Next Article in Journal
Posidonia-Based Compost and Dredged Sediment in Growing Media Improve Tolerance and Nutrient Uptake in Ornamental Plants
Previous Article in Journal
Subjective Norms or Psychological Empowerment? Moderation Effect of Power Distance on Knowledge Sharing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Potential and Green Chemistry Attributes of Biopesticides for Sustainable Agriculture

1
World Bank Africa Centre of Excellence, Centre for Oilfield Chemical Research, University of Port Harcourt, Port Harcourt 500272, Nigeria
2
Institute for the Development of Energy for Africa Sustainability, University of South Africa, Johannesburg 1709, South Africa
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(21), 14417; https://doi.org/10.3390/su142114417
Submission received: 6 September 2022 / Revised: 21 October 2022 / Accepted: 25 October 2022 / Published: 3 November 2022

Abstract

:
Chemotherapy has advanced modern agriculture with costly side effects such as the extinction of beneficial species, resistant pest resurgence, environmental pollution, tainted food consumption, and health implications. Attention is now focused on biopesticides as a solution to the abovementioned disadvantages. Additionally, there is a growing need to understand the range and relative effectiveness of biopesticides in controlling pests and promoting sustainable agriculture. The latter is the major driver of the Sustainable Development Goals (SDGs). In comparison to synthetic pesticides, biopesticides offer nearly similar protection against the most notorious pests, except Albugo candida (oomycetes), Ustilago maydis (fungi), Phytomonas spp. (protozoa), Nacobbus aberrans (nematode), and Cyperus rotundus (weed). This study shows that viruses are more vulnerable to essential oils, nematodes and weeds to natural enemies, herbivorous insects to biochemical insecticides, and plant pathogens to plant-incorporated protectants and microbial pesticides. This work also demonstrates that it is preferable to use plant-derived biopesticides in a field concurrently. Incorporating these findings into large-scale farming via the integrated pest management method would improve the outcome of sustainable agriculture (SA), which connects 11 of the 17 SDGs. Despite their proven efficacy and sustainable attributes, biopesticides have some deficiencies, such as slow action and a short shelf life span, which can be improved by omics, RNA interference, and nano-based technologies. This field of technologies provides relevant prospects for improving existing biopesticides and discovering and developing new bio-controlling agents (BCA).

1. Introduction

Biopesticides are green pesticides that adhere to the 12 Green Chemistry Principles (GCPs), as opposed to synthetic pesticides (SPs). The GCPs propose that (1) safe production processes begin with renewable resources and are powered by minimal energy; (2) waste is prevented or reduced, and (3) end products are harmless [1]. Biopesticides adhere to the Green Chemistry criteria by protecting plants, being innocuous to the environment, being safe to non-target organisms, and being effective in small amounts. The minimal energy condition of the GPCs is met through the use of transaminases, reductases, oxidases, and other biocatalysts in a production process [2]. Thus, biopesticides bridge the gap between green technology and sustainable agriculture (SA), a concept understood as large-scale farming that causes less environmental damage while being socially and economically viable. Crop productivity is lowered by pests; hence, they represent the main adversaries of (sustainable) agriculture. These pests include small vertebrates, nematodes, parasitic plants, microorganisms, pathogens, and insects; and act as biotrophs, necrotrophs, and hemibiotrophs [3]. Their impacts result in a decreased nutritional value, interference with biomass development, the generation of toxic metabolites, decreased market value, unfitness of food for ingestion, decreased yield, destruction of farmlands, widespread epidemics, mass displacement of humans, and economic loss [4,5,6,7,8,9]. The reversal of these pest-imposed constraints and the quest for food sufficiency and economic growth led to the intensive use of synthetic pesticides (SPs) during the Green Revolution in the 20th century. Notable examples of effective SPs used in this industrial era are dichlorodiphenyltrichloroethane (DDT), carbofuran, aldicarb, fenobucarb (carbamates), atrazine, simazine (triazines), and deltamethrin (pyrethroids), all of which work primarily by neurotoxicity [10]. Despite the proven efficiency of these SPs, they have drawbacks that are antithetical to sustainability. This calls for the use of biopesticides as alternative pesticides.
Biopesticides are now recognised as the future of agriculture. This is because of the many advantages they have over SPs, including but not limited to productivity, health promotion, and environmental promotion. These factors have led to increased efforts in the development of various biopesticides. The number of commercially available biopesticides is estimated to be 1400 worldwide, with similar efficacy to SPs, particularly in the management of insect pests [11]. Despite their advantages, it has been shown that biopesticides are not as effective as their conventional counterparts [12]. The efficiency of biopesticides relies on their multiple mechanisms of action (MoA).
These multiple MoA restrict the development of resistant pests, which are very common in the overuse of SPs. The strategic relevance of biopesticides has made them essential components of pest management (PM) systems. The Integrated PM (IPM) system promotes the adoption of joint solutions that cause the least disruption to agroecosystems and promote the safety of the environment and its receptors, including humans [13]. It is important to mention that biopesticides cover more than 75% of the IPM pyramid and offer the following benefits: they are innocuous to biofertilizers and pollinators; they promote agro-productivity, producing healthy food; and tend to ecosystem conservation, which are all integral parts of SA [14]. Biopesticide-linked SA is connected to 11 of the 17 Sustainable Development Goals, including SDG 1 (closing the global poverty gap), SDG 2 (ending hunger), SDG 3 (improving nutrition), SDG 6 (water supply), SDG 8 (economic engagement), SDG 12 (eliminating waste of produce), SDG 14 (aquaculture farming), and SDG 15 (land farming) [15]. Though biopesticides play significant roles in SDGs, their full potential has yet to be reached. This frontier can be exploited via nanotechnology, omics technology, gene editing and RNA interference (RNAi) technology.
The implication of biopesticides in SDGs and SA is due to their many benefits and advantages. However, it is important to review the varying level of efficacies and scope of the different categories of existing biopesticides, and the best methods for using them to achieve better outcomes. To achieve this goal, we explore current papers on biopesticides to understand their range of scope and effectiveness over 60 plant pests (excluding vertebrate pests), extracted from the plant pathology database. In addition, this study touched on areas in which the limitations of biopesticides can be improved. This review does not attempt to delve into biopesticide control of a particular group of plant pests. The outcome of this work shows that biopesticides offer a broad range of scope over plant pests with microbial pesticides ranking first and the most effective among the different categories of biopesticides. By the same measure, essential oils ranked second, followed by botanicals, then by biochemical pesticides, PIPs and finally natural enemies. However, the effectiveness of biopesticides against Albugo candida (oomycetes), Ustilago maydis (fungi), Phytomonas spp. (protozoa), Nacobbus aberrans (nematode), and Cyperus rotundus (weed) are poorly understood. This study also demonstrates that some biopesticide categories are preferred over others in the removal of pests. Natural enemies, such as biopesticides, can be paired preferably against nematode pests and weeds. Similarly, biochemical pesticides can be paired against insect pests, PIPs against specific phytopathogenic microbes, essential oils against viruses, and microbial pesticides against a wide spectrum of pathogens. These biopesticide categories do not, however, have a monopoly on effectiveness against their partnered pest group(s). When combining biopesticides to improve efficacy, it is best to apply essential oils with other plant-derived pesticides simultaneously.

2. Biopesticides’ Definition and Suitability as Green Chemistry Agents

Biopesticides are organisms or natural formulations that control or eliminate pests via diverse modes of action (MoA). They cover a wide range of products and formulations, such as predatory and parasitic species, biochemical compounds (and their chemical equivalents), and plant-incorporated protectants (PIPs) [16]. Organic extracts from shells, crustaceans, and algae function as signal molecules to trigger a defensive response in plants and animals, producing long-lasting effects against biotic infections. These substances are known as semiochemicals [17,18]. So, biopesticides include semiochemicals and biochemical analogues. However, if an organic product exerts its toxicity on pest neurological systems, it becomes a poison [19]. Based on this criterion, rotenone, nicotine, sabadilla, and pyrethrins are regarded as poisonous substances even though they are effective botanicals [20]. As green compounds, biopesticides are of great interest to the 12 Green Chemistry Principles (GCPs).
The overarching 12 GCPs were articulated by Anastas, P. and Warner, J. C. in 1998. These principles are the universal gold standard for developing environmentally friendly processes and products (Figure 1). Green Chemistry advocates methods that lessen the use of dangerous substances and minimise the development of toxic intermediates during chemical transformations. The 12 GCPs require that products are made from renewable feedstock (the focus of Principle 7) under the following conditions:
  • Synthetic processes are safe, green, and use minimal energy;
  • The final products are environmentally friendly;
  • Wastes are prevented or minimized.
According to Fenibo et al. [21], the first condition is connected to the 3rd, 5th, 6th, 8th, 9th, and 12th GCPs. The second condition is addressed by Principles 4 and 10, while Principles 1, 2 and 11 ensure waste prevention and minimization (condition three). The surest means of securing condition three is to ensure that the starting materials are integrated into the finished product [22]. As a green product, biopesticides are effective at low concentrations, biodegradable, and harmless to non-target biota. The use of reductases, transaminases, oxidases, hydrolases and other biocatalysts not only meets the minimum energy requirement but also benefits the environment and the economy. In this sense, biopesticides align with green technology and SA. The latter is characterised by the responsible consumption of resources, biodegradability, and productivity [23]. One important outcome of resource consumption in agriculture is the eradication of pests, as shown in Table 1. The conscious attempt to maintain these sustainable attributes implies integrating green chemistry principles into agricultural practices that previously relied on synthetic pesticides (SPs) for pest eradication.

3. Effectiveness of Synthetic Pesticides and Their Disadvantages

For 60 years, farmers have consistently applied SPs in agriculture based on economic considerations. The application of pesticides started with sulphur, dating back to earlier than 2000 BC, to control rodents, insects, and fungi. In the 15th century, heavy metals such as lead, arsenic, chromium, cobalt, and nickel were also used. In 1940, Paul Muller received the first chemical pesticide patent for dichlorodiphenyltrichloroethane (DDT) [47]. Pesticides are compounds or materials that can inhibit, prevent, kill, or reduce pests of farmland and livestock. They are categorised as microbicides, herbicides, insecticides, nematocides, and rodenticides based on the domain of pests they eliminate or control. Synthetic pesticides are essential components of commercial agriculture that control farm pests, which diminish production, cause diseases, and compromise food quality. DDT is an effective organochlorine insecticide that eliminates a wide range of pests. DDT harms the insects’ nervous systems when they come into contact with it. The active compound binds to the voltage-gated sodium channel proteins in neurons, keeping them constantly open, leading to the excessive diffusion of sodium ions (Na+) through the phospholipid membrane, causing hyperpolarization and, eventually, spasms that can be fatal [48,49].
Organophosphates (phosphoric acid esters) and carbamates (N-methyl carbamic acid esters) are also widely used to target insects. Examples of organophosphates are malathion and parathion, while carbaryl and propoxur are examples of carbamates. They function by inhibiting acetylcholinesterase, thereby preserving the continuous increase in neurotransmitter concentration, duration, and function, which causes neurotoxicity and death to (non-) target insects [49]. Other carbamates (cyclodienes, phynylureas, and thiolcarbamates) inhibit CO2 generation in soil respiration by acting as 2,4-dinitrophenol analogues that shunt oxidative phosphorylation in microbial aerobic cellular respiration. Triazoles are a class of pesticides that are sensitive to a wide range of pests, including mildew, leaf spot, and rusting fungal diseases. Triazoles’ efficacy is attributed to their capacity to block the cytochrome P450-dependent sterol 14-demethylase enzyme, a vital component of the sterol biosynthesis pathways, leading to the creation of ergosterols thought to be essential in the regulation of membrane fluidity and structure. The absence of ergosterols causes the reduction of Fe3+ to Fe2+, which in turn disrupts sterols needed for cell wall synthesis [50], leading to the ultimate death of insects. Two independent investigations show triazole to be very effective in suppressing Botrytis cinerea at a 50% concentration (EC50). Itraconazole [51] and tebuconazole [52] effectively inhibited the fungus’s radial growth on PDA in these studies.
Although SPs have been effective in reducing crop pestilence and improving productivity, they cause pollution with characteristic long-term persistency in the ecosystem and on ecological receptors [53]. A study found that aldrin, dieldrin, and endosulfan were the most persistent pesticides in soil [54,55]. Pentachlorophenol inhibits microbial activities, damages soil fertility, distorts nutrient cycling, causes mutation, and has neurotoxic effects on aquatic organisms [56,57]. Cancer, respiratory conditions, Parkinson’s disease, and death are some of the consequences of being exposed to SPs [58]. The classification of these pesticides and their pest-controlling impacts are reflected in Table 2. It is pertinent to note that the impact of exposure is proportional to the increase in the concentration of the SPs. This was demonstrated in a study carried out in Southern India (Vellore) involving 100,000 people, which revealed that observable health impact effects were evidenced when farmers sprayed pesticides at mean peak flow rates of 346 L/min and 390 L/min [59]. Overall, traditional pesticides place the environment, diversity, plants, animals, and humans at risk (Figure 2). Thus, the call for the minimal use of SPs and the diversion to safer and more sustainable alternatives. Implementing large-scale farming in the context of alternative pesticides would help to define SA and its attendant benefits, including profitability, environmental protection, and social acceptability. The surest way to achieve this is by relying on biopesticides, which have the potential to fill the gaps associated with SPs.

4. Biopesticides as a Substitute for Conventional Pesticides

The performance of any pesticide is dependent on its effective pest-controlling and eradicating potential. For this section, SPs effectiveness is taken as a standard. With this in mind, we can compare biopesticides with SPs from many angles. More than 55,000 plant pests have been subjected to SPs, with fungi accounting for up to 50,000 [79]. Rajput et al. [80] suggested that the efficacy of biopesticides is comparable to traditional pesticides in terms of insect management (Figure 3). Biopesticides have also performed well in controlling weeds and pathogens [81]. Tripathi et al. [82] revealed that the available biopesticides on the market amount to approximately 1400 (see Table 3), with commensurate pest management success.
The efficacy of biopesticides depends on their pest elimination and control potential. This in turn depends on the multiple MoA and synergy. The performance of biopesticides over the top 60 plant pests (from the plant pathology database) showed that they possess the ability to eliminate a wide range of harmful biotic factors, excluding vertebrate pests. Microbial pesticides ranked first among the various biopesticide groups. Essential oils came second, followed by botanicals, biochemical biopesticides, PIPs, and natural enemies. However, not many studies have been made public to prove the efficacy of biopesticides over Cyperus rotundus (weed), Nacobbus aberrans (nematode), Albugo candida (oomycetes), Ustilago maydis (fungi), and Phytomonas spp (protozoa). Hence the need for research in this direction.
Though biopesticides control a good number of pests, there are instances where certain biopesticide groups are preferred over others. As illustrated in Figure 4, natural enemies eliminate weeds and nematode pests better than other classes of biopesticides [83]. Hence, they will be more useful as bioherbicides and bionematicides. As mentioned in the literature, herbivorous insects are more sensitive to biochemical pesticides and their chemical analogues. Plant viruses are better controlled by essential oils, while phytopathogens are more vulnerable to PIPs and microbial pesticides [84,85,86]. However, microbial pesticides cover a broader scope. When a combination of biopesticides becomes necessary, it is better to use plant-based pesticides simultaneously for a synergistic result. Combining a plant-based pesticide and any of the microbial pesticides, entomopathogenic nematodes, or natural enemies will not yield a synergistic result because of the multiple MoA inherent in phytochemicals. However, if they must be used together, the application of plant-based pesticides should come first.
Biopesticides display multiple mechanisms of action, as shown in Table 3. Photobleaching, growth inhibition, protein content reduction, dry mass reduction, phytotoxicity, allopathy, and necrosis are used against weeds. Bioinsecticides act through parasitism, gut solubilisation, resistance induction, gene disruption, disruption of a life cycle, ovicidal activity, oviposition deterrence, antifeedants, growth mal-regulation, repellence, electron transport inhibition, and neurotoxicity in controlling insect pests [87,88]. Antibiosis, hyperparasitism, antagonism, competitive exclusion, biostimulation, cellular wall disruption, hypersensitivity response, resistance gene activation, and lipid peroxidation are some of the mechanisms employed by biofungicites and bio-bactericides [89,90]. Mechanisms by which bionematicides protect plants against pests are egg parasitism, root colonization, physicochemical barricade, chemotoxicity, and resistance induction [91]. The control of viruses is achieved through particle inactivation, cytotoxicity, protein masking, and envelope structure modification [92,93]. Natural enemies adopt parasitism and predation. Plant-incorporated pesticides confer protection via gene transformation, suppression of pest resistance, target gene silencing, biological barricade, larval mortality, and viral replication obstruction [94]. The multiple MoA exhibited by biopesticides make it difficult for pests to develop resistance, which is common with SPs [95]. According to Dammos et al. [96], there are more than 500 resistant pests globally connected with conventional pesticide overuse. These problematic resistant pests are also susceptible to biopesticides.
Steps for the synthesis of this diagram. The first ten pests of each category (excluding weeds) are taken from the molecular plant pathology database. Then, the 60 plant pests are crosschecked for biopesticides sensitivity using open-access publications.
However, the routine and prolonged use of a particular biopesticide can lead to pest resistance [97]; however, resistance will not be as intense as SPs. To avoid such a phenomenon, a change of management approaches should be encouraged since there is more than one biological option for controlling a particular pest or pests.
Table 3. Selected examples of commercially available biopesticides.
Table 3. Selected examples of commercially available biopesticides.
ProductManufacturersActive AgentMode of ActionControlled PestProtected CropReference
Biochemical
Aza-DirectGowan (USA)AzadiractinGrowth and moulting disruptionEgg, larvae and pupae of beetles and sucking insectsCotton, Papua, vegetables[98]
Timorex GoldSTK Stockton (Israel)Tea tree oil (1,8-cineole and terpinen-4-ol)Antifeedant, disruption of the fungal cell wall and curative activitiesBlack Sigatoka, a leaf-spot fungal diseaseBananas, strawberries, tomatoes, grapes, lettuce.[99]
Regalia MAXXMarrone Bio Innovations (USA)Extract of Reynoutria sachalinensisBroad-spectrum antimicrobial activity, resistance against diseaseFungi and bacteriaHemp, cannabis, tomatoes, apples, blueberries[19]
Nema-QBrandt Consolidated (USA)SaponinsNematicidal effects, Root-knot nematodesBerries, Citrus, Pome Fruit Grapes, Nut Crops[100]
Microbials
LipelAgri-Life (India)Bacillus thuringiensisLethal action against eggs and larvae of diamondback mothDiamondback MothCruciferous vegetables: Collard greens, cauliflower[101]
CordaleneAgrichem (Australia)Bacillus thuringiensisKilling of midgut cells by Cry toxins through signal transductionLepidoptera insect pestsMaize, sugarcane, soybeans, peanuts, flax[102]
DamanInternational Panaacea Ltd. (India)Beauveria bassianaGrowth inhibition and larvicidal activitiesLarval, pupal and nymphal stages of SpodopteraRice, maize, sorghum[103]
MeloCon WGCertis (USA)Paecilomyces lilacinusColonization of plant roots, egg mass of nematodes and incapacitating second-stage instarsNematodesVegetables, citrus strawberries, grapevines, tomato[104]
Grasshopper SporeARBRICO OrganicsNosema locustaeInfects insects at the moulting stageGrasshopperVegetables, fruits, [105]
LittovirAndermatt Biocontrol AG (Switzerland)SpliNPV *Induction of adult, pupa and larval malformationAfrican cotton leaf worm (Spodoptera littoralis)Okro, onion, groundnut, beetroot, cabbage
PIPs
Bt-cottonChinese Academy of Agricultural Sciences.Cry1Ac, Cry2Ab toxinsAct as gut poison leading to pore formationDiptera, beetles, H. armigeraCotton[106]
5345Monsanto (USA)Cr y1Ac geneKills insets by pore insertion into gut membranesLepidopteran pests: Fruitworm, pinworm, hornworm
Tomatoes[107]
At+-PotatoResearch-limitedRNAiCauses post-transcriptional
silencing target genes responsible for infection and maintenance
Phytophthora infestansPotatoes[108]
Natural Enemies
Mealybug Destroyer; Convergent Lady Beetles; Whitefly Predator; Spider Mite Destroyer; Scale Predator; Fungus Gnat PredatorGreat Lakes IPM, Inc. (Vestaburg, MI, USA)Cryptolaemus montrouzieri; Lady beetle; Delphastus; Stethorus; Cybocephalus nipponicus; Rove beetle (Atheta coriaria)PredationA broad variety of insects fall victim including slugs, aphids, mealybugs, thrips, whiteflies, scales, spider mites, leafhoppers, fungal growthsCitrus, corn, ornamentals, vegetables, sweet potato[109,110]
Chinese Mantid; Green Lacewings; Aphid Predator Midge; Predatory MitesCrop King, (Lodi, OH, USA)Tenodera aridofolia; Chrysoperia spp. larva; Aphidoletes aphidimyza larva; PhytoseiidaePredationA broad range of insects including aphids,Pea, cabbage, cowpea, cucurbits, crucifers, eggplants, okra, lettuce[111,112]
* Spodoptera littoralis nucleopolyhedrovirus; Agrobacterium tumefaciens.
The most suitable, effective, and flexible method that can address the drawbacks of conventional farming is the integrated pest management (IPM) system. IPM is a management approach that combines and implements dimensions and guidelines in sequence to lessen the impact of pests and produce healthy crops with minimal disruption to the agroecosystem. The first two IPM system processes are pest characterization and pest load quantification. The latter determines the economic thresholds at which pests begin to affect the micro-economy (farm) and macro-economy (country’s GDP). Pest characterization can be accomplished through a molecular approach, which provides more reliable details than morphological identification techniques [113]. Currently, drone technology [114], image technology [115], and an artificial intelligence system combined with a pest database [116] are used to identify arthropods. The pest destruction guidelines are carried out in the following order: cultural practice, monitoring, decision-making, physical treatment, bioaugmentation, application of least toxic pesticides, and the use of SPs [117]. In reality, pest elimination tactics are pyramidal (Figure 5) and use biopesticides, except in the decision-making component. For example, kairomone- or pheromone-baited traps can serve the purpose of monitoring resistant cultivars and natural enemies for cultural techniques, while garlic soap and diatomaceous earth (biochemical biopesticide examples) application can assist in the physical approach [118]. Recently, climate-smart pest management (CSPM) has been used in IPM management to identify the ideal seasons for the cultivation of particular crops and pest prevention [48]. Additionally, laser-based technology (a physical technique) has recently been used to eradicate aphid species such as Acyrthosiphon pisum and Rhopalosiphum padi [119].
Augmentation (which is the same as biointensive pest management), the third stage of the pest eradication guidelines, is accomplished by using entomopathogenic nematodes and microorganisms. A soil moisture between 20 and 60% encourages entomopathogenic microbes to function effectively [120]. The next stage of the pest management guidelines, the employment of least hazardous pesticides, requires the application of biochemical biopesticides, botanicals, and essential oils. Some notable botanicals have neurotoxic properties (e.g., pyrethrins, rotenone, sabadilla, azadirachtin) but can still be used because they are less harmful than SPs [121], which come last in the IPM strategy. Recently, Umetsu and Shirai [122] published novel synthetic chemicals that are safe for the environment and humans. They include fungicides (flupyrimin, flupyradifurone) that inhibit succinate dehydrogenase, demethylation, flupyrimin, flupyradifurone, and quinone. The production of low-toxicity pesticides in commercial quantities can be achieved through the manufacture of biochemical pesticide analogues. The eradication of SPs is a possibility on this path, but it depends on financing, research, and time. As we wait for this auspicious moment, SPs remains the last resort in the IPM strategy. A critical examination reveals that biopesticides dominate the IPM pyramid. This demonstrates the importance of biopesticides to agriculture now and in the future. Other components that are key to IPM implementation are knowledge of pest ecology and biology, resources, organization and planning, and communication.

5. Sustainability Attributes of Biopesticides

There are key parameters that promote agriculture and its related subjects. Constructive manipulations of these parameters in large-scale agriculture translate into sustainable agriculture (SA). The latter is understood as the prudent and continuous use of land and other factors of production in meeting the food needs of the populace and other dimensions such as economic productivity, environmental protection and socio-cultural acceptability. Some of these parameters serve as renewable feedstock and products used in agriculture, including biopesticides. Beyond these roles, biopesticides drive economic productivity and are harmless to non-target organisms, including plant growth-promoting microorganisms and pollinators [123]. These beneficial organisms interact closely with plants for mutual benefits. Plants benefit from the essential nutrients supplied by the microorganisms, which in turn gain energy and protection. In some cases, these rhizomicrobes (rhizobacteria) antagonise soil pests using bioactive substances they produce and metabolise carbon-rich pollutants as an energy source [124]. The majority of these growth-promoting rhizobacteria (PGPR) perform nitrogen fixation, mineral solubilization, generation of active metabolites, and induction of systemic resistance for the advantage of the concerned plants [125]. The assimilation of H2 during nitrogen fixation improves the growth of the associated plant [126]. Examples of PGPR (biofertilizers) are Azospirillum, Allorhizobium, Bacillus, Bradyrhizobium, Arthrobacter, Agrobacterium, Mesorrhizobium, Caulobacter, Rhizobium, Azotobacter, Frankia, Erwinia, Micrococcus, Serratia, and Flavobacterium [127]. Inoculation of PGPR to plants improves yield, and as a practice is becoming popular. The use of biofertilizers (PGPR) and biopesticides is essential to organic farming, conserves biodiversity and increases crop yield.
An increase in crop yield is partly attributed to insect pollination. Pollinators contribute to more than 75% of food crops, and 90% of wild angiosperms found in agro-fields benefit from most insect–plant interactions [128]. In addition, the interaction between pollinators and plants results in genetic variability, yield stability, early reproduction, and strong traits [129]. Adamidis et al. [130] established that honeybee-driven pollination contributed to a yield increase in sesame and cotton by 62%. Other significant pollinators are solitary bees, bumble bees, moths, and butterflies. Studies have shown that pollination services raise crop value by USD 17,174 per hectare, boost worldwide food production by 9.5%, with a net worth of USD 405 billion, and account for 67% of Tanzania’s food crops [131,132]. Apart from food, pollinators also contribute to the production of timber, biofuel, fibre, and medication [133]. Garratt et al. [134] noted that the gains from pollination services can only make sense where fertilisers are not limited. This implies that in the presence of sufficient plant nutrients (biofertilizers) and thinned pests, the many advantages of pollination can be fully realized.
From the preceding sections, it is clear that pests, nutrients, and pollinators are critical factors in agricultural practice. Water, the atmosphere, and humans are also critical factors in agriculture. The involvement of pesticides (biopesticides and SPs) improves agro-yield by mitigating pests. However, the other remaining factors respond differently to pesticides in general. Apart from their pest-controlling efficacy, SPs defile the air, soil, groundwater, and surface water. The polluted state of the environment normally leads to health challenges, the abandonment of natural resources, biodiversity losses, and high costs of remediation. Environmental chemical-instigated health challenges can lead to discomfort, fatigue, Parkinson’s disease, cancer, or even death [135]. Attina et al. [136] revealed that the negative impact of organophosphate pesticides costs the United States USD 44.7 billion/year. Earlier, Nowak and Greenfield [137] estimated damages caused by by SPs in the following sectors: bird losses (USD 2.2 billion/year), crop losses (USD 1.4 billion/year), public health (USD 1.1 billion/year), pesticide resistance (USD 1.5 billion/year), and groundwater pollution (USD 2.0 billion/year). The high degree of soil and surface water contamination leads to abandonment or suspension of usage until remediation is carried out. Chemical remediation can be accomplished through conventional and bioremediation approaches. Though the bioremediation approach is considered the best option, both techniques cost money, time, and effort. The attenuation of a particular pesticide to a safe target level usually requires a period of 12 to 24 months. All these costs and negative effects of using CPs in agriculture could be avoided by utilising biopesticides instead. The application of biopesticides provides a safer environment and improves productivity. These are areas of concern for the sustainable development goals (SDGs).
The SDGs address 17 subject areas, poverty (SDG1), hunger, health, education, gender, clean water, energy, economic growth, infrastructure, inequality, communities, production and consumption, climate, life underwater, life on land, strong institutions and partnership (SDG 17) [138]. However, the ones that are of interest to agriculture are life under water (SDG 14) and on land (SDG 15). Each of them serves as a base for agricultural activities. Land, especially fallowed ones, can be cultivated while water bodies can be used for aquaculture for subsistence and commercial farming. Sustainable agriculture (SA) is driven by the conscious utilisation of water (SDG 6), nutrients, and energy (SDG 7) [14]. Agriculture-based skills and education (SDG 4) and responsible production of agro-products and consumption (SDG 12) optimise the impact and services that SA provides. Examples of services agriculture provide are the reduction of poverty (SDG 1), drastic hunger reduction (SDG 2), nutrient-linked health improvement (SDG 3) and the reduction of global warming (SDG 13) [15]. Vegetation helps reduce greenhouse gases by absorbing CO2. For this reason, research on CO2 capture and sequestration has been given priority attention in recent times to address global warming [139]. Strong institutions (SDG 16), which promote policies, peace, and justice, serve as enablers of SA improvement and support practitioners on an equal gender basis (SDG 5) to engage in agriculture as a decent profession for economic growth (SDG 8). In summary, it is instructive to conclude that SA platforms (SDG 14 and 15) provide services such as hunger eradication (SDG 2), poverty reduction (SDG 1), health improvement (SDG 3), and air purification (SDG 13). These services can be optimised by ensuring education (SDG 4) and judicious production and consumption (SDG 12) habits. Strong institutions serve as essential elements for all aspects. Conclusively, biopesticide-driven large-scale agriculture is linked to 11 of the 17 SDGs and guarantees food security, productivity, and environmental sustainability.

6. Emerging Frontiers of Biopesticide

The existing classes of biopesticides have their disadvantages, which require improvement to improve their technical performance. Omics technology, nanotechnology, and RNA interference technology are avenues through which these improvements can be made [140]. PIPs are employed to deliver biopesticides during genetic modification, which causes plants to synthesise compounds to ward off pests. It was established in 2014 that PIPs could reduce synthetic pesticide consumption by 37% and increase agricultural output by 22% [141]. Current studies explore omics tools to ascertain potential proteins and metabolites that can be integrated into plants to eliminate bacteria, fungi, viruses, pathogens, and insect pests [142]. Metagenomics, a branch of omics technology, is used to discover novel enzymes and encode genes from microbial communities. For example, a chitinase encoded by the Chi18H8 gene was active against Rhizoctonia solani and Fusarium graminearum [143]. Additionally, the CfHNNI1 gene was active against Cladosporium fulvum, a pest of tomatoes and tobacco [144]. Hjort et al. [145] developed a framework for the identification of Chi18H8 genes from soil microbial communities. The protocol involved creating a library by performing PCR amplification and sequencing the Chi18H8 gene in gene pools. These results demonstrate that metagenomics can be utilised to locate pesticidal genes that can endow plants with transgenerational protection. Metagenomics research has also discovered biological control agents in soils.
Metabarcoding, a genomic technique, is described as a method for identifying and labelling organisms within microbial communities with peculiar characteristics. Genomes are amplified and sequenced using specific primers (16S, 27F, and ITS1 and ITS4), thereby aiding the description of host–pathogen interactions and the recognition of phylotypes whose DNA segments are identical to those of a known species [146]. Data collected through these means are instrumental in developing biopesticides and biofertilizers for managing pathogens. Cobo-Diaz et al. [147] used metabarcoding to recognise biocontrol microbes capable of mitigating Fusarium spp. disease in maize stalks. Pseudomonas, Pedobacter, Sphingomonas, Flavobacterium, and Janthinobacterium (bacteria genera), Epicoccum, Articulospora, Exophiala, and Sarocladium (fungi genera) were identified as biocontrol agents. Combined consortia-metabarcoding and co-occurrence network analyses were used to quantitatively ascertain the antagonistic relationship between the biocontrol consortia and Fusarium spp. The research also discovered an antagonistic interaction between F. avenaceum, F. graminearum, Sarocladium and Epicoccum. This report corroborates the results of Kemp et al. [148], who demonstrated how a symbiotic relationship between Sarocladium zeae and Fusarium spp. could discourage head blight disease in wheat. An earlier study carried out by Jensen et al. [149] established the antagonistic effect of the wheat endophyte, Epicoccum nigrum on Fusarium graminearum. Epicoccum layuense is an agent of biocontrol against Phaeomoniella chlamydospora and Fomitiporia mediterranea in grape vines [150].
The genome editing (GE) technique is currently used for modifying plant immunity in diverse ways by altering targeted genes to create more resistant crop species [151]. Three mutagenesis tools are used to edit genes that confer resistance in plants. CRISPR (clustered regularly interspaced short palindrome repeats) is one of these. The CRISPR-associated system (CRISPR-Cas) is easier to implement and has a high efficiency in generating mutagenesis [152]. CRISPR-Cas9 has been employed to raise wheat that is resistant to powdery mildew and white rot [153]. The other two are ZFNs (zinc finger nucleases) and TALENs (transcription activator-like effector nucleases) [154]. Chen et al. [155] used CRISPR-Cas9 to transform the cotton genome using an Agrobacterium-delivering medium via the shoot apexes. CRISPR-Cas9 gene editing confers immunity against blight, cotton geminiviruses, and viruses [153]. ZFN was used to modify the AaegGr3 gene encoding the heteromeric CO2 receptor subunit from Ae. aegypti mosquitoes [156], while TALEN was employed to produce mutant mosquitoes that are hyper-vulnerable to Plasmodium parasites [157]. This opens an opportunity for malaria prevention and elimination. Genomics has been useful in the development of transgenic plants and improvement of cultivars by way of introducing resistant genes into plants.
Experiments have proven RNA interference (RNAi) technology as a viable pest control method. It is a post-transcriptional regulatory process in eukaryotic cells in which specific RNAs inhibit gene translation and expression by counteracting targeted mRNA molecules. RNase participates in a series of activation complexes that culminate in the formation of the RNA-induced silencing complex (RISC), a multi-protein complex composed of a micro RNA (miRNA) or small interfering RNA (siRNA) and RNase enzymes [158]. With the right delivery mechanism, the biopesticides’ shelf life in the open environment can be extended [159]. Emerging delivery technologies of RNAi are liposomes, polymeric nanoparticles, viral-like particles, and peptide-based delivery techniques [160]. The application of RNAi for pathogen and pest control is achieved in three different ways. The first approach involves the silencing of host-induced genes (HIG). HIG silencing causes the development of transgenic crops that express double-stranded RNA (dsRNA) specific to pathogens or pests. An RNAi-defined transgenic plant is now available (Tradename: SmartStax) on the market. It targets the snf7 gene of corn rootworm by the elicitation of a hairpin-dsRNA [161]. The second approach is the silencing of spray-induced genes (SIG), which could have a pronounced effect on root adsorption and trunk injections. Sucking and chewing insects can accrue expressed double-stranded RNAs. The final method is the silencing of virus-induced genes (VIG), which requires engineered microbes as vectors for the production of siRNA or dsRNA. Colorado potato beetle, Asian corn borer, brown plant hoppers (insects), alfalfa mosaic virus, tobacco etch virus, pepper mild motile virus, sugarcane mosaic virus (plant viruses), and Fusarium graminearum, Botrytis cinerea, and Sclerotina sclerotiorum (fungi) are all controlled by double-stranded RNA [159]. RNAi is limited by its narrow application range because of its instability, sensitivity variance in insects, and poor risk assessment of the environment, plants, and humans.
In pest management, nanotechnology is an evolving field of study because new properties (high reactivity, uniform conductance, high surface area, etc.) arising from nano-sized particles are quite different from the attributes of the bulk/parent material. Jasrotia et al. [162] noted that nanotechnology in plant protection has the following advantages: long-lasting protection of active ingredients (AIs); enhanced physicochemical attributes (solubility, wettability, dispersion, storage stability, penetrability, bioavailability uniformity) of the AIs; controlled release of the AIs; high surface area; improved efficiency, etc. Inorganic compounds, metals, and organic polymers are common materials used for nanoformulations (ending in the form of suspension, spheres, aerosols, particulates, powder, capsule, micelle, gel, fibre, and granules). Often used terminologies with nanoformulations are nanoparticles, nanoemulsion, nanosuspension, nanocapsules, nanoencapsulation, nanogel, and nanofiber [163]. Nanoparticles have a pesticidal effect on plants. For example, copper, silver, platinum, sulphur, graphene oxide, titanium dioxide, silver oxide, copper oxide, and chitosan nanoparticles have been proven to control pests such as weevils, mosquitoes, and fungal and bacterial diseases through their biocidal (oxidative stress, enzyme inhibition, protein synthesis inhibition, ill-developmental changes leading to reproductive failure, gene mal-regulation, insect dehydration, and repellent effect) properties [164]. Nanoparticles are now obtained from plants, including Anacardium occidentale (gold, silver, and copper), Acorus calamus (gold and silver nanoparticles), and Aloe secundiflora (silver). Nanosizing parent pesticides could have the same improved results. The nanostructuring of biological pesticides represents an advantage of nanotechnology in terms of biopesticides [165]. Nanobipoesticide efficacy has been demonstrated against H. armegera, Anopheles, insect larvae, bacteria, fungi, and stored grain protection. Thus, nanoparticles and biopesticide nanoparticles serve as emerging areas of interest for practitioners.
Nanoformulations share the following properties: higher specificity, increased persistency, and reduced mass in comparison to free biopesticides, some of which have poor water solubility properties and a predilection for oxidation, volatility, and instability [21]. A nanoemulsion (droplets with a radius of less than 100 nm) is a diphasic system in which its active parts are spread in oil-in-water phases with the aid of surfactants. This suggests that nanoemulsions are oil-in-water dispersions that improve the efficiency of pesticides by dissolving them into very small oil droplets. In addition, nanoemulsions are suitable for both hydrophobic and hydrophilic pesticides. This property of nanoemulsion helps in eliminating the need for toxic organic solvents and synthetic surfactants [166], apart from aiding uniformity. Like synthetic surfactants, nanoemulsion enhances spreadability, wettability, and permeability. Essential oils are a notable subject of interest in nanoemulsion. They exert their pesticidal activities through repellent activity, toxicity attributes, development inhibition, behavioural alterations, sterility properties, and infertility effects [167]. Nanoemulsions of essential oils improve biopesticide delivery in agriculture, especially for phytophagous insects. Palermo et al. [168] formulated eight commercial nanoemulsions of essential oils (anise, fennel, artemisia, sage, garlic, lavender, mint, and rosemary) which showed improved repellency, aerosol/gel, and acute toxicity against Tribolium confusum (pest of stored grain). Behara [169] stated that nanoemulsions possess the following advantages: higher pesticidal activity, lower cytotoxicity, smaller dynamic contact angle on leaves, pesticide delivery, and physical stability in comparison to microemulsions and emulsions. Details about the methods of nanoemulsion production can be found in [170,171], but low-energy methods are much greener. Apart from the popular nanoemulsion technique, nanoencapsulation is another popular nanotechnology critical for plant protection advances.
Nanoencapsulation is implemented using polymers and inorganic materials that aid the delivery of the active parts in a controlled manner, either with slow (sustained) or stimulus-response release to abiotic and biotic factors [172]. Apart from protecting the active components from degradation, hydrolysis, volatilization, leaching, photolysis, and hydrolysis, encapsulation prolongs the release duration and affects precision control to yield effective results [173], enhance the mobility and distribution of nematicide against soil parasitic nematodes (plant-viral nanoparticles) [41], and provide good biocompatibility (silica and calcium carbonate) with biopesticides [174]. Thus, nanoencapsulation serves as an effective delivery platform for pesticide delivery and incorporation of particles (nano-pesticides) into plants to increase the effectiveness of biopesticides against pests and pestilence [175]. Regulating nanoencapsulation pore structure enhances the effectiveness of the active agents [176]. Noisomes and liposomes are efficient carriers for pesticide delivery [177]. Nano-gel has a reputation for serving as a carrier for pheromones, which is thus protected from decomposition and decreased evaporation. Nassar [178] demonstrated that chitosan nanogel improves the performance of copper-based antifungal remedies. Similarly, nanofibers improve the efficacy of essential and pheromone molecules by evading release outbursts [179]. For example, thiamethoxam-loaded fibre proved effective against whitefly with a 50% improvement rate [180]. In essence, nanoencapsulation improves pest control via the controlled release of AIs and delivers them to the target pest, thereby increasing specificity.

7. Constraints and Outlook

Despite the many advantages of using biopesticides, they do have their disadvantages, including slow action. This limitation serves as a hindrance to their widespread adoption in commercial farming. Their use is currently popular only among organic farmers who cater to the elite and high-end consumers. Furthermore, these biopesticides are easily influenced by the environment, as demonstrated by controlled versus uncontrolled experiments, with frequent significant variations in outcomes once environmental factors are accounted for in field studies. This research is needed in the areas of stability, delivery methods, shelf-life, and packaging materials that will preserve efficacy. Though most of the limitations are addressed by nanotechnologies, some areas could benefit from further improvement. Nanotechnology improves biopesticide stability, high precision in biopesticide delivery to target pests, controlled release of AIs, and concentration minimization of biopesticide, thereby reducing their possible toxic impact. Some identified areas where improvement is needed in nanotechnology as it relates to pest management are mass production, low selective toxicity, and the non-biodegradability of inorganic nanoparticles. As an evolving technology, questions exist regarding the environmental impact of nanotechnology and its effects on receptors. Thus, further research is needed to provide answers, including for the production of environmentally safer nano-pesticides. The safety of food tainted with nanoparticles is still unknown, and information about the exact precautions to be taken (where necessary) is still lacking. In addition, research centred on ecotoxicity studies of nanoparticles, cheap nanoformulations, and legislative frameworks is urgently needed. Such information would bring to the fore issues of green nanotechnology, which will be of great benefit to SA and is marked by food safety, economic productivity, eco-friendliness, social acceptability, and cost-effectiveness.

8. Conclusions

Biopesticides are emerging pesticides and are a better option than synthetic pesticides due to their unique sustainability and green chemistry attributes. Different types of biopesticides respond preferentially to different pest categories. For instance, phytophagous insects are more susceptible to biochemical pesticides and botanicals. As are plant pathogens to microbial pesticides and plant-incorporated protectants. Natural enemies preferentially eliminate weeds and nematodes. Thus, it is imperative to systematically characterise pests before applying biopesticides. The pest elimination guidelines of integrated pest management (IPM) pave the way for pest characterization and accommodate the use of bio-controlling agents. The implementation of IPM principles in large-scale farming gives rise to sustainable agriculture, which promotes continuity, economic productivity, social acceptability, and environmental sustainability. Sustainable agriculture is the main driver of the Sustainable Development Goals (SDGs), and it is connected to 11 of the 17 SDGs. Consequently, biopesticides connect green chemistry and sustainable development through agriculture. Though the call for the increased use of biopesticides is evident, some shortcomings need to be addressed through nanotechnology, omics technology, and RNA interference (RNAi) technology. When these avenues are fully explored, the use of synthetic pesticides will become an obsolete practice.

Author Contributions

Development of original draft: E.O.F. and G.N.I.; review and editing: E.O.F. and G.N.I.; visualization: W.N.; supervision and funding acquisition: T.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no funding.

Acknowledgments

The authors wish to acknowledge the graphical representation contributions of Rosina Nkuna. We also acknowledge the critical appraisal of students and staff of IDEAS Biotechnology Unit at UNISA.

Conflicts of Interest

No conflict of interest to declare.

References

  1. Anastas, P.; Eghbali, N. Green chemistry: Principles and practice. Chem. Society Rev. 2010, 39, 301–312. [Google Scholar] [CrossRef] [PubMed]
  2. Saranya, G.; Ramachandra, T.V. Life cycle assessment of biodiesel from estuarine microalgae. Energy Convers. Manag. X 2020, 8, 100065. [Google Scholar] [CrossRef]
  3. Kabbage, M.; Yarden, O.; Dickman, M.B. Pathogenic attributes of Sclerotinia sclerotiorum: Switching from a biotrophic to necrotrophic lifestyle. Plant Sci. 2015, 233, 53–60. [Google Scholar] [CrossRef] [PubMed]
  4. Turner, R.S. After the famine: Plant pathology, Phytophthora infestans, and the late blight of potatoes, 1845–1960. Hist. Stud. Phys. Biol. Sci. 2005, 35, 341–370. [Google Scholar] [CrossRef]
  5. Temba, B.A. Occurrence of Mycotoxins in Harvested Maize in Kenya and Tanzania and Postharvest Control by Photosensitisation. Available online: https://www.aflatoxinpartnership.org/ (accessed on 13 August 2022).
  6. Odeyemi, O.; Ugwu, J. Fall armyworm, Spodoptera frugiperda outbreak in Nigeria: Impacts and management on maize fields-a review. Ethiop. J. Environ. Stud. Manag. 2021, 14, 353–369. [Google Scholar]
  7. Dutta, T.K.; Khan, M.R.; Phani, V. Plant-parasitic nematode management via biofumigation using brassica and non-brassica plants: Current status and future prospects. Curr. Plant Biol. 2019, 17, 17–32. [Google Scholar] [CrossRef]
  8. Heinrichs, E.A. Management of Rice Insect Pests. Radcliffe’s IPM World Textbook. Available online: https://ipmworld.umn.edu/heinrichs (accessed on 4 January 2021).
  9. Nicholls, C.I.; Altieri, M.A. Plant biodiversity enhances bees and other insect pollinators in agroecosystems. A review. Agron. Sustain. Dev. 2013, 33, 257–274. [Google Scholar] [CrossRef] [Green Version]
  10. Keifer, M.C.; Firestone, J. Neurotoxicity of pesticides. J. Agromedicine 2007, 12, 17–25. [Google Scholar] [CrossRef]
  11. Jhala, J.; Baloda, A.S.; Rajput, V.S. Role of bio-pesticides in recent trends of insect pest management: A review. J. Pharmacogn. Phytochem. 2020, 9, 2237–2240. [Google Scholar]
  12. Constantine, K.L.; Kansiime, M.K.; Mugambi, I.; Nunda, W.; Chacha, D.; Rware, H.; Makale, F.; Mulema, J.; Lamontagne-Godwin, J.; Williams, F.; et al. Why don’t smallholder farmers in Kenya use more biopesticides? Pest Manag. Sci. 2020, 76, 3615–3625. [Google Scholar] [CrossRef]
  13. Singh, J.K.; Yadav, K.K.; Kumar, V. Integrated pest management: Conservation practices for agriculture and environment. Int. J. Environ. Rehabil. Conserv. 2017, 8, 17–28. [Google Scholar]
  14. Rao, N.C.; Bathla, S.; Kumar, A.; Jha, G.K. Agriculture and sustainable development goals: An overview and issues. Agric. Econ. Res. Rev. 2018, 31, 1–7. [Google Scholar] [CrossRef] [Green Version]
  15. Hák, T.; Janoušková, S.; Moldan, B. Sustainable Development Goals: A need for relevant indicators. Ecol. Indic. 2016, 60, 565–573. [Google Scholar] [CrossRef]
  16. Copping, L.G. Report on the 5th EEN Biopesticide event. Outlooks Pest Manag. 2017, 28, 108–112. [Google Scholar] [CrossRef]
  17. Cataldo, E.; Fucile, M.; Mattii, G.B. Biostimulants in viticulture: A sustainable approach against biotic and abiotic stresses. Plants 2022, 11, 162. [Google Scholar] [CrossRef]
  18. Oguh, C.E.; Okpaka, C.O.; Ubani, C.S.; Okekeaji, U.; Joseph, P.S.; Amadi, E.U. Natural pesticides (biopesticides) and uses in pest management-a critical review. Asian J. Biotechnol. Genetic. Eng. 2019, 2, 1–18. [Google Scholar]
  19. Marrone, P.G. An effective biofungicide with novel modes of action. Pestic. Outlook 2002, 13, 193–194. [Google Scholar] [CrossRef]
  20. Kumar, S.; Singh, A. Biopesticides: Present status and the future prospects. J. Fertil. Pestic. 2015, 6, 100–129. [Google Scholar] [CrossRef]
  21. Fenibo, E.O.; Ijoma, G.N.; Matambo, T. Biopesticides in sustainable agriculture: A critical sustainable development driver governed by green chemistry principles. Front. Sustain. Food Syst. 2021, 5, 619058. [Google Scholar] [CrossRef]
  22. Rai, P.K.; Lee, S.S.; Zhang, M.; Tsang, Y.F.; Kim, K.H. Heavy metals in food crops: Health risks, fate, mechanisms, and management. Environ. Intl. 2019, 125, 365–385. [Google Scholar] [CrossRef]
  23. Bhandari, S.; Kasana, V. Fe3+-Montmorillonite K10 as an Efficient, green and reusable heterogeneous catalyst for synthesis of mannich type reaction under solvent-free condition. Int. Res. J. Pure Appl. Chem. 2018, 16, 1–11. [Google Scholar] [CrossRef]
  24. Mubeen, K.; Shehzad, M.; Sarwar, N.; Rehman, H.U.; Yasir, T.A.; Wasaya, A.; Ahmad, M.; Hussain, M.; Abbas, M.B.; Yonas, M.W.; et al. The impact of horse purslane (Trianthema portulacastrum L.) infestation on soybean [Glycine max (L.) Merrill] productivity in northern irrigated plains of Pakistan. PLoS ONE 2021, 16, e0257083. [Google Scholar] [CrossRef] [PubMed]
  25. Badhai, S.; Gupta, A.K.; Maurya, S.P.; Koiri, B. Ecological/cultural measures of weed management for sustainable agriculture. J. Wastes Biomass. Manag. 2021, 3, 36–38. [Google Scholar] [CrossRef]
  26. Schroeder, J.; Thomas, S.H.; Murray, L.W. Impacts of crop pests on weeds and weed–crop interactions. Weed Sci. 2005, 6, 918–922. [Google Scholar] [CrossRef]
  27. Liebman, M.; Mohler, C.L.; Staver, C.P. Ecological management of agricultural weeds. Cambridge University Press: Cambride, MA, USA, 2001. [Google Scholar]
  28. Marro, N.; Caccia, M.; López-Ráez, J.A. Are strigolactones a key in plant–parasitic nematodes interactions? An intriguing question. Plant Soil 2021, 462, 591–601. [Google Scholar] [CrossRef]
  29. Benzonan, N.C.; Dalisay, L.C.; Reponte, K.C.; Mapanao, C.P.; Alvarez, L.V.; Rendon, A.O.; Zurbano, L.Y. Plant-parasitic nematodes associated with pineapple (Ananas comosus) in selected Provinces in Luzon, Philippines. Eur. J. Mol. Clinical. Med. 2021, 8, 945–957. [Google Scholar]
  30. Charlier, J.; Thamsborg, S.M.; Bartley, D.J.; Skuce, P.J.; Kenyon, F.; Geurden, T.; Hoste, H.; Williams, A.R.; Sotiraki, S.; Höglund, J.; et al. Mind the gaps in research on the control of gastrointestinal nematodes of farmed ruminants and pigs. Transbound. Emerg. Dis. 2018, 65, 217–234. [Google Scholar] [CrossRef] [Green Version]
  31. Ahmad, M.; Ali, Q.; Hafeez, M.M.; Malik, A. Improvement for biotic and abiotic stress tolerance in crop plants. Biol. Clin. Sci. Res. J. 2021, 2021, 1. [Google Scholar] [CrossRef]
  32. Budiyanto, G.; Trisnawati, D.W.; Aisyah, S.N. Vegetation analysis of the middle slope geomorphic units on the southern flank of Mount Merapi. In IOP Conference Series: Earth and Environmental Science; IOP Publishing: Bristol, UK, 2021. [Google Scholar]
  33. Burrows, M.; Franc, G.; Rush, C.; Blunt, T.; Ito, D.; Kinzer, K.; Olson, J.; O’Mara, J.; Price, J.; Tande, C.; et al. Occurrence of viruses in wheat in the Great Plains region. Plant Health Prog. 2008, 10, 14. [Google Scholar] [CrossRef] [Green Version]
  34. Deniau, M.; Pihain, M.; Béchade, B.; Jung, V.; Brunellière, M.; Gouesbet, V.; Prinzing, A. Seeds and seedlings of oaks suffer from mammals and molluscs close to phylogenetically isolated, old adults. Ann. Bot. 2021, 127, 787–798. [Google Scholar] [CrossRef]
  35. Matevski, D.; Glatthorn, J.; Kriegel, P.; Schuldt, A. Non-native Douglas fir (Pseudotsuga menziesii) promotes sentinel prey attack rates in Central European forests. Forest Ecol. Manag. 2021, 489, 119099. [Google Scholar] [CrossRef]
  36. Witmer, G.W. The ecology of vertebrate pests and integrated pest management (IPM). USDA Natl. Wildl. Res. Cent.-Staff. Publ. 2007, 17, 730. [Google Scholar]
  37. Sandle, T.; Vijayakumar, R.; Saleh, A.l.; Aboody, M.; Saravanakumar, S. In vitro fungicidal activity of biocides against pharmaceutical environmental fungal isolates. J. Appl. Microbiol. 2014, 117, 1267–1273. [Google Scholar] [CrossRef] [PubMed]
  38. Paarlberg, R. African non-adopters. In Handbook on Agriculture, Biotechnology and Development; Edward Elgar Publishing: Chertenham, UK, 2014. [Google Scholar]
  39. Dean, R.; Van Kan, J.A.; Pretorius, Z.A.; Hammond-Kosack, K.E.; Di Pietro, A.; Spanu, P.D.; Rudd, J.J.; Dickman, M.; Kahmann, R.; Ellis, J.; et al. The top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 2012, 13, 414–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Zviahintseva, A.M. Major cucumber diseases and the crop immunity. Ukr. J. Ecol. 2021, 11, 46–54. [Google Scholar]
  41. Usta, C. Microorganisms in biological pest control—A review (bacterial toxin application and effect of environmental factors). Curr. Prog. Biol. Res. 2013, 13, 287–317. [Google Scholar]
  42. Holvoet, K.; Sampers, I.; Seynnaeve, M.; Jacxsens, L.; Uyttendaele, M. Agricultural and management practices and bacterial contamination in greenhouse versus open field lettuce production. Int. J. Environ. Res. Public Health 2015, 12, 32–63. [Google Scholar] [CrossRef] [Green Version]
  43. Sharma, A.; Shankhdhar, D.; Shankhdhar, S.C. Growth promotion of the rice genotypes by PGPRs isolated from rice rhizosphere. J. Soil. Sci. Plant Nutr. 2014, 14, 505–517. [Google Scholar] [CrossRef]
  44. Varanda, C.M.; Félix, M.D.R.; Campos, M.D.; Patanita, M.; Materatski, P. Plant viruses: From targets to tools for CRISPR. Viruses 2021, 13, 141. [Google Scholar] [CrossRef]
  45. Nicaise, V. Crop immunity against viruses: Outcomes and future challenges. Front. Plant Sci. 2014, 5, 660. [Google Scholar] [CrossRef]
  46. Murcia, N.; Serra, P.; Olmos, A.; Durán-Vila, N. A novel hybridization approach for detection of citrus viroids. Mol. Cellular. Probes. 2009, 23, 95–102. [Google Scholar] [CrossRef] [PubMed]
  47. Jarman, W.M.; Ballschmiter, K. From coal to DDT: The history of the development of the pesticide DDT from synthetic dyes till Silent Spring. Endeavour 2012, 36, 131–142. [Google Scholar] [CrossRef] [PubMed]
  48. Davies, T.G.E.; Field, L.M.; Usherwood, P.N.R.; Williamson, M.S. DDT, pyrethrins, pyrethroids and insect sodium channels. IUBMB Life 2007, 59, 151–162. [Google Scholar] [CrossRef] [Green Version]
  49. Abreu-Villaca, Y.; Levin, E.D. Developmental neurotoxicity of succeeding generations of insecticides. Environ. Int. 2017, 99, 55–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hargrove, T.Y.; Wawrzak, Z.; Liu, J.; Waterman, M.R.; Nes, W.D.; Lepesheva, G.I. Structural complex of sterol 14a-demethylase (CYP51) with 14a-methylenecyclopropyl-?7-24, 25-dihydrolanosterol. J. Lipid. Res. 2011, 53, 311–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Stehmann, C.; De Waard, M.A. Relationship between chemical structure and biological activity of triazole fungicides against Botrytis cinerea. Pestic. Sci. 1995, 44, 183–195. [Google Scholar] [CrossRef]
  52. Kim, D.H.; Rossi, J.J. RNAi mechanisms and applications. Biotechniques 2008, 44, 613–616. [Google Scholar] [CrossRef]
  53. Linde, C.D. Physico-Chemical Properties and Environmental Fate of Pesticides. 1994. Available online: https://agris.fao.org/agris-search/index.do (accessed on 10 August 2022).
  54. Jablonowski, N.D.; Linden, A.; Koppchen, S.; Thiele, B.; Hofmann, D.; Mittelstaedt, W.; Putz, T.; Burauel, P. Long-term persistence of various 14C-labeled pesticides in soils. Environ. Pollut. 2012, 168, 29–36. [Google Scholar] [CrossRef]
  55. Toundou, O.; Palanga, K.K.; Simalou, O.; Abalo, M.; Woglo, I.; Tozo, K. Biopesticide Plants species of the mining area of Tokpli (South-Togo) effects on Okra (Abelmoschus esculentus) protection against Aphtona spp. Int. J. Biol. Chem. Sci. 2020, 14, 225–238. [Google Scholar] [CrossRef]
  56. Fox, J.E.; Gulledge, J.; Engelhaupt, E.; Burow, M.E.; McLachlan, J. A Pesticides reduce symbiotic efficiency of nitrogen-fixing rhizobia and host plants. Proc. Natl. Acad. Sci. USA 2007, 104, 10282–10287. [Google Scholar] [CrossRef] [Green Version]
  57. Han, S.H.; An, J.Y.; Hwang, J.; Kim, S.; Bin, P.B.B. The effects of organic manure and chemical fertilizer on the growth and nutrient concentrations of yellow poplar (Liriodendron tulipifera Lin.) in a nursery system. Forest Sci. Technol. 2016, 12, 137–143. [Google Scholar] [CrossRef] [Green Version]
  58. Knipe, D.W. Pesticide exposure in Sri Lanka. Int. J. Epidemiol. 2016, 45, 327–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Mathew, P.; Jose, A.; Alex, R.G.; Mohan, V.R. Chronic pesticide exposure: Health effects among pesticide sprayers in Southern India. Indian J. Occup. Environ. Med. 2015, 19, 95–101. [Google Scholar] [PubMed] [Green Version]
  60. Crow, W.D.; Catchot, A.L.; Gore, J.; Dodds, D.M.; Cook, D.R.; Allen, T.W. Evaluation of tillage, at-planting treatment, and nematicide on tobacco thrips (Thysanoptera: Thripidae) and reniform nematode (Tylenchida: Hoplolamidae) management in cotton. Agronomy 2020, 10, 300. [Google Scholar] [CrossRef] [Green Version]
  61. Nguyen, T.T. Pesticide use in rice farming and its impacts on climbing perch (Anabas testudineus) in the Mekong Delta of Vietnam. Ph.D. Thesis, Stockholm University, Stockholm, Sweden, 2016. [Google Scholar]
  62. Ramesh, M.; Narmadha, S.; Poopal, R.K. Toxicity of furadan (carbofuran 3% g) in Cyprinus carpio: Haematological, biochemical and enzymological alterations and recovery response. Beni-Seuf Univ. J. Appl. Sci. 2015, 4, 314–326. [Google Scholar] [CrossRef] [Green Version]
  63. Buser, H.R.; Muuller, M.D.; Buerge, I.J.; Poiger, T. Composition of aldrin, dieldrin, and photodieldrin enantiomers in technical and environmental samples. J. Agri. Food Chem. 2009, 57, 7445–7452. [Google Scholar] [CrossRef]
  64. Manyilizu, B.W. Pesticides, anthropogenic activities, history and the health of our environment: Lessons from africa. In Pesticides-Use and Misuse and Their Impact in the Environment; IntechOpen: London, UK, 2019. [Google Scholar]
  65. Bacmaga, M.; Wyszkowska, J.; Kucharski, J. The influence of chlorothalonil on the activity of soil microorganisms and enzymes. Ecotoxicology 2018, 27, 1188–1202. [Google Scholar] [CrossRef]
  66. Swarupa, P.; Kumar, A. Impact of chlorpyrifos on plant growth promoting rhizobacteria isolated from Abelmoschus esculentus. J. Pure Appl. Microbiol. 2018, 12, 2149–2158. [Google Scholar] [CrossRef]
  67. Okoroiwu, H.U.; Iwara, I.A. Dichlorvos toxicity: A public health perspective. Interdiscip. Toxicol. 2018, 11, 129–137. [Google Scholar] [CrossRef] [Green Version]
  68. Aggarwal, V.; Deng, X.; Tuli, A.; Goh, K.S. Diazinon—Chemistry and environmental fate: A California perspective. In Reviews of Environmental Contamination and Toxicology; Springer: New York, NY, USA, 2013; pp. 107–140. [Google Scholar]
  69. Johnson, B.; Whitford, F.; Flakne, D.; Bauman, T.; Nice, G.; Frankenberger, J.; Hahn, L.; Bailey, T.; Donald, B.; Mann, C.; et al. Atrazine Use and Weed Management Strategies to Protect Surface Water Quality. 2004. Available online: https://ppp.purdue.edu/wp-content/uploads/2016/08/PPP-67.pdf (accessed on 26 December 2021).
  70. Hernandez, M.; Jia, Z.; Conrad, R.; Eeger, M. Simazine application inhibits nitrification and changes the ammonia-oxidizing bacterial communities in a fertilized agricultural soil. FEMS Microbiol. Ecol. 2011, 78, 511–519. [Google Scholar] [CrossRef] [Green Version]
  71. Narayanan, M.; Ranganathan, M.; Subramanian, S.M.; Kumarasamy, S.; Kandasamy, S. Toxicity of cypermethrin and enzyme inhibitor synergists in red hairy caterpillar Amsacta albistriga (Lepidoptera: Arctiidae). J. Basic Appl. Zool. 2020, 81, 1–8. [Google Scholar] [CrossRef]
  72. Paul, M.; Hossain, M.S.; Rahman, M.M.; Khaliq, Q.A.; Rahman, S. Chemodynamics of cypermethrin insecticide in summer country bean ecosystem in Bangladesh. Res. J. Environ. Toxicol. 2016, 10, 50–62. [Google Scholar]
  73. Ogunnupebi, T.A.; Oluyori, A.P.; Dada, A.O.; Oladeji, O.S.; Inyinbor, A.A.; Egharevba, G.O. Promising natural products in crop protection and food preservation: Basis, advances, and future prospects. Int. J. Agron. 2020, 2020, 8840046. [Google Scholar] [CrossRef]
  74. Ujvary, I. Delmethrin: Properties, mode of action and safety issues related to tsetse control. In Proceedings of the WHO/TDR Progress and Planning Meeting, Nairobi, Kenya, 6–9 July 2015. [Google Scholar]
  75. Islam, F.; Wang, J.; Farooq, M.A.; Khan, M.S.; Xu, L.; Zhu, J.; Zhao, M.; Munos, S.; Li, Q.X.; Zhou, W. Potential impact of the herbicide 2, 4-dichlorophenoxyacetic acid on human and ecosystems. Environ. Intl. 2018, 111, 332–351. [Google Scholar] [CrossRef] [PubMed]
  76. Meena, R.S.; Kumar, S.; Datta, R.; Lal, R.; Vijayakumar, V.; Brtnicky, M.; Sharma, M.P.; Yadav, G.S.; Jhariya, M.K.; Jangir, C.K.; et al. Impact of agrochemicals on soil microbiota and management: A review. Land 2020, 9, 34–56. [Google Scholar] [CrossRef] [Green Version]
  77. Fayinminnu, O.O.; Adesiyan, S.O.; Sosanya, S.O. Effect of paraquat as post emergence herbicide on yield of cowpea (Vigna unguiculata (l) walp). J. Agric. Social. Res. 2010, 10, 25–39. [Google Scholar] [CrossRef] [Green Version]
  78. Meyer, C.J.; Norsworthy, J.K. Timing and application rate for sequential applications of glufosinate are critical for maximizing control of annual weeds in LibertyLink® Soybean. Int. J. Agron. 2020, 2020, 1–7. [Google Scholar] [CrossRef]
  79. Thakur, N.; Kaur, S.; Tomar, P.; Thakur, S.; Yadav, A.N. Microbial biopesticides: Current status and advancement for sustainable agriculture and environment. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2020; pp. 243–282. [Google Scholar]
  80. Rajput, V.S.; Jhala, J.; Acharya, V.S. Biopesticides and their mode of action against insect pests: A review. Int. JChe Stud. 2020, 8, 2856–2862. [Google Scholar] [CrossRef]
  81. Harding, D.P.; Raizada, M.N. Controlling weeds with fungi, bacteria and viruses: A review. Front. Plant Sci. 2015, 6, 659. [Google Scholar] [CrossRef]
  82. Tripathi, A.N.; Meena, B.R.; Pandey, K.K.; Singh, J. Microbial bioagents in agriculture: Current status and prospects. In New Frontiers in Stress Management for Durable Agriculture; Rakshit, A., Singh, H., Singh, A., Singh, U., Fraceto, L., Eds.; Springer: Singapore, 2020. [Google Scholar]
  83. Khan, M.R.; Mohiddin, F.A.; Khan, S.M.; Khan, B. Effect of seed treatment with certain biopesticides on root-knot of chickpea [Cicer arietinum L.; India]. Nematol. Mediterr. 2005, 33, 107–112. [Google Scholar]
  84. Hammami, I.; Smaoui, S.; Hsouna, A.B.; Hamdi, N.; Triki, M.A. Ruta montana L. leaf essential oil and extracts: Characterization of bioactive compounds and suppression of crown gall disease. EXCLI J. 2015, 14, 83. [Google Scholar] [PubMed]
  85. Ichim, E.; Marutescu, L.; Popa, M.; Cristea, S. Antimicrobial efficacy of some plant extracts on bacterial ring rot pathogen, Clavibacter michiganensis spp. sepedonicus. EuroBiotech J. 2017, 1, 85–88. [Google Scholar] [CrossRef] [Green Version]
  86. Vuko, E.; Dunkic, V.; Rušcic, M.; Nazlic, M.; Mandic, N.; Soldo, B.; Šprung, M.; Fredotovic, Ž. Chemical composition and new biological activities of essential oil and hydrosol of Hypericum perforatum L. ssp. veronense (Schrank) H. Lindb. Plants 2021, 10, 1014. [Google Scholar] [CrossRef]
  87. Senthil-Nathan, S. A review of biopesticides and their mode of action against insect pests. In Environmental Sustainability; Springer: New Delhi, India, 2015; pp. 49–63. [Google Scholar]
  88. Idrees, A.; Afzal, A.; Qadir, Z.A.; Li, J. Bioassays of Beauveria bassiana isolates against the fall armyworm, Spodoptera frugiperda. J. Fungi 2022, 8, 717. [Google Scholar] [CrossRef] [PubMed]
  89. Zamanizadeh, H.R.; Hatami, N.; Aminaee, M.M.; Rakhshandehroo, F. Application of biofungicides in control of damping disease off in greenhouse crops as a possible substitute to synthetic fungicides. Int. J. Environ. Sci. Technol. 2011, 8, 129–136. [Google Scholar] [CrossRef] [Green Version]
  90. Kohl, J.; Kolnaar, R.; Ravensberg, W.J. Mode of action of microbial biological control agents against plant diseases: Relevance beyond efficacy. Front. Plant Sci. 2019, 10, 845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Krif, G.; Lahlali, R.; El Aissami, A.; Laasli, S.E.; Mimouni, A.; Serderidis, S.; Picaud, T.; Moens, A.; Dababat, A.A.; Fahad, K.; et al. Efficacy of authentic bio-nematicides against the root-knot nematode, Meloidogyne javanica infecting tomato under greenhouse conditions. Physiol. Mol. Plant Pathol. 2022, 118, 101803. [Google Scholar] [CrossRef]
  92. Astani, A.; Reichling, J.; Schnitzler, P. Screening for antiviral activities of isolated compounds from essential oils. Evidence-based complementary. Alt. Med. 2011, 2011, 253643. [Google Scholar]
  93. Ma, L.; Yao, L. Antiviral effects of plant-derived essential oils and their components: An updated review. Molecules 2020, 25, 2627. [Google Scholar] [CrossRef]
  94. Williamson, S. Understanding natural enemies; a review of training and information in the practical use of biological control. Biocontrol. News Infor. 1998, 19, 117N–125N. [Google Scholar]
  95. Dang, K.; Doggett, S.L.; Veera Singham, G.; Lee, C.Y. Insecticide resistance and resistance mechanisms in bed bugs, Cimex spp.(Hemiptera: Cimicidae). Parasites Vectors 2017, 10, 318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Damos, P.; Colomar, L.A.E.; Ioriatti, C. Integrated fruit production and pest management in Europe: The apple case study and how far we are from the original concept? Insects 2015, 6, 626–657. [Google Scholar] [CrossRef] [PubMed]
  97. Tangtrakulwanich, K.; Reddy, G.V. Development of insect resistance to plant biopesticides: An overview. In Advances in Plant Biopesticides; Springer: New Delhi, India, 2014; pp. 47–62. [Google Scholar]
  98. Sharma, R.; Kole, C. Utilization of neem and neem products in agriculture. In The Neem Genome; Springer: Cham, Switzerland, 2019; pp. 31–48. [Google Scholar]
  99. Reuveni, M.; Sanches, E.; Barbier, M. Curative and suppressive activities of essential tea tree oil against fungal plant pathogens. Agronomy 2020, 10, 609. [Google Scholar] [CrossRef]
  100. Lopez-Perez, J.A.; Edwards, S.; Ploeg, A. Control of root-knot nematodes on tomato in stone wool substrate with biological nematicides. J. Nematol. 2011, 43, 110–117. [Google Scholar] [PubMed]
  101. Rajamani, M.; Negi, A. Biopesticides for pest management. Sustain. Bioecon 2020, 13, 239–266. [Google Scholar]
  102. Bravo, A.; Gill, S.S.; Soberon, M. Mode of action of Bacillus thuringiensis Cry and Cyt toxins and their potential for insect control. Toxicon 2007, 49, 423–435. [Google Scholar] [CrossRef] [Green Version]
  103. Baskar, K.; Muthu, C.; Raj, G.A.; Kingsley, S.; Ignacimuthu, S. Ovicidal activity of Atalantia monophylla (L) Correa against Spodoptera litura Fab.(Lepidoptera: Noctuidae). Asian Pacific. J. Trop. Biomed. 2012, 2, 987–991. [Google Scholar] [CrossRef] [Green Version]
  104. Hano, P.; Khan, M.R. Evaluation of fungal (Paecilomyces lilacinus) formulations against root knot nematode infecting tomato. Bangladesh J. Bot. 2016, 45, 1003–1013. [Google Scholar]
  105. Mahmoud, B.S. Effect of X-ray treatments on inoculated Escherichia coli O157: H7, Salmonella enterica, Shigella flexneri and Vibrio parahaemolyticus in ready-to-eat shrimp. Food Microbiol. 2009, 26, 860–864. [Google Scholar] [CrossRef]
  106. Tian, J.C.; Wang, X.P.; Chen, Y.; Romeis, J.; Naranjo, S.E.; Hellmich, R.L.; Wang, P.; Shelton, A.M. Bt cotton producing Cry1Ac and Cry2Ab does not harm two parasitoids, Cotesia marginiventris and Copidosoma floridanum. Sci. Rep. 2018, 8, 307. [Google Scholar] [CrossRef] [Green Version]
  107. Baranski, R.; Klimek-Chodacka, M.; Lukasiewicz, A. Approved genetically modified (GM) horticultural plants: A 25-year perspective. Folia Hortic. 2019, 31, 3–49. [Google Scholar] [CrossRef] [Green Version]
  108. Jahan, S.N.; Åsman, A.K.; Corcoran, P.; Fogelqvist, J.; Vetukuri, R.R.; Dixelius., C. Plant-mediated gene silencing restricts growth of the potato late blight pathogen Phytophthora infestans. J. Exp. Bot. 2015, 66, 2785–2794. [Google Scholar] [CrossRef] [Green Version]
  109. Edirisinghe, H.M.; Leschen, R.A.B.; Dale, J.; Wignall, A.E. Insights into the establishment of introduced species using Coccinellines (Coleoptera: Coccinellidae) as a model system. Coleopt. Bull. 2021, 75, 121–149. [Google Scholar] [CrossRef]
  110. Ferguson, C.M.; Barratt, B.I.P.; Bell, N.; Goldson, S.L.; Hardwick, S.; Jackson, M.; Wilson, M. Quantifying the economic cost of invertebrate pests to New Zealand’s pastoral industry. New Zeal. J. Agr. Res. 2018, 62, 255–315. [Google Scholar] [CrossRef]
  111. Golmohammadi, M.; Hassankiadeh, M.N.; Zhang, L. Facile biosynthesis of SnO2/ZnO nanocomposite using Acroptilon repens flower extract and evaluation of their photocatalytic activity. Ceram. Int. 2021, 47, 29303–29308. [Google Scholar] [CrossRef]
  112. Volesky, N.; Schrumm, Z.R. High Tunnel Pest Management-Aphids. Utah Pest Fact Sheet. Utta State University. 2021. Available online: https://digitalcommons.usu.edu/cgi/viewcontent.cgi?article=3197&context=extension_curall (accessed on 5 July 2022).
  113. Ouedraogo, L.; Fuchs, D.; Schaefer, H.; Kiendrebeogo, M. Morphological and molecular characterization of Zanthoxylum zanthoxyloides (Rutaceae) from Burkina Faso. Plants 2019, 8, 353. [Google Scholar] [CrossRef] [Green Version]
  114. Filho, F.H.; Heldens, W.B.; Kong, Z.; de Lange, E.S. Drones: Innovative technology for use in precision pest management. J. Econ. Entomol. 2019, 113, 1–25. [Google Scholar] [CrossRef] [Green Version]
  115. Li, Y.; Cheng, L.; Bingru, L.; Zhihong, M. Trifluralin residues in soils from main cotton fields of China and associated ecological risk. Chemosphere 2021, 284, 131300. [Google Scholar] [CrossRef]
  116. Karar, M.E.; Alsunaydi, F.; Albusaymi, S.; Alotaibi, S. A new mobile application of agricultural pests recognition using deep learning in cloud computing system. Alex. Eng. J. 2021, 60, 4423–4432. [Google Scholar] [CrossRef]
  117. James, B.; Atcha-Ahowé, C.; Godonou, I.; Baimey, H.; Goergen, H.; Sikirou, R.; Toko, M. Integrated Pest Management in Vegetable Production: A Guide for Extension Workers in West Africa. IITA: Ibadan, Nigeria, 2010.
  118. Deguine, J.P.; Aubertot, J.N.; Flor, R.J.; Lescourret, F.; Wyckhuys, K.A.; Ratnadass, A. Integrated pest management: Good intentions, hard realities. A review. Agro. Sustain. Dev. 2021, 41, 38. [Google Scholar]
  119. Gaetani, R.; Lacotte, V.; Dufour, V.; Clavel, A.; Duport, G.; Gaget, K.; Calevro, F.; Da Silva, P.; Heddi, A.; Vincent, D.; et al. Sustainable laser-based technology for insect pest control. Sci. Rep. 2021, 11, 11068. [Google Scholar]
  120. Borowik, A.; Wyszkowska, J. Soil moisture as a factor affecting the microbiological and biochemical activity of soil. Plant Soil. Environ. 2016, 62, 250–255. [Google Scholar] [CrossRef] [Green Version]
  121. Acheuk, F.; Basiouni, S.; Shehata, A.A.; Dick, K.; Hajri, H.; Lasram, S.; Yilmaz, M.; Emekci, M.; Tsiamis, G.; Spona-Friedl, M.; et al. Current status and prospects of botanical biopesticides in Europe and mediterranean countries. Biomolecules 2022, 12, 311. [Google Scholar] [CrossRef] [PubMed]
  122. Umetsu, N.; Shirai, Y. Development of novel pesticides in the 21st century. J. Pestic. Sci. 2020, 45, 54–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Ahirwar, S.; Malik, M.S.; Ahirwar, R.; Shukla, J.P. Identification of suitable sites and structures for artificial groundwater recharge for sustainable groundwater resource development and management. Groundw. Sustain. Dev. 2020, 11, 100388. [Google Scholar] [CrossRef]
  124. Kalwani, M.; Chakdar, H.; Srivastava, A.; Pabbi, S.; Shukla, P. Effects of nanofertilizers on soil and plant-associated microbial communities: Emerging trends and perspectives. Chemosphere 2022, 287, 132107. [Google Scholar] [CrossRef]
  125. Lata, R.K.; Divjot, K.; Nath, Y.A. Endophytic microbiomes: Biodiversity, ecological significance and biotechnological applications. Res. J. Biotechnol. 2019, 14, 10. [Google Scholar]
  126. Welbaum, G.E.; Sturz, A.V.; Dong, Z.; Nowak, J. Managing soil microorganisms to improve productivity of agro-ecosystems. Crit. Rev. Plant Sci. 2004, 23, 175–193. [Google Scholar] [CrossRef]
  127. Gupta, G.; Parihar, S.S.; Ahirwar, N.K.; Snehi, S.K.; Singh, V. Plant growth promoting rhizobacteria (PGPR): Current and future prospects for development of sustainable agriculture. J. Microb. Biochem. Technol. 2015, 7, 096–102. [Google Scholar]
  128. Bartomeus, C. Contribution of Insect Pollinators to Crop Yield and Quality Varies with Agricultural Intensification Vanbergen. A Report to the Department for Environment, Food and Rural Affairs; Defra: London, UK, 2014. [Google Scholar]
  129. IPBES. Summary for Policymakers of the Assessment Report of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services on Pollinators, Pollination and Food Production; IPBES: Panama, Germany, 2016. [Google Scholar]
  130. Adamidis, G.C.; Cartar, R.V.; Melathopoulos, A.P.; Pernal, S.F.; Hoover, S.E. Pollinators enhance crop yield and shorten the growing season by modulating plant functional characteristics: A comparison of 23 canola varieties. Sci. Rep. 2019, 9, 14208. [Google Scholar] [CrossRef] [Green Version]
  131. Porto, R.G.; de Almeida, R.F.; Cruz-Neto, O.; Tabarelli, M.; Viana, B.F.; Peres, C.A.; Lopes, A.V. Pollination ecosystem services: A comprehensive review of economic values, research funding and policy actions. Food Secur. 2020, 12, 1425–1442. [Google Scholar] [CrossRef]
  132. Sawe, T.; Nielsen, A.; Eldegard, K. Crop pollination in small-scale agriculture in Tanzania: Household dependence, awareness and conservation. Sustain 2020, 12, 2228. [Google Scholar] [CrossRef] [Green Version]
  133. Fijen, T.P.; Scheper, J.A.; Boom, T.M.; Janssen, N.; Raemakers, I.; Kleijn, D. Insect pollination is at least as important for marketable crop yield as plant quality in a seed crop. Ecol. Lett. 2018, 21, 1704–1713. [Google Scholar] [CrossRef] [PubMed]
  134. Garratt, M.P.; Bishop, J.; Degani, E.; Potts, S.G.; Shaw, R.F.; Shi, A.; Roy, S. Insect pollination as an agronomic input: Strategies for oilseed rape production. J. Appl. Ecol. 2018, 55, 2834–2842. [Google Scholar] [CrossRef] [Green Version]
  135. Centner, T.J. Pesticide usage is compromising people’s health in the United States: Ideas for reducing damages. Agriculture 2021, 11, 486. [Google Scholar] [CrossRef]
  136. Attina, T.M.; Hauser, R.; Sathyanarayana, S.; Hunt, P.A.; Bourguignon, J.-P.; Myers, J.P.; Trasande, L. Exposure to endocrine-disrupting chemicals in the USA: A population-based disease burden and cost analysis. Lancet Diabetes Endocrinol. 2016, 4, 996–1003. [Google Scholar] [CrossRef]
  137. Nowak, D.J.; Greenfield, E.J. US urban forest statistics, values, and projections. J. For. 2018, 116, 164–177. [Google Scholar] [CrossRef] [Green Version]
  138. Çelık, D.; Meral, M.E.; Waseem, M. A New Area Towards to Digitalization of Energy Systems: Enables, Challenges and Solutions. In Proceedings of the 2022 14th International Conference on Electronics, Computers and Artificial Intelligence (ECAI), Ploiesti, Romania, 30 June–1 July 2022; IEEE; pp. 1–6. [Google Scholar]
  139. Madejski, P.; Chmiel, K.; Subramanian, N.; Kus, T. Methods and techniques for CO2 capture: Review of potential solutions and applications in modern energy technologies. Energies 2022, 15, 887. [Google Scholar] [CrossRef]
  140. Hofmann, T.; Lowry, G.V.; Ghoshal, S.; Tufenkji, N.; Brambilla, D.; Dutcher, J.R.; Gilbertson, L.M.; Giraldo, J.P.; Kinsella, J.M.; Landry, M.P..; et al. Technology readiness and overcoming barriers to sustainably implement nanotechnology-enabled plant agriculture. Nat. Food 2020, 1, 416–425. [Google Scholar] [CrossRef]
  141. Klumper, W.; Qaim, M. A Meta-Analysis of the Impacts of Genetically Modified Crops. PLoS ONE 2014, 9, e111629. [Google Scholar] [CrossRef] [Green Version]
  142. Doni, F.; Suhaimi, N.S.M.; Mispan, M.S.; Fathurrahman, F.; Marzuki, B.M.; Kusmoro, J.; Uphoff, N. Microbial contributions for rice production: From conventional crop management to the use of ‘omics’ technologies. Int. J. Mol. Sci. 2022, 23, 737. [Google Scholar] [CrossRef]
  143. Berini, F.; Casciello, C.; Marcone, G.L.; Marinelli, F. Metagenomics: Novel enzymes from non-culturable microbes. FEMS Microbiol. Lett. 2017, 364, fnx211. [Google Scholar] [CrossRef] [PubMed]
  144. Xu, Y.P.; Chen, H.Y.; Xin, Z.; Cai, X.Z. Induction of hypersensitive response and nonhost resistance by a Cladosporium fulvum elicitor CfHNNI1 is dose-dependent and negatively regulated by salicylic acid. J. Integr. Agric. 2012, 11, 1665–1674. [Google Scholar] [CrossRef]
  145. Hjort, K.; Presti, I.; Elväng, A.; Marinelli, F.; Sjooling, S. Bacterial chitinase with phytopathogen control capacity from suppressive soil revealed by functional metagenomics. Appl. Microbiol. Biotechnol. 2014, 98, 2819–2828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Prigigallo, M.I.; Ahmed, A.O.; Cacciola, R.; Simona, M.; Sanzani, L.; Schena, L. Metabarcoding analysis of Phytophthora diversity using genus-specific primers and 454 pyrosequencing. Phytopathology 2016, 106, 305–313. [Google Scholar] [CrossRef] [Green Version]
  147. Cobo-Díaz, J.F.; Baroncelli, R.; Le Floch, G.; Picot, A. A novel metabarcoding approach to investigate Fusarium species composition in soil and plant samples. FEMS Microbiol. Ecol. 2019, 95, fiz084. [Google Scholar] [CrossRef]
  148. Kemp, N.D.; Vaughan, M.M.; McCormick, S.P.; Brown, J.A.; Bakker, M.G. Sarocladium zeae is a systemic endophyte of wheat and an effective biocontrol agent against Fusarium head blight. Biol. Cont. 2020, 149, 104329. [Google Scholar] [CrossRef]
  149. Jensen, B.D.; Knorr, K.; Nicolaisen, M. In vitro competition between Fusarium graminearum and Epicoccum nigrum on media and wheat grains. Eur. J. Plant Pathol. 2016, 146, 657–670. [Google Scholar] [CrossRef]
  150. Del-Frari, G.; Cabral, A.; Nascimento, T.; Boavida-Ferreira, R.; Oliveira, H. Epicoccum layuense a potential biological control agent of esca-associated fungi in grapevine. PLoS ONE 2019, 14, e0213273. [Google Scholar] [CrossRef] [Green Version]
  151. Zaidi, S.S.E.A.; Mansoor, S.; Paterson, A. The rise of cotton genomics. Trends Plant Sci. 2018, 23, 953–955. [Google Scholar] [CrossRef]
  152. Wang, M.; Tu, L.; Yuan, D.; Zhu, D.; Shen, C.; Li, J.; Liu, F.; Pei, L.; Wang, P.; Zhao, G. Reference genome sequences of two cultivated allotetraploid cottons, Gossypium hirsutum and Gossypium barbadense. Nat. Genet. 2019, 51, 224–229. [Google Scholar] [CrossRef] [Green Version]
  153. Tyagi, S.; Kumar, R.; Kumar, V.; Won, S.Y.; Shukla, P. Engineering disease resistant plants through CRISPR-Cas9 technology. GM Crops Food 2021, 12, 125–144. [Google Scholar] [CrossRef] [PubMed]
  154. Tarazi, R.; Vaslin, M.F. The viral threat in cotton: How new and emerging technologies accelerate virus identification and virus resistance breeding. Front. Plant Sci. 2022, 13, 851939. [Google Scholar] [CrossRef] [PubMed]
  155. Chen, J.S.; Dagdas, Y.S.; Kleinstiver, B.P.; Welch, M.M.; Sousa, A.A.; Harrington, L.B.; Sternberg, S.H.; Joung, J.K.; Yildiz, A.; Doudna, J.A. Enhanced proofreading governs CRISPR–Cas9 targeting accuracy. Nature 2017, 550, 407–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. McMeniman, C.J.; Corfas, R.A.; Matthews, B.J.; Ritchie, S.A.; Vosshall, L.B. Multimodal integration of carbon dioxide and other sensory cues drives mosquito attraction on humans. Cell 2014, 156, 1060–1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Smidler, A.L.; Terenzi, O.; Soichot, J.; Levashina, E.A.; Marois, E. Targeted mutagenesis in the malaria mosquito using TALE nucleases. PLoS ONE 2013, 8, e74511. [Google Scholar] [CrossRef] [PubMed]
  158. Kim, J.O.; Shin, J.H.; Gumilang, A.; Chung, K.; Choi, K.Y.; Kim, K.S. Effectiveness of different classes of fungicides on Botrytis cinerea causing gray mold on fruit and vegetables. Plant Pathol. J. 2016, 32, 570–574. [Google Scholar] [CrossRef]
  159. Fletcher, S.J.; Reeves, P.T.; Hoang, B.T.; Mitter, N. A perspective on RNAi-based biopesticides. Front. Plant Sci. 2020, 11, 51. [Google Scholar] [CrossRef]
  160. Christiaens, O.; Whyard, S.; Vélez, A.M.; Smagghe, G. Double-Stranded RNA technology to control insect pests: Current status and challenges. Front. Plant Sci. 2020, 11, 451. [Google Scholar] [CrossRef]
  161. Bachman, P.M.; Bolognesi, R.; Moar, W.J.; Mueller, G.M.; Paradise, M.S.; Ramaseshadri, P. Characterization of the spectrum of insecticidal activity of a double-stranded RNA with targeted activity against western corn rootworm (Diabrotica virgifera LeConte). Transg. Res. 2013, 22, 1207–1222. [Google Scholar] [CrossRef] [Green Version]
  162. Jasrotia, P.; Nagpal, M.; Mishra, C.N.; Sharma, A.K.; Kumar, S.; Kamble, U.; Bhardwaj, A.K.; Kashyap, P.L.; Kumar, S.; Singh, G.P. Nanomaterials for postharvest management of insect pests: Current state and future perspectives. Front. Nanotechnol. 2022, 3, 811056. [Google Scholar] [CrossRef]
  163. Pateiro, M.; Gómez, B.; Munekata, P.E.; Barba, F.J.; Putnik, P.; Kovacevic, D.B.; Lorenzo, J.M. Nanoencapsulation of promising bioactive compounds to improve their absorption, stability, functionality and the appearance of the final food products. Molecules 2021, 26, 1547. [Google Scholar] [CrossRef] [PubMed]
  164. Janczak, K..; Kosmalska, D.; Kaczor, D.; Raszkowska-Kaczor, A.; Wedderburn, L.; Malinowski, R. Bactericidal and fungistatic properties of LDPE modified with a biocide containing metal nanoparticles. Materials 2021, 14, 4228. [Google Scholar] [CrossRef] [PubMed]
  165. Hazafa, A.; Murad, M.; Masood, M.U.; Bilal, S.; Khan, M.N.; Farooq, Q.; Iqbal, M.O.; Shakir, M.; Naeem, M. Nano-biopesticides as an emerging technology for pest management. In Insecticides; IntechOpen: York, UK, 2021. [Google Scholar]
  166. Kumar, A.; Thakur, A.; Panesar, P.S. A review on emulsion liquid membrane (ELM) for the treatment of various industrial effluent streams. Rev. Environ. Sci. Bio./Technol. 2019, 18, 153–182. [Google Scholar] [CrossRef]
  167. Chaudhari, A.K.; Singh, V.K.; Kedia, A.; Das, S.; Dubey, N.K. Essential oils and their bioactive compounds as eco-friendly novel green pesticides for management of storage insect pests: Prospects and retrospects. Environ. Sci. Pollution. Res. 2021, 28, 18918–18940. [Google Scholar] [CrossRef] [PubMed]
  168. Palermo, D.; Giunti, G.; Laudani, F.; Palmeri, V.; Campolo, O. Essential oil-based nano-biopesticides: Formulation and bioactivity against the confused flour beetle Tribolium Confusum. Sustainability 2021, 13, 9746. [Google Scholar] [CrossRef]
  169. Behera, A. Nanomaterials. In Advanced Materials; Springer: Cham, Switzerland, 2022; pp. 77–125. [Google Scholar]
  170. McClements, D.J. Food Emulsions: Principles, Practices, and Techniques. Boca Raton; CRC Press: London, UK, 2004. [Google Scholar]
  171. Vilasau, J.; Solans, C.; Gómez, M.; Dabrio, J.; Mújika-Garai, R.; Esquena, J. Phase behaviour of a mixed ionic/ nonionic surfactant system used to prepare stable oil-in-water paraffin emulsions. Colloids Surf. A Physicochem. Eng. Asp. 2011, 384, 473–481. [Google Scholar] [CrossRef]
  172. Maluin, F.N.; Hussein, M.Z. Chitosan-based agronanochemicals as a sustainable alternative in crop protection. Molecules 2020, 25, 1611. [Google Scholar] [CrossRef]
  173. Wani, T.A.; Masoodi, F.A.; Baba, W.N.; Ahmad, M.; Rahmanian, N.; Jafari, S.M. Nanoencapsulation of agrochemicals, fertilizers, and pesticides for improved plant production. In Advances in Phytonanotechnology; Academic Press: Cambridge, MA, USA, 2019; pp. 279–298. [Google Scholar]
  174. Abdollahdokht, D.; Gao, Y.; Faramarz, S.; Poustforoosh, A.; Abbasi, M.; Asadikaram, G.; Nematollahi, M.H. Conventional agrochemicals towards nano-biopesticides: An overview on recent advances. Chem. Biol. Technol. Agric. 2022, 9, 13. [Google Scholar] [CrossRef]
  175. Patir, K. Nanoemulsion and Its Application in Pesticide Formulation. In Handbook of Research on Nanoemulsion Applications in Agriculture, Food, Health, and Biomedical Sciences; IGI Global: Hershey, PA, USA, 2022; pp. 401–424. [Google Scholar]
  176. Manzocco, L.; Mikkonen, K.S.; García-González, C.A. Aerogels as porous structures for food applications: Smart ingredients and novel packaging materials. Food Struct. 2021, 28, 100188. [Google Scholar]
  177. Okey-Onyesolu, C.F.; Hassanisaadi, M.; Bilal, M.; Barani, M.; Rahdar, A.; Iqbal, J.; Kyzas, G.Z. Nanomaterials as nanofertilizers and nanopesticides: An overview. ChemistrySelect 2021, 6, 8645–8663. [Google Scholar] [CrossRef]
  178. Nassar, A.M.K. Pesticide alternatives use in Egypt: The concept and potential. In Sustainability of Agricultural Environment in Egypt: Part II; Springer: Cham, Switzerland, 2018; pp. 111–143. [Google Scholar]
  179. Anandhi, S.; Saminathan, V.R.; Yasotha, P.; Saravanan, P.T.; Rajanbabu, V. Nano-pesticides in pest management. J. Entomol. Zool. Stud. 2020, 8, 685–690. [Google Scholar]
  180. Assalin, M.R.; de Souza, D.R.C.; Rosa, M.A.; Duarte, R.R.M.; Castanha, R.F.; Vilela, E.S.D.; Tasic, L.; Durán, N. Thiamethoxam used as nanopesticide for the effective management of Diaphorina citri psyllid: An environmental-friendly formulation. Int. J. Pest. Manag. 2022, 24, 1–9. [Google Scholar] [CrossRef]
Figure 1. A simplified presentation of the production of a green product concerning the 12 Green Chemistry Principles (GCP). GCP 7: Renewable feedstock; GCP 3: Less harmful chemical synthesis; GCP 5: Benign solvent and auxiliary; GCP 6: Energy efficiency; GCP 8: Reduce derivatives; GCP 9: Catalysis; GCP 10: Degradation design; GCP 12: Inherently safer process; GCP 4: Safer chemical design; GCP 10: Design for degradation; GCP 1: Waste prevention; GCP 2: Atom economy; GCP 11: Real-time analysis.
Figure 1. A simplified presentation of the production of a green product concerning the 12 Green Chemistry Principles (GCP). GCP 7: Renewable feedstock; GCP 3: Less harmful chemical synthesis; GCP 5: Benign solvent and auxiliary; GCP 6: Energy efficiency; GCP 8: Reduce derivatives; GCP 9: Catalysis; GCP 10: Degradation design; GCP 12: Inherently safer process; GCP 4: Safer chemical design; GCP 10: Design for degradation; GCP 1: Waste prevention; GCP 2: Atom economy; GCP 11: Real-time analysis.
Sustainability 14 14417 g001
Figure 2. Negative impacts linked to the toxicological distribution chain of chemical pesticides. D = Deposition, P = Percolation, L = Leaching, Pr = Precipitation; Red lines = Negative impacts.
Figure 2. Negative impacts linked to the toxicological distribution chain of chemical pesticides. D = Deposition, P = Percolation, L = Leaching, Pr = Precipitation; Red lines = Negative impacts.
Sustainability 14 14417 g002
Figure 3. Susceptible target points of insect pest, mode of action and applicable biopesticide. EP = Entomopathogenic.
Figure 3. Susceptible target points of insect pest, mode of action and applicable biopesticide. EP = Entomopathogenic.
Sustainability 14 14417 g003
Figure 4. Biopesticides coverage against plant pests with preferred categories.
Figure 4. Biopesticides coverage against plant pests with preferred categories.
Sustainability 14 14417 g004
Figure 5. A simplified integrated pest management system. Adapted from [48,113,118]. Key: CSPM: Climate-smart pest management; LTS: Least toxic pesticides; CP: Chemical pesticide.
Figure 5. A simplified integrated pest management system. Adapted from [48,113,118]. Key: CSPM: Climate-smart pest management; LTS: Least toxic pesticides; CP: Chemical pesticide.
Sustainability 14 14417 g005
Table 1. Notable examples of plant pests.
Table 1. Notable examples of plant pests.
PestsHostsImpactReference
Weeds:
Annual: Ambrosia artemisiifolia, Abutilon theophrasti, Chenopodium album (Baconweed), Amaranthus spp. (Pigweeds), Digitaria spp.
Biannual: Ailanthus latissimus (Tree-of-heaven), Cirsium vulgare (Bull thistle),
Perennial: Convolvulus arvensis, Rubus spp. (Blackberries), Smilax spp. (Greenbrier), Phytolacca Americana (Pokeweed), Toxicodendron radicans (Poison Ivy).
Varieties of crops including grains, wheat, rice, maize, beans, chickpeas, potatoes, vegetables, and cotton.
  • Reduces sunlight to crops;
  • Aggressively competes for water and nutrients;
  • Tends to grow faster than crops and crowd out actual crops;
  • Can produce certain chemicals that are toxic to crops or grazing animals.
[24,25,26,27]
Nematodes:
Heterodera spp. and Globodera spp., (Plant-parasitic nematodes (PPN), Meloidogyne spp. (root-knot nematodes), Pratylenchus spp., Heterodera and Globodera spp. (cyst nematodes), Bursaphelenchus xylophilus (pine wilt nematode), Aphelenchoides besseyi, Radopholus similis (burrowing nematodes), Xiphinema index (virus vector nematode), Ditylenchus dipsaci, Nacobbus aberrans, Rotylenchulus reniformis (reniform nematode).
Varieties of crops including peaches, nectarines, tomato, pepper, cucumber, almonds, squash, eggplant, okra, sugarcane, beetroot, and pineapple.
  • They attack plant roots and other below-ground parts and sometimes manipulate the gene regulation and metabolism to their advantage;
  • They also attack stems and leaves reducing photosynthetic and water and nutrient translocation.
[28,29,30]
Insects
Aphids, Mexican fruit flies (Anastrepha ludens), grasshoppers, whiteflies, spider mites, silkworms, desert locust (Schistocerca gregaria), migratory locust (Locusta migratoria), screwworm fly (Cochliomyia), tsetse flies (Glossina), uzi fly (Exorista bombycis), potato beetle, Banana- spotting bug (Amblypelta lutescens), European corn borer (Pyrausta nubilalis), Japanese beetle (Popillia japonica), alfalfa weevil (Hypera postica), alfalfa aphid (Therioaphis maculata).
Varieties of crops including sugar beets and potatoes, maize, peanuts, chickpeas, and cotton.
  • They cause both direct and indirect injuries to growing crops;
  • Direct injury is caused by the insects eating the leaves, reducing photosynthetic activities and burrowing holes in shoots and stems which reduce nutrients and water translocation, burrowing holes in fruits and or/roots, thereby reducing product quality.
[31,32,33]
Small vertebrates:
Field mice, house mice, rats, feral cats, bats, foxes, wild dogs, pigs, rabbits, snakes, dogs, pigeons.
Varieties of crops including potatoes, grains, sugar beets, citrus and succulent fruits, peaches, plums, pears, strawberries, grapes, potatoes, and carrots.
  • Bites on fruits and root crops are entryways for pathogenic and spoilage microorganisms;
  • Their excrement and urine on food could be a source of poisoning and diseases;
  • Cause injury and kill cultivated plants.
[34,35,36]
Fungi
Pythium and Phytophthora infestans (Fungal-like organisms), Fusarium spp., Fusarium graminearum, Fusarium oxysporum, Rhizoctonia solani, Tilletia spp., Plasmopara viticola, Puccinia graminis var. tritici, Gaeumannomyces graminis var. tritici, Blumeria graminis, Mycosphaerella graminicola, Botrytis cinerea, Ascosphaera spp., Ustilago maydis, Aspergillus spp., Magnaporthe oryzae, Puccinia spp., Colletotrichum spp., Sclerotinia sclerotiorum, Verticillium dahlia, Armillaria spp., Melampsora lini, Phakopsora pachyrhizi Blumeria graminis.
Variety of crops including grains, rice, wheat, sorghum, potatoes, cassava, tomatoes, bananas, cucumber, grapes, strawberries, coffee, cacao, spices, mangos, and several nuts.
  • Fungi play converse roles of being pathogenic agents of diseases and some species are used as biopesticides;
  • Destruction of mature and senescent tissues of dicots including stored grains;
  • The growth of fungi on crops, fruits and tubers causes spoilage in the form of seedling damping-off, chlorosis, wilts, rots, rust, brown spots, black spots, smuts, mildew and dusty powder.
[37,38,39]
Protozoa:
Phytomonas leptovasorum, Phytomonas stahelii, Phytomonas françai, Phytomonas serpens
Variety of crops including, coffee beans, Coconut palm, oil palms, cassava, tomatoes.
  • These protists can cause phloem necrosis, hampering the transport and distribution of organic nutrients in plants. Visible symptoms are yellowing, drooping leaves, foliage desiccation, root die-back and eventual death of the tree.
[40,41]
Bacteria:
Xanthomonas campestris, Psuedomonas syringae, Pseudomonas corrugate, Clavibacter michiganensis, Pseudomonas spp., Erwinia spp., Ralstonia solanacearum, Rhizomonas suberifaciens, Erwinia carotovora, Agrobacterium tumefaciens, Spiroplasma citri
Variety of crops including, lettuce, cucurbits, cucumber, pumpkin, melon, tomatoes, chilli, potatoes, eggplant, rice, beans.
  • Bacterial infection occurs at points of injury on plants and is aided by free water;
  • They cause a variety of symptoms including, scabs, blights, soft rot, lesions, wilt, cankers, discolouration, disagreeable odours, and necrosis on infected areas.
[42,43]
Viruses:
Cassava mosaic begomovirus, Citrus Tristeza closterovirus, Barley yellow dwarf luteovirus, Plum pox potyvirus, Potato leafroll polerovirus, Cacao swollen shoot badnavirus, Potato potexvirus X, Tobacco mosaic tobamovirus, Turnip crinkle carmovirus, Tomato spotted wilt virus, Cucumber green mottle mosaic virus, Pepino mosaic virus
Variety of crops including, cassava, citrus fruits, barley, cucumber, lettuce, tomatoes, peppers
  • Viruses can initially be symptomless, or their symptoms may resemble other diseases of known aetiology making it difficult to differentiate and treat;
  • Some observable symptoms include chlorosis, mosaic leaf pattern, crinkled leaves, yellowed leaves.
[44,45,46]
Table 2. Notable examples of pests controlled by synthetic pesticides.
Table 2. Notable examples of pests controlled by synthetic pesticides.
Chemical GroupTrade NamePest ControlledPlant ProtectedToxicityReference
Carbamates
  AldicarbTemix, Standak, NamexEffective against: thrips, aphids, nematodesCotton, potatoes, soya beansAcute, environmental[60]
  FenobucarbWardam, Knock, BioStadtEffective against: hoppersRice field, cottonIrritant, environmental[61]
  CarbofuranFuradan, Curaterr, CarbosipEffective against: Mites, nematodesCorn, soybeans, potatoesAcute, environmental[62]
Organochlorine
  AldrinAldrec, Altox, OctaleneEffective against: Termite, weevil, hoppersCorn, potatoesAcute, health, environmental[63]
  LindaneAgrocide, benexane, IsotoxEffective against: Beetles, ants, locustCorn, rice, seedsIrritant, acute, health, environmental
  DDTAnofex, Cezarex, dicophaneEffective against: Armyworms, mites, soil larvaeCowpea, cottonIrritant, acute, health, environmental[64]
  ChlorothalonilBravo, Daconil, NopcocideEffective against: Mold, mildew, algaeVegetables, trees, ornamental cropsCorrosive, irritant, acute, health, environmental[65]
Organophosphates
  ChlorpyrifosBrodan, Scout, NufosEffective against: Borers, hoppers, termitesApples, soybeans, broccoliAcute, environmental[66]
  DichlorvosVapona, DiclogreenEffective against: Beetles, aphids, larvaeGrains, mushrooms, citrusIrritant, acute, environmental[67]
  DiazinonBasudin, Gardentox DazzelEffective against: Mites, aphis, wormsVegetables, nuts, fruit treesIrritant, environmental[68]
Triazines
  AtrazineAatrex, Fenamin, ProzineEffective against: Weeds, corn, sugarcaneIrritant, health, environmental[69]
  SimazineAquazine, Primatol, SimadexEffective against: WeedsSeedlingsHealth, environmental[70]
Pyrethroids
  CypermethrinAmmo, Basathrin, ArrivoSpiders, scorpions, bugsLettuce, cotton, cowpeaIrritant, environmental[71]
  FenvalerateSumicidin, Devifen, PydrinMites, tobacco budwormsCotton lettuceIrritant, acute, environmental[72]
  DeltamethrinDecis, Kordon, SadethrinMillipedes, fleas, silverfishStrawberries, ornamental gardensAcute, environmental[73]
Phenoxy-derivatives
  2,4-DHi-Dep, Weedar 64, Weed RHAPDandelions, clover, and chickweedVariety of plantsCorrosive, irritant, environmental[74]
  2,4,5-TDacamine, Inverton 245, ForronBroadleaf weedsMonocotyledonous plantsIrritant, environmental[75]
  CarbofuranFuradan, Curaterr, CarbosipMites, nematodesCorn, soybeans, potatoesAcute, environmental[76]
Dipyridyl-derivatives
  ParaquatGramoxone, Helmquat, FirestormWeedsSeedlingsCorrosive, irritant, acute, health, environmental[77]
Glycine derivatives
  GlufosinateBasta, Rely, IgniteWeedsSeedlingsIrritant, health[78]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fenibo, E.O.; Ijoma, G.N.; Nurmahomed, W.; Matambo, T. The Potential and Green Chemistry Attributes of Biopesticides for Sustainable Agriculture. Sustainability 2022, 14, 14417. https://doi.org/10.3390/su142114417

AMA Style

Fenibo EO, Ijoma GN, Nurmahomed W, Matambo T. The Potential and Green Chemistry Attributes of Biopesticides for Sustainable Agriculture. Sustainability. 2022; 14(21):14417. https://doi.org/10.3390/su142114417

Chicago/Turabian Style

Fenibo, Emmanuel O., Grace N. Ijoma, Weiz Nurmahomed, and Tonderayi Matambo. 2022. "The Potential and Green Chemistry Attributes of Biopesticides for Sustainable Agriculture" Sustainability 14, no. 21: 14417. https://doi.org/10.3390/su142114417

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