. It also helps to identify and knock-down the AMR genes and to detect the mode of action of novel antibiotics [181]. Two important drawbacks of the use of RNA as a therapeutic measure are the limited intracellular absorptions and chemistry-dependent toxicities [179].
Genome-editing tools can be utilized to combat AMR in bacteria. These tools use restriction enzymes to target and cut a specific DNA sequence; for example, restriction nucleases, zinc finger nucleases (ZFN), and transcription-activator-like effector nucleases (TALENS), are the initially developed genome-editing technologies that can be specifically engineered for target-specific DNA cleavage. These enzymes produce target-specific double-strand breaks (DSBs) in the genome and thus help to obtain knock-down, knock-in, and/or knock-out mutants. The cleavage domain present in these enzymes can bind to a customized DNA binding domain, which allows the DNA cleavage at the targeted binding site. Several studies have shown the utilization of both these enzymes in genome editing; yet, they are costly, laborious, time-consuming, and error-prone due to higher ratios of off-target mutations [183]. Moreover, TALENs are much larger in size than ZFNs, making them difficult to deliver and express in the host cells [184,185].
Many of the above-discussed genetic manipulation approaches are effective in genome editing, providing desirable gene deletions and mutations for antimicrobials; however, the limitations associated with these approaches necessitate the development and discoverer of novel alternative strategies with a more precise and target-oriented approach for genetic manipulation [186]. Recently, clustered regularly interspaced short palindromic repeats/CRISPR-associated protein (CRISPR-Cas) systems have become an effective tool to target drug-resistant bacteria and genes by targeting the specific genome sequences. CRISPR-Cas systems are being utilized to develop precise and effective antimicrobials for various infections; however, CRISPR-based approaches for AMR are still needed to be explored on a wider scale.

5. Clustered Regularly Interspaced Palindromic Repeats/CRISPR-Associated Protein (CRISPR-Cas) System

CRISPR-Cas constitutes an adaptive immune system as present in bacteria and archaea, which provides them with an effective shield against various viral and bacteriophagal attacks [187]. Discovered by Japanese scientists in 1987 and obtaining its name, CRIPSR, in 1990 by Francisco Mojica, the biological function and ability of these repetitive palindromic DNA sequences were long unknown [188]. Initially, in 2007, CRISPR was experimentally attributed as a crucial component of a prokaryote’s adaptive immune system in the fight against viruses [189]. Later, in 2012, Doudna and Charpentier discovered the role of CRISPR-Cas9 in DNA editing in a target-specific manner by using the appropriate template sequence. Moreover, they explained the processing mechanism of CRISPR-derived RNAs (crRNA) under the effect of transactivating CRISPR RNAs (tracrRNA) [190,191]. Since then, CRISPR-Cas has emerged as the most effective, precise, efficient, and powerful approach for editing the genome in all living cells and is used in a wide range of practical fields [190,192].
Three fundamental steps of the CRSIPR-Cas system, which are adaptation (spacer acquisition), crRNA synthesis (expression), and target interference, play a vital role in bacterial defense against viral attacks (Figure 1). The cas gene encodes the Cas protein or nuclease protein, which is responsible for the cleavage and destruction of foreign viral DNA [193]. Its features, namely that it is an affordable, rapid, accurate, effective, and efficient method of gene editing, make CRISPR-Cas most researched genome-editing tool in recent years, where it has demonstrated the ability to get rid of bacterial infections [194,195], correct genetic flaws [196,197], and eradicate dangerous infectious viruses [198,199].

6. Classification of CRISPR-Cas System

Two main classes of the CRISPR-Cas system, Class 1 and Class 2, are distinguished by their signature genes and the arrangement of the CRISPR loci [200]. Both classes are further divided into 6 types (I–VI) and 33 subtypes. Class 1 comprises three types (I, III, and IV) and sixteen subtypes, whereas Class 2 includes three types (II, V, and VI) and seventeen subtypes.
In Class 1, when complexed with the crRNA, multi-subunit Cas effector proteins (Cas5, Cas7, Cas8, and SS) carry out the processing and interference mechanisms (where type I, III, and IV use Cas3, Cas10, and Csf1, respectively) to degrade the foreign DNA [201]. Nearly 60% of the bacterial population contains the Type I CRISPR-Cas system, which carries a multi-subunit CRISPR-associated complex for antiviral defense (Cascade) [202]. Seven subtypes (A to G) exist for Type I, among which the Type I-F CRISPR-Cas system is most prevalent among bacteria [200,203]. In the case of Class 2, a single enzymatic crRNA-binding protein is responsible for the identification and cleavage of the target sequence [201,204,205]. These multidomain crRNA-binding effector proteins include Cas9, Cas12a (Cpf1), and Cas13, which are responsible for the interference and processing mechanisms in type II, V, and VI, respectively [206]. CRISPR-Cas9 is isolated from the bacterium Streptococcus pyogenes, and in Type II CRISPR-Cas systems, it has received substantial attention from researchers for gene editing, as it is the easiest and most effective, versatile, and specific system [200,204].

7. Importance of CRISPR-Cas System

7.1. Role of CRISPR-Cas System against AMR Bacteria

Genome-editing technologies have been completely transformed by CRISPR-Cas due to its ability to be engineered to target almost any sequence of interest. As well as being repurposed for potential antimicrobials, CRISPR-Cas systems are also being used to reduce the level of undesirable genetic traits in bacteria [207,208]. Moreover, the ability of nucleic acid destruction by RNA enables the CRISPR-Cas system to develop next-generation antimicrobial mechanisms against infectious diseases, particularly those caused by AMR pathogens [209,210,211].
The CRISPR-Cas system based on the target gene locations can be employed as an antibacterial agent in two different ways, namely a pathogen-focused approach and a gene-focused approach [212]. In a pathogen-focused approach, particular bacterial chromosomal regions are targeted, where pathogen strain identification and bacterial cell death are achieved. On the other hand, targeting the plasmids that carry the AMR genes is part of the gene-focused approach. This method eliminates the plasmid and makes the bacterium susceptible to antibiotics [213,214]. The pathogen-focused approach is applied to cure distinct infections and can kill target-specific bacterial strains in a mixed culture, whereas the role of gene-focused approaches is still vague. These can be applied to the treatment of bacterial infections and can generally lessen the prevalence of the AMR gene in bacteria [206].
The CRISPR-Cas system, in contrast to conventional antimicrobials, operates in a target-specific manner, allowing the differentiation between friendly and harmful bacteria. Guide RNAs can be designed to target important virulence genes, antibiotic resistance, or pathogenicity [215,216]. According to the experimental designs and objectives, targeting effects can range from cell death to growth inhibition of the specific bacteria and at the genetic level can also result in gene deletions, transcriptional inhibition, and loss of antibiotic-resistant plasmids [217,218,219]. The CRISPR-Cas system utilizes the following three general strategies to address AMR:
  • It can be used in target-specific cleavage of infection-causing genes, deploying the desired bacteria while leaving the host’s microbiome unaffected [220,221]. For example, chromosomal genes for cell division and metabolism were removed from the mixed cell cultures of E. coli and S. enterica strains using the Type I CRISPR-Cas system [222];
  • It can be applied to cleave drug-resistant genes by killing the pathogenic bacteria but not affecting wild-types [222,223]. Bikard et al. applied the RNA-guided nuclease Cas9 against the virulence genes in Staphylococcus aureus, which resulted in the specific killing of virulent strain without affecting avirulent staphylococci [224];
  • It can be engineered to modify or silence resistance genes, causing bacterial mutations where the functionality of resistance genes is halted, while bacterial viability is maintained, known as the re-sensitization process [225,226]. The re-sensitization of E. coli strains using ESBL-encoding plasmids was carried out by Kim et al. [220]. They used plasmids encoding for Cas9 and crRNAs against conserved areas in the ESBL genes to transform strains of E. coli that produce ESBLs. The CRISPR-Cas9 system effectively reduced the resistance in the transformants by targeting specific cleavage of resistant plasmids. The realization of the broad utility of the CRISPR-Cas system in gene editing accelerated the need to search for Cas protein variants with enhanced functions, including higher activity, potential for therapeutic delivery, nucleic acid detection, etc. [227]. Among various Cas proteins, the most frequently used Cas proteins are Cas9, which results in a double-strand break by specifically cleaving the targeted sequence [214]; dCas9, a catalytically “dead or defective” Cas9 protein that contrasts with Cas9 by not showing double-strand nuclease activity, but instead staying attached to the targeted sequence and obstructing the RNA polymerase binding to that specific region, thud hindering the transcription initiation [228]; nSpCas9:rAPOBEC1, a Cas9 protein without nuclease activity attached to a deaminase, resulting in the conversion of cytidine bases into thymine and hence forming a stop codon [229]; and Cas13a, an RNA-specific endonuclease that, when recognized by particular DNA sequence, causes the cleavage of RNA fragments [5]. Cas14 is also attracting scientists’ attention as it is small, has single-stranded (ss) DNA-targeting activity, and does not require protospacer adjacent motif (PAM) sequences to bind, as compared to Cas9 and Cas12 proteins [227,230].

7.2. Recent Studies on the Application of CRISPR-Cas System in AMR Bacteria

Several academic studies have confirmed the application of the CRISPR-Cas system and its effectiveness in controlling and/or stopping of AMR in bacteria [231,232] (Table 2). The use of CRISPR-Cas13a to achieve cell death in E. coli strains against carbapenems and colistin-resistance genes was achieved through engineering and transfer of CRISPR-Cas13a via M13 phages [5]. Similarly, a trans-conjugative delivery system known as CRISPR Cas13a-based killing plasmids (CKPs) was applied to kill the endogenous AMR genes in S. enterica serotype Typhimurium, where colonies of S. typhimurium showed a substantial decrease through CRISPR-Cas13a [233]. More research can be performed on this strategy of employing Cas13a to target RNA transcripts to target the pathogenic bacterial strains depending on the presence of particular virulence genes.
AMR genes were successfully targeted in E. coli and S. aureus strains using a plasmid expressing a Cas9-driven RNA [212]. Clinical isolates of S. aureus were treated by manipulating the Cas9 and crRNA for the methicillin-resistance gene (mecA), which showed a significant (~50%) decrease in the disease as compared to the control [209,234,235]. Using the CRISPR-Cas system, a 20-time decrease in the number of viable cells of E. coli strain O157:H7 for the eae gene was observed, which is an essential gene for intestinal infections [219]. The application of the CRISPR-Cas9 system significantly reduced the S. aureus colonies on mouse skin in comparison to alternative treatment strategies [236]. A non-viral delivery was performed using a polymer, branched polyethyleneimine (bPEI), Cas9, and a single guide RNA (SgRNA) to combat the methicillin resistance gene, i.e., mecA, in S. aureus. The S. aureus strain treated with Cas9-bPEI did not grow in culture media, and growth was decreased by up to 32% compared to the control. These findings can lead to the development of novel CRISPR-based antimicrobial medications since the delivery of CRISPR via polymers can prevent risks of immunogenicity and off-target effects. It can also easily produce phenotypic alterations by editing and modifying the bacterial genome [237]. E. coli strains with resistance to colistin were developed by transforming the mcr-1 harboring E. coli strains with CRISPR-Cas9 plasmid, which not only eliminated the mcr-1 gene but also prohibited horizontal gene transfer after transformation with CRISPR-Cas9 plasmid [238]. In E. faecalis, the CRISPR-Cas9 system successfully targeted the erythromycin-resistance gene ermB and hampered growth, reducing the intestinal infections caused by this bacterium [239,240,241,242]. Many studies have shown the effectiveness of the CRISPR-Cas system for curing plasmids that present a resistant phenotype. This strategy can avoid horizontal gene transfer, target resistance genes to prevent AMR, and get rid of plasmid-carrying drug-resistant genes [234,243,244,245].
The CRISPR-Cas system can be used to explore the functions of various genes that can contribute to the increased antibiotic resistance in bacteria. The role of different genes in K. pneumoniae against tigecycline and colistin resistance was identified via the CRISPR-Cas9 system [246]. The study showed the knock-out mutants for the tetA gene presented a decrease in minimum inhibitory concentration (MIC) for tigecycline, whereas when the mgrB gene was inactivated, it resulted in the activation of the PhoPQ two-component system, ultimately increasing the MIC for colistin. E. coli strain SE15, responsible for biofilm formation in urinary catheters, resulted in reduced biofilm formation by targeting the quorum-sensing (QS) gene luxS through a CRISPR-Cas9 plasmid [247,248]. The CRISPR interference (CRISPRi) system, based on dCas9 (lacks the endonuclease activity compared to Cas9), is also used to target AMR genes. dCas9 forms a complex with sgRNA to bind at the targeted DNA sequence to inactivate the transcription and lead to gene silencing. Instead of a gene knock-out, the CRISPRi method is utilized to knock-down the desired gene, allowing for reversibility [249]. This strategy has been utilized for gene silencing and knock-down for the AMR genes in E. coli [250,251,252,253], Enterococcus faecalis [254,255], Caulobacter crescentus [256], Campylobacter jejuni [257,258], and other bacteria.
Newer strategies are also being explored, such as creating bacteriophages with DNA-encoding Cas9 and guide RNA and eliminating all phage sequences essential for phage replication. This approach will result in the cleavage and degradation of DNA in bacteriophages-infected bacteria through the CRISPR-Cas9 system; however, a continuous evolution of bacterial cells against foreign DNA might lead to the development of resistance to these approaches. Hence, bacterial mutations need to be researched, and phages must be designed to specifically target those mutations using the same strategy [259]. Moreover, CRISPR-Cas systems can be employed to either target and remove the AMR pathogen or to eliminate the bacteria themselves, which harbor the drug-resistant genes. As the evolution of bacterial resistance cannot be prevented, various CRISPR-Cas system-usage strategies can be evaluated further because the system is practical and easily reprogrammable [260].
Table 2. Application of CRISPR-Cas-based genome-editing strategies in various bacteria, including A. baumannii.
Table 2. Application of CRISPR-Cas-based genome-editing strategies in various bacteria, including A. baumannii.
GenusBacterial StrainsGram
Staining
Targeted Gene/sResulted Modifications/OutcomesReferences
Actinomyces Gram +
Actinomycetes actIORF1 and actVBGenome modification and gene inactivation and replacement[261]
Acinetobacter Gram –
A. baumannii blaOXA-23, blaTEM-1D, and blaADC-25Genome editing and gene manipulation and deletions[262]
A. baumannii AB43 AbaIType I-F CRISPR-Cas system[263]
A. baumannii AdvA and ftsZCRISPRi, transposon mutagenesis, and gene editing[264]
A. baumannii gltA and β-lactamase genesMultiplex PCR and CRISPR-Cas12a[265]
A. baumannii AYE pyrFGenome editing, gene knock-out, and gene manipulation and deletions[266]
Actinoplanes Gram +
Actinoplanes sp. SE50/110 MelCGenome editing and gene deletions[267]
Bacillus Gram +
B. subtilis ku and ligDGenome alteration, DSB, and non-homologous end-joining (NHEJ) repair[268]
B. subtilis uppSCRISPRi and gene activity of essential genes[269]
B. subtilis ATCC 6051a amyE, aprE, nprE, spoIIAC, and srfCGenome editing and gene manipulation (up to 50%)[270]
B. subtilis 168 trpc2Genome alteration, gene deletions, and point mutations[271]
B. smithii pyrFGenome modification, gene deletions, and silencing and insertions (90%, 100%, and 20%, respectively)[272]
B. smithii ET 138 ldhLGenome editing, gene inactivation, and silencing with ThermoCas9 (active @ 55 °C)[273]
B. licheniformis yvmCGenome editing and gene knock-outs and integration[274]
Brucella Gram –
B. melitensis BE3Gene manipulation and 100% base replacement (C-T)[275]
Campylobacter Gram –
C. jejuni strains M1Cam and 81–176 flaA, flab, astA, and flgR, CRISPRi-based repression[257]
C. jejuni strains M1Cam and 81–176 flaA, flab, and flgR,CRISPRi-based gene repression[258]
Caulobacter Gram –
C. crescentus ctrA and gcrACRISPRi and gene knock-downs[256]
Clostridium Gram +
C. acetobutylicum ATCC 824 uppGenome editing and gene deletions, substitution, and insertions[276]
C. acetobutylicum DSM792 hprKGenome editing and gene deletion and modifications[277]
C. autoethanogenum adh and 2,3-bdhGenome editing and gene deletions[278]
C. acetobutylicum ATCC 824 and
C. beijerinckii NCIMB 8052
spoOACRISPRi and genome deletion (C. acetobutylicum = 20 bp) (C. beijerinckii = 20–1149 bp)[279]
C. beijerinckii ptaGenome modifications and single-nucleotide modification, deletion, and insertion[280]
C. beijerinckii Amylase geneCRISPRi and genetic manipulation (up to 97%)[281]
C. botulinum Genome alteration and CRISPR-system presence analysis[282]
C. cellulolyticum afpGenome editing and gene deletion and integration[283]
C. difficile Multiple genome-editing applications[284]
C. difficile JIR8094 selDGenome editing and ~20–50% site-specific mutations[285]
C. saccharoperbutylacetonicum N1–4pta and bukGenomic modifications, gene deletions (~75%), and butanol production[286]
C. pasteurianum cpaGenome editing and gene deletion and insertion[287]
Corynebacterium Gram +
C. glutamicum glgC, idsA, gltA, and pycCRISPRi[288]
C. glutamicum pyk and ldhABase editor at different loci[289]
C. glutamicum ldhAGenome modification, gene deletion and insertion (~60%), and 80% gene modification[290]
C. glutamicum crtYfGenome editing and 86–100% successful deletions[291]
C. glutamicum clpX, mepA, and porBGenome editing, deletion, insertion, and point mutation[292]
C. glutamicum ATCC 13032 argR, gabT, and gabP Genome editing and gene knock-out for gamma-aminobutyric acid (GABA) over-production[293]
C. glutamicum pgi, pck, and pykCRISPRi (~98%)[294]
Escherichia Gram –
E. coli talB, tktA, xylA, and xylBGenetic manipulation, CRISPR, and enhanced xylose production[295]
E. coli sad1, sdhA, sdhB, sucD, and sucCCRISPRi[296]
E. coli aroAGene replacements and insertions, point mutations, and deletions[297]
E. coli norVWProgrammable DNA looping[298]
E. coli galK, lacZ, and pyrFGenome editing and simultaneous integration of 03 heterologous genes[299]
E. coli ackA, adhE, ldhA, maeA, and ptaCRISPRi and increased malate production[300]
E. coli lacZGenome editing, and gene replacement and insertions[301]
E. coli gltA, cat1, sucD, 4hbd, cat2, bld, and bdhCRISPRI, gene knock-out and knock-in, and 1,4-butanediol production[302]
E. coli gltACRISPRi, genome modification, and n-butanol production[303]
E. coli arcAB and cpxRCRISPR-dCas9-based gene repression and multiple gene regulation[304]
E. coli soxRGenome engineering[250]
E. coli sul1CRISPRi[251]
E. coli AcrA, AcrB, and TolCCRISPRi[252]
E. coli luxSCRISPRi[253]
Enterobacter Gram –
E. hormaechei 34978
and E. xiangfangensis 34399
blaKPC-3Genome modifications and gene deletions[243]
E. hormaechei 4962 blaTEM-1Genome editing and gene manipulation[234]
Enterococcus Gram +
E. faecium E745 msrCGenome editing[305]
E. faecalis T11 pCF10CRISPR based genome editing[306]
E. faecalis V583 pCF10Genome manipulation[307]
E. faecalis CK135 and
E. faecalis OG1SSp
tetM and ermBGenome editing[242]
E. faecaliscroR and ebpACRISPRi and gene inactivation and silencing [254]
Klebsiella Gram –
K. pneumoniae Y4 mgrBGenome modification and gene inactivation[308]
K. pneumoniae Y17 tetA and ramRGenome modification and gene inactivation[308]
K. pneumoniae Kp97_58
and K. pneumoniae 13001
blaKPC-2Genome modification and gene deletion[243]
K. pneumoniae 492110 and
K. pneumoniae 5193
blaOXA-48 and blaOXA-48-likeGenome modification and gene deletion[243]
K. pneumoniae 3744 and 5573 pyrF, fepB, ramA, fosA, and fepBGenetic manipulation using site-specific base editing[229]
K. pneumoniae KPCRE23 blaKPC-2, blaSHV, and blaCTX-M-65Genetic manipulation using site-specific base editing[229]
Lactobacilli Gram +
L. casei LC2W_1326, LC2W_1628, and LC2W_2189Genome editing and gene deletions and integrations up to 25–60%[309]
L. gassen CRISPR-Cas activity analysis in multiple strains[310]
L. reuteri Efficient site-specific base alterations 90–100%[311]
Mycobacterium Gram +
M. tuberculosis pknB and sigHCRISPRi and genetic modifications[312]
M. tuberculosis sigACRISPRi and single/multiple targeted genetic modifications[313]
M. tuberculosis Sth1CRISPRi and gene inactivation[314]
Pseudomonas Gram –
P. aeruginosa PAO1 and
P. aeruginosa PAK
rhlB, rhlR, and prtR [315]
P. aeruginosa PA154197 mexB, mexF, mexH, mexR, mexT, and gyrA [138]
P. aeruginosa PAO1 and
P. aeruginosa PAK
algR, lasR, nalD, rhlB, rhlR, and rsaL [225]
P. putida KT2440 ldhLCRISPRi-based genome editing[273]
P. fluorescens Pf0-1, SBW25, and WH6 mNG, ftsZ, and mreBCRISPRi and gene silencing[255]
P. aeruginosa, P. putida, and
P. fluorescens
ftsZCRISPRi-based genome editing [316]
Staphylococcus Gram + Genome editing and gene inactivation
S. aureus agrA, cntA, and esaDGenome modification and base editing[317]
S. aureus RN4220 ermR and mecAGenome editing and gene deletions[318]
S. aureus rfpGenome alteration and gene knock-out, insertion, knock-in, and single-base editing[319]
S. aureus CCARM, 3798, 3803,
and 3877
mecA [237]
S. aureus 6538-GFP nuc [320]
S. aureus AH1 mecType III-A CRISPR-Cas system for gene editing[321]
S. aureus ATCC 29213 rpoBGenome modifications and gene deletions[322]
S. aureus USA300, USA300-∆mecA and RN4220mecAGenome editing and gene inactivation[5]
S. aureus USA300φ and S. aureus RNφ mecAGenome editing[224]
S. aureus ATCC 6538 tarH, tarG, and tarOGenome alteration and gene knock-out [228]
S. aureus CTH96 NucGenome editing and genetic manipulation and deletion[323]
Streptomyces Gram +
Streptomyces Multiple genesMultiplex gene disruption[324]
S. coelicolor Genome editing and gene knocked-outs[325]
S. lividans, S. albus,
S. roseosporus, S. venezuelae, and
S. viridochromogenes
Biosynthetic gene clusters (BGCs)Multiple genome editing and gene knock-in and gene insertion[326]
S. coelicolor M145 actI-ORF2Genome editing and gene deletion (~900 bp)[327]
S. avermitilis Ac(3)ⅣGenomic disruption using Type I-E CRSIPR-Cas system[328]
S. rimosus zwf2 and devBGenome editing, gene deletions, point mutations, and oxytetracycline production[329]
S. lividans,
S. viridochromogenes, and S. albus
sshg_05713Multiple genome editing and genome deletion (20 bp–30 kb)[330]
S. coelicolor A3(2) actIORF1 (SCO5087) and actVB (SCO5092)CRISPRi and gene deletion[261]
S. coelicolor actII-orf4, redD, and glnRGenome editing and single- and multiple-gene deletions[331]
Synechococcus Gram –
S. elongatus
UTEX 2973
nblaGenome editing and gene deletion[332]
Retrieved and modified from [333,334].

7.3. CRISPR-Cas System in A. baumannii

The existence of two endogenous CRISPR-Cas systems in the genomes of various Acinetobacter species has been verified by the analysis of the CRISPRCas database (CRISPRCasdb) [335,336,337]. Nearly 2500 genomes of A. baumannii subjected to a pangenome study also confirmed the presence of two CRISPR-Cas systems in Acinetobacter spp. [338]. The first system is present in the genome of clinical isolates of A. baumannii AYE, AB037, 4190, and AB0057, whereas the second system was discovered in A. baumannii type strain ATCC®19606TM and A. baylyi ADP1 [339,340,341]. The CRISPR system is found in a variety of prokaryotes [337]. However, only 36% of the bacteria comprise both CRISPR arrays and Cas genes. The CRISPRCasdb analysis revealed the presence of CRISPR arrays and Cas genes in nearly 20% of organisms in the Acinetobacter genus and 18% of isolates of A. baumannii spp. [337,342]. The CRISPR system has been identified in different A. baumannii strains by analyzing the large volume of sequencing data and by application of bioinformatical tools [340]. The trailer and spacer regions of the CRISPR system are generally conserved among various bacterial isolates. This helps in the grouping of isolates and the identification of common ancestors based on the presence of sequence arrays [343]. Several studies have shown that the CRISPR-Cas system not only provides immunity in A. baumannii but also regulates various virulence gene expressions, controls group behaviors, provides DNA repair, and dictates genome evolution [342].
The Type I CRISPR-Cas system is the most common in nature, comprising a multi-subunit effector complex [200,344]. This effector complex includes nine subtypes known as A, B, C, G, D, E, F1, F2, and F3 [200]. Type I-F CRISPR-Cas systems are the most common in A. baumannii [337], but the Type IV variant with genes csf3, csf4 (also named as dinG), and cas6e, along with CRISPR arrays at both ends, is also present in some Acinetobacter spp. [345]. Recently, several studies confirmed the presence of the Type I-F CRISPR-Cas system in various A. baumannii isolates from throughout the world [346,347,348,349,350]. Based on the presence or absence of the 14 common genes, the Acinetobacter genome can be divided into two groups. The first group comprises fewer common CRISPR genes and hence shows rarity in the presence of plasmids [338]. The existence of Type 1 and Type IV CRISPR-Cas systems in A. baumannii was also confirmed. In silico analysis of 4977 A. baumannii genomes from the NCBI Refseq database revealed nearly 14% of A. baumannii clinical isolates carried CRISPR-Cas systems [351,352,353]. Further classification of A. baumannii genomes presenting CRISPR-Cas systems showed that Type I-F1 CRISPR-Cas system was most abundant (~67%), followed by Type I-F2 (~28%), while both Type I-F1 and Type I-F2 were present in ~4% genomes. Various studies have reported the coexistence of different types of CRISPR-Cas systems in other bacteria [345,354,355,356]; however, the co-localization of Type I-F (I-F1 + I-F2) in A. baumannii was reported by Yadav and Singh [353].
The CRISPR-Cas system, specifically Type 1, is successfully employed for genome editing in various bacterial strains. There is still great potential to explore and implement it for genetic manipulation in A. baumannii by exploring the association of CRISPR-Cas systems with bacterial virulence and pathogenesis mechanisms [357,358]. Tyumentseva et al. confirmed the existence of CRISPR arrays and Cas genes related to Type I-F2 in clinical isolates of A. baumannii [342]. They also found a correlation between the AMR genotype/phenotype of A. baumannii with its type of CRISPR-Cas system. It was observed that a higher number of AMR genes was present in the isolates where both CRISPR arrays and active Cas genes were missing as compared to the isolates with only CRISPR arrays or both CRISPR arrays and Cas genes. This helps bacteria fight against phage infections and protect against the spreading of AMR genes in A. baumannii. Virulence factors were also found to be dependent on CRISPR-Cas systems in A. baumannii. Additionally, regarding the difference between Type I-F1 and Type I-F2 CRISPR-Cas systems among the isolates of A. baumannii, CRISPR arrays were lower in isolates with Type I-F1 CRISPR-Cas system as compared to Type I-F2 isolates, which have a stronger immune system. The presence of more AMR genes in Type I-F1 A. baumannii isolates supports easier adaptation to different environmental conditions, whereas Type I-F2 isolates may utilize the CRISPR-Cas system to control the distribution of AMR genes. Similarly, the Type I-F1 CRISPR-Cas system affects the acquisition of AMR plasmids in wild-type antimicrobial-susceptible E. coli isolates and sustains the susceptible profile of these E. coli isolates [359]. Other studies also confirmed the impact of the CRISPR-Cas system on the accretion of virulence and AMR-related genes, where it can prevent the accumulation of resistance genes but does not affect the mutations in cells to attain AMR. However, the A. baumannii genomes having CRISPR-Cas systems did not show any correlations with any specific antibiotic classes and with virulence genes [352], indicating no effect of this system on the acquisition of resistance and virulence genes in A. baumannii. The genomes possessing only co-localizing Type I-F1 + F2 CRISPR-Cas systems showed negative correlation for factors including biofilm-associated proteins (bap, bauA) and quorum-sensing (QS) genes (abaI and abaR) [353]. Similar results were also presented by other research groups [338,342,360]. These association analyses suggest that without affecting the phage-based memory, spacers can target plasmids through an unknown mechanism for the acquisition of virulence and resistance factors in the genomes of A. baumannii.

7.4. Recent Studies on the Application of the CRISPR-Cas System in A. baumannii

To apply an effective exogenous CRISPR-Cas system to A. baumannii, many researchers have generated and tested various models. An exogenous recombination system involving two plasmids carrying Cas9 from S. pyogenes and sgRNA was developed, where each plasmid could replicate in both E. coli and A. baumannii. The impact of the genes (blaOXA-23, blaADC-25, and blaTEM-1D) was reported on imipenem and sulbactam resistance in A. baumannii. Additionally, the researchers constructed single-, double-, and triple-gene mutants to explain the role of each gene in accomplishing AMR [262,361]. The application of CRISPR-Cas-mediated genome modifications in A. baumannii may be tricky, as the procurement of several resistance genes is related to mobile genetic elements such as plasmids and transposons. Hence, the loss of the plasmid carrying the CRISPR components and unwanted genetic combinations might be obtained [229]. In this regard, an alternate strategy to perform gene editing is cytidine-base editing (C to T replacement), which does not require a DSB and a donor template. In this process, single-base replacement (A/C to T) in CAA or CAG can produce stop codons (TAA or TAG). The successful application of cytidine-base editing was achieved in K. pneumoniae and A. baumannii ATCC®17978TM by constructing the plasmid vector pBECAb-apr with a sgRNA and a fusion protein expression [229]. Moreover, a two-plasmid-based CRISPR-Cas system to perform gene editing in K. pneumoniae [262] was tested in A. baumannii.
The relationship between drug resistance and the CRISPR-Cas system was analyzed in A. baumannii strain AB43 using the whole-genome sequencing (WGS) technique [362]. The authors identified the presence of the Type I-Fb CRISPR-Cas system in the strain AB43 and found that the cas gene in the studied strain has a higher similarity index with the similar subtype cas genes. The role of the CRISPR-Cas system in the AB43 strain against invasive bacteriophage and plasmids was confirmed, as 28 out of 105 CRISPR spacers in the genome of this strain showed similarity with the genes present in the bacteriophage genome and with the plasmid database. However, no matches for CRISPR spacers were found for AMR genes for A. baumannii strain AB43. The drug resistance mechanism in A. baumannii strain AB43 via the CRISPR-Cas system is still unclear, as the endogenous CRISPR-Cas system might be responsible for inhibition of drug resistance gene expression, which requires further research [342,362]. Another study using WGS identified the CRISPR-Cas system Subtype I-F in A. baumannii strain ATCC BAA1605, with a high number of spacers present in the CRISPR loci [363]. The abaI gene responsible for biofilm formation through the quorum sensing in A. baumannii [364] was targeted to develop gene knockouts by designing sgRNAs using various bioinformatical tools [365]. To target the essential genes in A. baumannii ATCC 17978, a CRISPRi-system was applied to develop gene knock-down mutants using the anhydrotetracycline (aTc)-inducible dcas9 gene and a constitutive sgRNA for adc β-lactamases [264]. A significant (30-fold) decrease in β-lactamase synthesis was observed after the induction of aTc by dcas9, indicating successful silencing of the adc gene. The researchers also developed gene knockdowns for essential genes for cell replication, i.e., ftsZ and advA, using the CRISPRi approach and observed substantial decrease in cell growth as compared to the control. The CRISPRi approach was also used to characterize the transcriptional factor RS03245 encoding AraC in A. baumannii, which is important for bacterial growth. CRISPR-Cas12a array along with multiplex polymerase chain reaction (PCR) was implemented to detect MDR A. baumannii, enabling simultaneous amplification of essential genes and β-lactamase genes. This study showed the accuracy and specificity of the CRISPR-Cas12a system to detect important drug-resistance genes in A. baumannii [265].
An effective, convenient, and quick gene-manipulation system comprising pyrF-based suicide plasmids and pyrF-deleted uracil-auxotrophic hosts was developed and tested to successfully to delete the sequences of cas genes (cas1, cas3, and cascade) and the CRISPR sequence (except the leader and a single repeat structure) in the I-F CRISPR-Cas system in A. baumannii AYE∆F. This system is more efficient for developing knockouts in model strains than the clinical strains of A. baumannii due to lower transformation rates and biofilm formation [266]. The role of OxyR as an oxidative-stress-resistance regulator was also discovered using the CRISPR-Cas9 system. The CRISPR-Cas9-based genome-editing strategy (pCasAb-pSGAb) involved Cas9 and RecAb from A. baumannii. The amount and the length of the repair template were also optimized, which resulted in significant improvement in editing efficiency. The genome-editing efficiency of strains ATCC®17978TM and ATCC®19606T in one clinical isolate, A. baumannii ABH2, was also tested. The researchers used multiple strains because of the anticipated variation in the effectiveness of CRISPR-Cas-based genome editing due to differences in genomic background. An exogenous CRISPR-Cas system was used to introduce point mutations into a clinical isolate of A. baumannii ABH2 to study the role of H2O2-sensing amino acid residues present in OxyR. As anticipated, the mutant strains did not exhibit any deficiencies with regard to H2O2 sensitivity. However, a known residue (C202) and three new residues (E130, S133, and S226) were identified to be important for OxyR function [262]. In another study, the use of RecAb from the A. baumannii IS-123 strain produced more colonies by deleting OxyR via CRISPR-Cas [366]. The CRISPR-based genome-editing system (pCasAb-pSGAb) in comparison to the gene-editing strategy (pyrF/5-FOA) developed by [266] showed higher deletion efficiency. However, the utilization of a single plasmid and its easier removal from cells in the pyrF/5-FOA system make it a more convenient method.

8. Conclusions

CRISPR-Cas-based gene editing has provided successful results for gene manipulation in various bacteria to combat AMR; affect various important physiological processes that cause infections and reduce virulence, pathogenicity, and biofilm formation; and can lead to bacterial death. However, the application of CRISPR-Cas in A. baumannii is quite recent and not fully explored. This revolutionary technique can be used to improve the genetic makeup of A. baumannii to combat AMR by targeting and disrupting specific genes associated with AMR, such as enzymes producing genes that protect the bacteria from antibiotics and biofilm-formation genes that provide antibiotic resistance. Moreover, the CRISPR can be utilized to modify the existing genes or introduce new genes into the genome of A. baumannii to reduce the bacteria’s ability to resist antibiotics or to increase its sensitivity to antibiotics. For example, a specific gene associated with a receptor that regulates the activity of an antibiotic could be introduced into the bacteria to enhance its susceptibility to that specific antibiotic. Similarly, the introduction of genes into the bacterial genome that produce specific molecules that interfere with AMR mechanisms and bacterial growth can be achieved through CRISPR-based genetic manipulation. Overall, CRISPR-based gene editing could be effectively used to combat AMR in A. baumannii, and the method still requires more research in this regard in the future.

Author Contributions

Conceptualization, M.J. and M.T.C.; writing—original draft preparation, M.J.; writing—review and editing, M.J., K.T., P.K., and M.T.C.; visualization, M.J.; supervision, M.T.C.; project administration, M.T.C.; funding acquisition, M.T.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from Mahidol University, Thailand, under the Grant number NDFR 12/2565 and International Postdoctoral Fellowship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zohra, T.; Numan, M.; Ikram, A.; Salman, M.; Khan, T.; Din, M.; Salman, M.; Farooq, A.; Amir, A.; Ali, M. Cracking the challenge of antimicrobial drug resistance with CRISPR/Cas9, nanotechnology and other strategies in ESKAPE pathogens. Microorganisms 2021, 9, 954. [Google Scholar] [CrossRef] [PubMed]
  2. Jansen, K.U.; Knirsch, C.; Anderson, A.S. The role of vaccines in preventing bacterial antimicrobial resistance. Nat. Med. 2018, 24, 10–19. [Google Scholar] [CrossRef] [PubMed]
  3. Martinez, J.L.; Fajardo, A.; Garmendia, L.; Hernandez, A.; Linares, J.F.; Martínez-Solano, L.; Sánchez, M.B. A global view of antibiotic resistance. FEMS Microbiol. Rev. 2008, 33, 44–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hutchings, M.I.; Truman, A.W.; Wilkinson, B. Antibiotics: Past, present and future. Curr. Opin. Microbiol. 2019, 51, 72–80. [Google Scholar] [CrossRef] [PubMed]
  5. Kiga, K.; Tan, X.-E.; Ibarra-Chávez, R.; Watanabe, S.; Aiba, Y.; Sato’o, Y.; Li, F.-Y.; Sasahara, T.; Cui, B.; Kawauchi, M. Development of CRISPR-Cas13a-based antimicrobials capable of sequence-specific killing of target bacteria. Nat. Commun. 2020, 11, 2934. [Google Scholar] [CrossRef]
  6. Eze, E.C.; Chenia, H.Y.; El Zowalaty, M.E. Acinetobacter baumannii biofilms: Effects of physicochemical factors, virulence, antibiotic resistance determinants, gene regulation, and future antimicrobial treatments. Infect. Drug Resist. 2018, 11, 2277. [Google Scholar] [CrossRef] [Green Version]
  7. Mahamat, A.; Bertrand, X.; Moreau, B.; Hommel, D.; Couppie, P.; Simonnet, C.; Kallel, H.; Demar, M.; Djossou, F.; Nacher, M. Clinical epidemiology and resistance mechanisms of carbapenem-resistant Acinetobacter baumannii, French Guiana, 2008–2014. Int. J. Antimicrob. Agents 2016, 48, 51–55. [Google Scholar] [CrossRef]
  8. Diancourt, L.; Passet, V.; Nemec, A.; Dijkshoorn, L.; Brisse, S. The population structure of Acinetobacter baumannii: Expanding multiresistant clones from an ancestral susceptible genetic pool. PLoS ONE 2010, 5, e10034. [Google Scholar] [CrossRef] [Green Version]
  9. Zarrilli, R.; Pournaras, S.; Giannouli, M.; Tsakris, A. Global evolution of multidrug-resistant Acinetobacter baumannii clonal lineages. Int. J. Antimicrob. Agents 2013, 41, 11–19. [Google Scholar] [CrossRef]
  10. de la Fuente-Nunez, C.; Torres, M.D.; Mojica, F.J.; Lu, T.K. Next-generation precision antimicrobials: Towards personalized treatment of infectious diseases. Curr. Opin. Microbiol. 2017, 37, 95–102. [Google Scholar] [CrossRef] [Green Version]
  11. Morris, F.C.; Dexter, C.; Kostoulias, X.; Uddin, M.I.; Peleg, A.Y. The mechanisms of disease caused by Acinetobacter baumannii. Front. Microbiol. 2019, 10, 1601. [Google Scholar] [CrossRef] [Green Version]
  12. Lim, S.M.S.; Abidin, A.Z.; Liew, S.; Roberts, J.; Sime, F. The global prevalence of multidrug-resistance among Acinetobacter baumannii causing hospital-acquired and ventilator-associated pneumonia and its associated mortality: A systematic review and meta-analysis. J. Infect. 2019, 79, 593–600. [Google Scholar]
  13. Dexter, C.; Murray, G.L.; Paulsen, I.T.; Peleg, A.Y. Community-acquired Acinetobacter baumannii: Clinical characteristics, epidemiology and pathogenesis. Expert Rev. Anti-Infect. Ther. 2015, 13, 567–573. [Google Scholar] [CrossRef]
  14. Fournier, P.E.; Richet, H.; Weinstein, R.A. The epidemiology and control of Acinetobacter baumannii in health care facilities. Clin. Infect. Dis. 2006, 42, 692–699. [Google Scholar] [CrossRef] [Green Version]
  15. Wisplinghoff, H. Pseudomonas spp., Acinetobacter spp. and miscellaneous Gram-negative bacilli. In Infectious Diseases; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1579–1599.e1572. [Google Scholar]
  16. Metan, G.; Alp, E.; Aygen, B.; Sumerkan, B. Acinetobacter baumannii meningitis in post-neurosurgical patients: Clinical outcome and impact of carbapenem resistance. J. Antimicrob. Chemother. 2007, 60, 197–199. [Google Scholar] [CrossRef] [Green Version]
  17. Chung, D.R.; Song, J.H.; Kim, S.H.; Thamlikitkul, V.; Huang, S.G.; Wang, H.; So, T.M.k.; Yasin, R.M.; Hsueh, P.R.; Carlos, C.C. High prevalence of multidrug-resistant nonfermenters in hospital-acquired pneumonia in Asia. Am. J. Respir. Crit. Care Med. 2011, 184, 1409–1417. [Google Scholar] [CrossRef]
  18. Santimaleeworagun, W.; Sumret, W.; Likitmongkonsuk, K.; Noo-in, P.; Cheeaboonkana, P.; Suphannavej, A.; Suphanklang, J.; Saelim, W. Treatment Outcomes and Risk Factors Related to Mortality and Treatment Failure of Patients Infected with Acinetobacter baumannii at a General Hospital. BKK Med. J. 2019, 15, 154. [Google Scholar] [CrossRef]
  19. Wong, D.; Nielsen, T.B.; Bonomo, R.A.; Pantapalangkoor, P.; Luna, B.; Spellberg, B. Clinical and pathophysiological overview of Acinetobacter infections: A century of challenges. Clin. Microbiol. Rev. 2017, 30, 409–447. [Google Scholar] [CrossRef] [Green Version]
  20. Rafei, R.; Dabboussi, F.; Hamze, M.; Eveillard, M.; Lemarié, C.; Mallat, H.; Rolain, J.M.; Joly-Guillou, M.L.; Kempf, M. First report of blaNDM-1-producing Acinetobacter baumannii isolated in Lebanon from civilians wounded during the Syrian war. Int. J. Infect. Dis. 2014, 21, 21–23. [Google Scholar] [CrossRef] [Green Version]
  21. Tao, C.; Kang, M.; Chen, Z.; Xie, Y.; Fan, H.; Qin, L.; Ma, Y. Microbiologic study of the pathogens isolated from wound culture among Wenchuan earthquake survivors. Diagn. Microbiol. Infect. Dis. 2009, 63, 268–270. [Google Scholar] [CrossRef]
  22. Masuda, N.; Sakagawa, E.; Ohya, S.; Gotoh, N.; Tsujimoto, H.; Nishino, T. Contribution of the MexX-MexY-OprM efflux system to intrinsic resistance in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2000, 44, 2242–2246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Randall, C.P.; Mariner, K.R.; Chopra, I.; O’Neill, A.J. The target of daptomycin is absent from Escherichia coli and other gram-negative pathogens. Antimicrob. Agents Chemother. 2013, 57, 637–639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kostyanev, T.; Can, F. The global crisis of antimicrobial resistance. In Antimicrobial Stewardship; Elsevier: Amsterdam, The Netherlands, 2017; pp. 3–12. [Google Scholar]
  25. Pang, Z.; Raudonis, R.; Glick, B.R.; Lin, T.-J.; Cheng, Z. Antibiotic resistance in Pseudomonas aeruginosa: Mechanisms and alternative therapeutic strategies. Biotechnol. Adv. 2019, 37, 177–192. [Google Scholar] [CrossRef] [PubMed]
  26. Salimiyan-Rizi, K.; Noghondar, M. Adaptive antibiotic resistance: Overview and perspectives. J. Infect. Dis. Ther. 2018, 6, 363. [Google Scholar] [CrossRef]
  27. Tenover, F.C. Mechanisms of antimicrobial resistance in bacteria. Am. J. Med. 2006, 119, S3–S10. [Google Scholar] [CrossRef]
  28. Blair, J.M.; Webber, M.A.; Baylay, A.J.; Ogbolu, D.O.; Piddock, L.J. Molecular mechanisms of antibiotic resistance. Nat. Rev. Microbiol. 2015, 13, 42–51. [Google Scholar] [CrossRef]
  29. Ferenci, T.; Phan, K. How porin heterogeneity and trade-offs affect the antibiotic susceptibility of Gram-negative bacteria. Genes 2015, 6, 1113–1124. [Google Scholar] [CrossRef] [Green Version]
  30. Reygaert, W.C. An overview of the antimicrobial resistance mechanisms of bacteria. AIMS Microbiol. 2018, 4, 482. [Google Scholar] [CrossRef]
  31. Wassef, M.; Abdelhaleim, M.; AbdulRahman, E.; Ghaith, D. The role of OmpK35, OmpK36 porins, and production of β-lactamases on imipenem susceptibility in Klebsiella pneumoniae clinical isolates, Cairo, Egypt. Microb. Drug Resist. 2015, 21, 577–580. [Google Scholar] [CrossRef]
  32. Sandoval-Motta, S.; Aldana, M. Adaptive resistance to antibiotics in bacteria: A systems biology perspective. Wiley Interdiscip. Rev. Syst. Biol. Med. 2016, 8, 253–267. [Google Scholar] [CrossRef]
  33. Fernández, L.; Breidenstein, E.B.; Hancock, R.E. Creeping baselines and adaptive resistance to antibiotics. Drug Resist. Updat. 2011, 14, 1–21. [Google Scholar] [CrossRef]
  34. Kang, I.B.; Seo, K.H. Variation of antibiotic resistance in Salmonella Enteritidis, Escherichia coli O157: H7, and Listeria monocytogenes after exposure to acid, salt, and cold stress. J. Food Saf. 2020, 40, e12804. [Google Scholar] [CrossRef]
  35. Ayoub Moubareck, C.; Hammoudi Halat, D. Insights into Acinetobacter baumannii: A review of microbiological, virulence, and resistance traits in a threatening nosocomial pathogen. Antibiotics 2020, 9, 119. [Google Scholar] [CrossRef] [Green Version]
  36. Kim, Y.; Kim, S.; Kim, Y.; Hong, K.; Wie, S.; Park, Y.; Jeong, H.; Kang, M. Carbapenem-resistant Acinetobacter baumannii: Diversity of resistant mechanisms and risk factors for infection. Epidemiol. Infect. 2012, 140, 137–145. [Google Scholar] [CrossRef]
  37. Vrancianu, C.O.; Gheorghe, I.; Czobor, I.B.; Chifiriuc, M.C. Antibiotic resistance profiles, molecular mechanisms and innovative treatment strategies of Acinetobacter baumannii. Microorganisms 2020, 8, 935. [Google Scholar] [CrossRef]
  38. Pakharukova, N.; Tuittila, M.; Paavilainen, S.; Malmi, H.; Parilova, O.; Teneberg, S.; Knight, S.D.; Zavialov, A.V. Structural basis for Acinetobacter baumannii biofilm formation. Proc. Natl. Acad. Sci. USA 2018, 115, 5558–5563. [Google Scholar] [CrossRef] [Green Version]
  39. Yang, C.-H.; Su, P.-W.; Moi, S.-H.; Chuang, L.-Y. Biofilm formation in Acinetobacter baumannii: Genotype-phenotype correlation. Molecules 2019, 24, 1849. [Google Scholar] [CrossRef] [Green Version]
  40. Esterly, J.S.; Richardson, C.L.; Eltoukhy, N.S.; Qi, C.; Scheetz, M.H. Genetic mechanisms of antimicrobial resistance of Acinetobacter baumannii. Ann. Pharmacother. 2011, 45, 218–228. [Google Scholar] [CrossRef]
  41. Ravasi, P.; Limansky, A.S.; Rodriguez, R.E.; Viale, A.M.; Mussi, M.A. ISAba825, a functional insertion sequence modulating genomic plasticity and blaOXA-58 expression in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2011, 55, 917–920. [Google Scholar] [CrossRef] [Green Version]
  42. Magiorakos, A.P.; Srinivasan, A.; Carey, R.B.; Carmeli, Y.; Falagas, M.; Giske, C.; Harbarth, S.; Hindler, J.; Kahlmeter, G.; Olsson-Liljequist, B. 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, 268–281. [Google Scholar] [CrossRef] [Green Version]
  43. Nordmann, P.; Poirel, L. Epidemiology and diagnostics of carbapenem resistance in gram-negative bacteria. Clin. Infect. Dis. 2019, 69, S521–S528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Piperaki, E.-T.; Tzouvelekis, L.; Miriagou, V.; Daikos, G. Carbapenem-resistant Acinetobacter baumannii: In pursuit of an effective treatment. Clin. Microbiol. Infect. 2019, 25, 951–957. [Google Scholar] [CrossRef] [PubMed]
  45. Upadhyay, S.; Khyriem, A.B.; Bhattacharya, P.; Bhattacharjee, A.; Joshi, S.R. High-level aminoglycoside resistance in Acinetobacter baumannii recovered from Intensive Care Unit patients in Northeastern India. Indian J. Med. Microbiol. 2018, 36, 43–48. [Google Scholar] [CrossRef] [PubMed]
  46. Basatian-Tashkan, B.; Niakan, M.; Khaledi, M.; Afkhami, H.; Sameni, F.; Bakhti, S.; Mirnejad, R. Antibiotic resistance assessment of Acinetobacter baumannii isolates from Tehran hospitals due to the presence of efflux pumps encoding genes (adeA and adeS genes) by molecular method. BMC Res. Notes 2020, 13, 543. [Google Scholar] [CrossRef]
  47. Kyriakidis, I.; Vasileiou, E.; Pana, Z.D.; Tragiannidis, A. Acinetobacter baumannii antibiotic resistance mechanisms. Pathogens 2021, 10, 373. [Google Scholar] [CrossRef]
  48. Martínez-Trejo, A.; Ruiz-Ruiz, J.M.; Gonzalez-Avila, L.U.; Saldaña-Padilla, A.; Hernández-Cortez, C.; Loyola-Cruz, M.A.; Bello-López, J.M.; Castro-Escarpulli, G. Evasion of Antimicrobial Activity in Acinetobacter baumannii by Target Site Modifications: An Effective Resistance Mechanism. Int. J. Mol. Sci. 2022, 23, 6582. [Google Scholar] [CrossRef]
  49. Araujo Lima, A.V.; da Silva, S.M.; do Nascimento Júnior, J.A.A.; Correia, M.d.S.; Luz, A.C.; Leal-Balbino, T.C.; da Silva, M.V.; Lima, J.L.d.C.; Maciel, M.A.V.; Napoleao, T.H. Occurrence and diversity of intra-and interhospital drug-resistant and biofilm-forming Acinetobacter baumannii and Pseudomonas aeruginosa. Microb. Drug Resist. 2020, 26, 802–814. [Google Scholar] [CrossRef]
  50. Arhoune, B.; Oumokhtar, B.; Hmami, F.; El Fakir, S.; Moutaouakkil, K.; Chami, F.; Bouharrou, A. Intestinal carriage of antibiotic resistant Acinetobacter baumannii among newborns hospitalized in Moroccan neonatal intensive care unit. PLoS ONE 2019, 14, e0209425. [Google Scholar] [CrossRef]
  51. D’Onofrio, V.; Conzemius, R.; Varda-Brkić, D.; Bogdan, M.; Grisold, A.; Gyssens, I.C.; Bedenić, B.; Barišić, I. Epidemiology of colistin-resistant, carbapenemase-producing Enterobacteriaceae and Acinetobacter baumannii in Croatia. Infect. Genet. Evol. 2020, 81, 104263. [Google Scholar] [CrossRef]
  52. Makke, G.; Bitar, I.; Salloum, T.; Panossian, B.; Alousi, S.; Arabaghian, H.; Medvecky, M.; Hrabak, J.; Merheb-Ghoussoub, S.; Tokajian, S. Whole-genome-sequence-based characterization of extensively drug-resistant Acinetobacter baumannii hospital outbreak. mSphere 2020, 5, e00934-19. [Google Scholar] [CrossRef] [Green Version]
  53. Simo Tchuinte, P.L.; Rabenandrasana, M.A.N.; Kowalewicz, C.; Andrianoelina, V.H.; Rakotondrasoa, A.; Andrianirina, Z.Z.; Enouf, V.; Ratsima, E.H.; Randrianirina, F.; Collard, J.-M. Phenotypic and molecular characterisations of carbapenem-resistant Acinetobacter baumannii strains isolated in Madagascar. Antimicrob. Resist. Infect. Control 2019, 8, 31. [Google Scholar] [CrossRef] [Green Version]
  54. Tooke, C.L.; Hinchliffe, P.; Bragginton, E.C.; Colenso, C.K.; Hirvonen, V.H.; Takebayashi, Y.; Spencer, J. β-lactamases and β-lactamase inhibitors in the 21st century. J. Mol. Biol. 2019, 431, 3472–3500. [Google Scholar] [CrossRef]
  55. Caneiras, C.; Calisto, F.; Jorge da Silva, G.; Lito, L.; Melo-Cristino, J.; Duarte, A. First description of colistin and tigecycline-resistant Acinetobacter baumannii producing KPC-3 carbapenemase in Portugal. Antibiotics 2018, 7, 96. [Google Scholar] [CrossRef] [Green Version]
  56. Martinez, T.; Martinez, I.; Vazquez, G.J.; Aquino, E.E.; Robledo, I.E. Genetic environment of the KPC gene in Acinetobacter baumannii ST2 clone from Puerto Rico and genomic insights into its drug resistance. J. Med. Microbiol. 2016, 65, 784. [Google Scholar] [CrossRef]
  57. Benamrouche, N.; Lafer, O.; Benmahdi, L.; Benslimani, A.; Amhis, W.; Ammari, H.; Assaous, F.; Azzam, A.; Rahal, K.; Tali Maamar, H. Phenotypic and genotypic characterization of multidrug-resistant Acinetobacter baumannii isolated in Algerian hospitals. J. Infect. Dev. Ctries. 2020, 14, 1395–1401. [Google Scholar] [CrossRef]
  58. Smiline, A.; Vijayashree, J.; Paramasivam, A. Molecular characterization of plasmid-encoded blaTEM, blaSHV and blaCTX-M among extended spectrum β-lactamases [ESBLs] producing Acinetobacter baumannii. Br. J. Biomed. Sci. 2018, 75, 200–202. [Google Scholar] [CrossRef]
  59. Jeon, J.H.; Lee, J.H.; Lee, J.J.; Park, K.S.; Karim, A.M.; Lee, C.R.; Jeong, B.C.; Lee, S.H. Structural basis for carbapenem-hydrolyzing mechanisms of carbapenemases conferring antibiotic resistance. Int. J. Mol. Sci. 2015, 16, 9654–9692. [Google Scholar] [CrossRef] [Green Version]
  60. Queenan, A.M.; Bush, K. Carbapenemases: The versatile β-lactamases. Clin. Microbiol. Rev. 2007, 20, 440–458. [Google Scholar] [CrossRef] [Green Version]
  61. Giuseppe, C.; Helen, G.; Gian-Maria, R. Metallo-β-lactamases: A last frontier for β-lactams. Lancet Infect. Dis. 2011, 11, 381–393. [Google Scholar]
  62. Amin, M.; Navidifar, T.; Saleh Shooshtari, F.; Goodarzi, H. Association of the genes encoding metallo-β-lactamase with the presence of integrons among multidrug-resistant clinical isolates of Acinetobacter baumannii. Infect. Drug Resist. 2019, 12, 1171–1180. [Google Scholar] [CrossRef] [Green Version]
  63. López, C.; Ayala, J.A.; Bonomo, R.A.; González, L.J.; Vila, A.J. Protein determinants of dissemination and host specificity of metallo-β-lactamases. Nat. Commun. 2019, 10, 3617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Moulana, Z.; Babazadeh, A.; Eslamdost, Z.; Shokri, M.; Ebrahimpour, S. Phenotypic and genotypic detection of metallo-beta-lactamases in Carbapenem resistant Acinetobacter baumannii. Caspian J. Intern. Med. 2020, 11, 171. [Google Scholar] [PubMed]
  65. Hamidian, M.; Hall, R.M. Tn6168, a transposon carrying an ISAba1-activated ampC gene and conferring cephalosporin resistance in Acinetobacter baumannii. J. Antimicrob. Chemother. 2014, 69, 77–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Lopes, B.; Amyes, S. Role of ISAba1 and ISAba125 in governing the expression of blaADC in clinically relevant Acinetobacter baumannii strains resistant to cephalosporins. J. Med. Microbiol. 2012, 61, 1103–1108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Lee, C.-R.; Lee, J.H.; Park, M.; Park, K.S.; Bae, I.K.; Kim, Y.B.; Cha, C.-J.; Jeong, B.C.; Lee, S.H. Biology of Acinetobacter baumannii: Pathogenesis, antibiotic resistance mechanisms, and prospective treatment options. Front. Cell. Infect. Microbiol. 2017, 7, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Chagas, T.P.G.; Carvalho, K.R.; de Oliveira Santos, I.C.; Carvalho-Assef, A.P.D.A.; Asensi, M.D. Characterization of carbapenem-resistant Acinetobacter baumannii in Brazil (2008–2011): Countrywide spread of OXA-23–producing clones (CC15 and CC79). Diagn. Microbiol. Infect. Dis. 2014, 79, 468–472. [Google Scholar] [CrossRef] [Green Version]
  69. Evans, B.A.; Amyes, S.G. OXA β-lactamases. Clin. Microbiol. Rev. 2014, 27, 241–263. [Google Scholar] [CrossRef] [Green Version]
  70. Hu, W.S.; Yao, S.-M.; Fung, C.-P.; Hsieh, Y.-P.; Liu, C.-P.; Lin, J.-F. An OXA-66/OXA-51-like carbapenemase and possibly an efflux pump are associated with resistance to imipenem in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2007, 51, 3844–3852. [Google Scholar] [CrossRef] [Green Version]
  71. Mammina, C.; Bonura, C.; Aleo, A.; Calà, C.; Caputo, G.; Cataldo, M.; Benedetto, A.D.; Distefano, S.; Fasciana, T.; Labisi, M. Characterization of Acinetobacter baumannii from intensive care units and home care patients in Palermo, Italy. Clin. Microbiol. Infect. 2011, 17, E12–E15. [Google Scholar] [CrossRef] [Green Version]
  72. Pagano, M.; Martins, A.; Machado, A.; Barin, J.; Barth, A. Carbapenem-susceptible Acinetobacter baumannii carrying the ISAba1 upstream blaOXA-51-like gene in Porto Alegre, southern Brazil. Epidemiol. Infect. 2013, 141, 330–333. [Google Scholar] [CrossRef]
  73. Pajand, O.; Hojabri, Z.; Nahaei, M.R.; Hajibonabi, F.; Pirzadeh, T.; Aghazadeh, M.; Fasciana, T.; Bonura, C.; Mammina, C. In vitro activities of tetracyclines against different clones of multidrug-resistant Acinetobacter baumannii isolates from two Iranian hospitals. Int. J. Antimicrob. Agents 2014, 43, 476–478. [Google Scholar] [CrossRef] [Green Version]
  74. Wong, M.H.Y.; Chan, B.K.W.; Chan, E.W.C.; Chen, S. Over-expression of ISAba1-linked intrinsic and exogenously acquired OXA type carbapenem-hydrolyzing-class D-β-lactamase-encoding genes is key mechanism underlying carbapenem resistance in Acinetobacter baumannii. Front. Microbiol. 2019, 10, 2809. [Google Scholar] [CrossRef] [Green Version]
  75. Xu, C.; Bilya, S.; Xu, W. adeABC efflux gene in Acinetobacter baumannii. New Microbes New Infect. 2019, 30, 100549. [Google Scholar] [CrossRef]
  76. Ramirez, M.S.; Tolmasky, M.E. Aminoglycoside modifying enzymes. Drug Resist. Updat. 2010, 13, 151–171. [Google Scholar] [CrossRef] [Green Version]
  77. Foong, W.E.; Wilhelm, J.; Tam, H.-K.; Pos, K.M. Tigecycline efflux in Acinetobacter baumannii is mediated by TetA in synergy with RND-type efflux transporters. J. Antimicrob. Chemother. 2020, 75, 1135–1139. [Google Scholar] [CrossRef]
  78. He, T.; Wang, R.; Liu, D.; Walsh, T.R.; Zhang, R.; Lv, Y.; Ke, Y.; Ji, Q.; Wei, R.; Liu, Z. Emergence of plasmid-mediated high-level tigecycline resistance genes in animals and humans. Nat. Microbiol. 2019, 4, 1450–1456. [Google Scholar] [CrossRef]
  79. Li, L.; Hassan, K.A.; Tetu, S.G.; Naidu, V.; Pokhrel, A.; Cain, A.K.; Paulsen, I.T. The transcriptomic signature of tigecycline in Acinetobacter baumannii. Front. Microbiol. 2020, 11, 565438. [Google Scholar] [CrossRef]
  80. Savari, M.; Ekrami, A.; Shoja, S.; Bahador, A. Plasmid borne carbapenem-hydrolyzing class D β-lactamases (CHDLs) and AdeABC efflux pump conferring carbapenem-tigecycline resistance among Acinetobacter baumannii isolates harboring TnAbaRs. Microb. Pathog. 2017, 104, 310–317. [Google Scholar] [CrossRef]
  81. Wang, L.; Liu, D.; Lv, Y.; Cui, L.; Li, Y.; Li, T.; Song, H.; Hao, Y.; Shen, J.; Wang, Y. Novel plasmid-mediated tet(X5) gene conferring resistance to tigecycline, eravacycline, and omadacycline in a clinical Acinetobacter baumannii isolate. Antimicrob. Agents Chemother. 2019, 64, e01326-19. [Google Scholar] [CrossRef]
  82. Cho, Y.J.; Moon, D.C.; Jin, J.S.; Choi, C.H.; Lee, Y.C.; Lee, J.C. Genetic basis of resistance to aminoglycosides in Acinetobacter spp. and spread of armA in Acinetobacter baumannii sequence group 1 in Korean hospitals. Diagn. Microbiol. Infect. Dis. 2009, 64, 185–190. [Google Scholar] [CrossRef]
  83. Vázquez-López, R.; Solano-Gálvez, S.G.; Juárez Vignon-Whaley, J.J.; Abello Vaamonde, J.A.; Padró Alonzo, L.A.; Rivera Reséndiz, A.; Muleiro Álvarez, M.; Vega López, E.N.; Franyuti-Kelly, G.; Álvarez-Hernández, D.A. Acinetobacter baumannii resistance: A real challenge for clinicians. Antibiotics 2020, 9, 205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zaki, M.E.S.; Abou ElKheir, N.; Mofreh, M. Molecular study of quinolone resistance determining regions of gyrA gene and parC genes in clinical isolates of Acintobacter baumannii resistant to fluoroquinolone. Open Microbiol. J. 2018, 12, 116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Cheng, Y.; Yang, S.; Jia, M.; Zhao, L.; Hou, C.; You, X.; Zhao, J.; Chen, A. Comparative study between macrolide regulatory proteins MphR (A) and MphR (E) in ligand identification and DNA binding based on the rapid in vitro detection system. Anal. Bioanal. Chem. 2016, 408, 1623–1631. [Google Scholar] [CrossRef] [PubMed]
  86. Lima, W.G.; Alves, M.C.; Cruz, W.S.; Paiva, M.C. Chromosomally encoded and plasmid-mediated polymyxins resistance in Acinetobacter baumannii: A huge public health threat. Eur. J. Clin. Microbiol. Infect. Dis. 2018, 37, 1009–1019. [Google Scholar] [CrossRef] [PubMed]
  87. Potron, A.; Poirel, L.; Croizé, J.; Chanteperdrix, V.; Nordmann, P. Genetic and biochemical characterization of the first extended-spectrum Carb-type β-lactamase, RTG-4, from Acinetobacter baumannii. Antimicrob. Agents Chemother. 2009, 53, 3010–3016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Freitas, D.Y.; Araújo, S.; Folador, A.R.; Ramos, R.T.; Azevedo, J.S.; Tacão, M.; Silva, A.; Henriques, I.; Baraúna, R.A. Extended spectrum beta-lactamase-producing gram-negative bacteria recovered from an Amazonian lake near the city of Belém, Brazil. Front. Microbiol. 2019, 10, 364. [Google Scholar] [CrossRef] [Green Version]
  89. Gupta, R.; Malik, A.; Rizvi, M.; Ahmed, M. Presence of metallo-beta-lactamases (MBL), extended-spectrum beta-lactamase (ESBL) & AmpC positive non-fermenting Gram-negative bacilli among Intensive Care Unit patients with special reference to molecular detection of blaCTX-M & blaAmpC genes. Indian J. Med. Res. 2016, 144, 271. [Google Scholar]
  90. Mayanskiy, N.; Chebotar, I.; Alyabieva, N.; Kryzhanovskaya, O.; Savinova, T.; Turenok, A.; Bocharova, Y.; Lazareva, A.; Polikarpova, S.; Karaseva, O. Emergence of the uncommon clone ST944/ST78 carrying blaOXA-40-like and blaCTX-M-like genes among carbapenem-nonsusceptible Acinetobacter baumannii in Moscow, Russia. Microb. Drug Resist. 2017, 23, 864–870. [Google Scholar] [CrossRef]
  91. Al-Hassan, L.; El Mahallawy, H.; Amyes, S. First report of bla (PER-3) in Acinetobacter baumannii. Int. J. Antimicrob. Agents 2012, 41, 93–94. [Google Scholar] [CrossRef]
  92. Aly, M.; Abu Alsoud, N.; Elrobh, M.; Al Johani, S.; Balkhy, H. High prevalence of the PER-1 gene among carbapenem-resistant Acinetobacter baumannii in Riyadh, Saudi Arabia. Eur. J. Clin. Microbiol. Infect. Dis. 2016, 35, 1759–1766. [Google Scholar] [CrossRef]
  93. Bonnin, R.A.; Nordmann, P.; Potron, A.; Lecuyer, H.; Zahar, J.-R.; Poirel, L. Carbapenem-hydrolyzing GES-type extended-spectrum β-lactamase in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2011, 55, 349–354. [Google Scholar] [CrossRef] [Green Version]
  94. Alkasaby, N.M.; El Sayed Zaki, M. Molecular study of Acinetobacter baumannii isolates for metallo-β-lactamases and extended-spectrum-β-lactamases genes in intensive care unit, Mansoura University Hospital, Egypt. Int. J. Microbiol. 2017, 2017, 3925868. [Google Scholar] [CrossRef] [Green Version]
  95. Naas, T.; Namdari, F.; Réglier-Poupet, H.; Poyart, C.; Nordmann, P. Panresistant extended-spectrum β-lactamase SHV-5-producing Acinetobacter baumannii from New York City. J. Antimicrob. Chemother. 2007, 60, 1174–1176. [Google Scholar] [CrossRef] [Green Version]
  96. Huang, L.-Y.; Chen, T.-L.; Lu, P.-L.; Tsai, C.-A.; Cho, W.-L.; Chang, F.-Y.; Fung, C.-P.; Siu, L. Dissemination of multidrug-resistant, class 1 integron-carrying Acinetobacter baumannii isolates in Taiwan. Clin. Microbiol. Infect. 2008, 14, 1010–1019. [Google Scholar] [CrossRef] [Green Version]
  97. Poirel, L.; Mugnier, P.D.; Toleman, M.A.; Walsh, T.R.; Rapoport, M.J.; Petroni, A.; Nordmann, P. IS CR2, another vehicle for bla VEB gene acquisition. Antimicrob. Agents Chemother. 2009, 53, 4940–4943. [Google Scholar] [CrossRef] [Green Version]
  98. Safari, M.; Nejad, A.S.M.; Bahador, A.; Jafari, R.; Alikhani, M.Y. Prevalence of ESBL and MBL encoding genes in Acinetobacter baumannii strains isolated from patients of intensive care units (ICU). Saudi J. Biol. Sci. 2015, 22, 424–429. [Google Scholar] [CrossRef] [Green Version]
  99. Abdar, M.H.; Taheri-Kalani, M.; Taheri, K.; Emadi, B.; Hasanzadeh, A.; Sedighi, A.; Pirouzi, S.; Sedighi, M. Prevalence of extended-spectrum beta-lactamase genes in Acinetobacter baumannii strains isolated from nosocomial infections in Tehran, Iran. GMS Infect. Dis. 2019, 14, Doc02. [Google Scholar]
  100. Agoba, E.E.; Govinden, U.; Peer, A.K.C.; Osei Sekyere, J.; Essack, S.Y. ISAba1 regulated OXA-23 carbapenem resistance in Acinetobacter baumannii strains in durban, South Africa. Microb. Drug Resist. 2018, 24, 1289–1295. [Google Scholar] [CrossRef]
  101. Asgin, N.; Otlu, B.; Cakmakliogullari, E.K.; Celik, B. High prevalence of TEM, VIM, and OXA-2 beta-lactamases and clonal diversity among Acinetobacter baumannii isolates in Turkey. J. Infect. Dev. Ctries. 2019, 13, 794–801. [Google Scholar] [CrossRef] [Green Version]
  102. Poirel, L.; Corvec, S.; Rapoport, M.; Mugnier, P.; Petroni, A.; Pasteran, F.; Faccone, D.; Galas, M.; Drugeon, H.; Cattoir, V. Identification of the novel narrow-spectrum β-lactamase SCO-1 in Acinetobacter spp. from Argentina. Antimicrob. Agents Chemother. 2007, 51, 2179–2184. [Google Scholar] [CrossRef] [Green Version]
  103. Al-Agamy, M.H.; Jeannot, K.; El-Mahdy, T.S.; Shibl, A.M.; Kattan, W.; Plésiat, P.; Courvalin, P. First detection of GES-5 carbapenemase-producing Acinetobacter baumannii isolate. Microb. Drug Resist. 2017, 23, 556–562. [Google Scholar] [CrossRef] [PubMed]
  104. Hammoudi, D.; Moubareck, C.A.; Hakime, N.; Houmani, M.; Barakat, A.; Najjar, Z.; Suleiman, M.; Fayad, N.; Sarraf, R.; Sarkis, D.K. Spread of imipenem-resistant Acinetobacter baumannii co-expressing OXA-23 and GES-11 carbapenemases in Lebanon. Int. J. Infect. Dis. 2015, 36, 56–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Mabrouk, A.; Grosso, F.; Botelho, J.; Achour, W.; Ben Hassen, A.; Peixe, L. GES-14-producing Acinetobacter baumannii isolates in a neonatal intensive care unit in Tunisia are associated with a typical Middle East clone and a transferable plasmid. Antimicrob. Agents Chemother. 2017, 61, e00142-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Ribeiro, P.C.S.; Monteiro, A.S.; Marques, S.G.; Monteiro, S.G.; Monteiro-Neto, V.; Coqueiro, M.M.M.; Marques, A.C.G.; de Jesus Gomes Turri, R.; Santos, S.G.; Bomfim, M.R.Q. Phenotypic and molecular detection of the blaKPC gene in clinical isolates from inpatients at hospitals in São Luis, MA, Brazil. BMC Infect. Dis. 2016, 16, 737. [Google Scholar] [CrossRef] [Green Version]
  107. Pollini, S.; Maradei, S.; Pecile, P.; Olivo, G.; Luzzaro, F.; Docquier, J.-D.; Rossolini, G.M. FIM-1, a new acquired metallo-β-lactamase from a Pseudomonas aeruginosa clinical isolate from Italy. Antimicrob. Agents Chemother. 2013, 57, 410–416. [Google Scholar] [CrossRef] [Green Version]
  108. Girija, S.A.; Jayaseelan, V.P.; Arumugam, P. Prevalence of VIM-and GIM-producing Acinetobacter baumannii from patients with severe urinary tract infection. Acta Microbiol. Immunol. Hung. 2018, 65, 539–550. [Google Scholar] [CrossRef] [Green Version]
  109. Cayô, R.; Rodrigues-Costa, F.; Matos, A.P.; Carvalhaes, C.G.; Jové, T.; Gales, A.C. Identification of a new integron harboring blaIMP-10 in carbapenem-resistant Acinetobacter baumannii clinical isolates. Antimicrob. Agents Chemother. 2015, 59, 3687–3689. [Google Scholar] [CrossRef] [Green Version]
  110. Shakibaie, M.R.; Azizi, O.; Shahcheraghi, F. Insight into stereochemistry of a new IMP allelic variant (IMP-55) metallo-β-lactamase identified in a clinical strain of Acinetobacter baumannii. Infect. Genet. Evol. 2017, 51, 118–126. [Google Scholar] [CrossRef]
  111. Bonnin, R.; Poirel, L.; Naas, T.; Pirs, M.; Seme, K.; Schrenzel, J.; Nordmann, P. Dissemination of New Delhi metallo-β-lactamase-1-producing Acinetobacter baumannii in Europe. Clin. Microbiol. Infect. 2012, 18, E362–E365. [Google Scholar] [CrossRef] [Green Version]
  112. Kumar, M. Identification of a novel NDM variant, blaNDM-3, from a multidrug-resistant Acinetobacter baumannii. Infect. Control Hosp. Epidemiol. 2016, 37, 747–748. [Google Scholar] [CrossRef] [Green Version]
  113. Voulgari, E.; Politi, L.; Pitiriga, V.; Dendrinos, J.; Poulou, A.; Georgiadis, G.; Tsakris, A. First report of an NDM-1 metallo-β-lactamase-producing Acinetobacter baumannii clinical isolate in Greece. Int. J. Antimicrob. Agents 2016, 6, 761–762. [Google Scholar] [CrossRef]
  114. Gholami, M.; Moshiri, M.; Ahanjan, M.; Salimi Chirani, A.; Hasannejad Bibalan, M.; Asadi, A.; Eshaghi, M.; Pournajaf, A.; Abbasian, S.; Kouhsari, E.; et al. The diversity of class B and class D carbapenemases in clinical Acinetobacter baumannii isolates. Infez. Med. 2018, 26, 329–335. [Google Scholar]
  115. Toleman, M.A.; Simm, A.M.; Murphy, T.A.; Gales, A.C.; Biedenbach, D.J.; Jones, R.N.; Walsh, T.R. Molecular characterization of SPM-1, a novel metallo-β-lactamase isolated in Latin America: Report from the SENTRY antimicrobial surveillance programme. J. Antimicrob. Chemother. 2002, 50, 673–679. [Google Scholar] [CrossRef] [Green Version]
  116. Papa, A.; Koulourida, V.; Souliou, E. Molecular epidemiology of carbapenem-resistant Acinetobacter baumannii in a newly established Greek hospital. Microb. Drug Resist. 2009, 15, 257–260. [Google Scholar] [CrossRef]
  117. Ramadan, R.A.; Gebriel, M.G.; Kadry, H.M.; Mosallem, A. Carbapenem-resistant Acinetobacter baumannii and Pseudomonas aeruginosa: Characterization of carbapenemase genes and E-test evaluation of colistin-based combinations. Infect. Drug Resist. 2018, 11, 1261. [Google Scholar] [CrossRef] [Green Version]
  118. Jia, H.; Chen, Y.; Wang, J.; Ruan, Z. Genomic characterisation of a clinical Acinetobacter baumannii ST1928 isolate carrying a new ampC allelic variant blaADC-196 gene from China. J. Glob. Antimicrob. Resist. 2019, 19, 43–45. [Google Scholar] [CrossRef]
  119. Kumburu, H.H.; Sonda, T.; van Zwetselaar, M.; Leekitcharoenphon, P.; Lukjancenko, O.; Mmbaga, B.T.; Alifrangis, M.; Lund, O.; Aarestrup, F.M.; Kibiki, G.S. Using WGS to identify antibiotic resistance genes and predict antimicrobial resistance phenotypes in MDR Acinetobacter baumannii in Tanzania. J. Antimicrob. Chemother. 2019, 74, 1484–1493. [Google Scholar] [CrossRef] [Green Version]
  120. Uddin, F.; McHugh, T.D.; Roulston, K.; Platt, G.; Khan, T.A.; Sohail, M. Detection of carbapenemases, AmpC and ESBL genes in Acinetobacter isolates from ICUs by DNA microarray. J. Microbiol. Methods 2018, 155, 19–23. [Google Scholar] [CrossRef]
  121. Hujer, K.M.; Hamza, N.S.; Hujer, A.M.; Perez, F.; Helfand, M.S.; Bethel, C.R.; Thomson, J.M.; Anderson, V.E.; Barlow, M.; Rice, L.B. Identification of a new allelic variant of the Acinetobacter baumannii cephalosporinase, ADC-7 β-lactamase: Defining a unique family of class C enzymes. Antimicrob. Agents Chemother. 2005, 49, 2941–2948. [Google Scholar] [CrossRef] [Green Version]
  122. Giannouli, M.; Tomasone, F.; Agodi, A.; Vahaboglu, H.; Daoud, Z.; Triassi, M.; Tsakris, A.; Zarrilli, R. Molecular epidemiology of carbapenem-resistant Acinetobacter baumannii strains in intensive care units of multiple Mediterranean hospitals. J. Antimicrob. Chemother. 2009, 63, 828–830. [Google Scholar] [CrossRef] [Green Version]
  123. Mosqueda, N.; Espinal, P.; Cosgaya, C.; Viota, S.; Plasensia, V.; Álvarez-Lerma, F.; Montero, M.; Gómez, J.; Horcajada, J.P.; Vila, J. Globally expanding carbapenemase finally appears in Spain: Nosocomial outbreak of Acinetobacter baumannii producing plasmid-encoded OXA-23 in Barcelona, Spain. Antimicrob. Agents Chemother. 2013, 57, 5155–5157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Santimaleeworagun, W.; Samret, W.; Preechachuawong, P.; Kerdsin, A.; Jitwasinkul, T. Emergence of co-carbapenemase genes, blaOXA23, blaVIM, and blaNDM in carbapenem resistant Acinetobacter baumannii clinical isolates. Southeast Asian J. Trop. Med. Public Health 2016, 47, 1001–1007. [Google Scholar] [PubMed]
  125. Wibberg, D.; Salto, I.P.; Eikmeyer, F.G.; Maus, I.; Winkler, A.; Nordmann, P.; Pühler, A.; Poirel, L.; Schlüter, A. Complete genome sequencing of Acinetobacter baumannii strain K50 discloses the large conjugative plasmid pK50a encoding carbapenemase OXA-23 and extended-spectrum β-lactamase GES-11. Antimicrob. Agents Chemother. 2018, 62, e00212-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Acosta, J.; Merino, M.; Viedma, E.; Poza, M.; Sanz, F.; Otero, J.R.; Chaves, F.; Bou, G. Multidrug-resistant Acinetobacter baumannii harboring OXA-24 carbapenemase, Spain. Emerg. Infect. Dis. 2011, 17, 1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Merino, M.; Acosta, J.; Poza, M.; Sanz, F.; Beceiro, A.; Chaves, F.; Bou, G. OXA-24 carbapenemase gene flanked by XerC/XerD-like recombination sites in different plasmids from different Acinetobacter species isolated during a nosocomial outbreak. Antimicrob. Agents Chemother. 2010, 54, 2724–2727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Oteo, J.; Hernández, J.M.; Espasa, M.; Fleites, A.; Sáez, D.; Bautista, V.; Pérez-Vázquez, M.; Fernández-García, M.D.; Delgado-Iribarren, A.; Sánchez-Romero, I. Emergence of OXA-48-producing Klebsiella pneumoniae and the novel carbapenemases OXA-244 and OXA-245 in Spain. J. Antimicrob. Chemother. 2013, 68, 317–321. [Google Scholar] [CrossRef] [Green Version]
  129. Potron, A.; Rondinaud, E.; Poirel, L.; Belmonte, O.; Boyer, S.; Camiade, S.; Nordmann, P. Genetic and biochemical characterisation of OXA-232, a carbapenem-hydrolysing class D β-lactamase from Enterobacteriaceae. Int. J. Antimicrob. Agents 2013, 41, 325–329. [Google Scholar] [CrossRef]
  130. Aly, M.; Tayeb, H.; Al Johani, S.; Alyamani, E.; Aldughaishem, F.; Alabdulkarim, I.; Balkhy, H. Genetic diversity of OXA-51-like genes among multidrug-resistant Acinetobacter baumannii in Riyadh, Saudi Arabia. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 1223–1228. [Google Scholar] [CrossRef]
  131. Rafei, R.; Pailhoriès, H.; Hamze, M.; Eveillard, M.; Mallat, H.; Dabboussi, F.; Joly-Guillou, M.-L.; Kempf, M. Molecular epidemiology of Acinetobacter baumannii in different hospitals in Tripoli, Lebanon using blaOXA-51-like sequence based typing. BMC Microbiol. 2015, 15, 103. [Google Scholar] [CrossRef] [Green Version]
  132. Mathlouthi, N.; Ben Lamine, Y.; Somai, R.; Bouhalila-Besbes, S.; Bakour, S.; Rolain, J.-M.; Chouchani, C. Incidence of OXA-23 and OXA-58 carbapenemases coexpressed in clinical isolates of Acinetobacter baumannii in Tunisia. Microb. Drug Resist. 2018, 24, 136–141. [Google Scholar] [CrossRef]
  133. Gionco, B.; Pelayo, J.S.; Venancio, E.J.; Cayô, R.; Gales, A.C.; Carrara-Marroni, F.E. Detection of OXA-231, a new variant of blaOXA-143, in Acinetobacter baumannii from Brazil: A case report. J. Antimicrob. Chemother. 2012, 67, 2531–2532. [Google Scholar] [CrossRef] [Green Version]
  134. Mostachio, A.K.; Levin, A.S.; Rizek, C.; Rossi, F.; Zerbini, J.; Costa, S.F. High prevalence of OXA-143 and alteration of outer membrane proteins in carbapenem-resistant Acinetobacter spp. isolates in Brazil. Int. J. Antimicrob. Agents 2012, 39, 396–401. [Google Scholar] [CrossRef]
  135. Sarikhani, Z.; Nazari, R.; Rostami, M.N. First report of OXA-143-lactamase producing Acinetobacter baumannii in Qom, Iran. Iran. J. Basic Med. Sci. 2017, 20, 1282. [Google Scholar]
  136. Boyd, D.A.; Mataseje, L.F.; Pelude, L.; Mitchell, R.; Bryce, E.; Roscoe, D.; Embree, J.; Katz, K.; Kibsey, P.; Lavallee, C. Results from the Canadian Nosocomial Infection Surveillance Program for detection of carbapenemase-producing Acinetobacter spp. in Canadian hospitals, 2010–16. J. Antimicrob. Chemother. 2019, 74, 315–320. [Google Scholar] [CrossRef] [Green Version]
  137. Higgins, P.G.; Pérez-Llarena, F.J.; Zander, E.; Fernández, A.; Bou, G.; Seifert, H. OXA-235, a novel class D β-lactamase involved in resistance to carbapenems in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2013, 57, 2121–2126. [Google Scholar] [CrossRef] [Green Version]
  138. Xu, Z.; Li, M.; Li, Y.; Cao, H.; Miao, L.; Xu, Z.; Higuchi, Y.; Yamasaki, S.; Nishino, K.; Woo, P.C. Native CRISPR-Cas-mediated genome editing enables dissecting and sensitizing clinical multidrug-resistant P. aeruginosa. Cell Rep. 2019, 29, 1707–1717.e1703. [Google Scholar] [CrossRef]
  139. Rumbo, C.; Gato, E.; López, M.; Ruiz de Alegría, C.; Fernández-Cuenca, F.; Martínez-Martínez, L.; Vila, J.; Pachón, J.; Cisneros, J.M.; Rodríguez-Baño, J. Contribution of efflux pumps, porins, and β-lactamases to multidrug resistance in clinical isolates of Acinetobacter baumannii. Antimicrob. Agents Chemother. 2013, 57, 5247–5257. [Google Scholar] [CrossRef] [Green Version]
  140. Smani, Y.; Fàbrega, A.; Roca, I.; Sánchez-Encinales, V.; Vila, J.; Pachón, J. Role of OmpA in the multidrug resistance phenotype of Acinetobacter baumannii. Antimicrob. Agents Chemother. 2014, 58, 1806–1808. [Google Scholar] [CrossRef] [Green Version]
  141. Mirshekar, M.; Shahcheraghi, F.; Azizi, O.; Solgi, H.; Badmasti, F. Diversity of class 1 integrons, and disruption of carO and dacD by insertion sequences among Acinetobacter baumannii isolates in Tehran, Iran. Microb. Drug Resist. 2018, 24, 359–366. [Google Scholar] [CrossRef]
  142. Rajamohan, G.; Srinivasan, V.B.; Gebreyes, W.A. Molecular and functional characterization of a novel efflux pump, AmvA, mediating antimicrobial and disinfectant resistance in Acinetobacter baumannii. J. Antimicrob. Chemother. 2010, 65, 1919–1925. [Google Scholar] [CrossRef] [Green Version]
  143. Sheikhalizadeh, V.; Hasani, A.; Rezaee, M.A.; Rahmati-Yamchi, M.; Hasani, A.; Ghotaslou, R.; Goli, H.R. Comprehensive study to investigate the role of various aminoglycoside resistance mechanisms in clinical isolates of Acinetobacter baumannii. J. Infect. Chemother. 2017, 23, 74–79. [Google Scholar] [CrossRef] [PubMed]
  144. Costello, S.E.; Gales, A.C.; Morfin-Otero, R.; Jones, R.N.; Castanheira, M. Mechanisms of resistance, clonal expansion, and increasing prevalence of Acinetobacter baumannii strains displaying elevated tigecycline MIC values in Latin America. Microb. Drug Resist. 2016, 22, 253–258. [Google Scholar] [CrossRef] [PubMed]
  145. Hasani, A.; Sheikhalizadeh, V.; Ahangarzadeh Rezaee, M.; Rahmati-Yamchi, M.; Hasani, A.; Ghotaslou, R.; Goli, H.R. Frequency of aminoglycoside-modifying enzymes and ArmA among different sequence groups of Acinetobacter baumannii in Iran. Microb. Drug Resist. 2016, 22, 347–353. [Google Scholar] [CrossRef] [PubMed]
  146. Lin, M.-F.; Liou, M.-L.; Tu, C.-C.; Yeh, H.-W.; Lan, C.-Y. Molecular epidemiology of integron-associated antimicrobial gene cassettes in the clinical isolates of Acinetobacter baumannii from northern Taiwan. Ann. Lab. Med. 2013, 33, 242–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Salimizand, H.; Zomorodi, A.R.; Mansury, D.; Khakshoor, M.; Azizi, O.; Khodaparast, S.; Baseri, Z.; Karami, P.; Zamanlou, S.; Farsiani, H. Diversity of aminoglycoside modifying enzymes and 16S rRNA methylases in Acinetobacter baumannii and Acinetobacter nosocomialis species in Iran; wide distribution of aadA1 and armA. Infect. Genet. Evol. 2018, 66, 195–199. [Google Scholar] [CrossRef]
  148. Dönhöfer, A.; Franckenberg, S.; Wickles, S.; Berninghausen, O.; Beckmann, R.; Wilson, D.N. Structural basis for TetM-mediated tetracycline resistance. Proc. Natl. Acad. Sci. USA 2012, 109, 16900–16905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Bojkovic, J.; Richie, D.L.; Six, D.A.; Rath, C.M.; Sawyer, W.S.; Hu, Q.; Dean, C.R. Characterization of an Acinetobacter baumannii lptD deletion strain: Permeability defects and response to inhibition of lipopolysaccharide and fatty acid biosynthesis. J. Bacteriol. 2016, 198, 731–741. [Google Scholar] [CrossRef] [Green Version]
  150. Martins-Sorenson, N.; Snesrud, E.; Xavier, D.E.; Cacci, L.C.; Iavarone, A.T.; McGann, P.; Riley, L.W.; Moreira, B.M. A novel plasmid-encoded mcr-4.3 gene in a colistin-resistant Acinetobacter baumannii clinical strain. J. Antimicrob. Chemother. 2020, 75, 60–64. [Google Scholar] [CrossRef]
  151. Moffatt, J.H.; Harper, M.; Boyce, J.D. Mechanisms of polymyxin resistance. In Polymyxin Antibiotics: From Laboratory Bench to Bedside; Springer: Berlin/Heidelberg, Germany, 2019; pp. 55–71. [Google Scholar]
  152. Trebosc, V.; Gartenmann, S.; Tötzl, M.; Lucchini, V.; Schellhorn, B.; Pieren, M.; Lociuro, S.; Gitzinger, M.; Tigges, M.; Bumann, D. Dissecting colistin resistance mechanisms in extensively drug-resistant Acinetobacter baumannii clinical isolates. mBio 2019, 10, e01083-19. [Google Scholar] [CrossRef] [Green Version]
  153. Srinivasan, V.B.; Rajamohan, G.; Gebreyes, W.A. Role of AbeS, a novel efflux pump of the SMR family of transporters, in resistance to antimicrobial agents in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2009, 53, 5312–5316. [Google Scholar] [CrossRef] [Green Version]
  154. Doi, Y.; Murray, G.L.; Peleg, A.Y. Acinetobacter baumannii: Evolution of antimicrobial resistance—Treatment options. Semin. Respir. Crit. Care Med. 2015, 36, 085–098. [Google Scholar]
  155. Perez, F.; Hujer, A.M.; Hujer, K.M.; Decker, B.K.; Rather, P.N.; Bonomo, R.A. Global challenge of multidrug-resistant Acinetobacter baumannii. Antimicrob. Agents Chemother. 2007, 51, 3471–3484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lin, M.F.; Lan, C.-Y. Antimicrobial resistance in Acinetobacter baumannii: From bench to bedside. World J. Clin. Cases 2014, 2, 787. [Google Scholar] [CrossRef] [PubMed]
  157. Lucidi, M.; Visaggio, D.; Prencipe, E.; Imperi, F.; Rampioni, G.; Cincotti, G.; Leoni, L.; Visca, P. New shuttle vectors for real-time gene expression analysis in multidrug-resistant Acinetobacter species: In vitro and in vivo responses to environmental stressors. Appl. Environ. Microbiol. 2019, 85, e01334-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Lima, P.G.; Oliveira, J.T.; Amaral, J.L.; Freitas, C.D.; Souza, P.F. Synthetic antimicrobial peptides: Characteristics, design, and potential as alternative molecules to overcome microbial resistance. Life Sci. 2021, 278, 119647. [Google Scholar] [CrossRef]
  159. Gallagher, L.A. Methods for Tn-Seq analysis in Acinetobacter baumannii. In Acinetobacter baumannii: Methods and Protocols; Humana: New York, NY, USA, 2019; pp. 115–134. [Google Scholar]
  160. Roy, R.; You, R.I.; Lin, M.-D.; Lin, N.-T. Mutation of the carboxy-terminal processing protease in Acinetobacter baumannii affects motility, leads to loss of membrane integrity, and reduces virulence. Pathogens 2020, 9, 322. [Google Scholar] [CrossRef]
  161. Sun, B.; Liu, H.; Jiang, Y.; Shao, L.; Yang, S.; Chen, D. New mutations involved in colistin resistance in Acinetobacter baumannii. mSphere 2020, 5, e00895-19. [Google Scholar] [CrossRef] [Green Version]
  162. Dehbanipour, R.; Ghalavand, Z. Anti-virulence therapeutic strategies against bacterial infections: Recent advances. Germs 2022, 12, 262–275. [Google Scholar] [CrossRef]
  163. De Silva, P.M.; Patidar, R.; Graham, C.I.; Brassinga, A.K.C.; Kumar, A. A response regulator protein with antar domain, avnr, in Acinetobacter baumannii ATCC 17978 impacts its virulence and amino acid metabolism. Microbiology 2020, 166, 554–566. [Google Scholar] [CrossRef]
  164. Godeux, A.S.; Svedholm, E.; Lupo, A.; Haenni, M.; Venner, S.; Laaberki, M.-H.; Charpentier, X. Scarless removal of large resistance island AbaR results in antibiotic susceptibility and increased natural transformability in Acinetobacter baumannii. Antimicrob. Agents Chemother. 2020, 64, e00951-20. [Google Scholar] [CrossRef]
  165. Biswas, I. Genetic tools for manipulating Acinetobacter baumannii genome: An overview. J. Med. Microbiol. 2015, 64, 657–669. [Google Scholar] [CrossRef]
  166. Fels, U.; Gevaert, K.; Van Damme, P. Bacterial genetic engineering by means of recombineering for reverse genetics. Front. Microbiol. 2020, 11, 548410. [Google Scholar] [CrossRef]
  167. Gordillo Altamirano, F.L.; Barr, J.J. Phage therapy in the postantibiotic era. Clin. Microbiol. Rev. 2019, 32, e00066-18. [Google Scholar] [CrossRef] [Green Version]
  168. Kortright, K.E.; Chan, B.K.; Koff, J.L.; Turner, P.E. Phage therapy: A renewed approach to combat antibiotic-resistant bacteria. Cell Host Microbe 2019, 25, 219–232. [Google Scholar] [CrossRef] [Green Version]
  169. Hetta, H.F.; Ramadan, Y.N.; Al-Harbi, A.I.; Ahmed, E.A.; Battah, B.; Abd Ellah, N.H.; Zanetti, S.; Donadu, M.G. Nanotechnology as a Promising Approach to Combat Multidrug Resistant Bacteria: A Comprehensive Review and Future Perspectives. Biomedicines 2023, 11, 413. [Google Scholar] [CrossRef]
  170. Makabenta, J.M.V.; Nabawy, A.; Li, C.-H.; Schmidt-Malan, S.; Patel, R.; Rotello, V.M. Nanomaterial-based therapeutics for antibiotic-resistant bacterial infections. Nat. Rev. Microbiol. 2021, 19, 23–36. [Google Scholar] [CrossRef]
  171. Munir, M.U.; Ahmad, M.M. Nanomaterials aiming to tackle antibiotic-resistant bacteria. Pharmaceutics 2022, 14, 582. [Google Scholar] [CrossRef]
  172. Egorov, A.; Ulyashova, M.; Rubtsova, M.Y. Bacterial enzymes and antibiotic resistance. Acta Nat. 2018, 10, 33–48. [Google Scholar] [CrossRef] [Green Version]
  173. Gontijo, A.V.L.; Pereira, S.L.; de Lacerda Bonfante, H. Can drug repurposing be effective against carbapenem-resistant Acinetobacter baumannii? Curr. Microbiol. 2022, 79, 13. [Google Scholar] [CrossRef]
  174. Koh Jing Jie, A.; Hussein, M.; Rao, G.G.; Li, J.; Velkov, T. Drug Repurposing Approaches towards Defeating Multidrug-Resistant Gram-Negative Pathogens: Novel Polymyxin/Non-Antibiotic Combinations. Pathogens 2022, 11, 1420. [Google Scholar] [CrossRef]
  175. Liu, Y.; Tong, Z.; Shi, J.; Li, R.; Upton, M.; Wang, Z. Drug repurposing for next-generation combination therapies against multidrug-resistant bacteria. Theranostics 2021, 11, 4910. [Google Scholar] [CrossRef] [PubMed]
  176. Buchy, P.; Ascioglu, S.; Buisson, Y.; Datta, S.; Nissen, M.; Tambyah, P.A.; Vong, S. Impact of vaccines on antimicrobial resistance. Int. J. Infect. Dis. 2020, 90, 188–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Micoli, F.; Bagnoli, F.; Rappuoli, R.; Serruto, D. The role of vaccines in combatting antimicrobial resistance. Nat. Rev. Microbiol. 2021, 19, 287–302. [Google Scholar] [CrossRef] [PubMed]
  178. Rosini, R.; Nicchi, S.; Pizza, M.; Rappuoli, R. Vaccines against antimicrobial resistance. Front. Immunol. 2020, 11, 1048. [Google Scholar] [CrossRef]
  179. Kole, R.; Krainer, A.R.; Altman, S. RNA therapeutics: Beyond RNA interference and antisense oligonucleotides. Nat. Rev. Drug Discov. 2012, 11, 125–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Murugaiyan, J.; Kumar, P.A.; Rao, G.S.; Iskandar, K.; Hawser, S.; Hays, J.P.; Mohsen, Y.; Adukkadukkam, S.; Awuah, W.A.; Jose, R.A.M. Progress in alternative strategies to combat antimicrobial resistance: Focus on antibiotics. Antibiotics 2022, 11, 200. [Google Scholar] [CrossRef] [PubMed]
  181. Good, L.; Stach, J.E. Synthetic RNA silencing in bacteria–antimicrobial discovery and resistance breaking. Front. Microbiol. 2011, 2, 185. [Google Scholar] [CrossRef] [Green Version]
  182. Kotil, S.; Jakobsson, E. Rationally designing antisense therapy to keep up with evolving bacterial resistance. PLoS ONE 2019, 14, e0209894. [Google Scholar] [CrossRef]
  183. Gaj, T.; Sirk, S.J.; Shui, S.-l.; Liu, J. Genome-editing technologies: Principles and applications. Cold Spring Harb. Perspect. Biol. 2016, 8, a023754. [Google Scholar] [CrossRef] [Green Version]
  184. Li, H.; Yang, Y.; Hong, W.; Huang, M.; Wu, M.; Zhao, X. Applications of genome editing technology in the targeted therapy of human diseases: Mechanisms, advances and prospects. Signal Transduct. Target. Ther. 2020, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  185. Liu, Y.; Zhao, H.; Cheng, C.H. Mutagenesis in Xenopus and zebrafish using TALENs. In TALENs: Methods and Protocols; Humana: New York, NY, USA, 2016; pp. 207–227. [Google Scholar]
  186. Zhang, H.-X.; Zhang, Y.; Yin, H. Genome editing with mRNA encoding ZFN, TALEN, and Cas9. Mol. Ther. 2019, 27, 735–746. [Google Scholar] [CrossRef] [Green Version]
  187. Hille, F.; Charpentier, E. CRISPR-Cas: Biology, mechanisms and relevance. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2016, 371, 20150496. [Google Scholar] [CrossRef] [Green Version]
  188. Ishino, Y.; Krupovic, M.; Forterre, P. History of CRISPR-Cas from encounter with a mysterious repeated sequence to genome editing technology. J. Bacteriol. 2018, 200, e00580-17. [Google Scholar] [CrossRef] [Green Version]
  189. Ibrahim, A.; ÖZSÖZ, M.; Tirah, G.; Gideon, O. Genome engineering using the CRISPR Cas9 system. J. Biomed. Pharm. Sci. 2019, 2, 2. [Google Scholar]
  190. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA–guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  191. O’Connell, M.R.; Oakes, B.L.; Sternberg, S.H.; East-Seletsky, A.; Kaplan, M.; Doudna, J.A. Programmable RNA recognition and cleavage by CRISPR/Cas9. Nature 2014, 516, 263–266. [Google Scholar] [CrossRef] [Green Version]
  192. Adli, M. The CRISPR tool kit for genome editing and beyond. Nat. Commun. 2018, 9, 1911. [Google Scholar] [CrossRef] [Green Version]
  193. Rath, D.; Amlinger, L.; Rath, A.; Lundgren, M. The CRISPR-Cas immune system: Biology, mechanisms and applications. Biochimie 2015, 117, 119–128. [Google Scholar] [CrossRef]
  194. Palacios Araya, D.; Palmer, K.L.; Duerkop, B.A. CRISPR-based antimicrobials to obstruct antibiotic-resistant and pathogenic bacteria. PLoS Pathog. 2021, 17, e1009672. [Google Scholar] [CrossRef]
  195. Yeh, T.-K.; Jean, S.-S.; Lee, Y.-L.; Lu, M.-C.; Ko, W.-C.; Lin, H.-J.; Liu, P.-Y.; Hsueh, P.-R. Bacteriophages and phage-delivered CRISPR-Cas system as antibacterial therapy. Int. J. Antimicrob. Agents 2022, 59, 106475. [Google Scholar] [CrossRef]
  196. Chen, Y.; Wen, R.; Yang, Z.; Chen, Z. Genome editing using CRISPR/Cas9 to treat hereditary hematological disorders. Gene Ther. 2022, 29, 207–216. [Google Scholar] [CrossRef] [PubMed]
  197. Mani, I. CRISPR-Cas9 for treating hereditary diseases. Prog. Mol. Biol. Transl. Sci. 2021, 181, 165–183. [Google Scholar] [PubMed]
  198. Fuziwara, C.S.; de Mello, D.C.; Kimura, E.T. Gene Editing with CRISPR/Cas Methodology and Thyroid Cancer: Where Are We? Cancers 2022, 14, 844. [Google Scholar] [CrossRef] [PubMed]
  199. Kim, T.H.; Lee, S.-W. Therapeutic Application of Genome Editing Technologies in Viral Diseases. Int. J. Mol. Sci. 2022, 23, 5399. [Google Scholar] [CrossRef]
  200. Makarova, K.S.; Wolf, Y.I.; Iranzo, J.; Shmakov, S.A.; Alkhnbashi, O.S.; Brouns, S.J.; Charpentier, E.; Cheng, D.; Haft, D.H.; Horvath, P. Evolutionary classification of CRISPR–Cas systems: A burst of class 2 and derived variants. Nat. Rev. Microbiol. 2020, 18, 67–83. [Google Scholar] [CrossRef]
  201. Wan, F.; Draz, M.S.; Gu, M.; Yu, W.; Ruan, Z.; Luo, Q. Novel strategy to combat antibiotic resistance: A sight into the combination of CRISPR/Cas9 and nanoparticles. Pharmaceutics 2021, 13, 352. [Google Scholar] [CrossRef]
  202. Hidalgo-Cantabrana, C.; Goh, Y.J.; Barrangou, R. Characterization and repurposing of type I and type II CRISPR–Cas systems in bacteria. J. Mol. Biol. 2019, 431, 21–33. [Google Scholar] [CrossRef]
  203. Cady, K.; White, A.; Hammond, J.; Abendroth, M.; Karthikeyan, R.; Lalitha, P.; Zegans, M.; O’Toole, G. Prevalence, conservation and functional analysis of Yersinia and Escherichia CRISPR regions in clinical Pseudomonas aeruginosa isolates. Microbiology 2011, 157, 430. [Google Scholar] [CrossRef] [Green Version]
  204. Gholizadeh, P.; Köse, Ş.; Dao, S.; Ganbarov, K.; Tanomand, A.; Dal, T.; Aghazadeh, M.; Ghotaslou, R.; Ahangarzadeh Rezaee, M.; Yousefi, B. How CRISPR-Cas system could be used to combat antimicrobial resistance. Infect. Drug Resist. 2020, 13, 1111–1121. [Google Scholar] [CrossRef] [Green Version]
  205. Zhang, S.; Shen, J.; Li, D.; Cheng, Y. Strategies in the delivery of Cas9 ribonucleoprotein for CRISPR/Cas9 genome editing. Theranostics 2021, 11, 614. [Google Scholar] [CrossRef]
  206. Aslam, B.; Rasool, M.; Idris, A.; Muzammil, S.; Alvi, R.F.; Khurshid, M.; Rasool, M.H.; Zhang, D.; Ma, Z.; Baloch, Z. CRISPR-Cas system: A potential alternative tool to cope antibiotic resistance. Antimicrob. Resist. Infect. Control 2020, 9, 131. [Google Scholar] [CrossRef]
  207. Barrangou, R.; Ousterout, D. Repurposing CRISPR-Cas systems as DNA-based smart antimicrobials. Cell Gene Ther. Insights 2017, 3, 63–72. [Google Scholar] [CrossRef]
  208. Serajian, S.; Ahmadpour, E.; Oliveira, S.M.R.; Pereira, M.d.L.; Heidarzadeh, S. CRISPR-cas technology: Emerging applications in clinical microbiology and infectious diseases. Pharmaceuticals 2021, 14, 1171. [Google Scholar] [CrossRef]
  209. Duan, C.; Cao, H.; Zhang, L.-H.; Xu, Z. Harnessing the CRISPR-Cas systems to combat antimicrobial resistance. Front. Microbiol. 2021, 12, 716064. [Google Scholar] [CrossRef]
  210. Gleerup, J.L.; Mogensen, T.H. CRISPR-Cas in diagnostics and therapy of infectious diseases. J. Infect. Dis. 2022, 226, 1867–1876. [Google Scholar] [CrossRef]
  211. Shim, H. Investigating the genomic background of CRISPR-Cas genomes for CRISPR-based antimicrobials. Evol. Bioinform. 2022, 18, 11769343221103887. [Google Scholar] [CrossRef]
  212. Shabbir, M.A.B.; Shabbir, M.Z.; Wu, Q.; Mahmood, S.; Sajid, A.; Maan, M.K.; Ahmed, S.; Naveed, U.; Hao, H.; Yuan, Z. CRISPR-cas system: Biological function in microbes and its use to treat antimicrobial resistant pathogens. Ann. Clin. Microbiol. Antimicrob. 2019, 18, 21. [Google Scholar] [CrossRef] [Green Version]
  213. Ekwebelem, O.C.; Aleke, J.; Ofielu, E.; Nnorom-Dike, O. CRISPR-Cas9 system: A revolutionary tool in the fight against antimicrobial resistance. Infect. Microbes Dis. 2021, 3, 51–56. [Google Scholar] [CrossRef]
  214. Getahun, Y.A.; Ali, D.A.; Taye, B.W.; Alemayehu, Y.A. Multidrug-Resistant Microbial Therapy Using Antimicrobial Peptides and the CRISPR/Cas9 System. Vet. Med. Res. Rep. 2022, 13, 173–190. [Google Scholar] [CrossRef]
  215. Greene, A.C. CRISPR-based antibacterials: Transforming bacterial defense into offense. Trends Biotechnol. 2018, 36, 127–130. [Google Scholar] [CrossRef]
  216. Li, Y.; Peng, N. Endogenous CRISPR-Cas system-based genome editing and antimicrobials: Review and prospects. Front. Microbiol. 2019, 10, 2471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Alavi, M.; Rai, M. Antisense RNA, the modified CRISPR-Cas9, and metal/metal oxide nanoparticles to inactivate pathogenic bacteria. Cell. Mol. Biomed. Rep. 2021, 1, 52–59. [Google Scholar] [CrossRef]
  218. Cañez, C.; Selle, K.; Goh, Y.J.; Barrangou, R. Outcomes and characterization of chromosomal self-targeting by native CRISPR-Cas systems in Streptococcus thermophilus. FEMS Microbiol. Lett. 2019, 366, fnz105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Citorik, R.J.; Mimee, M.; Lu, T.K. Sequence-specific antimicrobials using efficiently delivered RNA-guided nucleases. Nat. Biotechnol. 2014, 32, 1141–1145. [Google Scholar] [CrossRef] [Green Version]
  220. Kim, J.-S.; Cho, D.-H.; Park, M.; Chung, W.-J.; Shin, D.; Ko, K.S.; Kweon, D.-H. CRISPR/Cas9-mediated re-sensitization of antibiotic-resistant Escherichia coli harboring extended-spectrum β-lactamases. J. Microbiol. Biotechnol. 2016, 26, 394–401. [Google Scholar] [CrossRef]
  221. Wang, Y.; Wang, D.; Wang, X.; Tao, H.; Feng, E.; Zhu, L.; Pan, C.; Wang, B.; Liu, C.; Liu, X. Highly efficient genome engineering in Bacillus anthracis and Bacillus cereus using the CRISPR/Cas9 system. Front. Microbiol. 2019, 10, 1932. [Google Scholar] [CrossRef] [Green Version]
  222. Gomaa, A.A.; Klumpe, H.E.; Luo, M.L.; Selle, K.; Barrangou, R.; Beisel, C.L. Programmable removal of bacterial strains by use of genome-targeting CRISPR-Cas systems. mBio 2014, 5, e00928-13. [Google Scholar] [CrossRef] [Green Version]
  223. Liu, Y.; Wan, X.; Wang, B. Engineered CRISPRa enables programmable eukaryote-like gene activation in bacteria. Nat. Commun. 2019, 10, 3693. [Google Scholar] [CrossRef] [Green Version]
  224. Bikard, D.; Euler, C.W.; Jiang, W.; Nussenzweig, P.M.; Goldberg, G.W.; Duportet, X.; Fischetti, V.A.; Marraffini, L.A. Exploiting CRISPR-Cas nucleases to produce sequence-specific antimicrobials. Nat. Biotechnol. 2014, 32, 1146–1150. [Google Scholar] [CrossRef] [Green Version]
  225. Chen, W.; Zhang, Y.; Zhang, Y.; Pi, Y.; Gu, T.; Song, L.; Wang, Y.; Ji, Q. CRISPR/Cas9-based genome editing in Pseudomonas aeruginosa and cytidine deaminase-mediated base editing in Pseudomonas species. iScience 2018, 6, 222–231. [Google Scholar] [CrossRef] [Green Version]
  226. Li, Q.; Sun, B.; Chen, J.; Zhang, Y.; Jiang, Y.; Yang, S. A modified pCas/pTargetF system for CRISPR-Cas9-assisted genome editing in Escherichia coli. Acta Biochim. Biophys. Sin. 2021, 53, 620–627. [Google Scholar] [CrossRef]
  227. Savage, D.F. Cas14: Big advances from small CRISPR proteins. Biochemistry 2019, 58, 1024–1025. [Google Scholar] [CrossRef] [Green Version]
  228. Wu, X.; Zha, J.; Koffas, M.A.; Dordick, J.S. Reducing Staphylococcus aureus resistance to lysostaphin using CRISPR-dCas9. Biotechnol. Bioeng. 2019, 116, 3149–3159. [Google Scholar] [CrossRef]
  229. Wang, Y.; Wang, S.; Chen, W.; Song, L.; Zhang, Y.; Shen, Z.; Yu, F.; Li, M.; Ji, Q. CRISPR-Cas9 and CRISPR-assisted cytidine deaminase enable precise and efficient genome editing in Klebsiella pneumoniae. Appl. Environ. Microbiol. 2018, 84, e01834-18. [Google Scholar] [CrossRef] [Green Version]
  230. Aquino-Jarquin, G. CRISPR-Cas14 is now part of the artillery for gene editing and molecular diagnostic. Nanomed. Nanotechnol. Biol. Med. 2019, 18, 428–431. [Google Scholar] [CrossRef]
  231. Uribe, R.V.; Rathmer, C.; Jahn, L.J.; Ellabaan, M.M.H.; Li, S.S.; Sommer, M.O.A. Bacterial resistance to CRISPR-Cas antimicrobials. Sci. Rep. 2021, 11, 17267. [Google Scholar] [CrossRef]
  232. Wu, Y.; Battalapalli, D.; Hakeem, M.J.; Selamneni, V.; Zhang, P.; Draz, M.S.; Ruan, Z. Engineered CRISPR-Cas systems for the detection and control of antibiotic-resistant infections. J. Nanobiotechnol. 2021, 19, 401. [Google Scholar] [CrossRef]
  233. Song, Z.; Yu, Y.; Bai, X.; Jia, Y.; Tian, J.; Gu, K.; Zhao, M.; Zhou, C.; Zhang, X.; Wang, H. Pathogen-Specific Bactericidal Method Mediated by Conjugative Delivery of CRISPR-Cas13a Targeting Bacterial Endogenous Transcripts. Microbiol. Spectr. 2022, 10, e01300-22. [Google Scholar] [CrossRef]
  234. Tagliaferri, T.L.; Guimarães, N.R.; Pereira, M.D.P.M.; Vilela, L.F.F.; Horz, H.P.; Dos Santos, S.G.; Mendes, T.A.D.O. Exploring the potential of CRISPR-Cas9 under challenging conditions: Facing high-copy plasmids and counteracting beta-lactam resistance in clinical strains of Enterobacteriaceae. Front. Microbiol. 2020, 11, 578. [Google Scholar] [CrossRef]
  235. Wongpayak, P.; Meesungnoen, O.; Saejang, S.; Subsoontorn, P. A highly effective and self-transmissible CRISPR antimicrobial for elimination of target plasmids without antibiotic selection. PeerJ 2021, 9, e11996. [Google Scholar] [CrossRef]
  236. Wang, P.; He, D.; Li, B.; Guo, Y.; Wang, W.; Luo, X.; Zhao, X.; Wang, X. Eliminating mcr-1-harbouring plasmids in clinical isolates using the CRISPR/Cas9 system. J. Antimicrob. Chemother. 2019, 74, 2559–2565. [Google Scholar] [CrossRef] [PubMed]
  237. Kang, Y.K.; Kwon, K.; Ryu, J.S.; Lee, H.N.; Park, C.; Chung, H.J. Nonviral genome editing based on a polymer-derivatized CRISPR nanocomplex for targeting bacterial pathogens and antibiotic resistance. Bioconjug. Chem. 2017, 28, 957–967. [Google Scholar] [CrossRef] [PubMed]
  238. Wan, P.; Cui, S.; Ma, Z.; Chen, L.; Li, X.; Zhao, R.; Xiong, W.; Zeng, Z. Reversal of mcr-1-mediated colistin resistance in Escherichia coli by CRISPR-Cas9 system. Infect. Drug Resist. 2020, 13, 1171–1178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Nath, A.; Bhattacharjee, R.; Nandi, A.; Sinha, A.; Kar, S.; Manoharan, N.; Mitra, S.; Mojumdar, A.; Panda, P.K.; Patro, S. Phage delivered CRISPR-Cas system to combat multidrug-resistant pathogens in gut microbiome. Biomed. Pharmacother. 2022, 151, 113122. [Google Scholar] [CrossRef] [PubMed]
  240. Neil, K.; Allard, N.; Roy, P.; Grenier, F.; Menendez, A.; Burrus, V.; Rodrigue, S. High-efficiency delivery of CRISPR-Cas9 by engineered probiotics enables precise microbiome editing. Mol. Syst. Biol. 2021, 17, e10335. [Google Scholar] [CrossRef]
  241. Price, V.J.; McBride, S.W.; Hullahalli, K.; Chatterjee, A.; Duerkop, B.A.; Palmer, K.L. Enterococcus faecalis CRISPR-Cas is a robust barrier to conjugative antibiotic resistance dissemination in the murine intestine. mSphere 2019, 4, e00464-19. [Google Scholar] [CrossRef] [Green Version]
  242. Rodrigues, M.; McBride, S.W.; Hullahalli, K.; Palmer, K.L.; Duerkop, B.A. Conjugative delivery of CRISPR-Cas9 for the selective depletion of antibiotic-resistant enterococci. Antimicrob. Agents Chemother. 2019, 63, e01454-19. [Google Scholar] [CrossRef]
  243. Hao, M.; He, Y.; Zhang, H.; Liao, X.-P.; Liu, Y.-H.; Sun, J.; Du, H.; Kreiswirth, B.N.; Chen, L. CRISPR-Cas9-mediated carbapenemase gene and plasmid curing in carbapenem-resistant Enterobacteriaceae. Antimicrob. Agents Chemother. 2020, 64, e00843-20. [Google Scholar] [CrossRef]
  244. He, Y.-Z.; Kuang, X.; Long, T.-F.; Li, G.; Ren, H.; He, B.; Yan, J.-R.; Liao, X.-P.; Liu, Y.-H.; Chen, L. Re-engineering a mobile-CRISPR/Cas9 system for antimicrobial resistance gene curing and immunization in Escherichia coli. J. Antimicrob. Chemother. 2022, 77, 74–82. [Google Scholar] [CrossRef]
  245. Mackow, N.A.; Shen, J.; Adnan, M.; Khan, A.S.; Fries, B.C.; Diago-Navarro, E. CRISPR-Cas influences the acquisition of antibiotic resistance in Klebsiella pneumoniae. PLoS ONE 2019, 14, e0225131. [Google Scholar] [CrossRef] [Green Version]
  246. Yao, S.; Wei, D.; Tang, N.; Song, Y.; Wang, C.; Feng, J.; Zhang, G. Efficient Suppression of Natural Plasmid-Borne Gene Expression in Carbapenem-Resistant Klebsiella pneumoniae Using a Compact CRISPR Interference System. Antimicrob. Agents Chemother. 2022, 66, e00890-22. [Google Scholar] [CrossRef]
  247. Kang, S.; Kim, J.; Hur, J.K.; Lee, S.-S. CRISPR-based genome editing of clinically important Escherichia coli SE15 isolated from indwelling urinary catheters of patients. J. Med. Microbiol. 2017, 66, 18–25. [Google Scholar] [CrossRef]
  248. Zuberi, A.; Ahmad, N.; Khan, A.U. CRISPRi induced suppression of fimbriae gene (fimH) of a uropathogenic Escherichia coli: An approach to inhibit microbial biofilms. Front. Immunol. 2017, 8, 1552. [Google Scholar] [CrossRef] [Green Version]
  249. Boettcher, M.; McManus, M.T. Choosing the right tool for the job: RNAi, TALEN, or CRISPR. Mol. Cell 2015, 58, 575–585. [Google Scholar] [CrossRef] [Green Version]
  250. Chen, C.; Choudhury, A.; Zhang, S.; Garst, A.D.; Song, X.; Liu, X.; Chen, T.; Gill, R.T.; Wang, Z. Integrating CRISPR-enabled trackable genome engineering and transcriptomic analysis of global regulators for antibiotic resistance selection and identification in Escherichia coli. Msystems 2020, 5, e00232-20. [Google Scholar] [CrossRef]
  251. Li, Q.; Zhao, P.; Li, L.; Zhao, H.; Shi, L.; Tian, P. Engineering a CRISPR interference system to repress a class 1 integron in Escherichia coli. Antimicrob. Agents Chemother. 2020, 64, e01789-19. [Google Scholar] [CrossRef] [Green Version]
  252. Wan, X.; Li, Q.; Olsen, R.H.; Meng, H.; Zhang, Z.; Wang, J.; Zheng, H.; Li, L.; Shi, L. Engineering a CRISPR interference system targeting AcrAB-TolC efflux pump to prevent multidrug resistance development in Escherichia coli. J. Antimicrob. Chemother. 2022, 77, 2158–2166. [Google Scholar] [CrossRef]
  253. Zuberi, A.; Misba, L.; Khan, A.U. CRISPR interference (CRISPRi) inhibition of luxS gene expression in E. coli: An approach to inhibit biofilm. Front. Cell. Infect. Microbiol. 2017, 7, 214. [Google Scholar] [CrossRef] [Green Version]
  254. Afonina, I.; Ong, J.; Chua, J.; Lu, T.; Kline, K.A. Multiplex CRISPRi system enables the study of stage-specific biofilm genetic requirements in Enterococcus faecalis. mBio 2020, 11, e01101-20. [Google Scholar] [CrossRef]
  255. Noirot-Gros, M.F.; Forrester, S.; Malato, G.; Larsen, P.E.; Noirot, P. CRISPR interference to interrogate genes that control biofilm formation in Pseudomonas fluorescens. Sci. Rep. 2019, 9, 15954. [Google Scholar] [CrossRef] [Green Version]
  256. Guzzo, M.; Castro, L.K.; Reisch, C.R.; Guo, M.S.; Laub, M.T. A CRISPR interference system for efficient and rapid gene knockdown in Caulobacter crescentus. mBio 2020, 11, e02415-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Coates, R.; Grant, A.; Floto, A.; Parkhill, J. Development of a CRISPR Interference System in Campylobacter jejuni. Ph.D. Dissertation, University of Cambridge, Cambridge, UK, 2022. [Google Scholar]
  258. Costigan, R.; Stoakes, E.; Floto, R.A.; Parkhill, J.; Grant, A.J. Development and validation of a CRISPR interference system for gene regulation in Campylobacter jejuni. BMC Microbiol. 2022, 22, 238. [Google Scholar] [CrossRef] [PubMed]
  259. Dolgin, E. The kill-switch for CRISPR that could make gene-editing safer. Nature 2020, 577, 308–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Kundar, R.; Gokarn, K. CRISPR-Cas System: A Tool to Eliminate Drug-Resistant Gram-Negative Bacteria. Pharmaceuticals 2022, 15, 1498. [Google Scholar] [CrossRef] [PubMed]
  261. Tong, Y.; Charusanti, P.; Zhang, L.; Weber, T.; Lee, S.Y. CRISPR-Cas9 based engineering of actinomycetal genomes. ACS Synth. Biol. 2015, 4, 1020–1029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Wang, Y.; Wang, Z.; Chen, Y.; Hua, X.; Yu, Y.; Ji, Q. A highly efficient CRISPR-Cas9-based genome engineering platform in Acinetobacter baumannii to understand the H2O2-sensing mechanism of OxyR. Cell Chem. Biol. 2019, 26, 1732–1742.e1735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Wang, Y.; Yang, J.; Sun, X.; Li, M.; Zhang, P.; Zhu, Z.; Jiao, H.; Guo, T.; Li, G. CRISPR-Cas in Acinetobacter baumannii Contributes to Antibiotic Susceptibility by Targeting Endogenous AbaI. Microbiol. Spectr. 2022, 10, e00829-22. [Google Scholar] [CrossRef]
  264. Bai, J.; Dai, Y.; Farinha, A.; Tang, A.Y.; Syal, S.; Vargas-Cuebas, G.; van Opijnen, T.; Isberg, R.R.; Geisinger, E. Essential gene analysis in Acinetobacter baumannii by high-density transposon mutagenesis and CRISPR interference. J. Bacteriol. 2021, 203, e00565-20. [Google Scholar] [CrossRef]
  265. Wang, Y.; Guo, Y.; Zhang, L.; Yang, Y.; Yang, S.; Yang, L.; Chen, H.; Liu, C.; Li, J.; Xie, G. Integration of multiplex PCR and CRISPR-Cas allows highly specific detection of multidrug-resistant Acinetobacter baumannii. Sens. Actuators B Chem. 2021, 334, 129600. [Google Scholar] [CrossRef]
  266. Wu, S.; Xu, R.; Su, M.; Gao, C.; Liu, Y.; Chen, Y.; Luan, G.; Jia, X.; Wang, R. A pyrF-Based Efficient Genetic Manipulation Platform in Acinetobacter baumannii to Explore the Vital DNA Components of Adaptive Immunity for IF CRISPR-Cas. Microbiol. Spectr. 2022, 10, e01957-22. [Google Scholar] [CrossRef]
  267. Wolf, T.; Gren, T.; Thieme, E.; Wibberg, D.; Zemke, T.; Pühler, A.; Kalinowski, J. Targeted genome editing in the rare actinomycete Actinoplanes sp. SE50/110 by using the CRISPR/Cas9 system. J. Biotechnol. 2016, 231, 122–128. [Google Scholar] [CrossRef]
  268. Bernheim, A.; Calvo-Villamañán, A.; Basier, C.; Cui, L.; Rocha, E.P.; Touchon, M.; Bikard, D. Inhibition of NHEJ repair by type II-A CRISPR-Cas systems in bacteria. Nat. Commun. 2017, 8, 2094. [Google Scholar] [CrossRef] [Green Version]
  269. Peters, J.M.; Colavin, A.; Shi, H.; Czarny, T.L.; Larson, M.H.; Wong, S.; Hawkins, J.S.; Lu, C.H.; Koo, B.-M.; Marta, E. A comprehensive, CRISPR-based functional analysis of essential genes in bacteria. Cell 2016, 165, 1493–1506. [Google Scholar] [CrossRef] [Green Version]
  270. Zhang, K.; Duan, X.; Wu, J. Multigene disruption in undomesticated Bacillus subtilis ATCC 6051a using the CRISPR/Cas9 system. Sci. Rep. 2016, 6, 27943. [Google Scholar] [CrossRef] [Green Version]
  271. Altenbuchner, J. Editing of the Bacillus subtilis genome by the CRISPR-Cas9 system. Appl. Environ. Microbiol. 2016, 82, 5421–5427. [Google Scholar] [CrossRef] [Green Version]
  272. Mougiakos, I.; Bosma, E.F.; Weenink, K.; Vossen, E.; Goijvaerts, K.; van der Oost, J.; van Kranenburg, R. Efficient genome editing of a facultative thermophile using mesophilic spCas9. ACS Synth. Biol. 2017, 6, 849–861. [Google Scholar] [CrossRef] [Green Version]
  273. Mougiakos, I.; Mohanraju, P.; Bosma, E.F.; Vrouwe, V.; Finger Bou, M.; Naduthodi, M.I.; Gussak, A.; Brinkman, R.B.; Van Kranenburg, R.; Van Der Oost, J. Characterizing a thermostable Cas9 for bacterial genome editing and silencing. Nat. Commun. 2017, 8, 1647. [Google Scholar] [CrossRef] [Green Version]
  274. Li, K.; Cai, D.; Wang, Z.; He, Z.; Chen, S. Development of an efficient genome editing tool in Bacillus licheniformis using CRISPR-Cas9 nickase. Appl. Environ. Microbiol. 2018, 84, e02608-17. [Google Scholar] [CrossRef] [Green Version]
  275. Zheng, K.; Wang, Y.; Li, N.; Jiang, F.-F.; Wu, C.-X.; Liu, F.; Chen, H.-C.; Liu, Z.-F. Highly efficient base editing in bacteria using a Cas9-cytidine deaminase fusion. Commun. Biol. 2018, 1, 32. [Google Scholar] [CrossRef] [Green Version]
  276. Wasels, F.; Jean-Marie, J.; Collas, F.; López-Contreras, A.M.; Ferreira, N.L. A two-plasmid inducible CRISPR/Cas9 genome editing tool for Clostridium acetobutylicum. J. Microbiol. Methods 2017, 140, 5–11. [Google Scholar] [CrossRef]
  277. Bruder, M.R.; Pyne, M.E.; Moo-Young, M.; Chung, D.A.; Chou, C.P. Extending CRISPR-Cas9 technology from genome editing to transcriptional engineering in the genus Clostridium. Appl. Environ. Microbiol. 2016, 82, 6109–6119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Nagaraju, S.; Davies, N.K.; Walker, D.J.F.; Köpke, M.; Simpson, S.D. Genome editing of Clostridium autoethanogenum using CRISPR/Cas9. Biotechnol. Biofuels 2016, 9, 219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Li, Q.; Chen, J.; Minton, N.P.; Zhang, Y.; Wen, Z.; Liu, J.; Yang, H.; Zeng, Z.; Ren, X.; Yang, J. CRISPR-based genome editing and expression control systems in Clostridium acetobutylicum and Clostridium beijerinckii. Biotechnol. J. 2016, 11, 961–972. [Google Scholar] [CrossRef] [PubMed]
  280. Wang, Y.; Zhang, Z.T.; Seo, S.O.; Lynn, P.; Lu, T.; Jin, Y.S.; Blaschek, H.P. Bacterial genome editing with CRISPR-Cas9: Deletion, integration, single nucleotide modification, and desirable “clean” mutant selection in Clostridium beijerinckii as an example. ACS Synth. Biol. 2016, 5, 721–732. [Google Scholar] [CrossRef] [PubMed]
  281. Wang, Y.; Zhang, Z.T.; Seo, S.O.; Lynn, P.; Lu, T.; Jin, Y.S.; Blaschek, H.P. Gene transcription repression in Clostridium beijerinckii using CRISPR-dCas9. Biotechnol. Bioeng. 2016, 113, 2739–2743. [Google Scholar] [CrossRef]
  282. Negahdaripour, M.; Nezafat, N.; Hajighahramani, N.; Rahmatabadi, S.S.; Ghasemi, Y. Investigating CRISPR-Cas systems in Clostridium botulinum via bioinformatics tools. Infect. Genet. Evol. 2017, 54, 355–373. [Google Scholar] [CrossRef] [PubMed]
  283. Xu, T.; Li, Y.; Shi, Z.; Hemme, C.L.; Li, Y.; Zhu, Y.; Van Nostrand, J.D.; He, Z.; Zhou, J. Efficient genome editing in Clostridium cellulolyticum via CRISPR-Cas9 nickase. Appl. Environ. Microbiol. 2015, 81, 4423–4431. [Google Scholar] [CrossRef] [Green Version]
  284. Hong, W.; Zhang, J.; Cui, G.; Wang, L.; Wang, Y. Multiplexed CRISPR-Cpf1-mediated genome editing in Clostridium difficile toward the understanding of pathogenesis of C. difficile infection. ACS Synth. Biol. 2018, 7, 1588–1600. [Google Scholar] [CrossRef]
  285. McAllister, K.N.; Bouillaut, L.; Kahn, J.N.; Self, W.T.; Sorg, J.A. Using CRISPR-Cas9-mediated genome editing to generate C. difficile mutants defective in selenoproteins synthesis. Sci. Rep. 2017, 7, 14672. [Google Scholar] [CrossRef] [Green Version]
  286. Wang, S.; Dong, S.; Wang, P.; Tao, Y.; Wang, Y. Genome editing in Clostridium saccharoperbutylacetonicum N1-4 with the CRISPR-Cas9 system. Appl. Environ. Microbiol. 2017, 83, e00233-17. [Google Scholar] [CrossRef] [Green Version]
  287. Pyne, M.E.; Bruder, M.; Moo-Young, M.; Chung, D.; Chou, C.P. Harnessing Heterologous and Endogenous CRISPR-Cas Machineries for Efficient Markerless Genome Editing in Clostridium. U.S. Patent 16/098,035, 16 May 2019. [Google Scholar]
  288. Park, J.; Shin, H.; Lee, S.-M.; Um, Y.; Woo, H.M. RNA-guided single/double gene repressions in Corynebacterium glutamicum using an efficient CRISPR interference and its application to industrial strain. Microb. Cell Factories 2018, 17, 4. [Google Scholar] [CrossRef] [Green Version]
  289. Wang, Y.; Liu, Y.; Liu, J.; Guo, Y.; Fan, L.; Ni, X.; Zheng, X.; Wang, M.; Zheng, P.; Sun, J. MACBETH: Multiplex automated Corynebacterium glutamicum base editing method. Metab. Eng. 2018, 47, 200–210. [Google Scholar] [CrossRef]
  290. Liu, J.; Wang, Y.; Lu, Y.; Zheng, P.; Sun, J.; Ma, Y. Development of a CRISPR/Cas9 genome editing toolbox for Corynebacterium glutamicum. Microb. Cell Factories 2017, 16, 205. [Google Scholar] [CrossRef] [Green Version]
  291. Jiang, Y.; Qian, F.; Yang, J.; Liu, Y.; Dong, F.; Xu, C.; Sun, B.; Chen, B.; Xu, X.; Li, Y. CRISPR-Cpf1 assisted genome editing of Corynebacterium glutamicum. Nat. Commun. 2017, 8, 15179. [Google Scholar] [CrossRef] [Green Version]
  292. Peng, F.; Wang, X.; Sun, Y.; Dong, G.; Yang, Y.; Liu, X.; Bai, Z. Efficient gene editing in Corynebacterium glutamicum using the CRISPR/Cas9 system. Microb. Cell Factories 2017, 16, 201. [Google Scholar] [CrossRef] [Green Version]
  293. Cho, J.S.; Choi, K.R.; Prabowo, C.P.S.; Shin, J.H.; Yang, D.; Jang, J.; Lee, S.Y. CRISPR/Cas9-coupled recombineering for metabolic engineering of Corynebacterium glutamicum. Metab. Eng. 2017, 42, 157–167. [Google Scholar] [CrossRef]
  294. Cleto, S.; Jensen, J.V.; Wendisch, V.F.; Lu, T.K. Corynebacterium glutamicum metabolic engineering with CRISPR interference (CRISPRi). ACS Synth. Biol. 2016, 5, 375–385. [Google Scholar] [CrossRef]
  295. Zhu, X.; Zhao, D.; Qiu, H.; Fan, F.; Man, S.; Bi, C.; Zhang, X. The CRISPR/Cas9-facilitated multiplex pathway optimization (CFPO) technique and its application to improve the Escherichia coli xylose utilization pathway. Metab. Eng. 2017, 43, 37–45. [Google Scholar] [CrossRef]
  296. Lv, L.; Ren, Y.-L.; Chen, J.-C.; Wu, Q.; Chen, G.-Q. Application of CRISPRi for prokaryotic metabolic engineering involving multiple genes, a case study: Controllable P (3HB-co-4HB) biosynthesis. Metab. Eng. 2015, 29, 160–168. [Google Scholar] [CrossRef]
  297. Yan, M.Y.; Yan, H.Q.; Ren, G.-X.; Zhao, J.P.; Guo, X.P.; Sun, Y.-C. CRISPR-Cas12a-assisted recombineering in bacteria. Appl. Environ. Microbiol. 2017, 83, e00947-17. [Google Scholar] [CrossRef] [Green Version]
  298. Hao, N.; Shearwin, K.E.; Dodd, I.B. Programmable DNA looping using engineered bivalent dCas9 complexes. Nat. Commun. 2017, 8, 1628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Ao, X.; Yao, Y.; Li, T.; Yang, T.T.; Dong, X.; Zheng, Z.T.; Chen, G.Q.; Wu, Q.; Guo, Y. A multiplex genome editing method for Escherichia coli based on CRISPR-Cas12a. Front. Microbiol. 2018, 9, 2307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Gao, C.; Wang, S.; Hu, G.; Guo, L.; Chen, X.; Xu, P.; Liu, L. Engineering Escherichia coli for malate production by integrating modular pathway characterization with CRISPRi-guided multiplexed metabolic tuning. Biotechnol. Bioeng. 2018, 115, 661–672. [Google Scholar] [CrossRef] [PubMed]
  301. Chung, M.E.; Yeh, I.H.; Sung, L.Y.; Wu, M.Y.; Chao, Y.P.; Ng, I.S.; Hu, Y.C. Enhanced integration of large DNA into E. coli chromosome by CRISPR/Cas9. Biotechnol. Bioeng. 2017, 114, 172–183. [Google Scholar] [CrossRef]
  302. Wu, M.Y.; Sung, L.Y.; Li, H.; Huang, C.H.; Hu, Y.C. Combining CRISPR and CRISPRi systems for metabolic engineering of E. coli and 1, 4-BDO biosynthesis. ACS Synth. Biol. 2017, 6, 2350–2361. [Google Scholar] [CrossRef] [PubMed]
  303. Heo, M.J.; Jung, H.M.; Um, J.; Lee, S.W.; Oh, M.K. Controlling citrate synthase expression by CRISPR/Cas9 genome editing for n-butanol production in Escherichia coli. ACS Synth. Biol. 2017, 6, 182–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Zhang, X.; Wang, J.; Cheng, Q.; Zheng, X.; Zhao, G.; Wang, J. Multiplex gene regulation by CRISPR-ddCpf1. Cell Discov. 2017, 3, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. de Maat, V.; Stege, P.B.; Dedden, M.; Hamer, M.; van Pijkeren, J.-P.; Willems, R.J.; van Schaik, W. CRISPR-Cas9-mediated genome editing in vancomycin-resistant Enterococcus faecium. FEMS Microbiol. Lett. 2019, 366, fnz256. [Google Scholar] [CrossRef]
  306. Price, V.J.; Huo, W.; Sharifi, A.; Palmer, K.L. CRISPR-Cas and restriction-modification act additively against conjugative antibiotic resistance plasmid transfer in Enterococcus faecalis. mSphere 2016, 1, e00064-16. [Google Scholar] [CrossRef] [Green Version]
  307. Hullahalli, K.; Rodrigues, M.; Palmer, K.L. Exploiting CRISPR-Cas to manipulate Enterococcus faecalis populations. eLlife 2017, 6, e26664. [Google Scholar] [CrossRef]
  308. Sun, Q.; Wang, Y.; Dong, N.; Shen, L.; Zhou, H.; Hu, Y.; Gu, D.; Chen, S.; Zhang, R.; Ji, Q. Application of CRISPR/Cas9-based genome editing in studying the mechanism of pandrug resistance in Klebsiella pneumoniae. Antimicrob. Agents Chemother. 2019, 63, e00113-19. [Google Scholar] [CrossRef] [Green Version]
  309. Song, X.; Huang, H.; Xiong, Z.; Ai, L.; Yang, S. CRISPR-Cas9D10A nickase-assisted genome editing in Lactobacillus casei. Appl. Environ. Microbiol. 2017, 83, e01259-17. [Google Scholar] [CrossRef] [Green Version]
  310. Sanozky-Dawes, R.; Selle, K.; O’Flaherty, S.; Klaenhammer, T.; Barrangou, R. Occurrence and activity of a type II CRISPR-Cas system in Lactobacillus gasseri. Microbiology 2015, 161, 1752–1761. [Google Scholar] [CrossRef]
  311. Oh, J.-H.; van Pijkeren, J.-P. CRISPR–Cas9-assisted recombineering in Lactobacillus reuteri. Nucleic Acids Res. 2014, 42, e131. [Google Scholar] [CrossRef]
  312. Singh, A.K.; Carette, X.; Potluri, L.-P.; Sharp, J.D.; Xu, R.; Prisic, S.; Husson, R.N. Investigating essential gene function in Mycobacterium tuberculosis using an efficient CRISPR interference system. Nucleic Acids Res. 2016, 44, e143. [Google Scholar] [CrossRef] [Green Version]
  313. Choudhary, E.; Thakur, P.; Pareek, M.; Agarwal, N. Gene silencing by CRISPR interference in mycobacteria. Nat. Commun. 2015, 6, 6267. [Google Scholar] [CrossRef] [Green Version]
  314. Rock, J.M.; Hopkins, F.F.; Chavez, A.; Diallo, M.; Chase, M.R.; Gerrick, E.R.; Pritchard, J.R.; Church, G.M.; Rubin, E.J.; Sassetti, C.M. Programmable transcriptional repression in mycobacteria using an orthogonal CRISPR interference platform. Nat. Microbiol. 2017, 2, 16274. [Google Scholar] [CrossRef] [Green Version]
  315. Xiang, L.; Qi, F.; Jiang, L.; Tan, J.; Deng, C.; Wei, Z.; Jin, S.; Huang, G. CRISPR-dCas9-mediated knockdown of prtR, an essential gene in Pseudomonas aeruginosa. Lett. Appl. Microbiol. 2020, 71, 386–393. [Google Scholar] [CrossRef]
  316. Tan, S.Z.; Reisch, C.R.; Prather, K.L. A robust CRISPR interference gene repression system in Pseudomonas. J. Bacteriol. 2018, 200, e00575-17. [Google Scholar] [CrossRef] [Green Version]
  317. Gu, T.; Zhao, S.; Pi, Y.; Chen, W.; Chen, C.; Liu, Q.; Li, M.; Han, D.; Ji, Q. Highly efficient base editing in Staphylococcus aureus using an engineered CRISPR RNA-guided cytidine deaminase. Chem. Sci. 2018, 9, 3248–3253. [Google Scholar] [CrossRef] [Green Version]
  318. Liu, Q.; Jiang, Y.; Shao, L.; Yang, P.; Sun, B.; Yang, S.; Chen, D. CRISPR/Cas9-based efficient genome editing in Staphylococcus aureus. Acta Biochim. Biophys. Sin. 2017, 49, 764–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Chen, W.; Zhang, Y.; Yeo, W.-S.; Bae, T.; Ji, Q. Rapid and efficient genome editing in Staphylococcus aureus by using an engineered CRISPR/Cas9 system. J. Am. Chem. Soc. 2017, 139, 3790–3795. [Google Scholar] [CrossRef] [PubMed]
  320. Cobb, L.H.; Park, J.; Swanson, E.A.; Beard, M.C.; McCabe, E.M.; Rourke, A.S.; Seo, K.S.; Olivier, A.K.; Priddy, L.B. CRISPR-Cas9 modified bacteriophage for treatment of Staphylococcus aureus induced osteomyelitis and soft tissue infection. PLoS ONE 2019, 14, e0220421. [Google Scholar] [CrossRef] [Green Version]
  321. Guan, J.; Wang, W.; Sun, B. Chromosomal targeting by the type III-A CRISPR-Cas system can reshape genomes in Staphylococcus aureus. mSphere 2017, 2, e00403-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Penewit, K.; Holmes, E.A.; McLean, K.; Ren, M.; Waalkes, A.; Salipante, S.J. Efficient and scalable precision genome editing in Staphylococcus aureus through conditional recombineering and CRISPR/Cas9-mediated counterselection. mBio 2018, 9, e00067-18. [Google Scholar] [CrossRef] [Green Version]
  323. Park, J.Y.; Moon, B.Y.; Park, J.W.; Thornton, J.A.; Park, Y.H.; Seo, K.S. Genetic engineering of a temperate phage-based delivery system for CRISPR/Cas9 antimicrobials against Staphylococcus aureus. Sci. Rep. 2017, 7, 44929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Wang, Y.; Cobb, R.; Zhao, H. High-efficiency genome editing of Streptomyces species by an engineered CRISPR/Cas system. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 2016; Volume 575, pp. 271–284. [Google Scholar]
  325. Li, L.; Wei, K.; Zheng, G.; Liu, X.; Chen, S.; Jiang, W.; Lu, Y. CRISPR-Cpf1-assisted multiplex genome editing and transcriptional repression in Streptomyces. Appl. Environ. Microbiol. 2018, 84, e00827-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Zhang, M.M.; Wong, F.T.; Wang, Y.; Luo, S.; Lim, Y.H.; Heng, E.; Yeo, W.L.; Cobb, R.E.; Enghiad, B.; Ang, E.L. CRISPR–Cas9 strategy for activation of silent Streptomyces biosynthetic gene clusters. Nat. Chem. Biol. 2017, 13, 607–609. [Google Scholar] [CrossRef]
  327. Zeng, H.; Wen, S.; Xu, W.; He, Z.; Zhai, G.; Liu, Y.; Deng, Z.; Sun, Y. Highly efficient editing of the actinorhodin polyketide chain length factor gene in Streptomyces coelicolor M145 using CRISPR/Cas9-CodA (sm) combined system. Appl. Microbiol. Biotechnol. 2015, 99, 10575–10585. [Google Scholar] [CrossRef]
  328. Qiu, Y.; Wang, S.; Chen, Z.; Guo, Y.; Song, Y. An active type IE CRISPR-Cas system identified in Streptomyces avermitilis. PLoS ONE 2016, 11, e0149533. [Google Scholar] [CrossRef] [Green Version]
  329. Jia, H.; Zhang, L.; Wang, T.; Han, J.; Tang, H.; Zhang, L. Development of a CRISPR/Cas9-mediated gene-editing tool in Streptomyces rimosus. Microbiology 2017, 163, 1148–1155. [Google Scholar] [CrossRef]
  330. Cobb, R.E.; Wang, Y.; Zhao, H. High-efficiency multiplex genome editing of Streptomyces species using an engineered CRISPR/Cas system. ACS Synth. Biol. 2015, 4, 723–728. [Google Scholar] [CrossRef] [Green Version]
  331. Huang, H.; Zheng, G.; Jiang, W.; Hu, H.; Lu, Y. One-step high-efficiency CRISPR/Cas9-mediated genome editing in Streptomyces. Acta Biochim. Biophys. Sin. 2015, 47, 231–243. [Google Scholar] [CrossRef] [Green Version]
  332. Wendt, K.E.; Ungerer, J.; Cobb, R.E.; Zhao, H.; Pakrasi, H.B. CRISPR/Cas9 mediated targeted mutagenesis of the fast growing cyanobacterium Synechococcus elongatus UTEX 2973. Microb. Cell Factories 2016, 15, 115. [Google Scholar] [CrossRef] [Green Version]
  333. González de Aledo, M.; González-Bardanca, M.; Blasco, L.; Pacios, O.; Bleriot, I.; Fernández-García, L.; Fernández-Quejo, M.; López, M.; Bou, G.; Tomás, M. CRISPR-cas, a revolution in the treatment and study of ESKAPE infections: Pre-clinical studies. Antibiotics 2021, 10, 756. [Google Scholar] [CrossRef]
  334. Liu, Z.; Dong, H.; Cui, Y.; Cong, L.; Zhang, D. Application of different types of CRISPR/Cas-based systems in bacteria. Microb. Cell Factories 2020, 19, 172. [Google Scholar] [CrossRef]
  335. Karah, N.; Samuelsen, Ø.; Zarrilli, R.; Sahl, J.W.; Wai, S.N.; Uhlin, B.E. CRISPR-cas subtype I-Fb in Acinetobacter baumannii: Evolution and utilization for strain subtyping. PLoS ONE 2015, 10, e0118205. [Google Scholar] [CrossRef] [Green Version]
  336. Karah, N.; Wai, S.N.; Uhlin, B.E. CRISPR-based subtyping to track the evolutionary history of a global clone of Acinetobacter baumannii. Infect. Genet. Evol. 2021, 90, 104774. [Google Scholar] [CrossRef]
  337. Pourcel, C.; Touchon, M.; Villeriot, N.; Vernadet, J.P.; Couvin, D.; Toffano-Nioche, C.; Vergnaud, G. CRISPRCasdb a successor of CRISPRdb containing CRISPR arrays and cas genes from complete genome sequences, and tools to download and query lists of repeats and spacers. Nucleic Acids Res. 2020, 48, D535–D544. [Google Scholar] [CrossRef]
  338. Mangas, E.L.; Rubio, A.; Álvarez-Marín, R.; Labrador-Herrera, G.; Pachón, J.; Pachón-Ibáñez, M.E.; Divina, F.; Pérez-Pulido, A.J. Pangenome of Acinetobacter baumannii uncovers two groups of genomes, one of them with genes involved in CRISPR/Cas defence systems associated with the absence of plasmids and exclusive genes for biofilm formation. Microb. Genom. 2019, 5, e000309. [Google Scholar] [CrossRef]
  339. Di Nocera, P.P.; Rocco, F.; Giannouli, M.; Triassi, M.; Zarrilli, R. Genome organization of epidemic Acinetobacter baumannii strains. BMC Microbiol. 2011, 11, 224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  340. Grissa, I.; Vergnaud, G.; Pourcel, C. The CRISPRdb database and tools to display CRISPRs and to generate dictionaries of spacers and repeats. BMC Bioinform. 2007, 8, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Hauck, Y.; Soler, C.; Jault, P.; Mérens, A.; Gérome, P.; Nab, C.M.; Trueba, F.; Bargues, L.; Thien, H.V.; Vergnaud, G. Diversity of Acinetobacter baumannii in four French military hospitals, as assessed by multiple locus variable number of tandem repeats analysis. PLoS ONE 2012, 7, e44597. [Google Scholar] [CrossRef] [PubMed]
  342. Tyumentseva, M.; Mikhaylova, Y.; Prelovskaya, A.; Tyumentsev, A.; Petrova, L.; Fomina, V.; Zamyatin, M.; Shelenkov, A.; Akimkin, V. Genomic and phenotypic analysis of multidrug-resistant Acinetobacter baumannii clinical isolates carrying different types of CRISPR/Cas systems. Pathogens 2021, 10, 205. [Google Scholar] [CrossRef] [PubMed]
  343. Horvath, P.; Romero, D.A.; Coûté-Monvoisin, A.-C.; Richards, M.; Deveau, H.; Moineau, S.; Boyaval, P.; Fremaux, C.; Barrangou, R. Diversity, activity, and evolution of CRISPR loci in Streptococcus thermophilus. J. Bacteriol. 2008, 190, 1401–1412. [Google Scholar] [CrossRef] [Green Version]
  344. Mohanraju, P.; Saha, C.; van Baarlen, P.; Louwen, R.; Staals, R.H.; van der Oost, J. Alternative functions of CRISPR–Cas systems in the evolutionary arms race. Nat. Rev. Microbiol. 2022, 20, 351–364. [Google Scholar] [CrossRef]
  345. Pinilla-Redondo, R.; Mayo-Muñoz, D.; Russel, J.; Garrett, R.A.; Randau, L.; Sørensen, S.J.; Shah, S.A. Type IV CRISPR–Cas systems are highly diverse and involved in competition between plasmids. Nucleic Acids Res. 2020, 48, 2000–2012. [Google Scholar] [CrossRef] [Green Version]
  346. Dong, J.F.; Feng, C.J.; Wang, P.; Li, R.Q.; Zou, Q.H. Comparative genomics analysis of Acinetobacter baumannii multi-drug resistant and drug sensitive strains in China. Microb. Pathog. 2022, 165, 105492. [Google Scholar] [CrossRef]
  347. Jalal, D.; Elzayat, M.G.; Diab, A.A.; El-Shqanqery, H.E.; Samir, O.; Bakry, U.; Hassan, R.; Elanany, M.; Shalaby, L.; Sayed, A.A. Deciphering multidrug-resistant Acinetobacter baumannii from a pediatric cancer hospital in Egypt. mSphere 2021, 6, e00725-21. [Google Scholar] [CrossRef]
  348. Montaña, S.; Vilacoba, E.; Fernandez, J.S.; Traglia, G.M.; Sucari, A.; Pennini, M.; Iriarte, A.; Centron, D.; Melano, R.G.; Ramírez, M.S. Genomic analysis of two Acinetobacter baumannii strains belonging to two different sequence types (ST172 and ST25). J. Glob. Antimicrob. Resist. 2020, 23, 154–161. [Google Scholar] [CrossRef]
  349. Mortensen, K.; Lam, T.J.; Ye, Y. Comparison of CRISPR–Cas immune systems in healthcare-related pathogens. Front. Microbiol. 2021, 12, 758782. [Google Scholar] [CrossRef]
  350. Shelenkov, A.; Petrova, L.; Zamyatin, M.; Mikhaylova, Y.; Akimkin, V. Diversity of international high-risk clones of Acinetobacter baumannii revealed in a Russian multidisciplinary medical center during 2017–2019. Antibiotics 2021, 10, 1009. [Google Scholar] [CrossRef]
  351. Pursey, E.; Dimitriu, T.; Paganelli, F.L.; Westra, E.R.; van Houte, S. CRISPR-Cas is associated with fewer antibiotic resistance genes in bacterial pathogens. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2022, 377, 20200464. [Google Scholar] [CrossRef]
  352. Shehreen, S.; Chyou, T.Y.; Fineran, P.C.; Brown, C.M. Genome-wide correlation analysis suggests different roles of CRISPR-Cas systems in the acquisition of antibiotic resistance genes in diverse species. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2019, 374, 20180384. [Google Scholar] [CrossRef] [Green Version]
  353. Yadav, G.; Singh, R. In silico analysis reveals the co-existence of CRISPR-Cas type I-F1 and type I-F2 systems and its association with restricted phage invasion in Acinetobacter baumannii. Front. Microbiol. 2022, 13, 909886. [Google Scholar] [CrossRef]
  354. Carte, J.; Christopher, R.T.; Smith, J.T.; Olson, S.; Barrangou, R.; Moineau, S.; Glover III, C.V.; Graveley, B.R.; Terns, R.M.; Terns, M.P. The three major types of CRISPR-Cas systems function independently in CRISPR RNA biogenesis in Streptococcus thermophilus. Mol. Microbiol. 2014, 93, 98–112. [Google Scholar] [CrossRef] [Green Version]
  355. Majumdar, S.; Zhao, P.; Pfister, N.T.; Compton, M.; Olson, S.; Glover, C.V.; Wells, L.; Graveley, B.R.; Terns, R.M.; Terns, M.P. Three CRISPR-Cas immune effector complexes coexist in Pyrococcus furiosus. RNA 2015, 21, 1147–1158. [Google Scholar] [CrossRef] [Green Version]
  356. Silas, S.; Lucas-Elio, P.; Jackson, S.A.; Aroca-Crevillén, A.; Hansen, L.L.; Fineran, P.C.; Fire, A.Z.; Sánchez-Amat, A. Correction: Type III CRISPR-Cas systems can provide redundancy to counteract viral escape from type I systems. eLife 2018, 7, e36853. [Google Scholar] [CrossRef]
  357. Nussenzweig, P.M.; Marraffini, L.A. Molecular mechanisms of CRISPR-Cas immunity in bacteria. Annu. Rev. Genet. 2020, 54, 93–120. [Google Scholar] [CrossRef]
  358. Sykes, E.M.; Deo, S.; Kumar, A. Recent advances in genetic tools for Acinetobacter baumannii. Front. Genet. 2020, 11, 601380. [Google Scholar] [CrossRef]
  359. Aydin, S.; Personne, Y.; Newire, E.; Laverick, R.; Russell, O.; Roberts, A.P.; Enne, V.I. Presence of Type IF CRISPR/Cas systems is associated with antimicrobial susceptibility in Escherichia coli. J. Antimicrob. Chemother. 2017, 72, 2213–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. Leungtongkam, U.; Thummeepak, R.; Kitti, T.; Tasanapak, K.; Wongwigkarn, J.; Styles, K.M.; Wellington, E.M.; Millard, A.D.; Sagona, A.P.; Sitthisak, S. Genomic analysis reveals high virulence and antibiotic resistance amongst phage susceptible Acinetobacter baumannii. Sci. Rep. 2020, 10, 16154. [Google Scholar] [CrossRef] [PubMed]
  361. Hsu, P.D.; Scott, D.A.; Weinstein, J.A.; Ran, F.A.; Konermann, S.; Agarwala, V.; Li, Y.; Fine, E.J.; Wu, X.; Shalem, O. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 2013, 31, 827–832. [Google Scholar] [CrossRef] [PubMed]
  362. Guo, T.; Yang, J.; Sun, X.; Wang, Y.; Yang, L.; Kong, G.; Jiao, H.; Bao, G.; Li, G. Whole-Genome Analysis of Acinetobacter baumannii Strain AB43 Containing a Type I-Fb CRISPR-Cas System: Insights into the Relationship with Drug Resistance. Molecules 2022, 27, 5665. [Google Scholar] [CrossRef]
  363. Ten, K.E.; Md Zoqratt, M.Z.H.; Ayub, Q.; Tan, H.S. Characterization of multidrug-resistant Acinetobacter baumannii strain ATCC BAA1605 using whole-genome sequencing. BMC Res. Notes 2021, 14, 83. [Google Scholar] [CrossRef]
  364. Mayer, C.; Muras, A.; Romero, M.; López, M.; Tomás, M.; Otero, A. Multiple quorum quenching enzymes are active in the nosocomial pathogen Acinetobacter baumannii ATCC17978. Front. Cell. Infect. Microbiol. 2018, 8, 310. [Google Scholar] [CrossRef]
  365. Karlapudi, A.P.; Venkateswarulu, T.; Tammineedi, J.; Srirama, K.; Kanumuri, L.; Kodali, V.P. In silico sgRNA tool design for CRISPR control of quorum sensing in Acinetobacter species. Genes Dis. 2018, 5, 123–129. [Google Scholar] [CrossRef]
  366. Tucker, A.T.; Nowicki, E.M.; Boll, J.M.; Knauf, G.A.; Burdis, N.C.; Trent, M.S.; Davies, B.W. Defining gene-phenotype relationships in Acinetobacter baumannii through one-step chromosomal gene inactivation. mBio 2014, 5, e01313-14. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustration of a three-stage adaptive immunity mechanism in bacteria using CRISPR-Cas machinery. (a) Adaptation/Acquisition: PAM sequence identification and protospacer integration into CRISPR-array by the Cas protein complex after viral invasion. (b) Transcription: production of CRISPR RNA (crRNA) molecules by transcription of CRISPR sequence. (c) Targeting/Interference: formation of crRNA + Cas nuclease complex, identification of the target invading sequence, and cleaving of foreign DNA to avoid infection.
Figure 1. Schematic illustration of a three-stage adaptive immunity mechanism in bacteria using CRISPR-Cas machinery. (a) Adaptation/Acquisition: PAM sequence identification and protospacer integration into CRISPR-array by the Cas protein complex after viral invasion. (b) Transcription: production of CRISPR RNA (crRNA) molecules by transcription of CRISPR sequence. (c) Targeting/Interference: formation of crRNA + Cas nuclease complex, identification of the target invading sequence, and cleaving of foreign DNA to avoid infection.
Pharmaceuticals 16 00920 g001
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

Junaid, M.; Thirapanmethee, K.; Khuntayaporn, P.; Chomnawang, M.T. CRISPR-Based Gene Editing in Acinetobacter baumannii to Combat Antimicrobial Resistance. Pharmaceuticals 2023, 16, 920. https://doi.org/10.3390/ph16070920

AMA Style

Junaid M, Thirapanmethee K, Khuntayaporn P, Chomnawang MT. CRISPR-Based Gene Editing in Acinetobacter baumannii to Combat Antimicrobial Resistance. Pharmaceuticals. 2023; 16(7):920. https://doi.org/10.3390/ph16070920

Chicago/Turabian Style

Junaid, Muhammad, Krit Thirapanmethee, Piyatip Khuntayaporn, and Mullika Traidej Chomnawang. 2023. "CRISPR-Based Gene Editing in Acinetobacter baumannii to Combat Antimicrobial Resistance" Pharmaceuticals 16, no. 7: 920. https://doi.org/10.3390/ph16070920

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