Next Article in Journal
Identification of Intervention Opportunities through Assessment of the Appropriateness of Antibiotic Prescribing in Surgical Patients in a UK Hospital Using a National Audit Tool: A Single Centre Retrospective Audit
Next Article in Special Issue
Prevalence and Antimicrobial Resistance of Bacterial Uropathogens Isolated from Dogs and Cats
Previous Article in Journal
The Effect of Dalbavancin in Moderate to Severe Hidradenitis Suppurativa
Previous Article in Special Issue
A Retrospective Study of Antimicrobial Resistant Bacteria Associated with Feline and Canine Urinary Tract Infection in Hong Kong SAR, China—A Case Study on Implication of First-Line Antibiotics Use
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antibiotic Resistance in the Finfish Aquaculture Industry: A Review

Faculty of Veterinary Medicine, Post-Graduate Specialization School in Food Inspection “G. Tiecco”, University of Teramo, Strada Provinciale 18, 64100 Teramo, Italy
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(11), 1574; https://doi.org/10.3390/antibiotics11111574
Submission received: 15 September 2022 / Revised: 25 October 2022 / Accepted: 6 November 2022 / Published: 8 November 2022
(This article belongs to the Special Issue Antibiotic Resistance in Companion and Food-Producing Animals)

Abstract

:
Significant challenges to worldwide sustainable food production continue to arise from environmental change and consistent population growth. In order to meet increasing demand, fish production industries are encouraged to maintain high growth densities and to rely on antibiotic intervention throughout all stages of development. The inappropriate administering of antibiotics over time introduces selective pressure, allowing the survival of resistant bacterial strains through adaptive pathways involving transferable nucleotide sequences (i.e., plasmids). This is one of the essential mechanisms of antibiotic resistance development in food production systems. This review article focuses on the main international regulations and governing the administering of antibiotics in finfish husbandry and summarizes recent data regarding the distribution of bacterial resistance in the finfish aquaculture food production chain. The second part of this review examines promising alternative approaches to finfish production, sustainable farming techniques, and vaccination that circumvents excessive antibiotic use, including new animal welfare measures. Then, we reflect on recent adaptations to increasingly interdisciplinary perspectives in the field and their greater alignment with the One Health initiative.

1. Introduction

Recent reports indicate that finfish and seafood consumption can be considered sustainable feeding sources. In a recent FAO Report (2020), it was observed that 156 million tons were for human consumption; this means roughly 20.5 kg/consumer per year. The remaining amount (23 million tons) was used for fish oil and fishmeal production [1]. In 2015, the world marine catch decreased by almost 2 million tons from 81.2 million tons. This decrease was justified by significant fish catching, which has led to reduced animal densities, causing environmental alterations due to the high impacts of anthropic activities [2].
Generally, fisheries have strategic importance for food production, human and animal nutrition, and the employment of millions of people (39 million people in the primary sector of capture fisheries and 20.5 million people in the aquaculture one) [3].
With the reduction in marine finfish populations and the increase in human consumers, aquaculture farmers worldwide have expanded productive systems from small fisheries to larger, more intensive ones; many countries (i.e., China, Thailand, India, etc.) have promoted the building of inland and mariculture husbandries [1,4]. In 2018, most reports indicated that global fish production had increased from 167 million tons in 2016 to 179 million tons, with 82 million tons being derived from aquaculture farms [1]. The intensively farmed finfish species present high mortality (50%), starting from the larval stage and continuing in sea cages. Losses are generally caused by bacterial or viral diseases; in major cases, bacterial pathologies are directly linked to the high density located in feces and sediments or to improper vaccination programs [4].
The high animal density, directly related to the welfare concept, is a crucial aspect because it provides epidemiological and environmental conditions that lead to possible infectious disease outbreaks that cause productive and economic losses. In most cases, antibiotics usually represent the first choice for treatments [5], which can be administered through different routes: feed, water immersion, or injection [6]. From a zootechnic perspective, it is also relevant to considerer two possible negative aspects: proper usage and the administering of unapproved (antibiotics in which usage is restricted only to the human medicine) or illegal molecules [i.e., chloramphenicol (banned in the EU member states)].
Comparing aquaculture to the terrestrial farms, based on pharmacological consumption, researchers have highlighted significant differences. Indeed, the World Health Organization classified aquaculture as an activity with a low environmental impact for antibiotic usage [7,8]. Directly linked to the above-explained concepts, veterinarians play key roles in pharmacological management. This consideration is justified because this professional figure firstly prescribes antibiotic therapies, avoiding the unnecessary administering (as metaphylactic one) of certain classes—in which usage is restricted to the human medicine: the so-called Critical Importance Antimicrobials WHO (CIA)—for food-producing animals; secondly, they must reduce administering to only restricted specific cases. The aim is to decrease a relevant selective pressure, which promotes resistant and pan-resistant bacterial strains survival [1]. These strains are named as antibiotic resistance bacteria (ARBs) that can be considered as “drivers” of antibiotic resistance genes (ARGs) with important repercussions on the environmental, animal, and human health [9]. It has been widely demonstrated that ARGs can be transferred to human intestinal microbiota and, consequently, to the ingestion of foods (numerous matrices: meat, dairy products, fish, etc.), which can drive commensal or pathogenic (Salmonella spp., Vibrio spp.) ARBs with extra chromosomal resistance forms. Finally, it is important to mention possible drugs residues due to the improper observation of legal limits [7]. Therefore, the European Regulations No. 470/2009 and No. 37/2010 established the residual limits of pharmacologically active substances in animal origin foodstuffs.
Furthermore, the aquatic environment (i.e., oceans, lakes, rivers) is also a possible reservoir of ARGs [1]. Generally, fin fish’s intestinal microbiota is characterized by bacterial populations that are like those ones detected in the aquatic environment. Therefore, microbiological water quality (influenced by wastewater management, fish industries, and other anthropic activities, etc.) can represent a critical environmental resistance factor that allows ARGs diffusion and preservation [1].

2. Antibiotic Usage: Regulations in Aquaculture Farms

A horizontal concept, that involves, at the same time, different cultured animal species, is represented by the therapeutic administering of antibiotic molecules related to the high subjects’ densities (per m2 or m3 of surface or water). Due to its crucial aspect, represented by the antimicrobial resistance (AMR) phenomenon, different nations have organized their respective legislations to prevent and decrease its diffusion. The WHO data report an alarmistic scenario: within 10 years, the antibiotic therapeutic efficacy, both from humans and animals, will be strongly reduced [7].
In this section, authors want to describe the main legislative measures adopted by different nations. During last 20 years, European and American (USA) public health institutions have produced lists of authorized molecules [10].
In 2000, the European Union, firstly, in the “White Paper on Food Safety”, identified the strict correlations between food and environmental safety concepts. From this document, the European Commission has evolved and, on this issue, has based the new regulations: EU Reg. No. 178/2002; EU Reg. No. 852/2004; EU Reg. No. 853/2004; EU Reg. No. 625/2017 [11].
The above-mentioned regulations supplement obligatory requirements for producing countries to follow the Council Directive 96/23/EC for aquaculture products to export to the EU States [12]. These legislative acts are supported by strict monitoring activities regarding usage and trade of veterinary antibiotics performed by law-designed control figures, i.e., the European Surveillance of Veterinary Antimicrobial Consumption (ESVAC) [13]. Indeed, the European Medicines Agency (EMA) banned the administering of certain molecules (cefuroxime, chloramphenicol, polymyxin B Sulphate, and Nystatin) to guarantee final consumer health [14]. More recently European agencies, including the European Commission, established maximum residue limits (MRL) of pharmacologically active substances in foodstuffs of animal origin (including finfish): EU Reg. No. 470/2009, and No. 37/2010 (see Table 1).
Moreover, the Food and Drug Administration (FDA) and the United States Department of Agriculture (USDA) worked on fish drugs, reporting a list of antibiotics that can be used in aquaculture [5]. As described above by Table 1 and Table 2, comparing developed countries legislations, the European limit values of active substances are expressed as µg/kg and mg/kg in the American one. Furthermore, in European legislation (i.e., EU Reg. No. 37/2010), more detailed and species-specific (including finfish and shellfish) limit values concerning a wide range of xenobiotic molecules than the North American ones are reported.
In the developing countries, there is a wide differentiation, which depends on their respective governmental agencies. This last sentence should not be considered redundant, because a clear legislation on antibiotic usage in veterinary medicine, more specifically on the aquaculture sector, is not yet well structured.
For the above-mentioned reasons, and in order to export finfish products, developing countries’ legislators adopted similar parameters to those reported in the European Union and USA laws, i.e., the Brazil [15], Vietnam [16,17,18], Chile [19], China [20,21,22,23], India (MPEDA) [23], the Philippines [1], and Thailand [24].
In China, Thailand, Vietnam, Brazil, Chile, Bangladesh, Norway, the Philippines, and India, governmental authorities have listed the authorized antibiotic compounds and banned other molecules for usage in the aquaculture sector [16,17,20,24,25,26]. In Asia, there are differences between geographic regions, depending on the antimicrobial usage (any farmers that still administer chloramphenicol that is banned in the aquaculture zootechnic) and local food safety regulatory agencies. However, this last issue has been justified by low pharma-surveillance programs, poor “Food Safety Legislation”, and inadequate monitoring systems’ control of drug usage [27]. These legal and surveillance gaps have contributed to the ARBs and ARGs in the aquatic environment [28,29,30].
To contrast the AMR phenomenon, there are multiple examples of collaboration between different legislative institutions that have improved the antimicrobial management. For instance, the FDA continues to detect nitrofurans and chloramphenicol in collaboration with Malaysian aquaculture producers, and, consequently, this country has banned them [31]. China realized innovative and specific “Applicative Guidelines”, which specify sulfonamides, tetracycline, and enrofloxacin usages, which were adopted in other Asian countries, i.e., Vietnam [23].
In conclusion, it is possible to affirm that the pleomorphic AMR phenomenon cannot be reduced by geographical limits, and the respective legislations difficulty could be organized with the same restrictions [11]. However, due to the environmental implications, which pose at-risk human and animals health, it is mandatory to align lists of antibiotic molecules that can be used for finfish disease treatments. Therefore, sanitary authorities in the international community should make an effort to contribute to a global reduction in their tons of consumption of finfish production, starting from the sharing of data and enforcing innovative pharmaco-surveillance systems. It implies that the realization of integrated tracing processes, which have origins from the pharma industries, arrive at the administering step in aquaculture farms and, consequently, involve the aquatic environment.

3. Aquatic Environment and Antibiotic Resistance Circulation

Infectious disease caused by ARBs are estimated to cause 10 million deaths worldwide by 2050 [32]. This issue is a horizontal problem involving humans, animals, plants, foods, and environments; due to these reasons, the One Health approach is essential to overcome this developing threat [33].
Previous studies have identified the aquatic environments (which include oceans, lakes, and rivers) as potential transmission routes and ARGs and ARBs reservoirs [10,24,33]. Water and, in particular, wastewater management are crucial steps in the so called “water-cycle” as vectors of antibiotic resistance forms. Indeed, it has been repeatedly reported that ARGs and ARBs detection in water samples collected from treatment plants and in numerous cases the public sanitary authorities (in accordance with specific national legislation cut-offs) have found high residual titers of antibiotic molecules (i.e., quinolones, tetracyclines, carbapenems, aminoglycosides, etc.) [34].
For this purpose, the usage of specific filters for wastewaters management could be useful for their contributions to the reduction in the environmental diffusion of microorganisms’ loads (bacteria, virus, etc.) [35]. Many research studies have analyzed the anthropic impact generated by hospitals, farms, domestic environments, and food industries by evaluating 79 wastewater samples in different geographic areas, identifying a limited AMR cluster encoding resistance against macrolides, quinolones, aminoglycosides (more than 30% of screened samples) in developed countries (Europe, North America, and Oceania). However, in developing continents (Asia, Africa, and South America) ARGs were mainly reported to encode resistance against sulfonamides and phenicols (especially chloramphenicol (40% of samples)) [34].
Animal origin manures, largely used in the agricultural sector, can hide notable risks for humans. It has been widely demonstrated that their fertilization usage is a reasonable source of further ARGs environmental diffusion. These substances are responsible for aquifers contaminations becoming an environmental concern and providing tangible evidence, where terrestrial and aquatic productive realties are strictly influenced [36]. Hatosy and Martiny [37] evidenced that the 28% of detected ARGs in marine water samples were transferred by freshwater and wastewaters flows. These findings also provide scientific evidence that the coastal runoff from terrestrial sources is one of the ARGs mechanisms of diffusion.
The selective pressures, caused by different anthropic activities, have further repercussions on mariculture farms, as demonstrated by Miranda et al. [38]. They discovered, in different Chilean salmonid farms, high circulation of tetracycline and quinolones ARGs. Authors justified these findings by the large antibiotic administering reported by public health institutions: an amount 363.4 antimicrobial tons were used by farmers [38]. To improve knowledge about this phenomenon, many researchers have studied other possible ARGs and ARBs reservoirs. Muziasari et al. [39] focused on the role of sediment and fish feces collected from mariculture and inland farms in the Baltic Sea. They observed ARGs presence in intestinal contents from rainbow trout (Oncorhynchus mykiss) specimens collected in different aquaculture systems. They discovered different resistance determinant amounts using the real time PCR assay: tetracycline and, in particular, tet genes (tetM: 6.25 10−2 copies); aminoglycosides target genes erm (ermB 3.13 10−1 copies); and sulfonamides, such as sul (sul3 3.13 10−1 copies). Furthermore, similar patterns were amplified from sediment samples. The phylogenetic analysis allowed us to demonstrate the same genomic source. These findings highlighted and enforced a fundamental concept concerning the antibiotic resistance phenomenon in which animal and environmental microbiomes are strictly connected to each other through horizontal gene transmissions.
From these considerations, the so-called One Health approach results are mandatory. The aquatic environment, characterized by various bacterial strains, can be considered as possible drivers of resistant forms. The aquatic creatures, including finfish ones, can be considered as entropic systems, where the intestinal microbial populations meet environmental ones involved in a fascinating “antibiotic resistant genes trade”. In this articulated “ARGs life-cycle”, final human consumers microbiomes could also be involved by the fish food commensal strains, harboring resistance genes; they can transmit them to the human intestinal bacterial species. These conditions are realistically responsible for the emergence of multidrug-resistant (MDR) or pan-resistant pathogenic or commensal microorganisms’ spreading [40].
In Asia, public sanitary authorities reported that fraudulent antimicrobial administering by farmers (in the order of tons) have conducted the selection of MDR microorganisms (Escherichia coli) isolated from aquaculture finfish. Phylogenetic analysis also demonstrated that other species belonging to the family Enterobacteriaceae, isolated from river water samples, presented the same phenotypic and genotypic resistance pattern and same codifying sequences [1].
Although antimicrobial molecules are administered in inland farms (closed ecosystems), the water’s turnover, performed by filter systems, represents a further potential source for environmental antibiotic diffusion, having critical repercussions on final consumers’ health [41]. Therefore, water’s microbiological quality is a crucial element that is influenced by different parameters and factors, i.e., temperature, salt content, space distance between coasts and catching areas (especially influenced by anthropic activities), natural presence of bacteria in the water environment, nutrition for fish, farming systems, catching methods, and technological aspects [1]. These parameters influence animal welfare (in accordance with the “Animal Health Law” EU Reg. No. 429/2016) and immunity, since stressed animals are more susceptible to infection and, in turn, require more application of antibiotic therapy. These molecules are especially used at larval stages, when farmers should register high mortality rates. To avoid this problem, they (farmers) often improperly administer antimicrobials, contributing to the ARBs and ARGs enforcement and diffusion.
From an ecological point of view, Reverter et al. [41] studied the effects and possible correlations between water temperature on aquatic animal mortality related to MDR microorganisms. They evaluated these aspects on bred aquatic animals (finfish), artificially infected with specific fish pathogens (i.e., Vibrio spp.). Researchers found a statistical significant correlation between high water temperature (simulating “global warming” phenomenon) and infected treated finfish. Warm water resulted responsible for high AMR diffusion. At the same time, they also discovered calculating multiple antibiotic resistance (MAR) indices through the phylogenetic analysis and metagenomic evaluations, a further statistical correlation to the bacteria isolated in human pathogenic specimens reservoirs. These findings were justified by the high human activities’ impact (agrochemical substances, toxic metals, abattoirs, and wastewaters managements) [42,43]. Furthermore, based on chemical characteristics, any molecules (including antibiotics) released in the marine environments are low bio-degradable and, for this reason, have been named as “pseudo-persistent” [44]. In polluted coastal areas, any researchers discovered a high direct correlation between heavy metal residue titers (as lead, cadmium, and mercury) and the horizontal ARGs transmission (more specifically observed for tetracycline molecules: tet genes) [45].
In conclusion, human-impacted aquatic areas (i.e., oceans, lakes, rivers, etc.) are responsible for the maintenance and diffusion of MDR bacteria, which have been defined as strains resistant to at least three antibiotic classes. For these reasons, they represent a public health issue [1]. In this way, the environment involves the role of one of the main reservoirs regarding MRB, and it represents “le fil rouge” between antibiotic residues (due to agricultural runoffs, sewage discharges, and leaching from nearby farms) and public health [46].

4. Global Antibiotic Administering in the Aquaculture Sector

The global antibiotic consumption is considered a dynamic value due to its annual variability. In 2017, scientists estimated that a total amount of 10,259 antibiotic tons were administered to food-producing animals [1]. The continent of Asia represents the largest producer and consumer (93.8%) of such molecules, larger than Africa (2.3%), and Europe (1.8%). Four countries present the highest consumption levels: China (57.9%), India (11.3%), Indonesia (8.6%), and Vietnam (5%) [11], as illustrated in Table 3.
EFSA and FDA have proposed a common aim that indicates an antibiotic usage reduction of 30% by 2030 [1]. Conversely to this purpose, there is an opposite trend reported in the BRICS countries, Brazil, Russia, India, China, and South Africa, where scientists estimate an increasing antibiotic consumption for terrestrial food-producing animals. The estimated total amounts could exceed human use [47]. This prospect for the next future has been also confirmed by Schar et al. [48] (Brazil (94%), Saudi Arabia (77%), Australia (61%), Russia (59%), and Indonesia (55%); these percentages are strictly correlated to the respective national usages) (See Table 3).
Table 3. Antibiotics and their distribution in finfish aquaculture from different geographical regions.
Table 3. Antibiotics and their distribution in finfish aquaculture from different geographical regions.
ContinentsCountriesAntibiotic Classes
Asia-Pacific: 9623 tonsChina: 5.572 tons [49]Tetracyclines: 3065 tons
Quinolones: 1393 tons
Beta-lactams: 836 tons
Sulfonamides (co-administered with phenicols): 278 tons
India: 1.087 tons [48]Tetracyclines: 706 tons
Beta-lactams: 195 tons
Quinolones: 186 tons
Indonesia: 827 tons [48]Tetracyclines: 645 tons
Beta-lactams: 182 tons
Vietnam: 481 tons [48]Tetracyclines: 370 tons
Quinolones: 62 tons
Beta-lactams: 49 tons
Africa: 236 tons [48]Egypt: 110 tonsTetracyclines: 86 tons
Beta-lactams: 13 tons
Quinolones: 11 tons
South Africa: 126 tonsTetracyclines: 107 tons
Sulfonamides: 19 tons
Europe: 185 tonsTurkey: 75 tons [50]Tetracyclines: 39 tons
Beta-lactams: 16 tons
Quinolones: 8 tons
Sulfonamides: 7 tons
Phenicols (Chloramphenicol): 5 tons
Norway: 45 tons [51]Tetracyclines: 30 tons
Sulfonamides: 10 tons
Quinolones: 5 tons
Scotland: 32 tons [51]Tetracyclines: 28 tons
Beta-lactams: 4 tons
Italy: 13 tons [51]Tetracyclines: 7 tons
Beta-lactams: 4 tons
Sulfonamides: 2 tons
The International Pharma Agencies report that global animal antimicrobial administering and consumption, and more specifically for the aquaculture sector, involve the following finfish species: 8.3% catfish (Ameiurus melas), 3.4% tilapia (Tilapia spp.), 2.7% shrimp (Penaeus spp.), 0.8% trout (Oncorhynchus mykiss), and 0.7% salmon (Salmo salar) [48].
Concerning the global antibiotic consumption volumes, innovative digital tracing systems will involve fundamental roles in the pharma-surveillance programs. Indeed, it is difficult to provide global monitoring with standardized data due to the multiple variables that can influence the results (i.e., legal antimicrobial administering, anthropic environmental impact, pollution, microbiological water quality, etc.) [52]. All these mentioned affirmations are examples of future challenges that cannot be postponed anymore.
Generally, the most frequently prescribed and detected antimicrobials are the following antibiotic classes: quinolones, tetracyclines, sulfonamides, and amphenicols [11,27].
Among the quinolone class, enrofloxacin, nalidixic acid, and ofloxacin are widely administered due to their chemical characteristics that permit to these molecules to be stable in water and sediment [52], resulting in them being easy to manage in the aquaculture [53]. There are also two different scenarios: in the developed countries, Lulijwa and coworkers found that 55% of global major aquaculture-producing countries have used enrofloxacin and less frequently ciprofloxacin and norfloxacin [11]. Indeed, since 2015 in China, the administering of norfloxacin has been banned for aquaculture [23].
Tetracyclines represent another important antibiotic class widely used in aquaculture farms due to their low costs in association with their high efficacy as a broad spectrum for treatment and prevention of infectious disease [10,54]. The WHO reported that doxycycline, oxytetracycline, and chlortetracycline have been administered for a long time in finfish farms. Due to the increasing antimicrobial resistance patterns, the WHO has suggested a further restriction of these molecules for veterinary usage [55,56]. More specifically, in Asia, oxytetracycline is the most commonly allowed-by-law and administered molecule in the aquaculture farms, and several studies have detected residues in water samples in numerous countries. Oxytetracycline residues have also been detected in the European water and sediment specimens, although, since 2006, it has been banned by all EU member states [57,58]. These considerations are justified by chemical characteristics, which confer a higher environmental resistance to oxytetracycline than the other molecules (belonging to the same antibiotic class) [59,60]. Regarding the tetracyclines, scientists estimated the following persistence periods: 21–25 min in aquaculture water, 2 days in freshwater, 12 days in seawater, 150 days in marine sediment (depending on chemical and environmental parameters as pH, temperature, salinity, and light) [61,62].
The sulfonamides class is largely administered in the finfish farms, and, in particular, veterinarians have prescribed sulfamethoxazole or the combined form sulfamethoxazole/trimethoprim. Generally, these molecules represent the third most prevalent antibiotic used in aquaculture after tetracyclines and quinolones [27]. They are largely used due to their low costs, high water solubility, and due to the high floating characteristics, allowing easy transport and distribution in the aquatic environments and adsorbed by finfish through the gills [63].

5. ARBs Isolation and ARGs Detection from Aquaculture Finfish Samples

The AMR phenomenon has been generally defined as the failure of growth’s inhibition or the killing capacity of an antimicrobial molecule beyond the normal susceptible bacteria [32,64].
The finfish aquaculture zootechnic sector has been characterized by a wide range of farming techniques as the embankment ponds or the watershed ones (as observed in the catfish (Clarias spp., Ictalurus spp., and Pangasius spp.) culture) [65], mariculture systems (i.e., for Salmo spp., Sparus spp., [66], intensive or semi-intensive inland pond systems for Tilapia spp. [65]), and other animals finfish species, etc.
In the above-mentioned systems the high animal densities have induced the necessity of antibiotic administering for therapeutic purposes. This last-explained concept was associated with possible inappropriate usages and has selected resistant pathogens or commensal bacterial strains. More in detail, tetracyclines, beta-lactams, quinolones, and sulfonamides antibiotic classes have been largely prescribed by veterinarians. Therefore, biomolecular diagnostic procedures have coupled the next generation sequencing to the bacterial whole genome analysis. This last cited method has permitted us to discover new oligonucleotide resistance determinants [63]. Oligonucleotide sequences are the main actors involved in the ARGs circulation and are, consequently, responsible for the presence of vector bacteria (not usually resulting in pathogens for humans) while constituting crucial environmental reservoirs for the human and animal host microbiota [64]. For these reasons, animal origin foodstuffs have acquired more attention from scientists. The reasons were firstly related to possible residual concentration, but more specifically the main concern is represented by the possibility of horizontal resistance genes transmission between alimentary commensal and opportunistic strains with the human microbiota. Indeed, microorganisms have elaborated numerous mechanisms to disseminate the ability to survive by mobile genetic elements, such as integrons, plasmids, insertion sequences, transposons, and gene cassettes [64,67], and the inappropriate antibiotic usage has produced a selective pressure and the consequential survival of resistant microorganisms (engendering multiple resistances).
Every year, bacterial genome sequencing allows the identification of emerging and re-emerging ARGs, and the most frequent examples are amplified from aquaculture seafood products, i.e., sul (sulfonamides resistance genes), tet (tetracyclines resistance genes), aa (aminoglycosides resistance genes), and bla (β-lactams resistance genes) [68,69,70,71]. Indeed, molecular biology, through the sequencing assays, constantly discovers different mutations among ARGs. There are numerous cases in which there is no matching between discovered phenotypic resistances results with the genotypic ones. This sentence offers explanations based on the concept of nucleotide sequences’ mutations that produce different DNA transcriptions (improper enzymes’ actions), or it is possibly correlated to an intrinsic resistance, which is typical of certain bacterial families against specific antibiotic molecules or classes. The scientific community, during these years, has investigated the AMR diffusion in various finfish species, especially in aquaculture systems, and the respective numerous genera of pathogenic and opportunistic bacteria that are generally implicated in seafood-borne diseases are Vibrio spp. (i.e., V. parahaemolyticus, V, vulnificus), Listeria monocytogenes, Clostridium botulinum, Aeromonas spp., Salmonella spp., Escherichia coli, Campylobacter jejuni, Shigella spp., Yersinia eneterocolitica, Bacillus cereus [41], Pseudomonas spp. [72], and Enterococcus faecium [73] (see Table 4). As previously mentioned, quinolones, tetracyclines, amphenicol, and sulfonamides are major antimicrobial classes used in aquaculture on a global scale [72].

5.1. Quinolones

Quinolone resistances are characterized by the involvement of DNA gyrase and topoisomerase IV, which are bacterial enzymes and quinolones target proteins. These two enzymes are encoded, respectively, by the gyrA and gyrB genes for DNA gyrase, and by the parC and parE genes for topoisomerase IV [38]. Chromosomal mutations in topoisomerases genes decrease drug accumulation and possible resistance driven by mobile elements, such as plasmid-mediated quinolone resistance (PMQR) (Qnr proteins, aac(6)-lb-cr aminoglycoside acetyltransferases and QepA and OqxAB efflux pumps), causing the constitutive or the acquired resistance to these antibiotic molecules. Increased mutations in DNA gyrase and topoisomerase IV, and in quinolone-resistant fish pathogens (Yersinia ruckeri, Flavobacterium psychrophilum, and V. anguillarum), are linked to the extensive administering of these antimicrobic classes worldwide [86,87,88]. Their wide usage was justified to reduce the hatching losses caused by Vibrio spp. infectious outbreaks. The wide detections of modified plasmids have been discovered from aquaculture finfish fillets [89]. The detected bacterial pathogens were A. hydrophila, V. anguillarum, and V. parahaemolyticus, which showed mutations in the quinolone resistance codifying sequences in specific gene regions belonging to gyrA and/or parC [90,91].
Quinolone resistance genes included in the so-called PMQR are: six qnr genes (qnrA, qnrB, qnrC, qnrD, qnrS, and qnrVC) encoding gyrase-protection repetitive peptides; oqxAB, qepA, and qaqBIII encoding efflux pumps; and aac(60)-Ib-cr encoding an aminoglycoside and quinolone inactivating acetyl-transferase [92]. The majority of PMQRs detection was largely amplified from finfish products worldwide; in China Yan et al. [93] found qepA and aac-(6′)-Ib genes as dominant among PMQR genes in aquatic environments and the possibility of co-emergence of resistance to β-lactams; Jiang and coworkers [94] detected qnrB, qnrS, and qnrD, with aac(6′)-Ib-cr in gut samples of farmed fish. Dobiasova et al. [95] found qnrS2, aac(6)-Ib-cr or qnrB17 genes in Aeromonas spp. isolated from tropical freshwater ornamental fish and coldwater ornamental (koi) carps. In Egypt, scientists reported the occurrence of qnr and aac(6)-Ib-cr resistance from fish farm water sample [1].
In Chile, genes qnrA, qnrB, and qnrS were detected both in the chromosomes of marine bacteria and the same genes in human pathogenic ones [96]. Furthermore, it has been demonstrated that the same plasmid plays an important role for different classes. Indeed, gene cassettes can be considered as multiple ARGs drivers, which conduce to the phenotypical expression of resistance against quinolones, β-lactams, and aminoglycosides [92,93]. Indeed, qnr genes are loaded with β-lactamase determinants on the same plasmids. Khajanchi and coworkers [97] considered aquaculture and the aquatic environment as possible sources of aac(6)-Ib-cr and qnrB2 and Enterobacteriaceae as hosts. They also detected Aeromonas spp. as a vector for qnrS2. Hence, Gram-negative hosts may a be reservoir of plasmid-mediated Qnr-like determinants that seem closely relate to the species V. splendidus [98].
From an environmental perspective, there is a strict correlation between remarkable anthropic activities as polluted water areas and quinolones ARGs diffusion [99]. Indeed, in Asia (especially in China), fish farmers normally use biofertilizers to improve production [100]. There is a real possibility that these organic molecules are vectors of antibiotic resistance genes. Zhao et al. [101] examined biofertilizers normally used in Chinese shrimp aquaculture systems and studied the correlation between fluoroquinolone resistance genes’ diffusion and biofertilizers. In this research project, they also screened the PMQR gene that includes: qnrA, qnrB, qnrC, qnrD, qnrS, qepA, oqxA, oqxB, and aaa(6′)-Ib genes. They screened 20 biofertilizer samples collected from shrimp farms and isolated 20 bacterial strains that were vectors of PMQR genes: 10 Escherichia coli, 9 Enterococcus faecalis, and 1 Enterococcus faecium. About 30% of biofertilizers samples presented qnrB, qnrD, and qepA resistance genes. This study was the first one which discovered the ARGs environmental repercussions due to the usage of contaminated manures on seafood farming systems. Similar patterns were also observed in terrestrial mammals, i.e., domestic swine and chicken manure (widely used in agriculture) [101]. Nowadays, there are not available data regarding finfishes, and it could be interesting to perform further investigations about possible statistical correlations between farming environments and possible agricultural implications. Therefore, these studies have confirmed ARGs diffusion and circulation in different environments through the fecal bacteria detected in common biofertilizer molecules. From these data, it can be seen that quinolones have presented reasonable risks due to the increase of their therapeutic failure. Their inclusion in the Critically Important Antimicrobials list has attracted more attention from pharma surveillance organizations.

5.2. Tetracyclines

Tetracyclines action consists of reversibly binding the 70S ribosome of cells blocking protein synthesis [1]. They are largely used in human and animal treatment as broad-spectrum antimicrobials. For the first time in Japan, it was observed that their improper administering conducted to the discovery of high nucleotide similarities of tetracycline genes between isolated bacteria from finfish aquaculture and from human clinical facilities. The phylogenetic analysis confirmed the same origin [3].
Evolution has selected different strategical and survival pathways and, in particular, four strategies: efflux pumps activation, ribosomal protection inducing a limit to the access, ribosomal RNA mutations avoiding tetracycline molecules binding, and tetracycline inactivation through enzymes [102,103]. In finfish aquaculture products, tet group responsible for proton-dependent efflux pumps encoding was mainly associated with tetracycline resistance [104]. Tet genes have been detected in several bacterial strains isolated from different animal species located in various geographical regions. There are multiple examples: tetB, tetM, tetW were firstly isolated in the intestine and rearing water of red seabream (Pagrus major) [105]; tetA, tetB, tetE, tetH, tetl, tet34, tet35 and 10 others had unknown tet genes isolated from Chilean salmon (Salmo salar) farms [106]; furthermore, Higuera-Llanten and coworkers [107] also detected the presence of tet34, tet35, tetA, tetB, tetE, tetH, tetL, and tetM genes in the same matrixes. Among seafoods (including finfish and crustaceans), Concha et al. [108] discovered tetX gene in Epilithonimonas strains from rainbow trout (Oncorhynchus mykiss) and Han et al. [109] amplified, in shrimp samples, that the tetB gene was carried in a single copy plasmid, named pTetB-VA1, comprising 5162-bp. The whole genome analysis revealed that this plasmid consists of 9 ORFs (overlapping open reading frames) encoding tetracycline-resistant repressor proteins, transcriptional regulatory proteins, and transposases and showed a 99% sequence identity to other tet gene plasmids (pIS04-68 and pAQU2). Furthermore, in terms of tet genes, with special regard to tetE, Agersø et al. [110] discovered tetE horizontal transmission between Aeromonas spp. and Escherichia coli strains, isolated from aquaculture Danish farms. TetA gene diffusion has been demonstrated to be realized through plasmids and transposons named Tn1721 and those that are Tn1721-like. Another similar example is represented by Tn5706, which is involved in tetH dissemination (amplified from Moraxella spp. and Acinetobacter spp. strains isolated from salmon farms) [111]. Due to the expanding of the AMR phenomenon among different bacterial strains, tet genes have been widely amplified from Enterobacteriaceae [112,113], Photobacterium spp., Vibrio spp., Alteromonas spp., Pseudomonas spp., and other marine commensal bacteria. Consequently, the possibility of transferring ARGs from marine microbiota to the human one is considered reasonable. Indeed, many biomolecular investigations have highlighted the possible cross-species ARGs transmission through the foodstuffs ingestion [102,114].
The wide oligonucleotide diversities, as described above, are expressions of the mass administering of tetracycline. Any mammalian zootechnic sectors (i.e., domestic swine) have improperly used this antibiotic class, inducing multiplication and genetic transmissions to the next generations of bacterial isolates (from pathogen to commensal strains, and vice versa).

5.3. Sulfonamides

In aquaculture, sulfonamides are commonly co-administered with trimethoprim, ormethoprim, and florfenicol [115]. The dihydropteroate synthase (DHPS) enzyme, in the folic acid pathway, represents the biochemical target reaction [114]. Sulfonamide’s resistance mechanisms derive from mutations in the chromosomal folP gene that provides varying degrees of trade-off between resistance and efficient folate synthesis, decreasing DHPS affinity for the antimicrobial molecule [114].
Among the discovered ARGs, four different sul gene determinants have been described to encode antibiotic resistance. Sul1 gene has been founded in class 1 integrons and linked to other resistance genes [116]; sul2 is associated with non-conjugative plasmids of the IncQ group and to large transmissible plasmids, such as pBP1 [117]. Sul3 is characterized in the Escherichia coli conjugative plasmid pVP440; sul4 gene has been recently mobilized and phylogenetic inference pinpoints its putative origin as part of the folate synthesis cluster in the Chloroflexi phylum [118]. All described ARGs have a common action, which is represented by the reduction in strategical bacterial structural expression. The transmembrane architectures are widely involved in the cyto-chemical interaction between strains and antibiotic molecules.
The genome and proteome analyses revealed that a gene cluster, containing a flavin-dependent monooxygenase and a flavin reductase, is highly upregulated in response to sulfonamides action, as reported by Kim et al. [119]. Indeed, the biochemical analysis showed that the two-components (belonging to the monooxygenase system) were key enzymes for the initial sulfonamides cleavage. It was observed that the co-expression of the two-component system in Escherichia coli conferred decreased susceptibility to sulfamethoxazole, indicating that the genes encoding drug inactivating enzymes are potential resistance determinants. Comparative genomic analysis revealed that this cluster gene, containing sulfonamide monooxygenase (renamed as sulX) and flavin reductase (sulR), is highly conserved in genomic islands. These ones are shared among sulfonamide-degrading Actinobacteria, all of which also contained sul1-carrying class 1 integrons [119].
Sulfonamide’s ARGs distribution has been widely found in numerous fish and environmental specimens, i.e., Muziasari and coworkers [120] discovered sul1, sul2, and intI1 genes detection in all analyzed samples and the dfrA1 gene in most samples in aquatic farm sediment in the Baltic Sea [39]. Domínguez et al. [121] detected sul1, sul2, class 1 integron-integrase gene intI1, dfrA1, dfrA12, and dfrA14 from a salmon farm in Chile and revealed the occurrence of transferable integrons and sul and dfr genes among sulfonamide- and/or trimethoprim-resistant bacteria, as amplified from Actinobacter spp., Bacillus spp., Proteus spp., and Pseudomanas spp. isolates [88].
ARGs for sulfonamides resistance were also discovered in many commensal bacterial strains in Japanese mariculture areas [122], in Vietnamese freshwater farms [108], China Hainan, Guangdong, Tianjin, Hangzhou, Yantai, and Taihu Lake [113,123,124,125].
These last considerations highlight that environmental stimuli can be responsible for increased or reduced ARGs transcriptions. The deduction leads to the consideration that in the AMR phenomenon, “the environment” plays a crucial role, while human and animal health are only “direct consequences”.

5.4. Thiamphenicol and Florfenicol

Thiamphenicol and florfenicol belong to the amphenicol antibiotic class and have been largely administered in aquaculture farms. Due to the possible chemical residual persistence in finfish muscular tissues, various studies have demonstrated possible sanitary implications on humans, animals, and environments [48,126].
Focusing on risk-based approach (in accordance with the EU Reg. No. 852/2004 and No. 37/2010), the European agencies EFSA and EMA published maximum residue limits and respective daily intakes for final human consumers [48].
Veterinary practitioners normally treat infectious disease (caused by, i.e., Vibrio spp., etc.) and relative possible septicemia cases using the above-mentioned molecules [127]. These have pharma-dynamic synergic effects (binding the 50S ribosomal subunit) if they were coupled with other antibiotic classes as tetracyclines. Both molecules have become widely prescribed because they have broad spectrum effects and low costs [128].
Amphenicol illegal administering has induced an intense evolutive pressure, determining the spreading of resistant strains harboring florfenicol-resistance genes (FRGs). These FRGs are plasmid determinants and have presented high genetic trades (through horizontal transmission) across different bacterial phyla, identifying strong correlations (p values < 0.05), as observed by Zeng et al. [127].
Among amplified FRGs, cat, cfr, cml, fexA, fexB, florB, and optrA have been discovered from animal origin food matrices (including finfish ones). Their biochemical actions are involved in several pathways, i.e., protein synthesis inhibition, exporter ability, methyltransferase activation, efflux pumps, etc. [129,130].
From a microbiological perspective applied to the veterinary clinical aspects, amphenicol administering has demonstrated biochemical repercussions on intestinal microbiota. It induces shifts among bacterial biodiversity acting as strong stressor [127].
This last consideration finds explanations from cyto-chemical interactions directly associated with the consequential expression of transmissible oligonucleotide sequences. Among the above-mentioned amphenicol-resistant determinants, the metagenomic technology, coupled with next generation sequencing, has identified multiple mutations on open reading frames regions, which encode resistant mechanisms, i.e., efflux pumps, new binding epitopes, etc. [126].
Innovative biomolecular technologies, combining thiamphenicol and florfenicol administering, has permitted us to reduce their respective dosages but preserve their therapeutic efficacy [131].
The notable ARGs heterogeneity and their extreme variabilities pose the basis for further diagnostic and One Health clinical challenges. Aminophenols, as with other previously mentioned antibiotic classes, are widely used in the aquaculture zootechnic sector. Therefore, it is mandatory to preserve their therapeutic actions.

6. Antibiotic Substitutions

6.1. Vaccination

FAO reports numerous administered vaccines against different bacterial or viral diseases among finfish species. The most frequently used provide seroconversion against Vibrio salmonicida, Vibrio anguillarum, Photobacterium damselae, Aeromonas salmonicida, Yersinia ruckeri, etc. [3].
Conversely, there are few vaccines for viral diseases, in which usage is highly recommended in marine finfish farms [132]. In the aquaculture farms, vaccines can be administered through different methods: injection, in bath, or through the orofecal route [3]. Injection, through the intraperitoneal route, provides powerful and durable protection, but, on the other hand, this procedure influences animal welfare, inducing a relevant stress condition. It is commonly used for Salmo salar finfish species but is not applicable for other species, i.e., Pangasius spp. and Tilapia spp. Conversely, oral administering reduces stress (due to animal handling), since animals receive immunization through food ingestion. The main difference between these two above-mentioned methods is represented by the need for large amounts of antigens in the ingestion method to obtain an adequate immunity [133]. There are contrasting opinions on vaccines’ efficacy regarding finfish farms. Usually, after vaccination, fish farmers must administer antibiotics to control infectious disease outbreaks [134]. This condition is related to an incomplete understanding of the vaccination type and the immune system’s reaction to the “antigenic stimuli”. It is improper to compare fish immune reactions with the generated response in mammalians [135].
However, in any species, such as farmed Atlantic salmons, vaccination represents an important preventive tool [3]. Farmed salmonids (Salmo salar) receive immune protection through the injection of a pentavalent vaccine against vibriosis, furunculosis, piscrickettsiosis, infectious pancreatic necrosis, and infection salmon anemia. The vaccine has permitted a reduced usage of antibiotics [135]. In tilapia’s farms, the mucosal administering route replaces the injective method. In this fish species, evidence supports a competent immune stimulation of the antigen-presenting cells (similarly to mammalians). In this way, fish farmers reduce antibiotic administering [136].
A new frontier is represented by nano-material vaccines, which use virus-like particles, immune-stimulating complexes, liposomes, polymeric, etc. These molecules drive antigens and can drive protective responses in fish. Furthermore, nanoparticles permit antigens’ release, and, for this reason, booster vaccinations are not necessary [137], but live attenuated vaccines’ employment in aquaculture is not allowed by the European Commission. Their usage has not yet been allowed due to the wide gap of knowledge concerning possible implications on human consumers [138].

6.2. Structural Improvements

Innovative production systems have become popular among fish farmers, i.e., catfish aquaculture in the USA [65]. New fish farming systems that provide more space and efficient wastewater management allow an avoidance of the large usage of antibiotic molecules [6].
Therefore, in the USA, fish farmers introduced an innovative system called “spilt-pond” to optimize fish’s sanitary conditions and productive levels. This new system is realized through the division of the traditional ponds in two areas: an algal growth basin and a fish holding area. In this way, the growth of production is allowed by the high animals’ density in the same period of production, and the reduction in antimicrobial use is due to the continuous water filtration [139]. Conversely, in Malaysia and other Asian-Pacific regions, fish are farmed by using pond culture, ex-mining pools, cement tanks, and freshwater pen culture systems. In these structures, there is low water filtration. Animal catabolites and feces remain for all productive cycles, producing a functional substrate for any bacterial species (i.e., Enterobacteriaceae) proliferation. Furthermore, in such countries of this continent (China, Vietnam, Philippines, India, etc.), the usage of antibiotics in aquaculture is not well regulated by national law [6]. Therefore, a new approach to the aquaculture systems of production is required. Indeed, Brunton et al. [140] generated a mapping system obtained through the stakeholders’ collaboration. Correlating ecological aspects to the new above-mentioned fish farms realities. It identifies hotspots and risk points related to antibiotic usage in the aquaculture food chain. The platform provides a quantitative risk analysis at different steps of production. Therefore, these maps allow us to understand the molecules’ flows, ARGs, and ARB. In this way, it is possible to monitor antibiotic resistance factors. From these elaborated data, food safety authorities may program control activities through surveillance measures.

6.3. Probiotics

Probiotics have different effects on fish farming issues, i.e., they reduce animal mortality (especially at the larval stage) [141,142,143], improve animal welfare through the immune system’s stimulation, and reduce the antibiotic therapies’ necessity. Fish farmers introduce these bacteria through the finfish diet, as supplementary feed [3]. Bacillus spp. Is largely used in numerous fish farms realities for its probiotic properties. This genus can mitigate pathogenic microorganisms’ growth and can eliminate ARB [143]. These capacities are related to the bioactive peptides’ synthesis (bacteriocin) [144], but there are also nonpeptidic molecules, i.e., phospholipids, polyketides, etc., that are classified as bacteriocin [145]. Bacillus spp. produces CAMT2; this molecule is a recent example of bacteriocin that inhibits the proliferation of different bacterial strains, i.e., Vibrio spp., Staphylococcus aureus, and Listeria monocytogenes [146]. Another interesting microbiological aspect of this genus is represented by bacterial competition. It includes competition for energy (obtained from substrates), nutrients, and adhesion sites [5]. Indeed, Bacillus spp. Can rapidly colonize organic substrates with strong adhesion capacities (due to hydrophobic and steric forces) [147]. Therefore, pathogenic bacteria find an inadequate micro-environment that results in hostility to their proliferation. Furthermore, Bacillus spp. also stimulates fish’s cell-mediated immune response. Indeed, bacterial pathogens decrease their virulence because the animal host presents a resistant and competent immune system [148].
Healthy animals require few antibiotic therapies, leading to the reduction in antibiotic consumption in the next ten years [3]. Thanks to these preventive measures, different finfish species (i.e., parrotfish) have become resistant to Vibrio alginolyticus infection. These considerations are strictly related to the concept that powerful immune systems reduce pathogen bacterial proliferation [60].

7. Conclusions

The AMR is a real public health issue [3]. This review aims to provide a current scenario about the circulation of ARGs in bacterial isolates, usually identified from aquaculture food industry chains. The authors also want to highlight that the AMR phenomenon is a dynamic concept, differing in time and regional areas among years. In this case, a particular and not limited ecosystem represents the main challenge. Oceans’, lakes’, rivers’ health requires periodical screenings.
Molecular biology has a key role in studying bacterial genomic mutations and ARGs horizontal transmission flows. In the future, marine currents’ studies will also have a crucial role filling any gaps of knowledge. They will allow the estimation of any geographical areas where ARGs will be widely diffused and improve the introduction of efficient corrective measures. In this way, a new concept of “Environmental Medicine” will produce factual results based on a holistic point of view.
Vaccination programs, probiotics’ administering, and structural improvements are three examples that represent valid alternative measures to reduce antibiotic usage.
These measures will produce satisfactory results if all nations of the world adopt them according to their ecosystems’ peculiarities.
These efforts are necessary due to the growing demand for animal origin finfish proteins and the increasing global demographic. However, these physiological necessities must be satisfied involving sustainable and innovative production systems.

Author Contributions

Conceptualization, G.F. and A.V.; methodology, G.F.; software, G.F.; validation, G.F., C.L. and A.V.; formal analysis, G.F.; investigation, G.F.; resources, G.F.; data curation, G.F.; writing—original draft preparation, G.F.; writing—review and editing, G.F., C.L. and A.V.; visualization, G.F.; supervision, A.V.; project administration, A.V.; funding acquisition, G.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

All authors wanted to express their appreciation to the post graduate specialization school in food inspection “G. Tiecco” that enabled this review to be prepared. This review article was intellectually supported by the EFS-MUR (European Social Fund–Italian University Ministry)–PON (Innovation and Research 2014–2020) CUP: C44I19001650006; Ph.D. program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Food and Agriculture Organization (FAO). Approved Drugs for Use in Seafoods; FAO: Rome, Italy, 2020; Available online: http://www.fao.org/3/ca9229en/ca9229en.pdf (accessed on 12 March 2021).
  2. United Nations (UN). The Sustainable Development Goals Report; UN: New York, NY, USA, 2018; Available online: https://unstats.un.org/sdgs/files/report/2018/TheSustainableDevelopmentGoalsReport2018-EN.pdf (accessed on 15 May 2021).
  3. Food and Agriculture Organization (FAO). The State of World Fisheries and Aquaculture 2018. Meeting the Sustainable Development Goals; FAO: Rome, Italy, 2018; Available online: http://www.fao.org/3/i9540en/i9540en.pdf (accessed on 7 June 2021).
  4. European Union Market Observatory for Fishery and Agriculture Products (EUMOFA). Case Study: Dried Salted Cod in Norway. Price Structure in the Supply Chain from Norway to Portugal March. 2018. Available online: www.EUMOFA.eu (accessed on 30 April 2021).
  5. Food and Drug Administration (FDA). Approved Drugs for Use in Seafoods; FDA:, New York. 2020. Available online: https://www.fda.gov/animal-veterinary/aquaculture/approved-aquaculture-drugs (accessed on 1 June 2021).
  6. Food and Agriculture Organization (FAO). Search Aquacultured Fact Sheets: Cultured Aquatic Species. 2016. Available online: http://www.fao.org/fishery/culturedspecies/search/en (accessed on 18 February 2021).
  7. World Health Organization (WHO). Report on Second WHO Expert Meeting, Copenhagen, Denmark: Critically Important Antimicrobials for Human Medicine; Categorization for the Development of Risk Management Strategies to Contain Antimicrobial Resistance Due to Non-Human Use. 2007. Available online: https://apps.who.int/iris/bitstream/handle/10665/43765/9789241595742_eng.pdf;jsessionid=A14A12F19C1EDC617AD74724BBD65ED2?sequence=1 (accessed on 4 September 2021).
  8. Froehlich, H.E.; Runge, C.A.; Gentry, R.R.; Gaines, S.D.; Halpern, B.S. Comparative terrestrial feed and land use of aquaculture-dominant world. Proc. Nat. Acad. Sci. USA 2018, 115, 5295–5300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Diana, J.S.; Egna, H.S.; Chopin, T.; Peterson, M.S.; Cao, L.; Pomeroy, R.; Verdegem, M.; Slack, W.T.; Bondad-Reantaso, M.G.; Cabello, F. Responsible aquaculture in 2050: Valuing local conditions and human innovations will be key to success. BioScience 2013, 6, 255–262. [Google Scholar] [CrossRef] [Green Version]
  10. Mo, W.Y.; Chen, Z.; Leung, H.M.; Leung, A.O.W. Application of veterinary antibiotics in China’s aquaculture industry and their potential human health risks. Environ. Sci. Pollut. Res. 2017, 24, 8978–8989. [Google Scholar] [CrossRef] [PubMed]
  11. Lulijwa, R.; Rupia, E.J.; Alfaro, A.C. Antibiotic use in aquaculture, policies and regulation, health and environmental risks: A review of the top 15 major producers. Rev. Aquacul. 2020, 12, 640–663. [Google Scholar] [CrossRef]
  12. Council Directive 96/23/EC of 29 April 1996 on measures to monitor certain substances and residues thereof in live animals and animal products and repealing Directives 85/358/EEC and 86/469/EEC and Decisions 89/187/EEC and 91/664/EEC. Off. J. Eur. Union L 125 1996, 125, 10–32.
  13. European Medicines Agency (EMA). European Surveillance of Veterinary Antimicrobial Consumption. 2018. Available online: https://www.ema.europa.eu/en/veterinary-regulatory/overview/antimicrobial-resistance/european-surveillance-veterinary-antimicrobial-consumption-esvac (accessed on 25 July 2021).
  14. European Medicines Agency (EMA). Committee for Veterinary Medicinal Products. Update of the Position. Paper on Availability of Veterinary Medicines Agreed on 21 June 2000. Veterinary Medicines Information Technology. 2000. Available online: https://www.ema.europa.eu/en/documents/position/position-paper-regarding-availability-products-minor-uses-minor-species-mums_en.pdf (accessed on 22 January 2022).
  15. Brasil. Ministerio da Agricultura, Pecuaria e Abastecimento. Secretaria de Defesa Agropecuaria. Instrucao Normativa n° 13. Aprova os Programas de Controle de Resıduos e Contaminantes em Carnes (Bovina, Aves, Suına e Equina), Leite, Pescado, Mel e Ovos para o exercıcio de 2015. Diario Oficial da Uniao. 2015. Available online: https://www.gov.br/agricultura/pt-br/assuntos/inspecao/produtos-animal/plano-de-nacional-de-controle-de-residuos-e-contaminantes/documentos-da-pncrc/pncrc-2015 (accessed on 26 February 2022).
  16. Ministry of Fishery (MF). Decision No. 07/2005/QD-BTS. “Issued by the Ministry of Fisheries Dated 24/2/2005 on the Lists of Chemicals and Antibiotics Banned or Restricted from Use in Fisheries Production and Trade (ed. Fisheries Mo)”. 2005. Available online: https://ouci.dntb.gov.ua/en/works/4rOyXvMl/ (accessed on 22 July 2022).
  17. Ministry of Agriculture and Rural Development (MARD). Circular No. 71/2009/TT-BNNPTNT. Issued by the Ministry of Agriculture and Rural Development dated 10/11/2009 on the Additional List of Products for Treating and Improving Aquaculture Environment Permitted for Circulation in Viet Nam. (ed. Development MoAaR). 2009. Available online: https://ouci.dntb.gov.ua/en/works/4rOyXvMl/ (accessed on 3 August 2022).
  18. Ministry of Agriculture and Rural Development (MARD). List of Drugs, Chemicals and Antibiotics of Banned or Limited Use for Aquaculture and Veterinary Purposes; Ministry of Agriculture and Rural Development: Hanoi, Vietnam, 2014. Available online: https://vanbanphapluat.co/circular-no-08-vbhn-bnnptnt-2014-list-of-banned-restricted-drugs-chemicals-antibiotics (accessed on 3 August 2022).
  19. National Service of Fishery and Aquaculture (NSFA). Report on the Use of Antimicrobials in National Salmon Farming 2015; National Service of Fishery and Aquaculture (NSFA): Valparaıso, Chile, 2016. [Google Scholar] [CrossRef]
  20. Ministry of Agriculture China (MAPRC). List of Banned Veterinary Drugs and Other Chemical Compounds for Food Animals. 2002. Available online: https://apps.fas.usda.gov/newgainapi/api/report/downloadreportbyfilename?filename=List%20of%20Veterinary%20Drugs%20Banned%20for%20Use%20for%20Food%20Animals_Beijing_China%20-%20Peoples%20Republic%20of_3-11-2011.pdf (accessed on 22 June 2022).
  21. Ministry of Agriculture China (MAPRC). Safety Food—Criterion for Usage of Fishery Drugs. 2002. Available online: https://www.fda.gov/media/80637/ (accessed on 12 April 2022).
  22. Ministry of Agriculture China (MAPRC). Administrative Regulation of Quality and Safety for Aquaculture. 2003. Available online: https://www.fda.gov/media/80637/ (accessed on 10 June 2022).
  23. Ministry of Agriculture China (MAPRC). Requirement for Water Discharge from Freshwater Aquaculture Pond. 2007. Available online: https://www.fda.gov/media/80637/ (accessed on 22 March 2022).
  24. Baoprasertkul, P.; Somsiri, T.; Boonyawiwat, V. Use of veterinary medicines in Thai aquaculture: Current status. In Improving Biosecurity through Prudent and Responsible Use of Veterinary Medicines in Aquatic Food Production; Bondad-Reantaso, M.G., Arthur, J.R., Subasinghe, R.P., Eds.; FAO: Rome, Italy, 2012; p. 83. Available online: https://www.fao.org/3/ba0056e/ba0056e.pdf (accessed on 22 March 2022).
  25. Ministry of Agriculture and Rural Development (MARD). Decision 06/2008/QD-BNN. Issued by the Ministry of Agriculture and Rural Development Dated 18/1/2008 on Amending and Supplementing the List of Animal Medicine for the Field of Fisheries and Allowed Materials Used in Improving the Aquaculture Environment. (ed. Development MoAaR). 2008. Available online: https://www.fao.org/3/ba0056e/ba0056e.pdf (accessed on 10 February 2022).
  26. Lozano, I.; Dıaz, N.F.; Munoz, S.; Riquelme, C. Antibiotics in Chilean Aquaculture: A Review. Antibiot. Use Anim. 2018, 3, 25–44. Available online: https://www.intechopen.com/chapters/57645 (accessed on 5 September 2022).
  27. Thiang, E.L.; Lee, C.W.; Takada, H.; Seki, K.; Takei, A.; Suzuki, S.; Wang, A.; Bong, C.W. Antibiotic residues from aquaculture farms and their ecological risks in Southeast Asia: A case study from Malaysia. Ecosyst. Health Sustain. 2021, 7, 1926337. [Google Scholar] [CrossRef]
  28. Lin, A.Y.C.; Yu, T.H.; Lin, C.F. Pharmaceutical Contamination in Residential, Industrial, and Agricultural Waste Streams: Risk to Aqueous Environments in Taiwan. Chemosphere 2008, 74, 131–141. [Google Scholar] [CrossRef]
  29. Xiong, W.; Sun, Y.; Zhang, T.; Ding, X.; Li, Y.; Wang, M.; Zeng, Z. Antibiotics, Antibiotic Resistance Genes, and Bacterial Community Composition in Fresh Water Aquaculture Environment in China. Microb. Ecol. 2015, 70, 425–432. [Google Scholar] [CrossRef]
  30. Lai, W.W.P.; Lin, Y.C.; Wang, Y.H.; Guo, Y.L.; Lin, A.Y.C. Occurrence of Emerging Contaminants in Aquaculture Waters: Cross-Contamination between Aquaculture Systems and Surrounding Waters. Water Air Soil Pollut. 2018, 229, 249. Available online: https://link.springer.com/article/10.1007/s11270-018-3901-3 (accessed on 2 March 2022). [CrossRef]
  31. Food and Drug Administration (FDA). U.S. Food and Drug Administration Import Refusals Report. 2018. Available online: https://www.accessdata.fda.gov/scripts/importrefusals/ (accessed on 2 March 2022).
  32. O’Neill, J. Antimicrobial Resistance: Tackling a Crisis for the Health and Wealth of Nations. The Review on Antimicrobial Resistance. 2014. Available online: https://amr-review.org/Publications.html (accessed on 2 March 2022).
  33. Amarasiri, M.; Sano, D.; Suzuki, S. Understanding human health risks caused by antibiotic resistant bacteria (ARB) and antibiotic resistance genes (ARG) in water environments: Current knowledge and questions to be answered. Crit. Rev. Environ. Sci. Technol. 2020, 50, 2016–2059. [Google Scholar] [CrossRef]
  34. Hendriksen, R.S.; Munk, P.; Njage, P.; van Bunnik, B.; McNally, L.; Lukjancenko, O.; Röder, T.; Nieuwenhuijse, D.; Pedersen, S.K.; Kjeldgaard, J.; et al. Global monitoring of antimicrobial resistance based on metagenomics analyses of urban sewage. Nat. Commun. 2019, 10, 1124. Available online: https://www.nature.com/articles/s41467-019-08853-3 (accessed on 2 March 2022). [CrossRef] [PubMed] [Green Version]
  35. Ferri, G.; Vergara, A. Hepatitis E Virus in the Food of Animal Origin: A Review. Foodborne Pathog. Dis. 2021, 18, 368–377. [Google Scholar] [CrossRef] [PubMed]
  36. O’Flaherty, E.; Borrego, C.M.; Balcazar, J.L.; Cummins, E. Human exposure assessment to antibiotic-resistant Escherichia coli through drinking water. Sci. Total. Environ. 2018, 616–617, 1356–1364. [Google Scholar] [CrossRef]
  37. Hatosy, S.M.; Martiny, A.C. The ocean as a global reservoir of antibiotic resistance genes. J. Appl. Environ. Microbiol. 2015, 81, 7593–7599. [Google Scholar] [CrossRef] [Green Version]
  38. Miranda, C.D.; Godoy, F.A.; Lee, M.R. Current status of the use of antibiotics and the antimicrobial resistance in the Chilean salmon farms. Front. Microbiol. 2018, 9, 1284. [Google Scholar] [CrossRef]
  39. Muziasari, W.I.; Pitkanen, L.K.; Sørum, H.; Stedtfeld, R.D.; Tiedje, J.M.; Virta, M. The resistome of farmed fish feces contributes to the enrichment of antibiotic resistance genes in sediments below Baltic Sea fish farms. Front. Microbiol. 2017, 7, 2137. [Google Scholar] [CrossRef] [Green Version]
  40. Tortorella, E.; Tedesco, P.; Esposito, F.P.; January, G.G.; Fani, R.; Jaspars, M.; de Pascale, D. Antibiotics from Deep-Sea Microorganisms: Current Discoveries and Perspectives. Marine Drugs 2018, 16, 355. [Google Scholar] [CrossRef] [Green Version]
  41. Reverter, M.; Sarter, S.; Caruso, D.; Avarre, J.C.; Combe, M.; Pepey, E.; Pouyaud, L.; Vega-Heredía, S.; de Verdal, H.; Gozlan, R.E. Aquaculture at the crossroads of global warming and antimicrobial resistance. Nat. Commun. 2020, 11, 1870. [Google Scholar] [CrossRef] [Green Version]
  42. Igbinosa, I.H.; Beshiru, A.; Igbinosa, E.O. Antibiotic resistance profile of Pseudomonas aeruginosa isolated from aquaculture and abattoir environments in urban communities. Asian Pac. J. Trop. Dis. 2017, 7, 47–52. [Google Scholar] [CrossRef]
  43. Reilly, A.; Käferstein, F. Food safety hazard and the application of the principles of hazard analysis and critical control point (HACCP) system for their control in aquaculture production. Aquacul. Res. 1997, 28, 735–752. [Google Scholar] [CrossRef]
  44. Vikesland, P.J.; Pruden, A.; Alvarez, P.; Aga, D.; Bürgmann, H.; Li, X.D.; Manaia, C.M.; Nambi, I.; Wigginton, K.; Zhang, T.; et al. Toward a Comprehensive Strategy to Mitigate Dissemination of Environmental Sources of Antibiotic Resistance. Environ. Sci. Technol. 2017, 51, 13061–13069. [Google Scholar] [CrossRef] [Green Version]
  45. Han, Q.F.; Zhao, S.; Zhang, X.R.; Wang, X.L.; Song, C.; Wang, S.G. Distribution, combined pollution and risk assessment of antibiotics in typical marine aquaculture farms surrounding the Yellow Sea, North China. Environ. Int. 2020, 138, 105551. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, Z.; Yu, D.; He, S.; Ye, H.; Zhang, L.; Wen, Y.; Zhang, W.; Shu, L.; Chen, S. Prevalence of antibiotic-resistant Escherichia coli in drinking water sources in Hangzhou city. Front. Microbiol. 2017, 8, 1133. [Google Scholar] [CrossRef] [Green Version]
  47. Van Boeckel, T.P.; Glennon, E.E.; Chen, D.; Gilbert, M.; Robinson, T.P.; Grenfell, B.T.; Levin, S.A.; Bonhoeffer, S.; Laxminarayan, R. Reducing antimicrobial use in food animals. Science 2017, 357, 1350–1352. [Google Scholar] [CrossRef] [Green Version]
  48. Schar, D.; Klein, E.Y.; Laxminarayan, R.; Gilbert, M.; Van Boeckel, T.P. Global trends in antimicrobial use in aquaculture. Sci. Rep. 2020, 10, 21878. [Google Scholar] [CrossRef] [PubMed]
  49. Shao, Y.; Wang, Y.; Yuan, Y.; Xie, Y. A systematic review on antibiotics misuse in livestock and aquaculture and regulation implications in China. Sci. Total Environ. 2021, 798, 149205. [Google Scholar] [CrossRef] [PubMed]
  50. Hossain, A.; Habibullah-Al-Mamun, M.; Nagano, I.; Masunaga, S.; Kitazawa, D.; Matsuda, H. Antibiotics, antibiotic-resistant bacteria, and resistance genes in aquaculture: Risks, current concern, and future thinking. Environ. Sci. Pollut. Res. Int. 2022, 29, 11054–11075. [Google Scholar] [CrossRef]
  51. Gravningen, K.; Sorum, H.; Horsberg, T.E. The future of therapeutic agents in aquaculture. Rev. Sci. Tech. 2019, 38, 641–651. [Google Scholar] [CrossRef]
  52. Romero, J.; Feijoó, C.G.; Navarrete, P. Antibiotics in Aquaculture-Use, Abuse and Alternatives. In Health and Environment in Aquaculture; Carvalho, E.D., David, G.S., Silva, R.J., Eds.; InTech: Rang-Du-Fliers, France, 2012; Volume 6, Available online: https://books.google.it/books?hl=it&lr=&id=f9CPDwAAQBAJ&oi=fnd&pg=PA159&dq=Antibiotics+in+Aquaculture-Use,+Abuse+and+Alternatives&ots=4xDCv9_EpS&sig=V1CoGUTkCRdKzkGdtJ (accessed on 22 July 2022).
  53. Kümmerer, K. Resistance in the Environment. J. Antimicrob. Chemother. 2004, 54, 311–320. [Google Scholar] [CrossRef] [Green Version]
  54. Hazrat, A.; Rico, A.; Murshed-e-jahan, K.; Belton, B. An Assessment of Chemical and Biological Product Use in Aquaculture in Bangladesh. Aquaculture 2016, 454, 199–209. [Google Scholar] [CrossRef]
  55. Shamsuzzaman, M.M.; Biswas, T.K. Aqua Chemicals in Shrimp Farm: A Study from South-west Coast of Bangladesh. Egypt J. Aquatic. Res. 2012, 38, 275–285. [Google Scholar] [CrossRef] [Green Version]
  56. Hassali, M.A.; Ho, R.Y.; Verma, A.K.; Hussain, R.; Sivaraman, S. Antibiotic Use in Food Animals: Malaysia Overview; Universiti Sains Malaysia: George Town, Malaysia, 2018; Available online: https://www.reactgroup.org/wp-content/uploads/2018/11/Antibiotic_Use_in_Food_Animals_Malaysia_Overview_2018web.pdf (accessed on 22 June 2022).
  57. Nonaka, L.; Ikeno, K.; Suzuki, S. Distribution of Tetracycline Resistance Gene, tet(M), in Gram-positive and Gram-negative Bacteria Isolated from Sediment and Seawater at a Coastal Aquaculture Site in Japan. Microb. Environ. 2007, 22, 355–364. [Google Scholar] [CrossRef] [Green Version]
  58. Suzuki, S.; Hoa, P.T.P. Distribution of Quinolones, Sulfonamides, Tetracyclines in Aquatic Environment and Antibiotic Resistance in Indochina. Front. Microbiol. 2012, 3, 67. [Google Scholar] [CrossRef] [Green Version]
  59. Chang, B.V.; Ren, Y.L. Biodegradation of Three Tetracyclines in River Sediment. Ecol. Engin. 2015, 75, 272–277. [Google Scholar] [CrossRef]
  60. Yang, Y.; Zeng, G.; Huang, D.; Zhang, C.; He, D.; Zhou, C.; Wang, W. Molecular Engineering of Polymeric Carbon Nitride for Highly Efficient Photocatalytic Oxytetracycline Degradation and H2O2 Production. Appl. Catal. Environ. 2020, 272, 118970. [Google Scholar] [CrossRef]
  61. Brooks, B.W.; Maul, J.D.; Belden, J.B. Antibiotics in Aquatic and Terrestrial Ecosystems. Encyclop. Ecol. 2008, 1, 210–217. [Google Scholar] [CrossRef]
  62. Leal, J.F.; Esteves, V.I.; Santos, E.B.H. Use of Sunlight to Degrade Oxytetracycline in Marine Aquaculture’s Waters. Environ. Pollut. 2016, 213, 932–939. [Google Scholar] [CrossRef]
  63. Liu, X.; Steele, J.C.; Meng, X.Z. Usage, Residue, and Human Health Risk of Antibiotics in Chinese Aquaculture: A Review. Environ. Pollut. 2017, 223, 161–169. [Google Scholar] [CrossRef]
  64. Igbinosa, I.H.; Igbinosa, E.O.; Okoh, A.I. Molecular detection of metallo-β-lactamase and putative virulence genes in environmental isolates of Pseudomonas species. Pol. J. Environ. Stud. 2014, 23, 2327–2331. Available online: http://www.pjoes.com/pdf-89415-23293?filename=Molecular%20Detection%20of.pdf (accessed on 19 June 2022).
  65. Chuah, L.O.; Effarizah, M.E.; Goni, A.M.; Rusul, G. Antibiotic application and emergence of multiple antibiotic resistance (MAR) in global catfish aquaculture. Curr. Environ. Health Rep. 2016, 3, 118–127. [Google Scholar] [CrossRef] [PubMed]
  66. Oyinlola, M.A.; Reygondeau, G.; Wabnitz, C.; Frölicher, T.L.; Lam, V.; Cheung, W. Projecting global mariculture production and adaptation pathways under climate change. Glob. Chang. Biol. 2022, 28, 1315–1331. [Google Scholar] [CrossRef] [PubMed]
  67. Ferri, M.; Ranucci, E.; Romagnoli, P.; Giaccone, V. Book Antibiotico-Resistenza: Una Minaccia Globale per la Sanità Pubblica e i Sistemi Sanitari; Chapter 2; Piano Prevenzione 2014–2018 ASL Roma 1; 2017 Crit. Rev. Food Sci. Nutr. Available online: https://www.veterinariapreventiva.it/wp-content/uploads/2019/11/027-050_CAPITOLO-2.2.pdf (accessed on 3 March 2022).
  68. Devarajan, N.; Laffite, A.; Mulaji, C.K.; Otamonga, J.P.; Mpiana, P.T.; Mubedi, J.I.; Prabakar, K.; Ibelings, B.W.; Pote, J. Occurrence of antibiotic resistance genes and bacterial markers in a tropical river receiving hospital and urban wastewaters. PLoS ONE 2016, 11, e0149211. [Google Scholar] [CrossRef] [PubMed]
  69. Nnadozie, C.F.; Odume, O.N. Freshwater environments as reservoirs of antibiotic resistant bacteria and their role in the dissemination of antibiotic resistance genes. Environ. Pollut. 2019, 254, 113067. [Google Scholar] [CrossRef] [PubMed]
  70. Piedra-Carrasco, N.; Fàbrega, A.; Calero-Cáceres, W.; Cornejo-Sánchez, T.; Brown-Jaque, M.; Mir-Cros, A.; Muniesa, M.; González-López, J.J. Carbapenemase-producing Enterobacteriaceae recovered from a Spanish river ecosystem. PLoS ONE 2017, 12, e0175246. [Google Scholar] [CrossRef] [Green Version]
  71. Szekeres, E.; Chiriac, C.M.; Baricz, A.; Szőke-Nagy, T.; Lung, I.; Soran, M.L.; Rudi, K.; Dragos, N.; Coman, C. Investigating antibiotics, antibiotic resistance genes, and microbial contaminants in groundwater in relation to the proximity of urban areas. Environ. Pollut. 2018, 236, 734–744. [Google Scholar] [CrossRef]
  72. Ginovyan, M.; Hovsepyan, V.; Sargsyan, M.; Grigoryan, K.; Trchounian, A. Antibiotic resistance of Pseudomonas species isolated from Armenian fish farms. Process Union Res Inst Marine Fish Oceanograph 2017, 167, 163–173. Available online: http://aquacultura.org/upload/files/pdf/library/trudy/167-20.pdf (accessed on 22 March 2022).
  73. Sacramento, A.G.; Fernandes, M.R.; Sellera, F.P.; Dolabella, S.S.; Zanella, R.C.; Cerdeira, L.; Lincopan, N. VanA-type vancomycin-resistant Enterococcus faecium ST1336 isolated from mussels in an anthropogenically impacted ecosystem. Marine Pollut. Bull. 2019, 142, 533–536. [Google Scholar] [CrossRef]
  74. Ferreira, A.; Pavelquesi, S.; Monteiro, E.; Rodrigues, L.; Silva, C.; Silva, I.; Orsi, D.C. Prevalence and Antimicrobial Resistance of Salmonella spp. in Aquacultured Nile Tilapia (Oreochromis niloticus) Commercialized in Federal District, Brazil. Foodborne Pathog. Dis. 2021, 18, 778–783. [Google Scholar] [CrossRef]
  75. Araújo, A.; Grassotti, T.T.; Frazzon, A. Characterization of Enterococcus spp. isolated from a fish farming environment in southern Brazil. Brazil. J. Biol. 2021, 81, 954–961. [Google Scholar] [CrossRef]
  76. Fu, H.; Yu, P.; Liang, W.; Kan, B.; Peng, X.; Chen, L. Virulence, Resistance, and Genomic Fingerprint Traits of Vibrio cholerae Isolated from 12 Species of Aquatic Products in Shanghai, China. Microb. Drug Res. 2020, 26, 1526–1539. [Google Scholar] [CrossRef] [PubMed]
  77. Feng, Y.; Bai, M.; Geng, Y.; Chen, D.; Huang, X.; Ouyang, P.; Guo, H.; Zuo, Z.; Huang, C.; Lai, W. The potential risk of antibiotic resistance of Streptococcus iniae in sturgeon cultivation in Sichuan, China. Environ. Sci. Pollut. Res. Int. 2021, 28, 69171–69180. [Google Scholar] [CrossRef] [PubMed]
  78. Zhu, W.; Zhou, S.; Chu, W. Comparative proteomic analysis of sensitive and multi-drug resistant Aeromonas hydrophila isolated from diseased fish. Microb. Pathog. 2020, 139, 103930. [Google Scholar] [CrossRef] [PubMed]
  79. Preena, P.G.; Dharmaratnam, A.; Swaminathan, T.R. Antimicrobial Resistance analysis of Pathogenic Bacteria Isolated from Freshwater Nile Tilapia (Oreochromis niloticus) Cultured in Kerala, India. Cur. Microbiol. 2020, 77, 3278–3287. [Google Scholar] [CrossRef] [PubMed]
  80. Basha, K.A.; Kumar, N.R.; Das, V.; Reshmi, K.; Rao, B.M.; Lalitha, K.V.; Joseph, T.C. Prevalence, molecular characterization, genetic heterogeneity and antimicrobial resistance of Listeria monocytogenes associated with fish and fishery environment in Kerala, India. Lett. Appl. Microbiol. 2019, 69, 286–293. [Google Scholar] [CrossRef]
  81. Labella, A.; Gennari, M.; Ghidini, V.; Trento, I.; Manfrin, A.; Borrego, J.J.; Lleo, M.M. High incidence of antibiotic multi-resistant bacteria in coastal areas dedicated to fish farming. Marine Pollut. Bull. 2013, 70, 197–203. [Google Scholar] [CrossRef]
  82. Nguyen, H.N.; Van, T.T.; Nguyen, H.T.; Smooker, P.M.; Shimeta, J.; Coloe, P.J. Molecular characterization of antibiotic resistance in Pseudomonas and Aeromonas isolates from catfish of the Mekong Delta, Vietnam. Vet. Microbiol. 2014, 171, 397–405. [Google Scholar] [CrossRef]
  83. Boss, R.; Overesch, G.; Baumgartner, A. Antimicrobial Resistance of Escherichia coli, Enterococci, Pseudomonas aeruginosa, and Staphylococcus aureus from Raw Fish and Seafood Imported into Switzerland. J. Food Protect. 2016, 79, 1240–1246. [Google Scholar] [CrossRef]
  84. Koudou, A.A.; Kakou-Ngazoa, S.; Konan, K.F.; Aka, E.; Addablah, A.; Coulibaly N’Golo, D.; Kouassi, S.; Sina, M.K.; Atta Diallo, H.; Guessend, N.; et al. Occurrence of multidrug-resistant bacteria in aquaculture farms in Côte d’Ivoire (West Africa). Afr. J. Microbiol. Res. 2020, 14, 182–188. Available online: https://academicjournals.org/journal/AJMR/article-full-text-pdf/D4E869B63678 (accessed on 2 July 2022).
  85. Rezai, R.; Ahmadi, E.; Salimi, B. Prevalence and Antimicrobial Resistance Profile of Listeria Species Isolated from Farmed and On-Sale Rainbow Trout (Oncorhynchus mykiss) in Western Iran. J. Food Protect. 2018, 81, 886–891. [Google Scholar] [CrossRef]
  86. Colquhoun, D.J.; Aarflot, L.; Melvold, C.F. gyrA and parC mutations and associated quinolone resistance in Vibrio anguillarum serotype O2b strains isolated from farmed Atlantic cod (Gadus morhua) in Norway. Antimicrob. Agents Chemother. 2007, 51, 2597–2599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Izumi, S.; Ouchi, S.; Kuge, T.; Arai, H.; Mito, T.; Fujii, H. PCR–RFLP genotypes associated with quinolone resistance in isolates of Flavobacterium psy- chrophilum. J. Fish Dis. 2007, 30, 141–147. [Google Scholar] [CrossRef] [PubMed]
  88. Shah, S.Q.; Nilsen, H.; Bottolfsen, K.; Colquhoun, D.J.; Sørum, H. DNA gyrase and topoisomerase IV mutations in quinolone-resistant Flavobacterium psychrophilum isolated from dis- eased salmonids in Norway. Microb. Drug Res. 2012, 18, 207–214. [Google Scholar] [CrossRef] [PubMed]
  89. Lukkana, M.; Wongtavatchai, J.; Chuanchuen, R. Class 1 integrons in Aeromonas hydrophila isolates from farmed Nile tilapia (Oreochromis nilotica). J. Vet. Med. Sci. 2012, 74, 435–440. [Google Scholar] [CrossRef] [PubMed]
  90. Goñi-Urriza, M.; Arpin, C.; Capdepuy, M.; Dubois, V.; Caumette, P.; Quentin, C. Type II topoisomerase quinolone resistance- determining regions of Aeromonas caviae, A. hydrophila and A. sobria complexes and mutations associated with quinolone resistance. Antimicrob. Agents Chemother. 2002, 46, 350–359. [Google Scholar] [CrossRef] [Green Version]
  91. Kim, S.R.; Nonaka, L.; Suzuki, S. Occurrence of tetracycline resistance genes tet(M) and tet(S) in bacteria from marine aqua- culture sites. FEMS Microbiol. 2004, 237, 147–156. [Google Scholar] [CrossRef]
  92. Pepi, M.; Focardi, S. Antibiotic-Resistant Bacteria in Aquaculture and Climate Change: A Challenge for Health in the Mediterranean Area. Int. J. Environ. Res. Public Health 2021, 18, 5723. [Google Scholar] [CrossRef]
  93. Yan, L.; Liu, D.; Wang, X.H.; Wang, Y.; Zhang, B.; Wang, M.; Xu, H. Bacterial plasmid-mediated quinolone resistance genes in aquatic environments in China. Sci. Rep. 2017, 7, 40610. [Google Scholar] [CrossRef] [Green Version]
  94. Jiang, H.; Tang, D.; Liu, Y.; Zhang, X.; Zeng, Z.; Xu, L.; Hawkey, P. Prevalence and characteristics of β-lactamase and plasmid-mediated quinolone resistance genes in Escherichia coli isolated from farmed fish in China. J. Antimicrob. Chemother. 2012, 67, 2350–2353. [Google Scholar] [CrossRef]
  95. Dobiasova, H.; Kutilova, I.; Piackova, V.; Vesely, T.; Cizek, A.; Dolejska, M. Ornamental fish as a source of plasmid-mediated quinolone resistance genes and antibiotic resistance plasmids. Vet. Microbiol. 2014, 171, 413–421. [Google Scholar] [CrossRef]
  96. Tomova, A.; Ivanova, L.; Buschmann, A.H.; Godfrey, H.P.; Cabello, F.C. Plasmid-Mediated Quinolone Resistance (PMQR) Genes and Class 1 Integrons in Quinolone-Resistant Marine Bacteria and Clinical Isolates of Escherichia coli from an Aquacultural Area. Microb. Ecol. 2017, 75, 104–112. [Google Scholar] [CrossRef] [PubMed]
  97. Khajanchi, B.K.; Fadl, A.A.; Borchardt, M.A.; Berg, R.L.; Horneman, A.J.; Stemper, M.E.; Joseph, S.W.; Moyer, N.P.; Sha, J.; Chopra, A.K. Distribution of Virulence Factors and Molecular Fingerprinting of Aeromonas Species Isolates from Water and Clinical Samples: Suggestive Evidence of Water-to-Human Transmission. Appl. Environ. Microbiol. 2010, 76, 2313–2325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Cattoir, V.; Poirel, L.; Mazel, D.; Soussy, C.J.; Nordmann, P. Vibrio splendidus as the source of plasmid-mediated QnrS-Like quinolone resistance determinants. Antimicrob. Agents Chemother. 2007, 51, 2650–2651. [Google Scholar] [CrossRef] [Green Version]
  99. Mukwabi, D.M.; Okemo, P.O.; Otieno, S.A.; Oduor, R.O.; Okwany, Z.W. Antibiotic Resistant Pathogenic Bacteria Isolated from Aquaculture System in Bungoma County, Kenya. J. Appl. Environ. Microbioliol. 2019, 7, 25–37. Available online: https://ir-library.ku.ac.ke/bitstream/handle/123456789/21128/Antibiotic%20resistant%20pathogenic%20bacteria%20isolated%20from%20Aquaculture%20systems.pdf?sequence=1 (accessed on 3 March 2022).
  100. Heuer, H.; Schmitt, H.; Smalla, K. Antibiotic resistance gene spread due to manure application on agricultural fields. Curr. Opin. Microbiol. 2011, 14, 236–243. [Google Scholar] [CrossRef]
  101. Zhao, S.; Wei, W.; Fu, G.; Zhou, J.; Wang, Y.; Li, X.; Ma, L.; Fang, W. Application of biofertilizers increases fluoroquinolone resistance in Vibrio parahaemolyticus isolated from aquaculture environments. Marine Pollut. Bull. 2020, 150, 110592. [Google Scholar] [CrossRef]
  102. Speer, B.S.; Shoemaker, N.B.; Salyers, A.A. Bacterial resistance to tetracycline: Mechanisms, transfer, and clinical significance. Clin. Microbiol. Rev. 1992, 5, 387–399. [Google Scholar] [CrossRef] [PubMed]
  103. Taylor, D.E.; Chau, A. Tetracycline resistance mediated by ribosomal protection. Antimicrob. Agents Chemother. 1996, 40, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Roberts, M.C.; Schwarz, S.; Aarts, H.J.M. Erratum: Acquired antibiotic resistance genes: An overview. Front. Microbiol. 2012, 3, 384. [Google Scholar] [CrossRef] [Green Version]
  105. Obayashi, Y.; Kadoya, A.; Kataoka, N.; Kanda, K.; Bak, S.M.; Iwata, H.; Suzuki, S. Tetracycline Resistance Gene Profiles in Red Seabream (Pagrus major) Intestine and Rearing Water After Oxytetracycline Administration. Front. Microbiol. 2020, 11, 1764. [Google Scholar] [CrossRef]
  106. Roberts, M.C.; No, D.; Kuchmiy, E.; Miranda, C.D. Tetracycline resistance gene tet(39) identified in three new genera of bacteria isolated in 1999 from Chilean salmon farms. J. Antimicrob. Chemother. 1999, 70, 619–621. [Google Scholar] [CrossRef] [Green Version]
  107. Higuera-Llante´n, S.; Va´squez-Ponce, F.; Barrientos-Espinoza, B.; Mardones, F.O.; Marshall, S.H.; Olivares-Pacheco, J. Extended antibiotic treatment in salmon farms select multiresistant gut bacteria with a high prevalence of antibiotic resistance genes. PLoS ONE 2018, 13, e0203641. [Google Scholar] [CrossRef] [Green Version]
  108. Concha, C.; Miranda, C.D.; Santander, J.; Roberts, M.C. Genetic Characterization of the Tetracycline Resistance Gene tet(X) Carried by Two Epilithonimonas Strains Isolated from Farmed Diseased Rainbow Trout, Oncorhynchus mykiss in Chile. Antibiotics 2021, 10, 1051. [Google Scholar] [CrossRef] [PubMed]
  109. Han, J.E.; Mohney, L.L.; Tang, K.F.J.; Pantoja, C.R.; Lightner, D.V. Plasmid mediated tetracycline resistance of Vibrio parahaemolyticus associated with acute hepatopancreatic necrosis disease (AHPND) in shrimps. Aquacul. Rep. 2015, 2, 17–21. [Google Scholar] [CrossRef] [Green Version]
  110. Agersø, Y.; Bruun, M.S.; Dalsgaard, I.; Larsen, J.L. The tetracycline resistance gene tet(E) is frequently occurring and present on large horizontally transferable plas- mids in Aeromonas spp. from fish farms. Aquaculture 2007, 266, 47–52. [Google Scholar] [CrossRef]
  111. Miranda, C.D.; Kehrenberg, C.; Ulep, C.; Schwarz, S.; Roberts, M.C. Diversity of tetracycline resistance genes in bacteria from Chilean salmon farms. Antimicrob. Agents Chemother. 2003, 47, 883–888. [Google Scholar] [CrossRef] [Green Version]
  112. Su, H.C.; Ying, G.G.; Tao, R.; Zhang, R.Q.; Fogarty, L.R.; Kolpin, D.W. Occurrence of antibiotic resistance and characterization of resistance genes and integrons in Enterobacteriaceae isolated from integrated fish farms in South China. J. Environ. Monit. 2011, 13, 3229–3236. [Google Scholar] [CrossRef]
  113. Gao, P.; Mao, D.; Luo, Y.; Wang, L.; Xu, B.; Xu, L. Occurrence of sulfonamide and tetracycline-resistant bacteria and resistance genes in aquaculture environment. Water Res. 2012, 46, 2355–2364. [Google Scholar] [CrossRef]
  114. Sørum, H. Antimicrobial drug resistance in fish pathogens. In Antimicrobial Resistance in Bacteria of Animal Origin; Wiley Online Library: Hoboken, NJ, USA, 2006; pp. 213–238. [Google Scholar] [CrossRef]
  115. Serrano, P.H. Responsible Use of Antibiotics in Aquaculture; FAO Fisheries Technical Paper; FAO: Rome, Italy, 2005; p. 469. Available online: https://books.google.it/books?hl=it&lr=&id=f3Eql-OUOrAC&oi=fnd&pg=PR9&dq=Responsible+use+of+antibiotics+in+aquaculture&ots=by9D2xmv0K&sig=7LtKD2EFrqNSS8spEK5nA1pAVho&redir_esc=y#v=onepage&q=Responsible%20use%20of%20antibiotics%20in%20aquaculture&f=false (accessed on 2 March 2021).
  116. Rådström, P.; Swedberg, G.; Sköld, O. Genetic analyses of sulfonamide resistance and its dissemination in gram-negative bacteria illustrate new aspects of R plasmid evolution. Antimicrob. Agents Chemother. 1991, 35, 1840–1848. [Google Scholar] [CrossRef] [Green Version]
  117. Enne, V.I.; Livermore, D.M.; Stephens, P.; Hall, L.M. Persistence of sulphonamide resistance in Escherichia coli in the UK despite national prescribing restriction. Lancet 2001, 357, 1325–1328. [Google Scholar] [CrossRef]
  118. Nakayama, T.; Tran, T.T.H.; Harada, K.; Warisaya, M.; Asayama, M.; Hinenoya, A.; Lee, J.W.; Phu, T.M.; Ueda, S.; Sumimura, Y. Water metagenomic analysis reveals low bacterial diversity and the presence of antimicrobial residues and resistance genes in a river containing wastewater from backyard aquacultures in the Mekong Delta, Vietnam. Environ. Pollut. 2017, 222, 294–306. [Google Scholar] [CrossRef] [PubMed]
  119. Kim, D.W.; Thawng, C.N.; Lee, K.; Wellington, E.M.H.; Cha, C.J. A novel sulfonamide resistance mechanism by two-component flavin-dependent monooxygenase system in sulfonamide-degrading actinobacteria. Environ. Int. 2019, 127, 206–215. [Google Scholar] [CrossRef] [PubMed]
  120. Muziasari, W.I.; Managaki, S.; Pärnänen, K.; Karkman, A.; Lyra, C.; Tamminen, M. Sulfonamide and Trimethoprim Resistance Genes Persist in Sediments at Baltic Sea Aquaculture Farms but Are Not Detected in the Surrounding Environment. PLoS ONE 2014, 9, e92702. [Google Scholar] [CrossRef] [PubMed]
  121. Domínguez, M.; Miranda, C.D.; Fuentes, O.; de la Fuente, M.; Godoy, F.A.; Bello-Toledo, H.; González-Rocha, G. Occurrence of Transferable Integrons and sul and dfr Genes Among Sulfonamide-and/or Trimethoprim-Resistant Bacteria Isolated From Chilean Salmonid Farms. Front. Microbiol. 2019, 10, 748. [Google Scholar] [CrossRef] [PubMed]
  122. Monika, H.; Ewa, K.; Iwona, G. The impact of a freshwater fish farm on the community of tetracycline resistant bacteria and the structure of tetracycline resistance genes in river water. Chemosphere 2014, 128, 134–141. [Google Scholar] [CrossRef]
  123. Wu, J.J.; Su, Y.L.; Deng, Y.Q.; Guo, Z.X.; Mao, C.; Liu, G.F.; Xu, L.W.; Cheng, C.H.; Bei, L.; Feng, J. Prevalence and distribution of antibiotic resistance in marine fish farming areas in Hainan, China. Sci. Total Environ. 2019, 653, 605–611. [Google Scholar] [CrossRef]
  124. Yuan, J.L.; Ni, M.; Liu, M.; Zheng, Y.; Gu, Z.M. Occurrence of antibiotics and antibiotic resistance genes in a typical estuary aquaculture region of Hangzhou Bay, China. Marine Pollut. Bull. 2019, 138, 376–384. [Google Scholar] [CrossRef]
  125. Liang, X.M.; Nie, X.P.; Shi, Z. Preliminary studies on the occurrence of antibiotic resistance genes in typical aquaculture area of the Pearl River Estuary. Chin. J. Environ. Sci. 2013, 34, 4073–4080. Available online: https://europepmc.org/article/med/24364333 (accessed on 25 August 2022).
  126. Jung, H.N.; Park, D.H.; Choi, Y.J.; Kang, S.H.; Cho, H.J.; Choi, J.M.; Shim, J.H.; Zaky, A.A.; Abd El-Aty, A.M.; Shin, H.C. Simultaneous Quantification of Chloramphenicol, Thiamphenicol, Florfenicol, and Florfenicol Amine in Animal and Aquaculture Products Using Liquid Chromatography-Tandem Mass Spectrometry. Front. Nutr. 2022, 8, 812803. [Google Scholar] [CrossRef]
  127. Zeng, Q.; Liao, C.; Terhune, J.; Wang, L. Impacts of florfenicol on the microbiota landscape and resistome as revealed by metagenomic analysis. Microbiome 2019, 7, 155. [Google Scholar] [CrossRef] [Green Version]
  128. Xie, X.; Wang, B.; Pang, M.; Zhao, X.; Xie, K.; Zhang, Y.; Wang, Y.; Guo, Y.; Liu, C.; Bu, X.; et al. Quantitative analysis of chloramphenicol, thiamphenicol, florfenicol and florfenicol amine in eggs via liquid chromatography-electrospray ionization tandem mass spectrometry. Food Chem. 2018, 269, 542–548. [Google Scholar] [CrossRef] [PubMed]
  129. Wang, Y.; Lv, Y.; Cai, J.; Schwarz, S.; Cui, L.; Hu, Z.; Zhang, R.; Li, J.; Zhao, Q.; He, T.; et al. A novel gene, optrA, that confers transferable resistance to oxazolidinones and phenicols and its presence in Enterococcus faecalis and Enterococcus faecium of human and animal origin. J. Antimicrob. Chemother. 2015, 70, 2182–2190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Figueroa, J.; Castro, D.; Lagos, F.; Cartes, C.; Isla, A.; Yáñez, A.J.; Avendaño-Herrera, R.; Haussmann, D. Analysis of single nucleotide polymorphisms (SNPs) associated with antibiotic resistance genes in Chilean Piscirickettsia salmonis strains. J. Fish Dis. 2019, 42, 1645–1655. [Google Scholar] [CrossRef]
  131. Assaneab, I.M.; Suemi, K.; Moraes, G.G.; Valladãoac, R.; Pilarskia, F. Combination of antimicrobials as an approach to reduce their application in aquaculture: Emphasis on the use of thiamphenicol/florfenicol against Aeromonas hydrophila. Aquaculture 2019, 507, 238–245. [Google Scholar] [CrossRef]
  132. Rodger, H.D. Fish Disease Causing Economic Impact in Global Aquaculture. In Fish Vaccines. Birkhäuser Advances in Infectious Diseases; Adams, A., Ed.; Springer: Basel, Switzerland, 2016; Available online: https://link.springer.com/chapter/10.1007/978-3-0348-0980-1_1 (accessed on 3 March 2022).
  133. Lakshmi, B.; Syed, S.; Buddolla, V. Chapter 10—Current advances in the protection of viral diseases in aquaculture with special reference to vaccination. In Recent Developments in Applied Microbiology and Biochemistry; Elsevier: Amsterdam, The Netherlands, 2019; pp. 127–146. [Google Scholar] [CrossRef]
  134. Center for Disease Control and Prevention (CDC). Get Smart: Know when Antibiotics Work; Center for Disease Control: Atlanta, GA, USA, 2010; Available online: https://apua.org/cdc-get-smart-week (accessed on 12 May 2022).
  135. Flores-Kossack, C.; Montero, R.; Köllner, B.; Maisey, K. Chilean aquaculture and the new challenges: Pathogens, immune response, vaccination and fish diversification. Fish Shellfish Immunol. 2020, 98, 52–67. [Google Scholar] [CrossRef]
  136. Munang’andu, H.M.; Mutoloki, S.; Evensen, Ø.A. Review of the Immunological Mechanisms Following Mucosal Vaccination of Finfish. Front. Immunol. 2015, 6, 427. [Google Scholar] [CrossRef] [Green Version]
  137. Ji, J.; Torrealba, D.; Ruyra, À.; Roher, N. Nanodelivery Systems as New Tools for Immunostimulant or Vaccine Administration: Targeting the Fish Immune System. Biology 2015, 4, 664–696. [Google Scholar] [CrossRef] [Green Version]
  138. Brudeseth, B.E.; Wiulsrød, R.; Fredriksen, B.N.; Lindmo, K.; Løkling, K.E.; Bordevik, M.; Steine, N.; Klevan, A.; Gravningen, K. Status and future perspectives of vaccines for industrialised fin-fish farming. Fish Shellfish Immunol. 2013, 35, 1759–1768. [Google Scholar] [CrossRef]
  139. United States Department of Agriculture (UNDA). Performance Evaluation of Intensive, Pond-Based Culture Systems for Catfish Production; Southern Regional Aquaculture Center (USA, New York): 2014. Available online: https://srac.msstate.edu/pdfs/APRS%20and%20Summary/27th%20APR%20web-%20indivual%20reports/Intensive%20Systems%20CR%20for%202014.pdf (accessed on 2 February 2022).
  140. Brunton, L.A.; Desbois, A.P.; Garza, M.; Wieland, B.; Mohan, C.V.; Häsler, B.; Tam, C.C.; Le, P.; Phuong, N.T.; Van, P.T.; et al. Identifying hotspots for antibiotic resistance emergence and selection, and elucidating pathways to human exposure: Application of a systems-thinking approach to aquaculture systems. Sci. Total Environ. 2019, 687, 1344–1356. [Google Scholar] [CrossRef]
  141. Lauzon, H.L.; Gudmundsdottir, S.; Steinarsson, A.; Oddgerisson, M.; Peturdottir, S.K.; Reynisson, E.; Bjorndottir, R.; Gudumundsdottir, B.K. Effects of bacterial treatment at early stages of Atalntic cod (Gadus morhua L.) on larval survival and development. J. Appl. Microbiol. 2010, 108, 624–632. [Google Scholar] [CrossRef]
  142. Makridis, P.; Martins, S.; Reis, J.; Dinis, M.T. Use of probiotic bacteria in the rearing of Senegalese sole (Solea senegalensis) larvae. Aquacul. Res. 2008, 39, 627–634. [Google Scholar] [CrossRef]
  143. Plante, S.; Pernet, F.; Hache, R.; Ritchie, R.; Ji, B.J.; McIntosh, D. Ontogenetic variations in lipid class and fatty acid composition of haddock larvae Melanogrammus aeglefinus in relation to changes in diet and microbial environment. Aquaculture 2007, 263, 107–121. [Google Scholar] [CrossRef]
  144. Zou, J.; Jiang, H.; Cheng, H.; Fang, J.; Huang, G. Strategies for screening, purification and characterization of bacteriocins. Int. J. Biol. Macromol. 2018, 117, 781–789. [Google Scholar] [CrossRef] [PubMed]
  145. Stein, T. Bacillus subtilis antibiotics: Structures, syntheses and specific functions. Mol. Microbiol. 2005, 56, 845–857. [Google Scholar] [CrossRef]
  146. An, J.; Zhu, W.; Liu, Y.; Zhang, X.; Sun, L.; Hong, P.; Wang, Y.; Xu, C.; Liu, H. Purification and characterization of a novel bacteriocin CAMT2 produced by Bacillus amyloliquefaciens isolated from marine fish Epinephelus areolatus. Food Control 2015, 51, 278–282. [Google Scholar] [CrossRef]
  147. Mohapatra, S.; Chakraborty, T.; Kumar, V.; DeBoeck, G.; Mohanta, K.N. Aquaculture and stress management: A review of probiotic intervention. J. Animal Physiol. Animal Nutr. 2013, 97, 405–430. [Google Scholar] [CrossRef]
  148. Wilson, A.B. MHC and adaptive immunity in teleost fishes. Immunogen 2017, 69, 521–528. [Google Scholar] [CrossRef]
Table 1. Pharmacologically active substances and their classification regarding maximum residue limits (MRL) in foodstuffs of animal origin (from EU Reg. No. 37/2010).
Table 1. Pharmacologically active substances and their classification regarding maximum residue limits (MRL) in foodstuffs of animal origin (from EU Reg. No. 37/2010).
Pharmacologically Active SubstanceMarker ResidueAnimal SpeciesMRL*
BenzylpenicillinBenzylpenicillinAll other food-producing species.50 µg/kg
ChlortetracyclineSum of parent drug and its 4-epimerFin fish (all other food-producing species).100 µg/kg
CloxacillinCloxacillinFin fish (all other food-producing species).300 µg/kg
ColistinColistinFin fish (all other food-producing species).150 µg/kg
DanofloxacinDanofloxacinFin fish (all other food-producing species).100 µg/kg
DicloxacillinDicloxacillinFin fish (all other food-producing species).300 µg/kg
DifloxacinDifloxacinFin fish (all other food-producing species).300 µg/kg
EnrofloxacinEnrofloxacinFin fish100 µg/kg
ErythromycinErythromycin AFin fish200 µg/kg
FlorfenicolSum of florfenicol and its metabolites measured as florfenicol amineFin fish1000 µg/kg
FlumequineFlumequineFin fish600 µg/kg
LincomycinLincomycinFin fish (all other food-producing species).1000 µg/kg
Neomycin (including Framycetin)Neomycin BFin fish (all other food-producing species).500 µg/kg
OxacillinOxacillinFin fish (all other food-producing species).300 µg/kg
Oxolinic acidOxolinic acidFin fish (all other food-producing species).100 µg/kg
OxytetracyclineSum of parent drug and its 4-epimerFin fish (all other food-producing species).100 µg/kg
ParomomycinParomomycinFin fish (all other food-producing species).500 µg/kg
SarafloxacinSarafloxacinSalmonidae30 µg/kg
SpectinomycinSpectinomycinFin fish (all other food-producing species).300 µg/kg
Sulfonamides (all substances belonging to the Sulfonamides group)Parent groupFin fish (all other food-producing species).100 µg/kg
TetracyclineSum of parent drug and its 4-epimerFin fish (all other food-producing species).100 µg/kg
ThiamphenicolThiamphenicolFin fish (all other food-producing species).50 µg/kg
TilmicosinTilmicosinFin fish (all other food-producing species).50 µg/kg
TrimethoprimTrimethoprimFin fish (all other food-producing species).50 µg/kg
TylosinTylosinFin fish (all other food-producing species).100 µg/kg
*MRL: Maximum residue limit. It represents the length of time necessary to assure the absence or below-defined values of drug molecules in animals’ tissues. Target tissue: muscle (related to “muscle and skin”, as reported by art.14(7) EU Reg. No. 470/2009). Therapeutic classification: Anti-infectious agents/antibiotics.
Table 2. FDA-approved aquaculture drugs.
Table 2. FDA-approved aquaculture drugs.
Antimicrobials/Chemical MoleculesUseDoseWithdrawal Time and Other Limitations (Useful for MRL*)
Chloramine-TFor the control of mortality in:
Freshwater-reared salmonids infected by Flavobacterium spp.
Walleye due to Flavobacterium columnare.
12–20 mg/L (administered as a static bath every day for three treatments).0 day
Formalin (37%)The use of formalin is possible to be expanded as a parasiticide for all finfish and penaeid shrimp and as a fungicide to the eggs of all finfishAdministered in tanks and raceways for up 1 h (µL/L):
Salmon and trout → up to 170 µL/L with a temperature above 10 °C/50 °F, or → up to 250 µL/L with a temperature below 10 °C/50 °F.
All other finfish → up to 250 µL/L.
0 day
Hydrogen peroxide (35%)For the control of mortality in finfish’s eggs and other losses caused by Flavobacterium branchiophilum and F. columnare.Freshwater-reared finfish eggs:
500 to 1000 mg/L for 15 min in a continuous flow system (consecutive or alternate days) until hatch.
Freshwater-reared salmonids:
100 mg/L for 30 min or 50–100 mg/L for 60 min once per day on alternate days for three treatments in a continuous flow.
0 day
Oxytetraycline hydrochlorideFor the marking of skeletal tissues in finfish fry and fingerlings.200–700 mg/2 L of water for 2 to 6 h.0 day
FlorfenicolFor control mortality caused by Edwardsiela ictaluri (enteric septicemia) and Flavobacterium columnare.10 mg/kg of body weight for 10 consecutive days.12 days (under veterinarian prescription)
Oxytetracycline dehydrateControl Aeromonas liquifaciens and Pseusomonas spp. disease (they cause hemorrhagic septicemia), especially in Oncorhynchus spp. and Salmo spp.10 mg/kg of body weight for 10 consecutive days.21 days to catfish and 30 days to lobster
Sulfadimethoxine/ormetoprimControl of E. ictulari50 mg/kg of body weight for 5 days.3 days.
*MRL: Maximum residue limit. It represents the length of time necessary to assure the absence or below-defined values of drug molecules in animals’ tissues.
Table 4. AMR* and MDR* of bacterial strains isolated from aquaculture finfish sample tissues.
Table 4. AMR* and MDR* of bacterial strains isolated from aquaculture finfish sample tissues.
CountryFinfish Samples
n.
Isolated Bacterial StrainsPhenotypic AMR*/MDR*References
Braziln. 101
Oreochromis niloticus
Salmonella spp.
(46 isolates)
Amoxicillin/Clavulanic acid (87.7%)
Tetracycline (82.5%)
Sulfonamide (57.9%)
Chloramphenicol (26.3%)
56:1% of Salmonella spp. isolates were MDR:
Beta-lactam (blaCTX gene 66.7%)
Tetracycline (tetA gene 54.4%)
Chloramphenicol (floR gene 50.9%)
Sulfonamide (sul2 gene 49.1%)
[74]
n. 50
Cyprinus carpio
n. 50
Oreochromis niloticus
Enterococcus faecalis
(79 isolates)
Tetracycline (57.7% tetL and tetM)
Erythromycin (31.01% msrC)
[75]
Chinan. 50 fish samples:
Aristichthys nobilis
Carassius auratus
Ctenopharyngodon idellus
Parabramis pekinensis
Vibrio cholerae
(370 isolates)
MDR:
Streptomycin (62.2%) 230
Ampicillin (60.3%) 223
Rifampicin (53.8%) 199
[76]
n. 17
Acipenser spp.
Streptococcus iniae
(18 isolates)
Tetracycline (35.6% tetA-02)
Beta-lactams (25.3% blaTEM)
Aminoglycosides (22.1% aadA1)
[77]
n. 75
Carassius auratus
Aeromonas hydrophila
(n. 28 isolates)
MDR:
Penicillin (100%)
Ampicillin (100%)
Amoxicillin (96.4%)
Piperacillin (92.9%)
Cefalexin (78.6%)
Doxitard (75%)
Teicoplanin (67.9%)
[78]
Indian. 25
Oreochromis niloticus
Pseudomonas entomophila
Aeromonas hydrophila
MDR:
Bacitracin (100%)
Ampicillin (70%)
Cephalothin (60%)
Cafazolin (50%)
All resistant to:
Amoxicillin
Ampicillin
[79]
n. 97
Mugil cephalus
Listeria monocytogenes
(n. 21 isolates)
69% of Listeria isolates were MDR to:
Ampicillin
Penicillin
Erythromycin
Tetracycline
Clindamycin
[80]
Armenian. 25
Oncorhyncus mykiss
Pseudomonas spp.:
P. anguilliseptica
P. fluorescens
P. stutzeri
P.putida
P. aeruginosa
P. algaligenes
Resistance percentages:
Piperacillin (45.6%)
Pefloxacin (33.3%)
Ciprofloxacin (3.2%)
All susceptible to:
Chloramphenicol
[72]
Italyn. 300 fish samples:
n. 100 Dicentrarchus labrax
n. 100 Umbrina cirrose
n. 100 Sparus aurata
Vibrio spp.
Aeromonas spp.
Shewanella spp.
Photobacterium spp.
Resistance percentages:
Tetracycline (11.54%) (147/1274)
Trimethoprim/Sulfadiazine (7%) (89/1274)
[81]
Vietnamn. 50
Ictalurus spp.
Pseudomonas spp.
(n. 116 isolates)
Ampicillin (99.1%)
Sulfamethoxazole (93.1%)
Chloramphenicol (88.8%)
Nitrofurantoin (90.5%)
Nalidixic acid (90.5%)
Norfloxacin (9.5%)
Ciprofloxacin (8.6%)
Tetracycline (30.2%)
Doxycycline (25%)
[82]
Aeromonas spp.
(n. 92 isolates)
Ampicillin (93.5%)
Sulfamethoxazole (60.9%
Chloramphenicol (31.5%)
Nitrofurantoin (25%)
Nalidixic acid (52.2%)
Ciprofloxacin (7.6%)
Norfloxacin (4.4%)
Vietnam
Scotland
Denmark
Norway
France
Bangladesh
Thailand
Indonesia
Ecuador
n. 44 fish samples:
n. 12 Pangasiodon hypophthalmus
11 Salmo salar
10 Crassostrea gigas
11 Penaeus mongodon
Escherichia coli (n. 60)
Enterococcus spp. (n.69)
Pseudomonas spp. (n. 26)
Staphylococcus aureus (n. 9)
(246 isolates)
MDR strains:
n. 7 E. coli
resistant to:
Chloramphenicol
Ciprofloxacin
Ampicillin
Nalidixic acid
Sulfamethoxazole
Trimethoprim
n. 3 Enterococcus faecalis
resistant to:
Chloramphenicol
Gentamicine
Tetracycline
n. 4 Staphylococcus aureus
resistant to:
Chloramphenicol
Kanamycin
Tetracycline
[83]
Côte d’Ivoiren. 480
Oreochromis niloticus
n. 1696 strains:
Escherichia coli (15.9%)
Pseudomonas aeruginosa (10.4%)
Bacillus cereus (14.9%)
Enterococcus faecalis (14.2%)
Citrobacter freundii (13.5%)
Resistance percentages:
Amoxicillin/Clavulanic Acid (5.8%)
Piperacillin and Penicillin (8.7%)
Gentamycin (7.2%)
[84]
Irann. 240
Trota iridea
n. 86
Listeria spp. isolates
Tetracycline (62.79%)
Enrofloxacin (56.97%)
Ciprofloxacin (38.37%)
Penicillin (36.04%)
Ampicillin (34.88%)
[85]
*AMR: Antimicrobial resistance. *MDR: Multidrug resistance.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ferri, G.; Lauteri, C.; Vergara, A. Antibiotic Resistance in the Finfish Aquaculture Industry: A Review. Antibiotics 2022, 11, 1574. https://doi.org/10.3390/antibiotics11111574

AMA Style

Ferri G, Lauteri C, Vergara A. Antibiotic Resistance in the Finfish Aquaculture Industry: A Review. Antibiotics. 2022; 11(11):1574. https://doi.org/10.3390/antibiotics11111574

Chicago/Turabian Style

Ferri, Gianluigi, Carlotta Lauteri, and Alberto Vergara. 2022. "Antibiotic Resistance in the Finfish Aquaculture Industry: A Review" Antibiotics 11, no. 11: 1574. https://doi.org/10.3390/antibiotics11111574

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