Next Article in Journal
Impact of Thickness of Pd/Cu Membrane on Performance of Biogas Dry Reforming Membrane Reactor Utilizing Ni/Cr Catalyst
Previous Article in Journal
A Review of Catalyst Integration in Hydrothermal Gasification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advancements in Synthetic Biology for Enhancing Cyanobacterial Capabilities in Sustainable Plastic Production: A Green Horizon Perspective

1
Department of Biology and Microbiology, South Dakota State University, Brookings, SD 57007, USA
2
Department of Mechanical Engineering, J. J. Lohr College of Engineering/Brookings, South Dakota State University, Brookings, SD 57007, USA
3
Department of Agronomy, Abdul Wali Khan University Mardan, Mardan 23200, Pakistan
4
Department of Natural Sciences, Lebanese American University, Byblos P.O. Box 13-5053, Lebanon
5
College of Life Science, Linyi University, Linyi 276000, China
*
Authors to whom correspondence should be addressed.
Fuels 2024, 5(3), 394-438; https://doi.org/10.3390/fuels5030023
Submission received: 20 May 2024 / Revised: 26 June 2024 / Accepted: 1 August 2024 / Published: 27 August 2024

Abstract

:
This comprehensive review investigates the potential of cyanobacteria, particularly nitrogen-fixing strains, in addressing global challenges pertaining to plastic pollution and carbon emissions. By analyzing the distinctive characteristics of cyanobacteria, including their minimal growth requirements, high photosynthetic efficiency, and rapid growth rates, this study elucidates their crucial role in transforming carbon sequestration, biofuel generation, and biodegradable plastic production. The investigation emphasizes cyanobacteria’s efficiency in photosynthesis, positioning them as optimal candidates for cost-effective bioplastic production with minimized land usage. Furthermore, the study explores their unconventional yet promising utilization in biodiesel production, mitigating environmental concerns such as sulfur emissions and the presence of aromatic hydrocarbons. The resulting biodiesel exhibits significant combustion potential, establishing cyanobacteria as a viable option for sustainable biofuel production. Through a comprehensive assessment of both achievements and challenges encountered during the commercialization process, this review offers valuable insights into the diverse contributions of cyanobacteria. Its objective is to provide guidance to researchers, policymakers, and industries interested in harnessing bio-inspired approaches for structural and sustainable applications, thereby advancing global efforts towards environmentally conscious plastic and biofuel production.

1. Introduction

The escalating levels of toxicants in the atmosphere, coupled with the excessive use of fossil fuels and plastic derivatives, have led to an unprecedented increase in the carbon footprint [1]. The persistent accumulation of non-degradable plastics, discarded in landfills, poses a significant environmental challenge with implications lasting over the next century [2]. This scenario has given rise to new threats to the Earth’s ecosystems and the survival of living organisms [3]. Rapid climate change, exacerbated by the proliferation of billions of tons of plastic products, poses a serious global environmental hazard, impacting marine ecosystems and giving rise to phenomena such as the “plastisphere” [4,5]. The marine ecosystem faces severe consequences due to the accumulation of plastic debris, leading to the fragmentation of plastics into microplastics that infiltrate the food chain [6]. Additionally, the ongoing COVID-19 situation has contributed to an uncalculated surge in single-use plastics, adding to the environmental challenges. The depletion of fossil fuels and the ensuing energy crisis pose a potential misbalance in the world economy and progress, adding further challenges to achieving environmental goals [7,8]. The plastic industry, consuming approximately 6% of the world’s oil, is projected to grow up to 20% by 2050, highlighting the urgent need for sustainable alternatives [5,9]. Current assessments indicate a two-fold increase in global plastic production over the next 20 years, accompanied by substantial greenhouse gas emissions. The combustion of fossil fuels has led to a rise in atmospheric CO2 levels, contributing to glacial mass loss and rising sea levels. The need to identify sustainable substrates that fulfill both energy demands and contribute to biodegradable plastics has never been more crucial [10,11,12,13]. Calculations for achieving carbon neutrality emphasize the necessity of reducing fossil fuel usage by 6–7 percent annually until 2030 [14]. Fossil fuel consumption, both for energy and commercial resources, incurs significant environmental and economic costs. This has spurred widespread interest in biofuels and bioplastics as sustainable alternatives [15,16]. While biofuels and bioplastics have been successfully produced from terrestrial plants, their cultivation competes for valuable agricultural land [17]. Cyanobacteria offer a promising alternative for bioplastic production and offer advantages over traditional bacteria and yeast. Their ability to use sunlight to generate energy and convert atmospheric CO2 directly into organic compounds makes them more energy-efficient and environmentally friendly. Cyanobacteria can be genetically engineered to produce various bioplastics, require minimal nutrients, and are scalable for large-scale production. In contrast to bacterial cultivation, which often involves higher costs and energy-intensive processes, cyanobacteria offer a more economical approach to the sustainable production of bioplastics [16].
Cyanobacteria, often overlooked in the context of biofuel and bioplastic production, are now gaining attention due to their potential to efficiently sequester atmospheric CO2 [18]. Common cyanobacteria species such as Anabaena and Synechocystis have demonstrated their ability to produce industrial compounds, including biofuels and bioplastics [19]. Compared to other organisms used for similar purposes, such as eukaryotic algae and land plants, cyanobacteria offer several advantages. They have a high specific growth rate, can thrive in adverse conditions such as salt water and barren land, and have the ability to fix atmospheric nitrogen, reducing the need for nitrogen fertilizers. In addition, cyanobacteria are easier to genetically manipulate, allowing more efficient optimization of metabolic pathways for the production of biofuels and bioplastics. Their abundant fatty acid and oil content, as well as other active metabolites, make them an excellent alternative for the sustainable production of these industrial compounds. The biofixation efficiency of cyanobacteria is estimated to be approximately 10-fold higher than that of terrestrial plants [20,21]. The dynamic metabolic versatility of cyanobacteria presents an opportunity to overcome challenges associated with biofuel and bioplastic production [22]. Genetic engineering and targeted genetic modulations are being explored to enhance cyanobacterial strains for increased yield [23]. This review paper aims to shed light on the potential of cyanobacteria as a green sustainable methodology for biofuel and bioplastic production. It explores the genetic engineering approaches to enhance their yield and addresses the scope and constraints related to large-scale industrial production. The study underscores the urgent need to replace fossil fuel substrates with environmentally friendly alternatives and presents cyanobacteria as a promising step toward sustainability in biofuels and bioplastics.

2. Synthetic Biology Strategies for Enhancing Cyanobacterial Plastic Production

Synthetic biology methodologies have played a pivotal role in propelling the capacity of cyanobacteria for sustainable plastic production [24]. Through the utilization of techniques grounded in genetic manipulation and metabolic engineering, investigators have endeavored to refine cyanobacterial metabolic pathways to augment the synthesis of plastic precursors [25]. The application of synthetic biology tools, exemplified by CRISPR-Cas9-mediated genome editing and homologous recombination, facilitates meticulous manipulation of cyanobacterial genetic material [26]. By integrating exogenous genetic elements encoding plastic biosynthesis enzymes and intricately adjusting endogenous regulatory mechanisms, cyanobacterial strains exhibiting augmented plastic production capacities have been engineered [27]. Synthetic biology methodologies, exemplified by CRISPR-Cas9-mediated genome editing and homologous recombination, play a pivotal role in the genetic manipulation of cyanobacteria [28]. Metabolic engineering strategies are geared towards orchestrating cyanobacterial metabolism to facilitate the biosynthesis of plastic precursors [29]. This entails precise modulation of enzymatic activities and pathway flux to enhance the generation of key carbon intermediates crucial for polymer synthesis, notably acetyl-CoA and malonyl-CoA [30]. Moreover, the optimization of growth conditions, encompassing parameters such as light intensity, carbon source availability, and nutrient composition, serves to further bolster the accumulation of plastic precursors within cyanobacterial cells [31]. Pathway engineering assumes a central role in the meticulous design of customized biosynthetic pathways optimized for proficient polymer synthesis [32]. Through the deliberate engineering of synthetic gene circuits and genetic constructs, the orchestration of enzymatic activities and metabolic flux is achieved, thereby establishing a foundational framework for the intricate machinery of cellular plastic production [33]. Despite significant advancements, persistent challenges encompass heterologous pathway expression and metabolic burden [34]. Future research endeavors are directed toward surmounting these obstacles through the utilization of sophisticated synthetic biology tools and a comprehensive comprehension of cyanobacterial physiology [35]. This collaborative endeavor is directed towards fully harnessing the inherent potential of cyanobacteria as eco-friendly platforms for plastic synthesis [36].

3. Role of Cyanobacteria in Carbon Sequestration

Cyanobacteria, commonly known as blue-green algae, assume a pivotal role in carbon sequestration, a process crucial for mitigating climate change by capturing and storing carbon dioxide (CO2) from the atmosphere [37,38,39]. Synthetic biology has shown great potential in artificially constructing new CO2 assimilation pathways, and important research progress has been made in improving the carboxylation activity of the Rubisco enzyme, introducing a CO2 concentration mechanism, reducing carbon loss, and reducing photorespiration (Figure 1). These microscopic photosynthetic organisms engage in oxygenic photosynthesis, harnessing sunlight to convert carbon dioxide and water into organic compounds such as carbohydrates while releasing oxygen [40]. This photosynthetic activity enables cyanobacteria to act as primary producers and create a carbon sink in the form of biomass. Carbon fixation, a process wherein atmospheric CO2 is incorporated into organic molecules, is a fundamental step in their carbon sequestration mechanism [41]. Cyanobacteria often form dense microbial mats and biofilms in aquatic environments, providing protective niches where organic material accumulates and becomes buried in sediment layers [42]. Some cyanobacteria contribute to carbonate precipitation, forming calcium carbonate (CaCO3), which aids in long-term carbon storage [43]. Additionally, the capacity of cyanobacteria to induce algal blooms and subsequent sedimentation enhances the sequestration of carbon in both aquatic and terrestrial ecosystems [44]. A large-scale study across the Loess Plateau in China found that the total ecosystem carbon stock was 2.8 Pg, with 30% stored in soil (0–20 cm), 53% in above-ground biomass, and 17% in below-ground biomass [45]. This study demonstrates the significant carbon sequestration potential of large ecosystems. In symbiotic relationships with plants, cyanobacteria further enrich soils with organic matter, promoting carbon storage [46]. The global impact of cyanobacteria in regulating atmospheric CO2 levels underscores their significance in addressing climate change through effective carbon sequestration mechanisms [47].

3.1. Cyanobacteria as Photosynthetic Prokaryotes

Cyanobacteria, classified as photosynthetic prokaryotes, represent a unique group of microorganisms with a remarkable capacity for oxygenic photosynthesis [48]. Unlike eukaryotic algae and plants, cyanobacteria lack membrane-bound organelles, such as chloroplasts, and their genetic material is housed in a nucleoid region within the cell [49]. Nonetheless, they share the fundamental photosynthetic machinery with higher plants [50]. These microorganisms harness sunlight to drive the conversion of carbon dioxide and water into organic compounds, generating energy-rich molecules such as carbohydrates [51]. The photosynthetic pigments, including chlorophyll-a, phycocyanin, and phycoerythrin, allow cyanobacteria to capture light across a broad spectrum, enabling them to thrive in various environments, from freshwater to marine ecosystems [52]. The thylakoid membranes within their cells serve as the sites for the light-dependent reactions of photosynthesis [53]. Cyanobacteria exhibit a unique ability among prokaryotes to produce oxygen during photosynthesis, a process that played a crucial role in shaping Earth’s early atmosphere [54]. Their evolutionary significance is underscored by their role as primary producers, contributing substantially to global carbon fixation and the oxygenation of the atmosphere [55]. Moreover, cyanobacteria form diverse morphologies, ranging from unicellular to filamentous and colonial forms, allowing them to adapt to a wide range of ecological niches [56]. In addition to their role in carbon cycling, some cyanobacteria possess nitrogen-fixing capabilities, further influencing nutrient dynamics in ecosystems [46]. Overall, cyanobacteria’s role as photosynthetic prokaryotes is integral to ecological processes, with implications for global biogeochemical cycles and the understanding of the evolution of photosynthesis in the context of early life on Earth [57].

3.2. Unique Attributes: Minimal Growth Requirements, High Photosynthetic Efficiency, and Rapid Growth Rates

Cyanobacteria possess a suite of unique attributes that collectively contribute to their ecological success and environmental impact [58]. Their minimal growth requirements set them apart, allowing them to thrive in environments with limited nutrients [59]. Cyanobacteria can colonize diverse habitats, ranging from nutrient-poor soils to oligotrophic aquatic ecosystems, showcasing their adaptability to varying ecological conditions [60]. This adaptability is partly attributed to their ability to fix atmospheric nitrogen, a crucial nutrient often limiting in terrestrial and aquatic ecosystems [61]. By converting atmospheric nitrogen into forms usable by other organisms, cyanobacteria play a pivotal role in nutrient cycling and ecosystem fertility. The high photosynthetic efficiency of cyanobacteria is another distinctive feature that enhances their ecological significance [46,62]. Their photosynthetic pigments, particularly chlorophyll-a along with phycocyanin and phycoerythrin, allow them to absorb light efficiently across a broad spectrum. This adaptation enables cyanobacteria to harness solar energy effectively and convert it into chemical energy through oxygenic photosynthesis [63]. This process not only supports their own growth but also contributes to the primary production of organic compounds in ecosystems, forming the basis of food webs [64]. Rapid growth rates further characterize cyanobacteria, and this ability has both ecological implications and applications. In recent years, with the rapid growth of high-throughput sequencing technology, more and more cyanobacterial genomes have been sequenced and analyzed, which has made the overall understanding of cyanobacterial genome size more comprehensive and accurate [65,66]. From a scientific standpoint, there exists a dearth of research regarding the copy number of cyanobacterial genomes in comparison to other areas of study. The relevant studies on cyanobacterial genome copies that have been reported so far are summarized in (Table 1). In favorable conditions, cyanobacteria can undergo exponential growth, leading to the formation of visible algal blooms [67]. While such blooms can have detrimental effects on water quality and ecosystems, the rapid growth of cyanobacteria can also be harnessed for beneficial purposes [68]. Researchers explore the potential of cyanobacteria in biotechnological applications, such as biofuel production and wastewater treatment, leveraging their ability to proliferate rapidly under controlled conditions [69].
The intricate interplay of minimal growth requirements, high photosynthetic efficiency, and rapid growth rates underscores the ecological versatility of cyanobacteria [70]. Their contributions to nutrient cycling, carbon sequestration, and their ability to respond dynamically to environmental changes make them integral components of ecosystems worldwide. Moreover, understanding these unique attributes offers insights into the evolutionary success of cyanobacteria and their pivotal role in shaping the dynamics of the biosphere [71,72].
Table 1. Exploring ploidy levels and key parameters in cyanobacterial species.
Table 1. Exploring ploidy levels and key parameters in cyanobacterial species.
SpeciesTemperature (°C)Time (h)Genome Size (Mb)Genome Copy No.References
Anabaena cylindrica2818.725[73]
Anabaena variabilis307.15–8[74]
1450/10 Microcystis aeruginosa28stat.ph. a1–10[75]
Anabaena sp. Strain PCC 7120287.28.2[76]
Prochlorococcus1.7[77]
Synechococcus elongatus PCC 794229262.73.7/3.4[78]
Synechococcus elongatus PCC 6301)3510 to >502.92–7 to >1–2 d[79]
Synechococcus sp. (strain WH7803)262.53.7 e[78]
Synechococcus sp. Strain WH7803232.32–6 c[80]
Parasynechococcus subtropicalis WH 780526172.51 e[80]
Synechococcus sp. WH 810127153.31 e[81]
Synechocystis sp. PCC 6803 23233.7215/56/59 c[78]
Synechocystis sp. PCC 6803 22223.8143/46/47 c[78]
Synechocystis sp. PCC 6803 3115–243.916[82]
Trichodesmium erythraeumIMS10126LDC b7.77700[83]
a Stationary stage; b blight-dark period; c Log growth phase/linear phase/stationary phase; d Longer doubling time correlates with reduced genome copy number. e. Derived from comparing the fluorescent chromosome band and genome size of PCC 6301.

3.3. Lipid Content in Thylakoid Membranes and Its Significance

The lipid content within the thylakoid membranes of cyanobacteria represents a highly nuanced and critical aspect of their cellular architecture and functional dynamics [84]. Thylakoid membranes, situated within the chloroplasts of cyanobacteria, serve as the primary site for the execution of photosynthesis, a complex process involving the conversion of light energy into chemical energy. The diverse array of lipids presents in these membranes, including phospholipids, glycolipids, and sulfolipids, collaboratively contributes to the structural stability, organization, and adaptability of the photosynthetic machinery [85]. Phospholipids, with phosphatidylglycerol being a prominent constituent, are instrumental in anchoring and stabilizing photosynthetic protein complexes within the thylakoid membrane [86]. This lipid matrix ensures the ordered arrangement of components such as light-harvesting complexes and reaction centers, creating an optimal environment for the efficient flow of electrons during the light-dependent reactions of photosynthesis [87]. The spatial arrangement of these lipids, along with their interaction with protein complexes, is finely tuned to achieve the intricate dance of energy transfer within the thylakoid membrane [88]. Glycolipids, specifically monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG), contribute to the structural dynamics of thylakoid membranes and play a pivotal role in the organization of photosynthetic pigments [89]. These glycolipids encapsulate chlorophyll molecules, creating a specialized microenvironment that facilitates light absorption and energy transfer. The fluidity and flexibility of thylakoid membranes, crucial for their adaptability to changing environmental conditions, are influenced by the composition and arrangement of these glycolipids. This flexibility is vital for the optimization of photosynthetic efficiency, allowing cyanobacteria to acclimate to varying light intensities and other environmental factors [90,91,92].
Beyond their structural roles, the lipids within thylakoid membranes contribute to the resilience of cyanobacteria in the face of environmental stresses [93]. The presence of polyunsaturated fatty acids in these lipids acts as a buffer against oxidative stress generated during photosynthesis [94]. Cyanobacteria, often exposed to high light intensities and fluctuating temperatures, employ these lipids as a form of protection against reactive oxygen species [95]. The unsaturation of fatty acids enhances the fluidity and integrity of thylakoid membranes, contributing to the overall stress tolerance of cyanobacteria [96,97]. The complex relationship between the lipid makeup of thylakoid membranes and the physiological reactions of cyanobacteria highlights the evolutionary adaptations these organisms have undergone over extended periods [98]. This understanding not only enriches our comprehension of fundamental cellular processes but also has practical implications [99]. Manipulating lipid metabolism within thylakoid membranes holds potential for biotechnological applications, such as engineering cyanobacteria for enhanced biofuel production, stress resistance, or other tailored functionalities. Fundamentally, the lipid composition of thylakoid membranes plays a crucial role in governing the comprehensive operation and adaptability of cyanobacteria, thereby influencing their ecological prowess and prospective applications in biotechnology [28,100].

4. Cyanobacteria’s Contribution to Biofuel Production

Cyanobacteria’s contribution to biofuel production is a field of burgeoning research and innovation, presenting a sustainable approach to meet the global demand for renewable energy [101]. These photosynthetic microorganisms possess distinctive characteristics that make them particularly attractive for biofuel applications [102]. One of the key features is their high lipid content, especially triacylglycerols (TAGs), which serve as valuable precursors for biodiesel production [103]. Researchers have been exploring ways to enhance lipid accumulation in cyanobacteria by adjusting growth conditions, nutrient availability, and light exposure [104]. Genetic engineering techniques are also being employed to tailor cyanobacterial strains for increased biofuel productivity. In addition to biodiesel, cyanobacteria exhibit the potential for biohydrogen production through a process known as hydrogen photoproduction [105]. This process involves the direct conversion of solar energy into molecular hydrogen, offering a clean and sustainable alternative to traditional hydrogen production methods [106]. Cyanobacteria capable of biohydrogen production, such as certain strains of Anabaena and Synechocystis, are being investigated for their feasibility in large-scale biofuel applications [107]. Cyanobacteria’s adaptability to diverse environmental conditions further enhances their appeal for biofuel production. These microorganisms can thrive in various ecosystems, including saline environments and wastewater, demonstrating their potential to utilize marginal lands unsuitable for conventional agriculture [108,109]. Their relatively rapid growth rates and ability to fix atmospheric carbon dioxide contribute to their efficiency in converting solar energy and carbon into biomass, laying the groundwork for sustainable biofuel feedstock production [110]. Ongoing research aims to optimize cyanobacterial cultivation techniques, exploring novel photobioreactor designs, nutrient management strategies, and synthetic biology approaches. By fine-tuning these parameters, scientists seek to maximize biofuel yields and establish economically viable and environmentally friendly production systems. Moreover, advancements in genetic engineering offer the potential to tailor cyanobacteria for specific biofuel traits, enhancing their overall productivity and resilience [111,112].
As the global community amplifies endeavors toward transitioning to renewable energy sources, the utilization of cyanobacteria in biofuel production emerges as a promising prospect. The sustainable and carbon-neutral attributes of cyanobacterial biofuels position them as pivotal entities in the pursuit of cleaner energy solutions. Despite persistent obstacles, ongoing scientific investigations and advancements in technology aimed at exploiting the potential of cyanobacteria emphasize their significance in influencing the course of biofuel production and promoting a more environmentally sustainable energy landscape.

4.1. Overview of Biodiesel Production from Cyanobacteria

Biodiesel production from cyanobacteria represents a multifaceted and dynamic process that integrates various scientific disciplines, from microbiology and biochemistry to engineering [113]. The journey begins with the careful cultivation of cyanobacterial strains under controlled conditions to optimize their growth and lipid accumulation. This involves fine-tuning parameters such as light intensity, temperature, nutrient availability, and carbon dioxide concentration [114]. However, microalgae production is generally expensive compared to crops, photosynthetic growth requires light, carbon dioxide, water, and inorganic salts. The temperature is generally maintained between 20 and 30 °C. To minimize costs, biodiesel production can ignore seasonal and day-night changes in light and rely on natural light. In the culture process of microalgae, the inorganic elements required for their cell growth include nitrogen (N), phosphorus (P), iron (Fe), and in some cases the appropriate amount of silicon (Si), which must also be supplemented. The minimum requirements of these elements can be determined by the approximate molecular formula expressed by microalgae, namely CO0.48 H1.83 N0.11 P0.01. The formula is based on Grobelaar’s calculations [115]. Nutrients such as phosphorus, must be supplied in excess because the absorption of certain metal ions is coupled with phosphate ion absorption, and therefore, not all supplemental phosphorus is available to organisms. Grima et al. [116] successfully cultured microalgae using seawater supplemented with commercial nitrate, phosphate fertilizers, and several other nutrients, and seawater has become an increasingly common medium for marine microalgae. The theoretical production process of biodiesel from microalgae is shown in [Figure 2] [117]. Once the cyanobacterial biomass reaches an optimum lipid content, the next critical step is lipid extraction. Various methods are employed for this purpose, each with its own set of advantages and challenges [118]. Traditional solvent extraction methods involve the use of organic solvents to dissolve lipids, while mechanical pressing physically separates lipids from the biomass [119]. Additionally, emerging technologies such as supercritical fluid extraction are being explored for their potential to provide more efficient and environmentally friendly lipid extraction. Following lipid extraction, the isolated triacylglycerols (TAGs) undergo transesterification, a chemical process that transforms them into biodiesel [120]. In this process, the TAGs react with an alcohol, usually methanol or ethanol, in the presence of a catalyst. The result is the production of biodiesel and glycerol as a byproduct [121]. Catalyst selection, reaction conditions, and the overall efficiency of the transesterification process are crucial factors that influence the quality and yield of biodiesel [122]. The advantages of using cyanobacteria for biodiesel production are multifaceted. Their photosynthetic efficiency allows them to directly convert sunlight and carbon dioxide into lipids, making them highly efficient biofuel producers [123]. Furthermore, their ability to thrive in diverse environments, including non-arable land and wastewater, mitigates concerns related to competition with food crops for cultivation space. This aligns with the broader goal of developing sustainable and environmentally friendly biofuel feedstocks [109,124].
Despite the favorable characteristics displayed in cyanobacterial biodiesel production, persistent challenges persist. Ongoing scientific endeavors are focused on optimizing cultivation parameters, refining lipid extraction methodologies, and tackling economic and scalability hurdles. Additionally, researchers are exploring co-cultivation strategies, nutrient recycling techniques, and potential synergies with other biotechnological processes to augment efficiency and sustainability. As advancements continue, cyanobacterial biodiesel production underscores the inventive exploitation of natural reservoirs for sustainable energy solutions. The integration of biological research, engineering, and environmental science in this domain offers prospects for substantially advancing the global quest for renewable and environmentally sound energy alternatives.

4.2. Mitigation of Toxic Sulfur Release

The mitigation of toxic sulfur release in the context of biodiesel production from cyanobacteria is a multifaceted challenge that demands careful consideration and strategic intervention throughout the production process [125]. During transesterification, a critical step in biodiesel production, the reaction between triacylglycerols (TAGs) and alcohol is facilitated by catalysts, often sodium hydroxide or potassium hydroxide [126]. However, the potential release of toxic sulfur-containing compounds, stemming from impurities in the catalyst or contaminants in the starting materials, poses environmental and product quality concerns [127]. To tackle these challenges, scientists are vigorously investigating catalyst-centric approaches. Primarily, the identification of top-tier catalysts with low sulfur content is paramount [128]. Furthermore, rigorous purification processes are implemented to ensure that the catalysts used in transesterification are free from sulfur impurities. Investigating alternative catalysts, such as solid catalysts or enzymes, which inherently possess lower sulfur content, is also gaining traction [129]. These approaches collectively aim to reduce the introduction of sulfur into the biodiesel synthesis process, promoting cleaner and more sustainable production. Beyond catalyst-related strategies, researchers are examining the influence of feedstock selection and cultivation conditions on sulfur content [130]. The choice of cyanobacterial strains and optimization of growth conditions can impact the lipid composition and, consequently, the sulfur content in the resulting biodiesel [131]. Post-transesterification processes are integral to further purify biodiesel and diminish sulfur impurities. Washing and distillation are among the techniques employed to enhance the overall quality of biodiesel [132]. These purification steps contribute to meeting stringent fuel standards, ensuring that the final biodiesel product is environmentally compliant and suitable for widespread use [133].
The challenge of mitigating toxic sulfur release is not only a technical concern but also a regulatory and environmental imperative. Researchers and engineers are actively engaged in refining these processes to align with global biodiesel quality standards and environmental regulations [134]. Continuous advancements in alternative catalysts, feedstock optimization, and post-transesterification purification techniques underscore the commitment to producing high-quality biodiesel from cyanobacteria while minimizing the potential negative impacts associated with sulfur-containing compounds [135]. As the scientific discipline advances, addressing the emission of toxic sulfur emerges as a pivotal component within the overarching goal of promoting cyanobacterial biodiesel as a viable, eco-conscious alternative. Employing a comprehensive strategy encompassing catalyst optimization, efficient feedstock handling, and rigorous post-production refining, the biodiesel sector endeavors to foster a greener energy paradigm while diminishing reliance on conventional fossil fuel sources.

4.3. Reduction of Aromatic Hydrocarbons

The reduction of aromatic hydrocarbons is a key consideration in the production of biodiesel from cyanobacteria, particularly as it pertains to improving the fuel’s combustion characteristics and meeting stringent environmental standards [136]. Aromatic hydrocarbons, such as benzene, toluene, ethylbenzene, and xylene (BTEX), are undesirable components due to their association with toxicity, low cetane numbers, and elevated emissions during combustion [137]. Addressing the presence of aromatic hydrocarbons in biodiesel derived from cyanobacteria involves strategic interventions throughout the production process [123]. Cyanobacterial strains and cultivation conditions significantly influence the composition of lipids, including the fatty acid profile in the resulting biodiesel. Researchers are actively exploring the manipulation of these factors to optimize the fuel’s properties [138]. By selecting cyanobacterial strains with lower aromatic hydrocarbon content and fine-tuning growth conditions, efforts are underway to mitigate the incorporation of undesirable components into the biodiesel matrix [139]. The transesterification process itself can be tailored to minimize the formation of aromatic hydrocarbons [140]. The choice of alcohol for transesterification, such as methanol or ethanol, can impact the final composition of the biodiesel, and optimizing this parameter is part of the ongoing research efforts [141]. Post-transesterification purification procedures, including fractional distillation, liquid-liquid extraction, and adsorption, are pivotal for diminishing aromatic hydrocarbons in biodiesel, thereby enhancing its combustion attributes and environmental suitability through the selective elimination of undesired constituents [142]. For a variety of nitroaromatic hydrocarbons and their halogenated derivatives with different structures, microorganisms can evolve corresponding metabolic pathways in a short period, and show strong adaptability and diversity of metabolic capacity. Compared with the non-specific metabolism and transformation of fungi and anaerobic bacteria, bacteria can often use nitroaromatic compounds as the only carbon source, nitrogen source, and energy growth, which provides greater advantages in biological management and bioreaction of pollutants. Microbiologists have isolated many bacteria capable of degrading nitro-aromatic pollutants (Table 2).
Furthermore, there is ongoing exploration into enhancing the quality of lipids in cyanobacteria through feedstock engineering and genetic modification techniques, with a specific focus on reducing aromatic hydrocarbon levels [143]. This endeavor is driven by the goal of aligning cyanobacteria-derived biodiesel with the stringent requirements of contemporary combustion engines and environmental mandates. The endeavor to decrease aromatic hydrocarbons in cyanobacteria-derived biodiesel is not only a technical pursuit but also a crucial element in promoting sustainable and cleaner energy alternatives [123]. By integrating advancements in strain selection, cultivation optimization, transesterification processes, and post-production purification methods, researchers and engineers are striving to produce biodiesel of superior quality that not only meets industry standards but also contributes to a more eco-friendly energy landscape [144]. These evolving strategies highlight a dedicated effort to tackle the challenges associated with aromatic hydrocarbons and advance the feasibility of cyanobacteria-derived biodiesel within the broader framework of renewable energy solutions [145].

4.4. Combustion Potential and Sustainability

The combustion potential and sustainability of biodiesel produced from cyanobacteria are critical aspects that determine its viability as a renewable energy source [123]. Understanding the combustion characteristics and evaluating the overall sustainability of cyanobacterial biodiesel involves a comprehensive analysis of its chemical composition, combustion efficiency, and environmental impact [146]. Cyanobacteria-derived biodiesel exhibits favorable combustion properties owing to its composition of fatty acid methyl esters (FAMEs), primarily triacylglycerols (TAGs). The absence of sulfur, lower levels of aromatics, and higher cetane numbers contribute to improved combustion efficiency and reduced emissions compared to traditional fossil fuels [147]. The higher cetane number signifies better ignition quality, leading to smoother combustion and lower emissions of nitrogen oxides (NOx). The sustainability of cyanobacterial biodiesel extends beyond combustion efficiency to encompass environmental and social aspects [148]. Cyanobacteria offer advantages in terms of feedstock cultivation, as they can thrive in diverse environments, including non-arable land and wastewater. This reduces competition with food crops for resources, mitigating potential land-use conflicts associated with traditional biofuel feedstocks [149]. Moreover, cyanobacteria have the ability to sequester carbon dioxide during their growth, contributing to a closed carbon cycle when used as biodiesel feedstock. This carbon-neutral aspect aligns with the broader goal of mitigating greenhouse gas emissions and addressing climate change [150,151]. The potential integration of cyanobacterial cultivation in wastewater treatment processes further enhances the sustainability profile, providing a dual benefit of biofuel production and environmental remediation. To evaluate the combustion potential and sustainability comprehensively, life cycle assessments (LCAs) are employed [151,152]. LCAs consider the entire life cycle of biodiesel production, from cultivation and processing to distribution and combustion. Assessing factors such as energy input, greenhouse gas emissions, and resource use provides a holistic view of the environmental impact and sustainability of cyanobacterial biodiesel [153].
While cyanobacterial biodiesel demonstrates promising combustion potential and sustainability, ongoing research continues to refine cultivation techniques, optimize processing methods, and address challenges associated with scale-up. Integration with other biotechnological processes, such as nutrient recycling and co-cultivation strategies, further enhances the overall sustainability of cyanobacterial biodiesel production [154,155]. In conclusion, cyanobacteria-derived biodiesel exhibits favorable combustion characteristics, including higher cetane numbers and reduced emissions, making it a promising renewable energy option [147]. Its sustainability is underscored by its ability to thrive in diverse environments, carbon-neutral growth, and potential for environmental remediation. Cyanobacteria have shown success in environmental remediation, particularly in oil spill bioremediation, heavy metal removal, pesticide degradation, and wastewater treatment. Species such as Oscillatoria salina and Synechococcus sp. have been effectively used to clean up oil spills in aquatic environments. In wastewater treatment, cyanobacteria efficiently convert nutrients from dairy wastewater into biomass. While their potential for biofuel production is recognized due to their photosynthetic nature, most applications in this area are still in the research and development stage, focusing on improving efficiency and economic viability [156]. As the field advances, the continuous improvement of cultivation practices and processing methods ensures that cyanobacterial biodiesel remains a key player in the pursuit of sustainable and environmentally friendly energy alternatives [157].

5. Polyhydroxyalkanoates (PHAs) Production by Cyanobacteria

Polyhydroxyalkanoate (PHA) production by cyanobacteria represents a promising avenue within the field of bioplastics, offering a sustainable and environmentally friendly alternative to conventional plastics derived from fossil fuels [158]. The synthesis pathways for polyhydroxyalkanoates (PHAs) are real and well-established. Bacteria and archaea are the organisms that carry out the reactions to synthesize PHAs. As shown in Figure 3, there are at least 6 possible ways to synthesize PHAs from lignin depolymerization and degradation into intermediates. (1) Pyruvate is converted to propionyl-coA by lactose dehydrogenase (LDH) and propionyl-CoA transferase (PCT) [159], and then PHAs is synthesized by PHA synthetase (PhaC). (2) Conversion from SucD, 4-hydroxy-butyrate dehydrogenase (4 hbD), 4-hydroxy-butyrate CoA transferase (Cat 1, Cat 2) to 4-hydroxy-butyrate coA, Then, PHAs are synthesized by PHA synthase (PhaC) [160]. (3) Succinic acid CoA is catalyzed by methylmalonyl-CoA decarboxylase (YgfG) and keto reductase (phaB) to form hydroxyvaleryl CoA, and finally to synthesize PHAs [161]. (4) Oxaloacetic acid was catalyzed by threonine dehydrogenase (IIVA), propionic acid dehydrogenase (LDH), ketoacylthiolase (BktB), ketoreductase (phaB), and PHA synthetase to synthesize PHAs. (5) Acetyl coenzyme A is catalyzed to synthesize PHAs by ketoacylthiolysis (phaA), ketoreductase (phaB) and PHA synthetase [3]. (6) de novo synthesis pathway of fatty acids, first acetyl-CoA is converted to malonyl ACP, Then PHAs [162,163] are synthesized under the pressure of FabG, 3-hydroxyl ester-ACP-CoA transferase (phaG), ACyl CoA transferase (Alkk) and PHA synthase. PHAs are biopolymers synthesized by various microorganisms as a means of carbon storage, and cyanobacteria, being photosynthetic organisms, are particularly well-suited for this purpose. As shown in (Figure 3), lignin depolymerizes and degrades into intermediates to synthesize PHAs. Cyanobacteria can accumulate PHAs intracellularly, especially when subjected to conditions of nutrient limitation, such as nitrogen or phosphorus deprivation [164,165]. These stress conditions trigger the diversion of excess carbon fixed through photosynthesis into the production of PHAs, serving as carbon and energy storage compounds. This natural ability of cyanobacteria to synthesize PHAs makes them attractive candidates for the sustainable production of biodegradable plastics [16]. Researchers are actively engaged in optimizing cultivation conditions to enhance PHA production in cyanobacteria. This involves manipulating various factors such as light intensity, carbon dioxide availability, and nutrient levels to achieve optimal conditions for both biomass growth and PHA accumulation [166]. Additionally, genetic engineering approaches are employed to enhance the efficiency of PHAs biosynthesis in cyanobacterial strains, enabling the development of engineered strains with improved PHA production capabilities [167]. The bioplastic potential of PHAs produced by cyanobacteria lies not only in their biodegradability but also in their versatility. PHAs can be engineered to exhibit specific material properties suitable for various applications, ranging from packaging materials to medical devices [168]. The sustainable and renewable nature of cyanobacteria as the source of PHAs aligns with the growing global emphasis on reducing dependence on fossil fuels and mitigating the environmental impact of plastic waste [169]. The production of PHAs by cyanobacteria also presents an opportunity for the integration of bioplastic synthesis with other biotechnological processes. For instance, the utilization of wastewater or carbon dioxide from industrial emissions as nutrient sources for cyanobacterial cultivation enhances the overall sustainability of PHA production [170]. This approach not only addresses environmental concerns but also contributes to the circular economy by utilizing waste streams for valuable bioplastic production [171].
In summary, the synthesis of polyhydroxyalkanoates (PHAs) by cyanobacteria signifies a forefront and eco-conscious strategy toward achieving sustainable bioplastic production. The adeptness of cyanobacteria in photosynthesis, coupled with continual innovations in genetic manipulation methodologies, establishes them as pivotal entities in the advancement of PHAs for an eco-friendlier and more cyclical paradigm in plastics manufacturing. As research progresses, the potential of polyhydroxyalkanoates (PHAs) sourced from cyanobacteria to significantly impact the bioplastics industry is burgeoning, offering a promising solution to the challenges posed by conventional plastics.

5.1. Intracellular Energy Source and Carbon Sink

The fundamental physiological and ecological importance of cyanobacteria stems from their dual roles as intracellular energy generators and carbon sequesters. As photosynthetic organisms, cyanobacteria harness solar energy to transform carbon dioxide into organic molecules via the photosynthesis pathway [36,109]. This process not only serves as an energy source for the cells but also functions as a carbon sink, contributing to carbon sequestration and ecosystem dynamics [172]. The primary intracellular energy source for cyanobacteria is derived from photosynthesis, a complex biochemical process that involves the capture of sunlight by photosynthetic pigments, such as chlorophyll-a, phycocyanin, and phycoerythrin [173]. These pigments are embedded in the thylakoid membranes, where light-dependent reactions take place. [174]. This process generates carbohydrates, such as glucose, as an energy storage reservoir for the cyanobacterial cells. Simultaneously, cyanobacteria act as carbon sinks by fixing atmospheric carbon dioxide into organic compounds during photosynthesis [175]. Photosynthetic bacteria can decompose propionic acid and butyric acid into acetic acid through their own extracellular secretion of an extracellular protease under phototrophic conditions, and then use CO2 as an electron donor in the Calvin cycle to balance the redox potential of the cell, while acetic acid enters the tricarboxylic acid cycle (TCA) through the combined action of succinyl coenzyme A and glycine to synthesize ALA, the conversion pathway is shown in (Figure 4). This dual role as an energy source and carbon sink underscores their importance in global carbon cycling [55]. The incorporation of carbon into cellular biomass not only provides the necessary building blocks for the cell’s growth and metabolism but also contributes to the reduction of atmospheric carbon dioxide, playing a crucial role in mitigating greenhouse gas levels [176]. Furthermore, cyanobacteria exhibit the ability to store excess carbon in the form of glycogen or polyhydroxyalkanoates (PHAs) under certain conditions [168]. These intracellular reserves serve as energy and carbon storage pools that cyanobacteria can tap into during periods of environmental stress or nutrient limitation [177]. The storage and subsequent utilization of these reserves enhance the resilience of cyanobacteria in fluctuating environmental conditions, allowing them to thrive in various ecosystems, from freshwater to extreme environments such as deserts [56]. The comprehension of the complex equilibrium between serving as an internal provider of energy and a reservoir for carbon elucidates the adaptive tactics employed by cyanobacteria across various ecological habitats [67]. Additionally, the ability of cyanobacteria to sequester carbon holds significant ramifications for environmental sustainability, impacting the cycling of nutrients, the productivity of ecosystems, and the global carbon balance [46]. As scientists investigate the underlying molecular mechanisms dictating these phenomena, the distinctive function of cyanobacteria as pivotal agents in both energy fluxes and carbon modulation remains a compelling focus for researchers exploring the confluence of microbiology and environmental science.

5.2. Cyanobacteria’s Role in Bioplastic Production

Cyanobacteria play a significant and burgeoning role in the production of bioplastics, particularly through the synthesis of polyhydroxyalkanoates (PHAs). PHAs are biodegradable polyesters produced by various microorganisms as intracellular carbon and energy storage compounds [178]. Cyanobacteria, being photosynthetic organisms, offer a sustainable and environmentally friendly approach to bioplastic production [16]. The production of PHAs by cyanobacteria involves manipulating their metabolic pathways to divert carbon flux toward the synthesis of these biopolymers [179]. Under conditions of nutrient limitation, especially when deprived of nitrogen or phosphorus, cyanobacteria channel excess carbon, derived from photosynthesis, into the production of PHAs [180]. This serves as a natural strategy for these microorganisms to store carbon and energy for future use. Researchers are actively exploring the optimization of cultivation conditions and the genetic engineering of cyanobacterial strains to enhance PHA production [181]. Through precise adjustments of growth parameters such as light intensity, carbon dioxide concentration, and nutrient composition, researchers endeavor to establish an optimal milieu promoting both biomass proliferation and the synthesis of polyhydroxyalkanoates (PHAs) within cyanobacterial cellular structures [182]. Additionally, genetic manipulation allows the development of cyanobacterial strains with improved PHA biosynthesis capabilities, offering potential for increased bioplastic yields. The bioplastic potential of cyanobacterial PHAs lies not only in their biodegradability but also in their versatility [183]. PHAs can be engineered to exhibit specific material properties suitable for various applications, including packaging materials, agricultural films, and medical devices [184]. The sustainable and renewable nature of cyanobacteria as the source of PHAs aligns with the global imperative to reduce reliance on fossil fuels and address the environmental challenges posed by conventional plastics [185].
Moreover, cyanobacteria play a pivotal role in enhancing the sustainability of bioplastic manufacturing owing to their adaptability to various environments, including non-arable land and wastewater, thereby mitigating the competition with food crops for resources [186]. Progress in cyanobacterial bioplastic production involves ongoing investigations into refining cultivation methods, improving processing techniques, and overcoming scalability challenges, alongside integration with complementary biotechnological processes such as nutrient recycling and co-cultivation strategies to bolster overall sustainability [187]. Additionally, the prospect of utilizing carbon dioxide emissions from industrial sources as a nutrient for cyanobacteria offers an added environmental advantage to the bioplastic production paradigm [16]. In summary, cyanobacteria, leveraging their intrinsic capacity for polyhydroxyalkanoate (PHA) synthesis, emerge as promising candidates for sustainable bioplastic manufacturing, underscored by interdisciplinary endeavors aimed at optimizing cyanobacterial strains, cultivation parameters, and downstream processing protocols [170,180].

5.3. Comparative Analysis with Conventional Plastics

The comparative analysis between bioplastics derived from cyanobacteria, specifically polyhydroxyalkanoates (PHAs), and conventional plastics reveals nuanced environmental considerations [188]. PHAs, synthesized by cyanobacteria through photosynthesis, not only demonstrate inherent biodegradability but also contribute to a closed carbon cycle, making them environmentally advantageous [189]. In contrast, conventional plastics, originating from petrochemicals, pose challenges related to non-biodegradability, leading to persistent waste accumulation and environmental harm [190]. The renewable source of cyanobacterial bioplastics is a key differentiator [191]. Cyanobacteria utilize sunlight and carbon dioxide to produce PHAs, presenting a sustainable alternative to conventional plastics that rely on finite fossil fuel resources [16]. The transition towards renewable feedstocks reflects a global imperative to mitigate carbon emissions and embrace sustainability [192]. In this context, the carbon neutrality achieved in Polyhydroxyalkanoates (PHAs) production offers significant environmental benefits [193]. Through photosynthesis, cyanobacteria effectively sequester carbon dioxide, contributing to atmospheric carbon reduction, unlike the carbon-intensive processes involved in conventional plastic production from fossil fuels [194]. By diminishing reliance on non-renewable resources, particularly prevalent in the petrochemical industry, cyanobacterial feedstocks present a viable pathway toward eco-friendly plastic manufacturing [16]. Moreover, the energy dynamics favor cyanobacterial bioplastics, with solar-powered photosynthesis offsetting downstream processing energy requirements, unlike the energy-intensive processes in conventional plastic production [195]. Waste management implications further underscore the superiority of cyanobacterial bioplastics, as their biodegradability lessens the strain on waste disposal systems compared to the persistence of non-biodegradable plastics [196]. In essence, the comprehensive analysis highlights the manifold environmental advantages of cyanobacterial bioplastics, especially PHAs, signaling a promising shift towards sustainable alternatives to conventional plastics [197].

5.4. PHAs as Parallel Competitors to Petrochemical-Based Plastics

Polyhydroxyalkanoates (PHAs) have emerged as potent contenders against petrochemical-based plastics, presenting a sustainable and eco-friendly option within the dynamic realm of plastic manufacturing [198]. PHAs, synthesized by cyanobacteria through the process of photosynthesis, offer a promising resolution to the environmental quandaries associated with traditional plastics derived from fossil fuels [199]. As concurrent rivals, various pivotal factors underscore the potential of PHAs to transform the plastics sector. One paramount advantage of PHAs is their inherent biodegradability, which sharply contrasts with the enduring nature of petrochemical-based plastics [200]. PHAs possess the ability to naturally degrade, thereby mitigating the environmental impact of plastic waste [201]. This attribute positions them as a compelling alternative amidst mounting apprehensions regarding plastic pollution and its detrimental repercussions on ecosystems [202].
Moreover, the renewable and carbon-neutral characteristics of PHA production further solidify their status as adversaries to petrochemical-based plastics [203]. The utilization of cyanobacteria, which harness solar energy and carbon dioxide during photosynthesis, contributes to a closed carbon cycle [204]. This presents a stark contrast to the carbon emissions stemming from the extraction and processing of fossil fuels for conventional plastics, aligning with global endeavors to combat climate change and transition towards sustainable practices [205]. A critical aspect of the competition between PHAs and petrochemical-based plastics lies in diminishing reliance on finite fossil fuels. By employing renewable resources for PHA production, the dependence on non-renewable fossil fuel sources is reduced, thereby addressing concerns related to resource depletion and environmental impact [206,207].
This shift is in harmony with the broader transition towards a circular economy and sustainable resource management practices. Energy consumption considerations further bolster the competitiveness of PHAs [208]. Although energy is requisite for downstream processing, the utilization of solar energy during photosynthesis diminishes the overall environmental footprint of PHA production [209]. In contrast, the energy-intensive procedures involved in fossil fuel extraction and refinement for conventional plastics contribute substantially to environmental strain. The versatility of PHAs concerning material properties further augments their competitiveness. PHAs can be tailored to exhibit specific characteristics suitable for a myriad of applications, ranging from packaging materials to medical devices. This adaptability positions them as feasible alternatives to petrochemical-based plastics across various industries, thereby offering a wide array of eco-conscious solutions [210,211,212].
In the final analysis, PHAs emerge as parallel contenders to petrochemical-based plastics, presenting a sustainable and environmentally mindful substitute. The biodegradability, renewability, carbon neutrality, and versatility of PHAs form a compelling rationale for their role in reshaping the future of plastic production, thereby fostering a more sustainable and circular economy. As research and technology progress, the rivalry between PHAs and conventional plastics is poised to drive innovation and promote a more environmentally responsible approach to plastic utilization and disposal.

6. Environmental Benefits of Cyanobacteria

Cyanobacteria, or blue-green algae, are pivotal agents in environmental sustainability due to their diverse biological activities, profoundly impacting ecosystems [213]. Primarily, they undertake photosynthesis, utilizing sunlight to convert carbon dioxide into organic compounds while liberating oxygen [214]. This photosynthetic process is indispensable for regulating atmospheric oxygen levels, supporting aquatic organisms’ respiratory requirements, and enhancing overall air quality [215]. Moreover, cyanobacteria significantly contribute to carbon sequestration by assimilating atmospheric carbon dioxide during photosynthesis. This aids in mitigating greenhouse gas emissions, as carbon is integrated into their biomass, thereby reducing atmospheric carbon dioxide concentrations [216]. Certain cyanobacterial species exhibit the remarkable capability of fixing atmospheric nitrogen, converting it into bioavailable forms for other organisms [217]. Indigenous bacteria capable of degrading MICROCYSTINs were found in the study. The isolated and screened strains included Proteobacteria, firmicutes, actinomyces, and fungi, and different strains showed differences in MICROCYSTIN degradation (Table 3). Some of these strains can simultaneously control algae and remove MICROCYSTINs [218]. Nitrogen fixation plays a crucial role in nutrient cycling within ecosystems, particularly in nitrogen-deficient environments, and fosters the growth of diverse plant species [219]. Additionally, cyanobacteria play a vital role in soil stabilization, especially in arid and semi-arid regions, by forming crusts that bind soil particles, preventing erosion, and enhancing soil structure [220].
In wastewater treatment, cyanobacteria are utilized for nutrient removal, efficiently absorbing and assimilating nitrogen, and phosphorus compounds from wastewater. This application aids in mitigating nutrient pollution in aquatic environments [221]. Furthermore, ongoing research explores the potential of cyanobacteria in biofuel and bioplastic production. Their capacity to convert sunlight and carbon dioxide into biomass offers a sustainable source of biofuels, while genetic engineering facilitates the production of biodegradable bioplastics, presenting environmentally friendly alternatives to conventional fossil fuel-derived plastics [222].
Cyanobacteria also support biodiversity in aquatic ecosystems, serving as a primary food source for various organisms and contributing to the intricate balance of food webs. Their ability to form symbiotic relationships further emphasizes their ecological significance [56,223]. Overall, the multifaceted roles of cyanobacteria in oxygen production, carbon sequestration, nutrient cycling, soil stabilization, wastewater treatment, and sustainable biofuel and bioplastic production underscore their pivotal contribution to environmental health and sustainability [224]. As ongoing investigations elucidate the nuanced ecological functions of cyanobacteria, these microorganisms persist as fundamental constituents within the intricate fabric of ecosystems, highlighting the efficacy of nature-derived approaches in mitigating modern environmental adversities.
Table 3. Exploring microcystin-degrading strains and their characteristics.
Table 3. Exploring microcystin-degrading strains and their characteristics.
Strain ClassificationFunctional StrainsTarget MicrocystinsInitial Microcystins ConcentrationMicrocystins Degradation EfficiencyReferences
BacillotaBacillus subtilis (strain 168)Microcystin-LR6 μg/mL(48 h)[225]
BacillotaBacillus brevis LEw-1238Microcystin-LR6 μg/mL(48 h)[226]
BacillotaBacillus thuringiensis LEw-2010MICROCYSTIN-RR20 μg/mL73% (3 d)[226]
BacillotaLysinibacillus boronitolerans strain CQ5Microcystin-LR14.10 μg/L90% (24 h)[227]
ActinomycesRhodococcus cavernicola C1-24Microcystin-LR10 mg/L9, 10, 8, 6, days[228]
ActinomycesBrevibacterium F3MICROCYSTIN-RR10 mg/L9, 10, 7, 9, days[228]
ActinomycesArthrobacter sp. C6Microcystin-LR10 mg/L9, 10, 8, 5 days[228]
ActinomycesArthrobacter sp. F7MICROCYSTIN-LF10 mg/L9, 10, 7, 6, days[228]
ActinomycesArthrobacter sanguinis sp. F10Microcystin-LR6 μg/mL98% (4 days)[229]
ActinomycetesArthrobacter sp. strain R1Microcystin-LR6 μg/mL(4 days)[229]
ActinomycesArthrobacter sp. strain R6. Microcystin-LR6 μg/mL(4 days)[229]
ActinomycetesArthrobacter sp. strain R9Microcystin-LR6 μg/mL(4 days)[229]
ActinomycesArthrobacter sanguinis sp. 423Microcystin-LR4 μg/L25.88% (10 days)[230]
ActinomycetesArthrobacter sanguinis sp. 443Microcystin-LR4 μg/L17.91% (10 days)[230]
ActinomycetesBifidobacterium BB-12Microcystin-LR110μg/L59.12% (3 days)[231]
ActinomycetesBifidobacterium longum BB-46Microcystin-LR110 μg/L48.0% (3 days)[231]
ActinomycetesBifidobacterium BB-420Microcystin-LR110 μg/L48.8% (3 days)[231]
FungusTrichaptum abietinum 1302BGMicrocystin-LR0.07 mg/L(12 h)[232,233]
FungusSchizophyllum commune strain IUM1114-SS01Microcystin-LR20 mg/L(2 days)[234]
Fungusascomycete strain TrichodermaMICROCYSTINs2.9 mg/L(4 days)[67]
FungusWinter mucus d. ursingii strain EH5Microcystin-LR0.05 mg/L39% (3 days)[235]
FungusAureobasidium pullulans KKUY0701MICROCYSTINs3 mg/L58.8% (1 h)[236]
ProteobacteriaSphingomonas Paucimobilis
. CBA4
MICROCYSTIN-RR150 μg/L(3 days)[237]
AlphaproteobacteriaSphingomonas sp. ACM-3962 MICROCYSTIN-RR15 mg/L2.7 mg/(L·h)[238]
ProteobacteriaPseudomonas paucimobilisMICROCYSTIN-RR3 mg/L0.09, mg/(L·h)[239,240]
ProteobacteriaNovosphingobium sp. NV-3Microcystin-LR30 mg/L 3)0.39 mg/(L·h)[241]
ProteobacteriaSphingomonas parapaucimobilis
7CY
MICROCYSTIN-LW8 mg/L5 days[242]
ProteobacteriaSphingomonas Y2Microcystin-LR22 mg/L0.26 mg/(L·h)[243]
ProteobacteriaNovosphingobium sp. KKU15Microcystin-LR6 μg/mL(4 days)[244]
ProteobacteriaNovosphingobium sp. ERW19Microcystin-LR, 0.2 mg/L0.010 mg/(L·h)[245]
ProteobacteriaNovosphingobium sp. ERN07MICROCYSTIN-RR0.2 mg/L0.007 mg/(L·h)[245]
ProteobacteriaNovosphingobium sp. KKU25sMicrocystin-LR30 μg/L1.07 μg/(L·h)[246]
AlphaproteobacteriaSphingopyxis sp. a7Microcystin-LR14.5 mg/L3.44 mg/(L·h)[121]
AlphaproteobacteriaSphingopyxis sp. X20Microcystin-LR7 mg/L(2 days)[122]
AlphaproteobacteriaSphingopyxis sp. YF1Microcystin-LR15 μg/mL53.6 μg/(mL·h)[102]
AlphaproteobacteriaSphingopyxis sp. USTB-05MICROCYSTIN-LA3.7 μg/L(3 days)[247]
AlphaproteobacteriaSphingosinicella microcystinivorans strain B-9MICROCYSTIN-RR2.4 mg/L0.06, mg/(L·h)[239,240]
Alphaproteobacteriaalgicidal bacterium Ochrobactrum sp. FDT5Microcystin-LR460.9 mg/L(6 days)[248]
PseudomonadotaR. solanacearumMicrocystin-LR9.6 mg/L(2 days)[249]
PseudomonadotaBordetella sp. strain MC-LTH1MICROCYSTIN-RR11, 8 mg/L(3 days)[250]
Beta ProteobacteriaB pertussis.Microcystin-LR18 μg/L(2 days)[251]
PseudomonadotaMethylobacillus sp.MICROCYSTIN-RR3.7, 4.2 mg/L<0.26, >0.26 mg/(L·h)[252]
PseudomonadotaPaucibacter sp. KCTC 42545Microcystin-LR11.8 μg/mL(10 h)[253]
Beta ProteobacteriaPaucibacter toxinivorans DSM 16,998;MICROCYSTIN-LY15 mg/L8, days[228]
Beta ProteobacteriaPaucibacter toxinivorans IM-4)MICROCYSTIN-YR0.95, mg/L(3 days)[119]
PseudomonadotaP. toxinivorans (2C20 strain)MICROCYSTIN-YR,15, μg/mL(8 days)[254]
PseudomonadotaComamonasLEw-2Microcystin-LR10 μg/mL(3 days)[226]
PseudomonadotaStenotrophomonas maltophilia
. EMS
Microcystin-LR1.7 μg/mL(2 days)[255]
PseudomonadotaStenotrophomonas maltophila LEw-1278 Microcystin-LR10 μg/mL(2 days)[226]
PseudomonadotaStenotrophomonas maltophilaMicrocystin-LR10 μg/mL0.6 μg/(mL·h)[256]
GammaproteobacteriaXanthomonas beteliMicrocystin-LR39.4 mg/L31, mg/(L·h)[257]
PseudomonadotaStenotrophomonas mori sp.MICROCYSTIN-LF5.7 μg/mL(12 days),
(14 days),
(15 days)
[258]
PseudomonadotaMorganella morganiiMicrocystin-LR25 μg/L3.77 mg/(L·h)[259]
PseudomonadotaPseudomonas aeruginosa UCBPP-PA14MICROCYSTIN-LW350 μg/L 3)85% (30 days)[218]
PseudomonadotaArthrobacter siderocapsulatusMicrocystin-LR350 μg/L 3)37% (30 days)[218]
GammaproteobacteriaAchromobacter xylosoxidansMicrocystin-LR150 μg/L79.8% (7 days)[260]
GammaproteobacteriaEnterobacter kobeiMicrocystin-LR5 μg/mL0.388 μg/(mL·h)[261]
BacillotaLacticaseibacillus rhamnosusMicrocystin-LR150 μg/L(3 days)[231]
FirmicutesLacticaseibacillus rhamnosus Lc 705Microcystin-LR150 μg/L(3 days)[231]
BacillotaBacillus sp.MICROCYSTIN-RR15 mg/L(6 days)[262]
BacillotaBacillus subtilisMICROCYSTIN-RR15 mg/L(6 days)[262]
FirmicutesB. thuringiensis sp. MICROCYSTIN-LR20 mg/L85% (10 days)[263]
BacillotaBacillus atrophaeus sp. MICROCYSTIN-RR15 mg/L(5 days)[262]
BacillotaBacillus sp. Ak3MICROCYSTIN-RR25 μg/mL75%(5 days)[225]
FirmicutesBrevibacillus brevis LEw-1238MICROCYSTIN-LR15 μg/mL(3 days)[226]
BacillotaBacillus thuringiensis LEw-2010MICROCYSTIN-LR15 μg/mL(3 days)[228]
BacillotaLysinibacillus boronitolerans (T-10a, AB199591)MICROCYSTIN-LR14.15 μg/L(3 days)[227]
ActinomycetotaRhodococcus cavernicola C1-24MICROCYSTIN-LR, 15 mg/L5 days[228]
ActinomycetotaBrevibacterium F3MICROCYSTIN-LR, 15 mg/L6 days[228]
ActinomycetesArthrobacter sp. C6MICROCYSTIN-LR 15 mg/L5 days[228]
Arthrobacter sanguinis sp. F7MICROCYSTIN-LR15 mg/L6 days[228]
Actinomycetota MICROCYSTIN-LY10 mg/L97 days[228]
CorynebacteriumArthrobacter sanguinis sp. F10MICROCYSTIN-LR7 μg/mL(4 days)[229]
CorynebacteriumArthrobacter sanguinis sp. R1MICROCYSTIN-LR7 μg/mL(4 days)[229]
BifidobacteriumArthrobacter sanguinis sp. R6MICROCYSTIN-LR7 μg/mL(4 days)[229]
ActinomycetesArthrobacter sanguinis sp. R9MICROCYSTIN-LR7 μg/mL(4 days)[229]
BifidobacteriumArthrobacter sp. 423MICROCYSTIN-LR55 μg/L(10 days)[230]
ActinomycetesArthrobacter sp. JS443MICROCYSTIN-LR5 μg/L15.92% 10 days)[230]
BifidobacteriumBifidobacterium BB-12MICROCYSTIN-LR150 μg/L(2 days)[231]
BifidobacteriumB. longum strainsMICROCYSTIN-LR150 μg/L(2 days)[231]
BifidobacteriumBifidobacterium animalis subsp. lactis 420MICROCYSTIN-LR150 μg/L(2 days)[231]

6.1. Carbon Sequestration and Reduction of Carbon Footprint

Cyanobacteria play a crucial role in the sequestration of carbon dioxide, actively participating in the reduction of carbon emissions and alleviating the consequences of human-induced carbon dioxide release [264]. Through the mechanism of photosynthesis, cyanobacteria utilize solar energy to convert carbon dioxide into organic molecules, facilitating the assimilation of carbon into their cellular structure [265]. This process of carbon fixation offers a twofold environmental benefit: it diminishes the concentration of atmospheric carbon dioxide, a significant greenhouse gas, while simultaneously providing a means for carbon retention within cyanobacterial cells [266]. By acting as a natural carbon sink, cyanobacteria contribute to the regulation of global carbon cycles, essential for mitigating the effects of climate change exacerbated by human activities such as fossil fuel combustion [267]. The efficiency of cyanobacteria in carbon sequestration underscores their significance in carbon offset endeavors and sustainable environmental strategies [268]. Moreover, ongoing research explores the potential of cyanobacteria for carbon capture and utilization technologies. By cultivating cyanobacteria in controlled settings such as photobioreactors and exposing them to carbon dioxide-rich industrial emissions, these microorganisms can absorb and convert emitted carbon dioxide into biomass [269,270]. This application not only aids in curbing greenhouse gas emissions but also holds promise to produce biofuels and other valuable bioproducts [271]. In the broader context of reducing carbon footprints, cyanobacteria demonstrate their potential as environmentally friendly agents of carbon sequestration and climate change mitigation [194]. Leveraging the innate carbon-fixing abilities of these microorganisms not only addresses environmental challenges but also emphasizes the importance of integrating nature-based solutions in the pursuit of sustainable, low-carbon pathways for the future [272]. As scientific inquiry progresses, the role of cyanobacteria in carbon sequestration emerges as a promising aspect of collective endeavors to combat climate change and minimize human-induced alterations to the global carbon equilibrium.

6.2. Sustainable Environmental Goals and Cyanobacteria’s Contribution

Cyanobacteria significantly contribute to the attainment of sustainable environmental objectives, aligning with broader initiatives targeted at mitigating pressing challenges such as climate change, biodiversity loss, and environmental degradation [273]. Their varied ecological functions and distinctive biochemical capabilities position cyanobacteria as valuable allies in the pursuit of sustainable environmental aims [274]. One notable contribution lies in carbon sequestration, as cyanobacteria actively participate in ameliorating climate change by sequestering atmospheric carbon dioxide during photosynthesis [194]. This directly aligns with global sustainability objectives aimed at reducing greenhouse gas emissions and achieving carbon neutrality. By assimilating carbon into their biomass, cyanobacteria serve as natural carbon sinks, playing a pivotal role in regulating atmospheric carbon levels and bolstering climate resilience efforts [275,276]. Cyanobacteria also play a crucial role in sustainable water management objectives, particularly in nutrient-rich environments [277]. Certain species exhibit nitrogen-fixing abilities, converting atmospheric nitrogen into accessible forms for other organisms [278]. This biological nitrogen fixation enhances nutrient cycling, fosters soil fertility, and bolsters the health of aquatic ecosystems. In wastewater treatment, cyanobacteria are utilized to eliminate surplus nutrients, addressing concerns associated with nutrient pollution and eutrophication [279]. Additionally, cyanobacteria contribute to sustainable agricultural practices by promoting soil stabilization [280]. In arid and semi-arid regions, cyanobacterial crusts aid in preventing soil erosion, maintaining soil structure, and fostering plant growth [281]. These ecological services play a vital role in promoting sustainable land use practices and combating desertification. The potential of cyanobacteria in biofuel and bioplastic production aligns with sustainable energy and resource management goals [282]. Through genetic manipulation and optimization of cultivation conditions, cyanobacteria can be engineered to produce biofuels, offering renewable alternatives to traditional fossil fuels [22]. Similarly, the production of biodegradable bioplastics from cyanobacteria addresses concerns related to plastic pollution and contributes to a circular economy approach [16]. In the context of the United Nations’ Sustainable Development Goals (SDGs), cyanobacteria’s multifaceted contributions align with several specific targets. These include SDG 13 (Climate Action) through carbon sequestration, SDG 6 (Clean Water and Sanitation) through nutrient cycling and water treatment, and SDG 15 (Life on Land) by promoting soil stability and biodiversity [68,283,284]. In summary, cyanobacteria emerge as crucial stakeholders in achieving sustainable environmental goals by addressing climate change, supporting nutrient cycling, contributing to water management, and providing alternatives in the production of biofuels and bioplastics. Their diverse ecological services underscore the importance of integrating cyanobacteria into comprehensive approaches for sustainable resource management and environmental conservation.

7. Efficiency of Cyanobacteria in Bioplastic Production

Cyanobacteria demonstrate remarkable efficiency in synthesizing bioplastics, particularly polyhydroxyalkanoates (PHAs), due to their inherent ability to conduct photosynthesis and possess unique metabolic pathways [285]. Their high photosynthetic efficiency enables the conversion of atmospheric carbon dioxide into biomass, serving as a sustainable carbon source for PHA synthesis [193]. When facing nutrient scarcity, cyanobacteria direct surplus carbon towards PHAs as intracellular reservoirs, highlighting their adeptness in carbon utilization. Genetic engineering advancements further enhance efficiency by enabling the customization of cyanobacterial strains to optimize PHAs production according to specific needs [16,286]. Various cultivation parameters such as light intensity, temperature, and nutrient availability significantly influence biomass and PHA yield. Strategic strain selection, considering both natural and engineered traits conducive to PHA production, also contributes to overall efficiency [287]. Effective downstream processing techniques for PHAs extraction and purification from cyanobacterial biomass are essential for bioplastic production efficiency [167]. Additionally, the utilization of renewable resources such as sunlight and carbon dioxide emphasizes the sustainability and environmental advantages of cyanobacteria-based bioplastics over conventional plastics [18]. Co-cultivation strategies leveraging symbiotic relationships and metabolic cooperation further enhance nutrient utilization and biomass production efficiency, positioning cyanobacteria as promising contributors to sustainable and renewable bioplastics development as research progresses [288].
Plastic degradation is mainly completed by microorganisms. Some insect larvae eat plastic and their intestinal microorganisms play a role [289]. Fungi that degrade plastic are mainly filamentous fungi, whose hyphal structure is conducive to secreting extracellular enzymes and acting on plastics. In addition, fungi have multiple strategies to degrade different compounds, including powerful enzyme systems, strong adsorption capacity, and generation of biosurfactants [290]. Fungal hydrophobin plays an important role in the process of biological in-situ repair. Its double-layer structure can form an amphipathic film at the hydrophobic/hydrophilic interface. As a biosurfactant, it increases the contact area of the substrate, thereby increasing the contact area of the substrate. Greatly improve the degradation efficiency [290]. Bacteria lack similar structures, so it is generally believed that the degradation ability of fungi is generally higher than that of bacteria (Table 4).

7.1. Photosynthetic Process and Increased Production Efficiency

The photosynthetic mechanism in cyanobacteria is a fundamental determinant of their productivity and effectiveness across various applications, notably in augmenting the yield of valuable compounds such as biofuels and bioplastics [327]. Cyanobacteria, as photosynthetic prokaryotes, employ pigments such as chlorophyll-a, phycocyanin, and phycoerythrin for light absorption. This absorbed solar energy is subsequently transduced into chemical energy through intricate biochemical pathways [328]. Understanding the nuances of the photosynthetic process is critical for optimizing production efficiency. The process encompasses two primary phases: the light-dependent reactions and the light-independent reactions (Calvin cycle) [329,330]. During the light-dependent reactions within thylakoid membranes, solar energy drives the synthesis of ATP (adenosine triphosphate) and NADPH (reduced nicotinamide adenine dinucleotide phosphate) [174]. These high-energy molecules are indispensable for carbon dioxide fixation in the Calvin cycle, the light-independent reactions, occurring in the chloroplast stroma [331]. Maximizing photosynthetic efficiency in cyanobacteria involves modulation of several key parameters such as light intensity, wavelength, and duration. Cyanobacteria exhibit adaptive mechanisms to varying light conditions, adjusting their pigment composition and photosynthetic machinery for optimal light capture [332,333]. Researchers can manipulate these factors through environmental regulation during cultivation or genetic engineering to enhance overall photosynthetic efficiency. Genetic engineering facilitates the customization of cyanobacterial strains for heightened production efficiency by modifying crucial enzymes and pathways associated with photosynthesis, thereby redirecting carbon flux toward desired product synthesis [334,335]. Furthermore, enhancing cyanobacterial tolerance to environmental stressors, such as light and nutrient fluctuations, promotes sustained and efficient production [336]. Improvements in photosynthetic efficiency extend to the design of cultivation systems such as photobioreactors, offering controlled environments for optimizing light, temperature, and nutrient conditions to enhance productivity [337]. Advances in comprehending the molecular intricacies of cyanobacterial photosynthesis enable the development of strategies to enhance light absorption, energy transfer, and carbon assimilation [338]. In essence, the photosynthetic process in cyanobacteria serves as a pivotal conduit for converting solar energy into valuable biomass and bioproducts [339]. Enhancing the efficiency of this process necessitates a comprehensive approach integrating environmental management, genetic engineering, and innovative cultivation techniques [340]. As scientific exploration into the molecular underpinnings of cyanobacterial photosynthesis advances, the potential for heightened production efficiency and sustainable applications expands, driving progress in bio-based industries [36].

7.2. Limited Land Input and Acceptable Cost Implications

The utilization of cyanobacteria in diverse biotechnological endeavors, such as the production of biofuels and bioplastics, presents distinct advantages characterized by minimal land requirements and favorable cost implications [341]. These features render cyanobacteria particularly promising for sustainable production processes [342]. Their adaptability to cultivation in environments with limited land availability distinguishes them from traditional crops used for biofuel production, as they do not necessitate arable land or fertile soil [343]. Cyanobacteria exhibit robust growth capabilities in varied settings, including non-arable land and wastewater, thereby reducing competition with food crops for cultivation space and mitigating concerns regarding the land-use impact of biofuel and bioplastic feedstocks [344]. Additionally, the cost-effectiveness of cyanobacteria cultivation is notable, especially when contrasted with conventional biofuel feedstocks, owing to their low land input requirements [107]. The economic feasibility of cyanobacteria cultivation is particularly increased by the low land requirement compared to conventional biofuel raw materials. By introducing a multi-product approach, focusing on high-quality compounds, and using cost-effective cultivation methods, overall economics can be significantly improved. By integrating processes and optimization strategies, for example, through the use of agro-industrial waste and wastewater, input costs can be further reduced. Additionally, advances in nanotechnology and recycling strategies can increase production efficiency. These joint efforts make cyanobacteria a promising and cost-effective alternative for the sustainable production of biofuels and bioproducts [345]. Technological advancements in cultivation methods, such as the development of economical photobioreactor systems, contribute to the economic viability of cyanobacteria-based processes [346]. Cyanobacteria’s efficient utilization of nutrients, including nitrogen and phosphorus, is pivotal for cost-effectiveness, reducing the necessity for excessive nutrient inputs [347]. Certain cyanobacteria species’ ability to fix atmospheric nitrogen aids in sustainable nutrient cycling, aligning with environmentally conscious and economically sound practices [348]. Furthermore, the utilization of wastewater streams as a nutrient source serves a dual purpose of wastewater treatment and biomass production, diminishing the environmental impact of nutrient-rich wastewater while lowering cultivation costs. Leveraging unconventional nutrient sources enhances the overall cost-effectiveness of cyanobacteria-based processes [349,350]. In conclusion, cyanobacteria’s minimal land requirements and favorable cost implications position them as promising candidates for sustainable biofuel and bioplastic production, in line with objectives aimed at minimizing environmental impact, mitigating resource competition, and advancing economically viable alternatives in the pursuit of a more sustainable and circular bio-based economy [158,351]. Continued research and technological innovations in cyanobacteria cultivation and bioprocess optimization further augment their potential for cost-effective and environmentally friendly applications in the bioenergy and bioplastics sectors.

7.3. Advantages over Other Sources of Bioplastics

Cyanobacteria present distinct advantages over alternative bioplastic sources, positioning them as promising contenders for sustainable and eco-friendly alternatives to conventional plastics. One notable advantage lies in their minimal land input demands [352]. Unlike certain bioplastic feedstocks derived from traditional crops, cyanobacteria do not necessitate arable land or fertile soil, thus mitigating concerns regarding land competition and fostering more sustainable bioplastic production. Another key advantage is their potential cost-effectiveness [353,354]. Cyanobacteria’s rapid growth rates and scalability, particularly in controlled environments such as photobioreactors, contribute to high biomass productivity, enhancing the overall cost-effectiveness of cyanobacteria-based bioplastic processes [355]. Moreover, their ability to thrive in various environments, including non-arable land and wastewater, reduces reliance on costly nutrient supplementation, further bolstering economic viability [356]. The photosynthetic capability of cyanobacteria represents a significant advantage. Through photosynthesis, these microorganisms directly convert solar energy into chemical energy, facilitating the synthesis of bioplastics from atmospheric carbon dioxide [16]. This renewable and carbon-neutral approach contrasts with certain bioplastic feedstocks reliant on agricultural inputs and processes contributing to a carbon footprint [357]. Cyanobacteria’s photosynthetic efficiency aligns with the broader goal of reducing dependence on finite fossil fuels and mitigating climate change impacts associated with conventional plastics [100]. Additionally, their genetic adaptability serves as a platform for tailored bioplastic production. By employing genetic engineering, researchers can optimize cyanobacterial strains to enhance the synthesis of specific bioplastics with desired properties, allowing thproduction of customized bioplastics suited to diverse applications [285,358].
Furthermore, cyanobacteria demonstrate versatility in growth conditions, thriving in diverse climates from arid regions to aquatic environments, surpassing certain crop-based bioplastic feedstocks in adaptability [359]. This adaptability contributes to the resilience and reliability of cyanobacteria-based bioplastic production systems [360]. In close, the advantages of cyanobacteria, including their low land input requirements, cost-effectiveness, photosynthetic efficiency, genetic malleability, and versatile growth conditions, position them as compelling and sustainable candidates for eco-friendly alternatives to conventional plastics. They offer a pathway to address environmental concerns associated with plastic pollution and contribute to the development of a circular bio-based economy.

8. Achievements and Constraints in Commercialization Pathway

The advancement of cyanobacteria-based technologies along the commercialization pathway has yielded significant progress, particularly in the domains of bioplastic and biofuel synthesis [286]. Cyanobacteria, functioning as photosynthetic microorganisms, have exhibited their capability in generating polyhydroxyalkanoates (PHAs), a class of biodegradable bioplastics, thereby holding promise for mitigating the environmental impact of conventional plastics [180]. Moreover, within the realm of biofuels, cyanobacteria have demonstrated efficacy in harnessing sunlight and carbon dioxide to produce renewable energy sources, contributing to the pursuit of sustainable alternatives [361]. Notably, they have been effectively employed in wastewater treatment processes, efficiently absorbing and metabolizing nutrients from wastewater, thereby offering a dual advantage of nutrient removal and biomass production [362]. This utilization aligns with sustainable environmental practices and presents a practical solution to challenges in wastewater management [363]. Genetic engineering advancements have significantly augmented cyanobacteria’s capacities for bioplastic and biofuel synthesis by optimizing metabolic pathways, resulting in enhanced yields and improved product characteristics [16]. These genetic engineering breakthroughs represent a significant step toward the commercial viability of cyanobacteria-based technologies [107]. However, despite these achievements, various constraints persist along this commercialization trajectory. Challenges such as elevated production costs, scalability limitations, and technological intricacies pose hurdles that necessitate careful consideration [364]. The refinement of cultivation systems, downstream processing techniques, and the resolution of scalability concerns are essential for bridging the gap between laboratory-scale successes and large-scale commercial applications [365]. Furthermore, ensuring regulatory compliance and garnering public acceptance are pivotal for the widespread adoption of cyanobacteria-based technologies [366]. While notable strides have been made in the commercialization of cyanobacteria-based technologies, addressing existing constraints is imperative for realizing their full potential [367]. Continued research, technological innovation, and strategic collaborations will be instrumental in overcoming these challenges and establishing cyanobacteria as viable contributors to sustainable bioplastic and biofuel production at a commercial scale.

Challenges Faced in Scaling up Cyanobacteria-Based Production

Expanding the production of bioplastics using cyanobacteria encounters a multitude of complex challenges that demand meticulous analysis and inventive solutions. Among these challenges, cost-effectiveness stands out as paramount [368,369]. As production scales up, operational expenses inevitably increase, underscoring the need for economical cultivation methods, downstream processing, and extraction techniques to ensure the economic viability of large-scale cyanobacteria-based production [370]. Additionally, scalability presents a significant hurdle in transitioning from small laboratory experiments to industrial production [371]. Optimizing growth conditions, light distribution, and nutrient availability becomes increasingly intricate as production scales [372]. Overcoming these challenges necessitates innovative design approaches for photobioreactors and cultivation strategies capable of maintaining efficiency and consistency at larger scales [373].
Furthermore, addressing the nutrient requirements of cyanobacteria, essential for their growth, poses challenges in sourcing and sustainability when scaling up [374]. Ensuring a stable and sustainable nutrient supply, especially in open pond systems, is crucial and intersects with broader environmental sustainability concerns. Balancing productivity and product quality remains an ongoing challenge, requiring constant refinement of growth conditions, light exposure, and nutrient utilization to enhance overall efficiency and economic feasibility [375,376]. Navigating regulatory compliance is also critical as cyanobacteria-based technology progresses toward commercial deployment. Adhering to regulations pertaining to environmental safety, product quality, and overall compliance is vital for successful large-scale production. Establishing clear regulatory guidelines and fostering collaboration between industry stakeholders and regulatory bodies are imperative for creating a conducive environment for large-scale implementation [377,378,379]. Moreover, downstream processing complexities, particularly in extracting and purifying bioplastics from cyanobacterial biomass at industrial scales, present additional challenges. Developing scalable and cost-effective separation and purification technologies is essential for maintaining economic viability [380,381,382]. Public perception and acceptance further complicate matters, particularly regarding safety and environmental impact. Educating the public and providing transparent information are crucial for overcoming societal resistance to large-scale implementation. Addressing public concerns and engaging in outreach efforts are essential for building acceptance and support [383,384]. However, the technological complexity associated with scaling up cyanobacteria-based production introduces challenges in monitoring, control, and automation. Innovations in process engineering and control systems are necessary to ensure the robustness and reliability of large-scale cultivation [385]. Collaborative efforts involving researchers, industry stakeholders, and regulatory bodies are vital for overcoming these challenges, fostering innovation, and realizing the full potential of cyanobacteria-based production at a commercial scale [386,387].

9. Cyanobacterial Species as Sources of Green and Clean Energy

Cyanobacterial species are attracting significant scientific interest as potent candidates for sustainable energy sources, leveraging their distinct photosynthetic abilities and metabolic pathways [388]. A primary focus of the investigation revolves around the potential of cyanobacteria for biofuel production [389]. These microorganisms serve as inherent biofuel factories, utilizing sunlight and carbon dioxide for the direct synthesis of biofuels such as biodiesel and bioethanol [390]. Researchers are actively involved in refining cyanobacterial strains through genetic manipulation to enhance the production of lipids and carbohydrates, crucial constituents in biofuel synthesis. This approach not only presents a sustainable alternative to traditional fossil fuels but also holds promise for carbon-neutral energy production [22,391]. In the domain of hydrogen production, specific cyanobacterial species demonstrate the remarkable capability to generate molecular hydrogen via photobiological processes. Through harnessing solar energy, cyanobacteria facilitate the breakdown of water molecules, releasing hydrogen as a clean and renewable byproduct. This photobiological hydrogen production method shows significant potential as an environmentally friendly approach to addressing the rising demand for hydrogen fuel, thereby contributing to a more sustainable energy landscape [392,393,394].
Furthermore, cyanobacteria exhibit potential in electricity generation through microbial fuel cells (MFCs) [395]. Within these systems, cyanobacteria serve as biocatalysts, utilizing their photosynthetic activity to produce electrons from organic matter, which can then be utilized to generate electrical energy [396]. Although the efficiency of cyanobacteria in MFCs is an active area of investigation, the concept of employing cyanobacteria as green catalysts for electricity production presents an intriguing avenue for sustainable power generation [397]. Apart from energy production, cyanobacteria also play a vital role in carbon capture and storage [398]. Through photosynthesis, cyanobacteria capture carbon dioxide from the atmosphere, aiding in the mitigation of greenhouse gas emissions [399]. This dual functionality highlights the versatility of cyanobacteria, positioning them not only as energy producers but also as crucial contributors to efforts aimed at combating climate change [400]. However, despite the promising potential of cyanobacteria in green energy production, challenges persist. Optimizing growth conditions, enhancing productivity, and addressing scalability concerns are ongoing research priorities. The economic viability of large-scale cyanobacteria-based energy production necessitates careful consideration of factors such as cultivation systems, nutrient availability, and downstream processing [401,402,403].
In essence, cyanobacterial species embody adaptable and eco-friendly reservoirs of renewable energy. Ranging from biofuel synthesis to hydrogen generation and electrical power generation, the distinctive attributes of cyanobacteria present diverse strategies for addressing the worldwide energy dilemma. Persistent exploration and advancements in technology are imperative for maximizing the complete capacity of cyanobacteria-mediated energy generation, thereby fostering the development of a more environmentally friendly and sustainable energy landscape.

9.1. Diversity of Cyanobacterial Species

Cyanobacteria, a diverse assembly of photosynthetic prokaryotes, exhibit a wide array of morphological, physiological, and ecological traits [48]. These microorganisms demonstrate notable versatility in adapting to various environments, thriving in environments ranging from freshwater bodies to extreme habitats such as hot springs and polar regions [404]. Their morphological diversity encompasses unicellular, filamentous, and colonial forms, each characterized by distinct structures such as trichomes or mucilaginous sheaths [405]. Renowned for their vibrant pigmentation, cyanobacteria employ pigments such as chlorophyll-a, phycocyanin, and phycoerythrin, resulting in the characteristic blue-green hue, though certain species manifest a spectrum of colors [406]. Primarily relying on photosynthesis, cyanobacteria played a crucial role in the evolution of oxygenic photosynthesis, contributing significantly to the oxygenation of Earth’s atmosphere [407]. Their metabolic repertoire extends to nitrogen fixation, facilitated by specialized cells such as heterocysts [408]. Ecologically, cyanobacteria contribute to primary production, nutrient cycling, and symbiotic relationships with plants, yet they can also form harmful algal blooms under specific conditions, generating toxins with potential ecological and health ramifications [409]. A thorough comprehension of cyanobacterial diversity is imperative for elucidating their ecological functions and potential applications across diverse ecosystems [410].

9.2. High Potential in Biofuel Production

Cyanobacteria demonstrate considerable promise for biofuel generation, offering a sustainable and renewable energy solution [123]. Capitalizing on their distinct photosynthetic prowess, these microorganisms efficiently convert sunlight and carbon dioxide into biofuels, presenting an eco-friendly alternative to conventional fossil fuels [411]. The primary focus centers on biodiesel and bioethanol production [412]. Through genetic engineering, cyanobacteria such as Synechocystis and Synechococcus have been modified to boost lipid and carbohydrate synthesis, essential precursors for biofuel production [413]. Lipids, especially, serve as valuable raw materials for biodiesel manufacture. Genetic manipulation endeavors strive to enhance metabolic pathways within cyanobacteria, augmenting their efficacy in accumulating these biofuel precursor compounds [147]. One of the principal advantages of cyanobacteria in biofuel generation lies in their capacity to thrive in non-arable land and diverse environmental conditions [186]. This adaptability mitigates resource competition with food crops and diminishes the environmental impact associated with land-use alterations [414]. Moreover, cyanobacteria can be cultivated in enclosed systems such as photobioreactors, enabling precise regulation of growth parameters and heightened productivity [415]. Ongoing research endeavors continuously explore cyanobacteria’s potential as a biofuel source by tackling challenges related to optimizing growth conditions, increasing biomass output, and refining downstream processing techniques for efficient biofuel retrieval [416]. Progress in synthetic biology and metabolic engineering is facilitating the development of tailored cyanobacterial strains with enhanced biofuel generation capabilities [417]. As these technologies advance, cyanobacteria are poised to play a crucial role in the transition toward sustainable and carbon-neutral biofuel production, significantly contributing to global efforts to mitigate climate change and secure a cleaner energy future [147,418].

9.3. Role in the Production of Biodegradable Plastics

Cyanobacteria are pivotal in the synthesis of biodegradable plastics, providing an ecologically sound substitute for conventional petroleum-derived plastics [419]. The utilization of cyanobacteria for this purpose is rooted in their innate capacity to produce polyhydroxyalkanoates (PHAs), a category of biopolymers suitable for biodegradable plastic production [420]. Through manipulation of their metabolic pathways, researchers have effectively engineered cyanobacterial strains to augment PHA synthesis [167]. Prominent cyanobacteria such as Synechocystis and Synechococcus have been subject to genetic modifications to enhance PHA production, showcasing the adaptability of these microorganisms in sustainable bioplastic manufacturing [16]. The advantage of employing cyanobacteria lies in their photosynthetic prowess, enabling them to harness solar energy and convert carbon dioxide into biomass rich in PHAs [421]. This process aligns with the ethos of a circular economy, utilizing renewable resources and diminishing reliance on fossil fuels [422]. Furthermore, cyanobacteria can be cultivated in diverse environments, including enclosed systems such as photobioreactors, affording control overgrowth parameters and facilitating scalable production of biodegradable plastics [423]. Endeavors are ongoing to optimize cyanobacteria-based bioplastic production by addressing variables such as strain selection, cultivation conditions, and nutrient availability [424]. Researchers are investigating methods to enhance the efficiency of PHA extraction and subsequent processing, ensuring an economically viable and environmentally sustainable bioplastic production pipeline [425]. With increasing demand for environmentally conscious alternatives to traditional plastics, cyanobacteria emerge as a promising solution for biodegradable plastics [16]. Their singular capacity to utilize sunlight and carbon dioxide for PHA synthesis situates cyanobacteria at the vanguard of sustainable bioplastic production, contributing to global efforts to mitigate the environmental ramifications of plastic waste [426,427]. Ongoing research and technological progress in this domain hold the potential to establish cyanobacteria as pivotal contributors to the advancement of biodegradable plastics, fostering a greener and more sustainable trajectory for the plastics industry.

10. Conclusions

In conclusion, cyanobacteria demonstrate a broad spectrum of metabolic functionalities, including but not limited to carbon fixation, biofuel synthesis, and biopolymer formation, thereby substantiating their crucial role in sustainable biotechnological pursuits. Advancements in genetic engineering and cultivation techniques are indispensable for maximizing the potential of cyanobacteria across these domains. In particular, directing cyanobacteria to synthesize polyhydroxyalkanoates (PHAs) for eco-friendly plastic production represents a groundbreaking tactic in line with circular economy principles. This strategy addresses ecological apprehensions while endorsing efficient waste treatment. Subsequent research endeavors should focus on refining cyanobacterial strains via genetic interventions, exploring sophisticated systems biology methodologies, and enhancing sustainable cultivation protocols to fully exploit their potential to promote sustainability. Future research directions could include developing advanced genetic tools to improve cyanobacteria metabolic pathways to enable higher yields of biofuels and biopolymers utilizing systems biology approaches to understand and optimize the complex interactions within cyanobacterial cells and their environment, and innovative cultivation techniques that minimize resource use and maximize productivity, using wastewater for nutrient supply and integrating renewable energy sources conducting comprehensive studies on the environmental impacts of large-scale cyanobacterial applications to ensure ecological balance and sustainability; and exploring the economic viability and market potential of cyanobacteria-derived products to facilitate their transition from research to commercial use. Ultimately, cyanobacteria represent innovative ways to promote a greener and more resilient future and contribute significantly to global sustainability efforts by providing renewable energy sources, biodegradable materials, and efficient waste management solutions.

Author Contributions

T.N. and L.G.: conceptualization, data curation, writing—original draft; Z.H., S.F., S.S., Z.H. and R.Z.: formal analysis, investigation, methodology, resources; supervision, validation, visualization, writing—review and editing; R.Z. and S.F.: supervision, validation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors express sincere gratitude to the reviewers for their invaluable constructive feedback, which has greatly contributed to enhancing the quality and impact of this review article. The authors appreciate the USDA Supported Hatch Project SD00H691-20 titled Engineering Solar-powered N2-fixing Cyanobacteria for Agricultural and Industrial Applications (to R.Z.). This acknowledgement recognizes the contribution of Biorender in creating the figures for the paper and highlights the platform’s role in enhancing the visual presentation of scientific information.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kibria, M.G.; Masuk, N.I.; Safayet, R.; Nguyen, H.Q.; Mourshed, M. Plastic waste: Challenges and opportunities to mitigate pollution and effective management. Int. J. Environ. Res. 2023, 17, 20. [Google Scholar] [CrossRef] [PubMed]
  2. Moshood, T.D.; Nawanir, G.; Mahmud, F.; Mohamad, F.; Ahmad, M.H.; AbdulGhani, A. Sustainability of biodegradable plastics: New problem or solution to solve the global plastic pollution? Curr. Res. Green Sustain. Chem. 2022, 5, 100273. [Google Scholar] [CrossRef]
  3. Nawaz, T.; Gu, L.; Fahad, S.; Saud, S.; Bleakley, B.; Zhou, R. Exploring Sustainable Agriculture with Nitrogen-Fixing Cyanobacteria and Nanotechnology. Molecules 2024, 29, 2534. [Google Scholar] [CrossRef] [PubMed]
  4. Davis, H. Plastic Matter; Duke University Press: Durham, NC, USA, 2022; ISBN 978-1-4780-1775-2. [Google Scholar] [CrossRef]
  5. Kumar, R.; Verma, A.; Shome, A.; Sinha, R.; Sinha, S.; Jha, P.K.; Kumar, R.; Kumar, P.; Shubham; Das, S. Impacts of plastic pollution on ecosystem services, sustainable development goals, and need to focus on circular economy and policy interventions. Sustainability 2021, 13, 9963. [Google Scholar] [CrossRef]
  6. Issac, M.N.; Kandasubramanian, B. Effect of microplastics in water and aquatic systems. Environ. Sci. Pollut. Res. 2021, 28, 19544–19562. [Google Scholar] [CrossRef] [PubMed]
  7. Ammendolia, J.; Saturno, J.; Brooks, A.L.; Jacobs, S.; Jambeck, J.R. An emerging source of plastic pollution: Environmental presence of plastic personal protective equipment (PPE) debris related to COVID-19 in a metropolitan city. Environ. Pollut. 2021, 269, 116160. [Google Scholar] [CrossRef]
  8. Mejjad, N.; Cherif, E.K.; Rodero, A.; Krawczyk, D.A.; El Kharraz, J.; Moumen, A.; Laqbaqbi, M.; Fekri, A. Disposal behavior of used masks during the COVID-19 pandemic in the Moroccan community: Potential environmental impact. Int. J. Environ. Res. Public Health 2021, 18, 4382. [Google Scholar] [CrossRef]
  9. Awogbemi, O.; Von Kallon, D.V. Achieving affordable and clean energy through conversion of waste plastic to liquid fuel. J. Energy Inst. 2023, 106, 101154. [Google Scholar] [CrossRef]
  10. Warning, U.N.E.P.D.o.E. UNEP Year Book 2011: Emerging Issues in Our Global Environment; UNEP/Earthprint: Nairobi, Kenya, 2011. [Google Scholar]
  11. Stegmann, P.; Daioglou, V.; Londo, M.; van Vuuren, D.P.; Junginger, M. Plastic futures and their CO2 emissions. Nature 2022, 612, 272–276. [Google Scholar] [CrossRef]
  12. Nawaz, T.; Gu, L.; Gibbons, J.; Hu, Z.; Zhou, R. Bridging Nature and Engineering: Protein-Derived Materials for Bio-Inspired Applications. Biomimetics 2024, 9, 373. [Google Scholar] [CrossRef]
  13. Ghosh, K.; Jones, B.H. Roadmap to biodegradable plastics—Current state and research needs. ACS Sustain. Chem. Eng. 2021, 9, 6170–6187. [Google Scholar] [CrossRef]
  14. Chen, L.; Msigwa, G.; Yang, M.; Osman, A.I.; Fawzy, S.; Rooney, D.W.; Yap, P.-S. Strategies to achieve a carbon neutral society: A review. Environ. Chem. Lett. 2022, 20, 2277–2310. [Google Scholar] [CrossRef] [PubMed]
  15. Nawaz, T.; Fahad, S.; Zhou, R. Protein Phosphorylation Nexus of Cyanobacterial Adaptation and Metabolism. Kinases Phosphatases 2024, 2, 209–223. [Google Scholar] [CrossRef]
  16. Agarwal, P.; Soni, R.; Kaur, P.; Madan, A.; Mishra, R.; Pandey, J.; Singh, S.; Singh, G. Cyanobacteria as a promising alternative for sustainable environment: Synthesis of biofuel and biodegradable plastics. Front. Microbiol. 2022, 13, 939347. [Google Scholar] [CrossRef]
  17. Chen, H.-G.; Zhang, Y.-H.P. New biorefineries and sustainable agriculture: Increased food, biofuels, and ecosystem security. Renew. Sustain. Energy Rev. 2015, 47, 117–132. [Google Scholar] [CrossRef]
  18. Żymańczyk-Duda, E.; Samson, S.O.; Brzezińska-Rodak, M.; Klimek-Ochab, M. Versatile applications of cyanobacteria in biotechnology. Microorganisms 2022, 10, 2318. [Google Scholar] [CrossRef]
  19. Singh, A.K.; Sharma, L.; Mallick, N.; Mala, J. Progress and challenges in producing polyhydroxyalkanoate biopolymers from cyanobacteria. J. Appl. Phycol. 2017, 29, 1213–1232. [Google Scholar] [CrossRef]
  20. Cheah, W.Y.; Show, P.L.; Chang, J.-S.; Ling, T.C.; Juan, J.C. Biosequestration of atmospheric CO2 and flue gas-containing CO2 by microalgae. Bioresour. Technol. 2015, 184, 190–201. [Google Scholar] [CrossRef]
  21. Cuellar-Bermudez, S.; Aguilar-Hernandez, I.; Cardenas-Chavez, D.; Ornelas-Soto, N.; Romero-Ogawa, M.; Parra-Saldivar, R. Extraction and purification of high-value metabolites from microalgae: Essential lipids, astaxanthin and phycobiliproteins. Microb. Biotechnol. 2015, 8, 190–209. [Google Scholar] [CrossRef]
  22. Khan, A.Z.; Bilal, M.; Mehmood, S.; Sharma, A.; Iqbal, H.M. State-of-the-art genetic modalities to engineer cyanobacteria for sustainable biosynthesis of biofuel and fine-chemicals to meet bio–economy challenges. Life 2019, 9, 54. [Google Scholar] [CrossRef]
  23. Johnson, T.J.; Gibbons, J.L.; Gu, L.; Zhou, R.; Gibbons, W.R. Molecular genetic improvements of cyanobacteria to enhance the industrial potential of the microbe: A review. Biotechnol. Prog. 2016, 32, 1357–1371. [Google Scholar] [CrossRef]
  24. Satta, A.; Esquirol, L.; Ebert, B.E. Current metabolic engineering strategies for photosynthetic bioproduction in cyanobacteria. Microorganisms 2023, 11, 455. [Google Scholar] [CrossRef] [PubMed]
  25. Carr, R. B-Pinene Synthesis in Synechococcus sp. pcc 7002 Cyanobacteria: Metabolic Engineering for High-Density Biofuel Applications. Master’s Thesis, The University of Wisconsin Oshkosh, Oshkosh, WI, USA, 2017. [Google Scholar]
  26. Jeong, S.H.; Lee, H.J.; Lee, S.J. Recent Advances in CRISPR-Cas Technologies for Synthetic Biology. J. Microbiol. 2023, 61, 13–36. [Google Scholar] [CrossRef] [PubMed]
  27. Hu, J.; Meng, W.; Su, Y.; Qian, C.; Fu, W. Emerging technologies for advancing microalgal photosynthesis and metabolism toward sustainable production. Front. Mar. Sci. 2023, 10, 1260709. [Google Scholar] [CrossRef]
  28. Kumar, G.; Shekh, A.; Jakhu, S.; Sharma, Y.; Kapoor, R.; Sharma, T.R. Bioengineering of microalgae: Recent advances, perspectives, and regulatory challenges for industrial application. Front. Bioeng. Biotechnol. 2020, 8, 914. [Google Scholar] [CrossRef] [PubMed]
  29. Koch, M. Metabolic Engineering Strategies for an Increased PHB Production in Cyanobacteria. Ph.D. Thesis, Universität Tübingen, Tübingen, Germany, 2020. [Google Scholar]
  30. Choi, Y.-N.; Lee, J.W.; Kim, J.W.; Park, J.M. Acetyl-CoA-derived biofuel and biochemical production in cyanobacteria: A mini review. J. Appl. Phycol. 2020, 32, 1643–1653. [Google Scholar] [CrossRef]
  31. Verdú-Navarro, F.; Moreno-Cid, J.A.; Weiss, J.; Egea-Cortines, M. The advent of plant cells in bioreactors. Front. Plant Sci. 2023, 14, 1310405. [Google Scholar] [CrossRef]
  32. Liu, Y.; Zhang, D.; Yang, T.; Chen, R.; Luo, X. Pathway engineering of plant-derived bioactive compounds in microbes. In Engineering Biology for Microbial Biosynthesis of Plant-Derived Bioactive Compounds; Elsevier: Amsterdam, The Netherlands, 2024; pp. 73–87. [Google Scholar]
  33. Hassan, A.I.; Saleh, H.M. Production of Amino Acids and Nucleic Acids from Genetically Engineered Microbial Cells and their Relevance to Biodegradation. Green Energy Environ. Technol. 2023. [Google Scholar] [CrossRef]
  34. Kastberg, L.L.B.; Ard, R.; Jensen, M.K.; Workman, C.T. Burden imposed by heterologous protein production in two major industrial yeast cell factories: Identifying sources and mitigation strategies. Front. Fungal Biol. 2022, 3, 827704. [Google Scholar] [CrossRef]
  35. Gudmundsson, S.; Nogales, J. Cyanobacteria as photosynthetic biocatalysts: A systems biology perspective. Mol. BioSystems 2015, 11, 60–70. [Google Scholar] [CrossRef]
  36. Janssen, P.J.; Lambreva, M.D.; Plumeré, N.; Bartolucci, C.; Antonacci, A.; Buonasera, K.; Frese, R.N.; Scognamiglio, V.; Rea, G. Photosynthesis at the forefront of a sustainable life. Front. Chem. 2014, 2, 36. [Google Scholar] [CrossRef] [PubMed]
  37. Nawaz, T.; Gu, L.; Fahad, S.; Saud, S.; Jiang, Z.; Hassan, S.; Harrison, M.T.; Liu, K.; Khan, M.A.; Liu, H. A comprehensive review of the therapeutic potential of cyanobacterial marine bioactives: Unveiling the hidden treasures of the sea. Food Energy Secur. 2023, 12, e495. [Google Scholar] [CrossRef]
  38. Gu, L.; Nawaz, T.; Qiu, Y.; Wu, Y.; Zhou, R. Photosynthetic conversion of CO2 and H2O to long-chain terpene alcohol by genetically engineered N2-fixing cyanobacteria. In Photosynthesis; Elsevier: Amsterdam, The Netherlands, 2023; pp. 451–461. [Google Scholar]
  39. Nawaz, T.; Gu, L.; Fahad, S.; Saud, S.; Hassan, S.; Harrison, M.T.; Liu, K.; Zhou, R. Unveiling the antioxidant capacity of fermented foods and food microorganisms: A focus on cyanobacteria. J. Umm Al-Qura Univ. Appl. Sci. 2023, 10, 232–243. [Google Scholar] [CrossRef]
  40. Archer, M.D.; Barber, J. Photosynthesis and photoconversion. Mol. Glob. Photosynth. 2004, 2, 1–42. [Google Scholar]
  41. Singh, S.K.; Sundaram, S.; Kishor, K. Photosynthetic Microorganisms: Mechanism for Carbon Concentration; Springer: Cham, Switzerland, 2014. [Google Scholar]
  42. Golubic, S.; Seong-Joo, L.; Browne, K.M. Cyanobacteria: Architects of sedimentary structures. In Microbial Sediments; Springer: Berlin/Heidelberg, Germany, 2000; pp. 57–67. [Google Scholar]
  43. Robles-Fernández, A.; Areias, C.; Daffonchio, D.; Vahrenkamp, V.C.; Sánchez-Román, M. The Role of microorganisms in the nucleation of carbonates, environmental implications and applications. Minerals 2022, 12, 1562. [Google Scholar] [CrossRef]
  44. Paerl, H.W. Mitigating harmful cyanobacterial blooms in a human-and climatically-impacted world. Life 2014, 4, 988–1012. [Google Scholar] [CrossRef]
  45. Yang, Y.; Liu, L.; Zhang, P.; Wu, F.; Wang, Y.; Xu, C.; Zhang, L.; An, S.; Kuzyakov, Y. Large-scale ecosystem carbon stocks and their driving factors across Loess Plateau. Carbon Neutrality 2023, 2, 5. [Google Scholar] [CrossRef]
  46. Bhardwaj, A.; Singh, P.; Gupta, N.; Bhattacharjee, S.; Srivastava, A.; Parida, A.; Mishra, A.K. Cyanobacteria: A key player in nutrient cycling. In Cyanobacteria; Elsevier: Amsterdam, The Netherlands, 2024; pp. 579–596. [Google Scholar]
  47. Ullah, I.; Dawar, K.; Tariq, M.; Sharif, M.; Fahad, S.; Adnan, M.; Ilahi, H.; Nawaz, T.; Alam, M.; Ullah, A. Gibberellic acid and urease inhibitor optimize nitrogen uptake and yield of maize at varying nitrogen levels under changing climate. Environ. Sci. Pollut. Res. 2022, 29, 6568–6577. [Google Scholar] [CrossRef]
  48. Stanier, R.; Cohen-Bazire, G. Phototrophic prokaryotes: The cyanobacteria. Annu. Rev. Microbiol. 1977, 31, 225–274. [Google Scholar] [CrossRef]
  49. Ligrone, R.; Ligrone, R. The Chloroplast and Photosynthetic Eukaryotes. Biological Innovations That Built the World: A Four-Billion-Year Journey through Life and Earth History; Springer: Berlin/Heidelberg, Germany, 2019; pp. 269–310. [Google Scholar]
  50. Nevo, R.; Charuvi, D.; Tsabari, O.; Reich, Z. Composition, architecture and dynamics of the photosynthetic apparatus in higher plants. Plant J. 2012, 70, 157–176. [Google Scholar] [CrossRef]
  51. Sekar, N.; Ramasamy, R.P. Recent advances in photosynthetic energy conversion. J. Photochem. Photobiol. C Photochem. Rev. 2015, 22, 19–33. [Google Scholar] [CrossRef]
  52. Assunção, J.; Amaro, H.M.; Malcata, F.X.; Guedes, A.C. Cyanobacterial pigments: Photosynthetic function and biotechnological purposes. In The Pharmacological Potential of Cyanobacteria; Elsevier: Amsterdam, The Netherlands, 2022; pp. 201–256. [Google Scholar]
  53. Johnson, M.P.; Wientjes, E. The relevance of dynamic thylakoid organisation to photosynthetic regulation. Biochim. Biophys. Acta (BBA)-Bioenerg. 2020, 1861, 148039. [Google Scholar] [CrossRef] [PubMed]
  54. Knoll, A.H. Cyanobacteria and earth history. In The Cyanobacteria: Molecular Biology, Genomics, and Evolution; Caister Academic Press: Wymondham, UK, 2008; Volume 484. [Google Scholar]
  55. Cole, J.J.; Hararuk, O.; Solomon, C.T. The carbon cycle: With a brief introduction to global biogeochemistry. In Fundamentals of Ecosystem Science; Elsevier: Amsterdam, The Netherlands, 2021; pp. 131–160. [Google Scholar]
  56. Gaysina, L.A.; Saraf, A.; Singh, P. Cyanobacteria in diverse habitats. In Cyanobacteria; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–28. [Google Scholar]
  57. Fischer, W.W.; Hemp, J.; Valentine, J.S. How did life survive Earth’s great oxygenation? Curr. Opin. Chem. Biol. 2016, 31, 166–178. [Google Scholar] [CrossRef]
  58. Dextro, R.B.; Andreote, A.P.; Vaz, M.G.; Carvalho, C.R.; Fiore, M.F. The pros and cons of axenic cultures in cyanobacterial research. Algal Res. 2024, 78, 103415. [Google Scholar] [CrossRef]
  59. Singhal, R.K.; Fahad, S.; Kumar, P.; Choyal, P.; Javed, T.; Jinger, D.; Singh, P.; Saha, D.; Md, P.; Bose, B. Beneficial elements: New Players in improving nutrient use efficiency and abiotic stress tolerance. Plant Growth Regul. 2023, 100, 237–265. [Google Scholar] [CrossRef]
  60. Wehr, J.; van Vuuren, S.J. Algae and Cyanobacteria Communities. In Wetzel’s Limnology; Elsevier: Amsterdam, The Netherlands, 2024; pp. 463–510. [Google Scholar]
  61. Rabalais, N.N. Nitrogen in aquatic ecosystems. AMBIO J. Hum. Environ. 2002, 31, 102–112. [Google Scholar] [CrossRef] [PubMed]
  62. Carey, C.C.; Ibelings, B.W.; Hoffmann, E.P.; Hamilton, D.P.; Brookes, J.D. Eco-physiological adaptations that favour freshwater cyanobacteria in a changing climate. Water Res. 2012, 46, 1394–1407. [Google Scholar] [CrossRef] [PubMed]
  63. Rodríguez-Zúñiga, D.; Méndez-Zavala, A.; Solís-Quiroz, O.; Morales-Oyervides, L.; Montañez-Saénz, J.C.; Benavente-Valdés, J.R. Biology and composition of microalgae and cyanobacteria. In Sustainable Industrial Processes Based on Microalgae; Elsevier: Amsterdam, The Netherlands, 2024; pp. 1–22. [Google Scholar]
  64. Creed, I.F.; Bergström, A.K.; Trick, C.G.; Grimm, N.B.; Hessen, D.O.; Karlsson, J.; Kidd, K.A.; Kritzberg, E.; McKnight, D.M.; Freeman, E.C. Global change-driven effects on dissolved organic matter composition: Implications for food webs of northern lakes. Glob. Change Biol. 2018, 24, 3692–3714. [Google Scholar] [CrossRef] [PubMed]
  65. Hagemann, M.; Hess, W.R. Systems and synthetic biology for the biotechnological application of cyanobacteria. Curr. Opin. Biotechnol. 2018, 49, 94–99. [Google Scholar] [CrossRef]
  66. Shih, P.M.; Wu, D.; Latifi, A.; Axen, S.D.; Fewer, D.P.; Talla, E.; Calteau, A.; Cai, F.; Tandeau de Marsac, N.; Rippka, R. Improving the coverage of the cyanobacterial phylum using diversity-driven genome sequencing. Proc. Natl. Acad. Sci. USA 2013, 110, 1053–1058. [Google Scholar] [CrossRef]
  67. Oliver, R.L.; Hamilton, D.P.; Brookes, J.D.; Ganf, G.G. Physiology, blooms and prediction of planktonic cyanobacteria. In Ecology of Cyanobacteria II: Their Diversity in Space and Time; Springer: Berlin/Heidelberg, Germany, 2012; pp. 155–194. [Google Scholar]
  68. Zahra, Z.; Choo, D.H.; Lee, H.; Parveen, A. Cyanobacteria: Review of current potentials and applications. Environments 2020, 7, 13. [Google Scholar] [CrossRef]
  69. Singh, R.; Parihar, P.; Singh, M.; Bajguz, A.; Kumar, J.; Singh, S.; Singh, V.P.; Prasad, S.M. Uncovering potential applications of cyanobacteria and algal metabolites in biology, agriculture and medicine: Current status and future prospects. Front. Microbiol. 2017, 8, 515. [Google Scholar] [CrossRef] [PubMed]
  70. Turnau, K.; Płachno, B.J.; Bień, P.; Świątek, P.; Dąbrowski, P.; Kalaji, H. Fungal symbionts impact cyanobacterial biofilm durability and photosynthetic efficiency. Curr. Biol. 2023, 33, 5257–5262.e3. [Google Scholar] [CrossRef] [PubMed]
  71. Mishra, A.K.; Tiwari, D.; Rai, A.N. Cyanobacteria: From Basic Science to Applications; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  72. Butman, D.; Stackpoole, S.; Stets, E.; McDonald, C.P.; Clow, D.W.; Striegl, R.G. Aquatic carbon cycling in the conterminous United States and implications for terrestrial carbon accounting. Proc. Natl. Acad. Sci. USA 2016, 113, 58–63. [Google Scholar] [CrossRef] [PubMed]
  73. Simon, R.D. Macromolecular composition of spores from the filamentous cyanobacterium Anabaena cylindrica. J. Bacteriol. 1977, 129, 1154–1155. [Google Scholar] [CrossRef] [PubMed]
  74. Lynn, M.E.; Bantle, J.A.; Ownby, J.D. Estimation of gene expression in heterocysts of Anabaena variabilis by using DNA-RNA hybridization. J. Bacteriol. 1986, 167, 940–946. [Google Scholar] [CrossRef]
  75. Kurmayer, R.; Kutzenberger, T. Application of real-time PCR for quantification of microcystin genotypes in a population of the toxic cyanobacterium Microcystis sp. Appl. Environ. Microbiol. 2003, 69, 6723–6730. [Google Scholar] [CrossRef]
  76. Hu, B.; Yang, G.; Zhao, W.; Zhang, Y.; Zhao, J. MreB is important for cell shape but not for chromosome segregation of the filamentous cyanobacterium Anabaena sp. PCC 7120. Mol. Microbiol. 2007, 63, 1640–1652. [Google Scholar] [CrossRef]
  77. Vaulot, D.; Marie, D.; Olson, R.J.; Chisholm, S.W. Growth of Prochlorococcus, a photosynthetic prokaryote, in the equatorial Pacific Ocean. Science 1995, 268, 1480–1482. [Google Scholar] [CrossRef]
  78. Griese, M.; Lange, C.; Soppa, J. Ploidy in cyanobacteria. FEMS Microbiol. Lett. 2011, 323, 124–131. [Google Scholar] [CrossRef]
  79. Palacio, A.S.; Cabello, A.M.; García, F.C.; Labban, A.; Morán, X.A.G.; Garczarek, L.; Alonso-Sáez, L.; López-Urrutia, Á. Changes in population age-structure obscure the temperature-size rule in marine cyanobacteria. Front. Microbiol. 2020, 11, 2059. [Google Scholar] [CrossRef] [PubMed]
  80. Binder, B.J.; Chisholm, S.W. Cell cycle regulation in marine Synechococcus sp. strains. Appl. Environ. Microbiol. 1995, 61, 708–717. [Google Scholar] [CrossRef] [PubMed]
  81. Armbrust, E.; Bowen, J.; Olson, R.; Chisholm, S. Effect of light on the cell cycle of a marine Synechococcus strain. Appl. Environ. Microbiol. 1989, 55, 425–432. [Google Scholar] [CrossRef] [PubMed]
  82. Labarre, J.; Chauvat, F.; Thuriaux, P. Insertional mutagenesis by random cloning of antibiotic resistance genes into the genome of the cyanobacterium Synechocystis strain PCC 6803. J. Bacteriol. 1989, 171, 3449–3457. [Google Scholar] [CrossRef]
  83. Sargent, E.C.; Hitchcock, A.; Johansson, S.A.; Langlois, R.; Moore, C.M.; LaRoche, J.; Poulton, A.J.; Bibby, T.S. Evidence for polyploidy in the globally important diazotroph Trichodesmium. FEMS Microbiol. Lett. 2016, 363, fnw244. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, H.; Blankenship, R.E. On the interface of light-harvesting antenna complexes and reaction centers in oxygenic photosynthesis. Biochim. Biophys. Acta (BBA)-Bioenerg. 2019, 1860, 148079. [Google Scholar] [CrossRef]
  85. Zhou, Y.; vom Dorp, K.; Dörman, P.; Hölzl, G. Chloroplast lipids. In Chloroplasts. Current Research and Future Trends; Caister Academic Press: Wymondham, UK, 2016; pp. 1–24. [Google Scholar]
  86. Yang, Z.; Su, X.; Wu, F.; Gong, Y.; Kuang, T. Photochemical activities of plant photosystem I particles reconstituted into phosphatidylglycerol liposomes. J. Photochem. Photobiol. B Biol. 2005, 78, 125–134. [Google Scholar] [CrossRef]
  87. Duffy, C.; Valkunas, L.; Ruban, A. Light-harvesting processes in the dynamic photosynthetic antenna. Phys. Chem. Chem. Phys. 2013, 15, 18752–18770. [Google Scholar] [CrossRef]
  88. Puthiyaveetil, S.; Kirchhoff, H.; Höhner, R. Structural and functional dynamics of the thylakoid membrane system. In Chloroplasts: Current Research and Future Trends; Caister Academic Press: Wymondham, UK, 2016; pp. 59–88. [Google Scholar]
  89. Kobayashi, K. Role of membrane glycerolipids in photosynthesis, thylakoid biogenesis and chloroplast development. J. Plant Res. 2016, 129, 565–580. [Google Scholar] [CrossRef]
  90. Yip, L.X.; Wang, J.; Xue, Y.; Xing, K.; Sevencan, C.; Ariga, K.; Leong, D.T. Cell-derived nanomaterials for biomedical applications. Sci. Technol. Adv. Mater. 2024, 25, 2315013. [Google Scholar] [CrossRef] [PubMed]
  91. Allakhverdiev, S.I. Optimising photosynthesis for environmental fitness. Funct. Plant Biol. 2020, 47, iii–vii. [Google Scholar] [CrossRef] [PubMed]
  92. Huokko, T.; Ni, T.; Dykes, G.F.; Simpson, D.M.; Brownridge, P.; Conradi, F.D.; Beynon, R.J.; Nixon, P.J.; Mullineaux, C.W.; Zhang, P. Probing the biogenesis pathway and dynamics of thylakoid membranes. Nat. Commun. 2021, 12, 3475. [Google Scholar] [CrossRef] [PubMed]
  93. Douchi, D.; Si Larbi, G.; Fel, B.; Bonnanfant, M.; Louwagie, M.; Jouhet, J.; Agnely, M.; Pouget, S.; Maréchal, E. Dryland endolithic Chroococcidiopsis and temperate fresh water Synechocystis have distinct membrane lipid and photosynthesis acclimation strategies upon desiccation and temperature increase. Plant Cell Physiol. 2023, 65, pcad139. [Google Scholar] [CrossRef] [PubMed]
  94. He, M.; Ding, N.-Z. Plant unsaturated fatty acids: Multiple roles in stress response. Front. Plant Sci. 2020, 11, 562785. [Google Scholar] [CrossRef] [PubMed]
  95. Nawaz, T.; Saud, S.; Gu, L.; Khan, I.; Fahad, S.; Zhou, R. Cyanobacteria: Harnessing the Power of Microorganisms for Plant Growth Promotion, Stress Alleviation, and Phytoremediation in the Era of Sustainable Agriculture. Plant Stress 2024, 11, 100399. [Google Scholar] [CrossRef]
  96. Latifi, A.; Ruiz, M.; Zhang, C.-C. Oxidative stress in cyanobacteria. FEMS Microbiol. Rev. 2009, 33, 258–278. [Google Scholar] [CrossRef] [PubMed]
  97. Singh, S.C.; Sinha, R.P.; Hader, D.-P. Role of lipids and fatty acids in stress tolerance in cyanobacteria. Acta Protozool. 2002, 41, 297–308. [Google Scholar]
  98. Marchetto, F.; Santaeufemia, S.; Lebiedzińska-Arciszewska, M.; Śliwińska, M.A.; Pich, M.; Kurek, E.; Naziębło, A.; Strawski, M.; Solymosi, D.; Szklarczyk, M. Dynamic adaptation of the extremophilic red microalga Cyanidioschyzon merolae to high nickel stress. Plant Physiol. Biochem. 2024, 207, 108365. [Google Scholar] [CrossRef] [PubMed]
  99. Awais, M.; Javaid, A.; Khan, I. Study of biological processes using physical principles (eg, protein folding, molecular interactions). Worldw. J. Phys. 2022, 3, 7–15. [Google Scholar]
  100. Zorz, J. An Integrated Approach to Improving Efficiency in Photosynthetic Microbial Systems. Ph.D. Thesis, University of Calgary, Calgary, AB, Canada, 2021. [Google Scholar]
  101. Singh, S.; Morya, R.; Jaiswal, D.K.; Keerthana, S.; Kim, S.-H.; Manimekalai, R.; de Araujo Pereira, A.P.; Verma, J.P. Innovations and advances in enzymatic deconstruction of biomass and their sustainability analysis: A review. Renew. Sustain. Energy Rev. 2024, 189, 113958. [Google Scholar] [CrossRef]
  102. Lu, X. A perspective: Photosynthetic production of fatty acid-based biofuels in genetically engineered cyanobacteria. Biotechnol. Adv. 2010, 28, 742–746. [Google Scholar] [CrossRef] [PubMed]
  103. Hu, Q.; Sommerfeld, M.; Jarvis, E.; Ghirardi, M.; Posewitz, M.; Seibert, M.; Darzins, A. Microalgal triacylglycerols as feedstocks for biofuel production: Perspectives and advances. Plant J. 2008, 54, 621–639. [Google Scholar] [CrossRef] [PubMed]
  104. Eungrasamee, K.; Zhu, Z.; Liu, X.; Jantaro, S.; Lindblad, P. Lipid metabolism in cyanobacteria: Biosynthesis and utilization. In Cyanobacteria; Elsevier: Amsterdam, The Netherlands, 2024; pp. 85–116. [Google Scholar]
  105. Srirangan, K.; Pyne, M.E.; Chou, C.P. Biochemical and genetic engineering strategies to enhance hydrogen production in photosynthetic algae and cyanobacteria. Bioresour. Technol. 2011, 102, 8589–8604. [Google Scholar] [CrossRef] [PubMed]
  106. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  107. Sharma, N.K.; Stal, L.J. The economics of cyanobacteria-based biofuel production: Challenges and opportunities. In Cyanobacteria: An Economic Perspective; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 167–180. [Google Scholar]
  108. Ambrosio, R.; Rizza, L.S.; Do Nascimento, M.; Pacheco, H.G.J.; Ramos, L.M.M.; Hernandez, J.A.; Curatti, L. Promises and challenges for expanding the use of N2-fixing cyanobacteria as a fertilizer for sustainable agriculture. In Cyanobacterial Lifestyle and Its Applications in Biotechnology; Elsevier: Amsterdam, The Netherlands, 2022; pp. 99–158. [Google Scholar]
  109. Mazard, S.; Penesyan, A.; Ostrowski, M.; Paulsen, I.T.; Egan, S. Tiny microbes with a big impact: The role of cyanobacteria and their metabolites in shaping our future. Mar. Drugs 2016, 14, 97. [Google Scholar] [CrossRef]
  110. Weigmann, K. Fixing carbon: To alleviate climate change, scientists are exploring ways to harness nature’s ability to capture CO2 from the atmosphere. EMBO Rep. 2019, 20, e47580. [Google Scholar] [CrossRef]
  111. Johnson, T.J.; Katuwal, S.; Anderson, G.A.; Gu, L.; Zhou, R.; Gibbons, W.R. Photobioreactor cultivation strategies for microalgae and cyanobacteria. Biotechnol. Prog. 2018, 34, 811–827. [Google Scholar] [CrossRef]
  112. Ambaye, T.G.; Vaccari, M.; Bonilla-Petriciolet, A.; Prasad, S.; van Hullebusch, E.D.; Rtimi, S. Emerging technologies for biofuel production: A critical review on recent progress, challenges and perspectives. J. Environ. Manag. 2021, 290, 112627. [Google Scholar] [CrossRef]
  113. Çelekli, A.; Zariç, Ö.E. Plasma-Enhanced Microalgal Cultivation: A Sustainable Approach for Biofuel and Biomass Production. In Emerging Applications of Plasma Science in Allied Technologies; IGI Global: Hershey, PA, USA, 2024; pp. 243–263. [Google Scholar]
  114. Work, V.H. Metabolic and Physiological Engineering of Photosynthetic Microorganisms for the Synthesis of Bioenergy Feedstocks: Development, Characterization, and Optimization. Ph.D. Thesis, Colorado School of Mines, Golden, CO, USA, 2014. [Google Scholar]
  115. Chisti, Y. Biodiesel from microalgae. Biotechnol. Adv. 2007, 25, 294–306. [Google Scholar] [CrossRef]
  116. Fernández, F.A.; Camacho, F.G.; Chisti, Y. Photobioreactors: Light regime, mass transfer, and scaleup. In Progress in Industrial Microbiology; Elsevier: Amsterdam, The Netherlands, 1999; pp. 231–247. [Google Scholar]
  117. Santos-Merino, M.; Singh, A.K.; Ducat, D.C. New applications of synthetic biology tools for cyanobacterial metabolic engineering. Front. Bioeng. Biotechnol. 2019, 7, 33. [Google Scholar] [CrossRef]
  118. Enamala, M.K.; Enamala, S.; Chavali, M.; Donepudi, J.; Yadavalli, R.; Kolapalli, B.; Aradhyula, T.V.; Velpuri, J.; Kuppam, C. Production of biofuels from microalgae-A review on cultivation, harvesting, lipid extraction, and numerous applications of microalgae. Renew. Sustain. Energy Rev. 2018, 94, 49–68. [Google Scholar] [CrossRef]
  119. Egesa, D.; Plucinski, P. Efficient extraction of lipids from magnetically separated microalgae using ionic liquids and their transesterification to biodiesel. Biomass Convers. Biorefinery 2024, 14, 419–434. [Google Scholar] [CrossRef]
  120. Taher, H.; Al-Zuhair, S.; Al-Marzouqi, A.H.; Haik, Y.; Farid, M.M. A review of enzymatic transesterification of microalgal oil-based biodiesel using supercritical technology. Enzym. Res. 2011, 2011, 468292. [Google Scholar] [CrossRef] [PubMed]
  121. Okoye, P.; Hameed, B. Review on recent progress in catalytic carboxylation and acetylation of glycerol as a byproduct of biodiesel production. Renew. Sustain. Energy Rev. 2016, 53, 558–574. [Google Scholar] [CrossRef]
  122. Gülüm, M.; Yesilyurt, M.K.; Bilgin, A. The modeling and analysis of transesterification reaction conditions in the selection of optimal biodiesel yield and viscosity. Environ. Sci. Pollut. Res. 2020, 27, 10351–10366. [Google Scholar] [CrossRef]
  123. Singh, V.K.; Jha, S.; Rana, P.; Soni, R.; Lalnunpuii, R.; Singh, P.K.; Sinha, R.P.; Singh, G. Cyanobacteria as a Biocatalyst for Sustainable Production of Biofuels and Chemicals. Energies 2024, 17, 408. [Google Scholar] [CrossRef]
  124. Gaurav, K.; Neeti, K.; Singh, R. Microalgae-based biodiesel production and its challenges and future opportunities: A review. Green Technol. Sustain. 2023, 2, 100060. [Google Scholar] [CrossRef]
  125. Jiao, H.; Tsigkou, K.; Elsamahy, T.; Pispas, K.; Sun, J.; Manthos, G.; Schagerl, M.; Sventzouri, E.; Al-Tohamy, R.; Kornaros, M. Recent advances in sustainable hydrogen production from microalgae: Mechanisms, challenges, and future perspectives. Ecotoxicol. Environ. Saf. 2024, 270, 115908. [Google Scholar] [CrossRef]
  126. Salaheldeen, M.; Mariod, A.A.; Aroua, M.K.; Rahman, S.A.; Soudagar, M.E.M.; Fattah, I.R. Current state and perspectives on transesterification of triglycerides for biodiesel production. Catalysts 2021, 11, 1121. [Google Scholar] [CrossRef]
  127. Lin, F.; Xu, M.; Ramasamy, K.K.; Li, Z.; Klinger, J.L.; Schaidle, J.A.; Wang, H. Catalyst deactivation and its mitigation during catalytic conversions of biomass. ACS Catal. 2022, 12, 13555–13599. [Google Scholar] [CrossRef]
  128. Trunschke, A. Prospects and challenges for autonomous catalyst discovery viewed from an experimental perspective. Catal. Sci. Technol. 2022, 12, 3650–3669. [Google Scholar] [CrossRef]
  129. Yahaya, M.S. Development of Solid Acid Catalysts for Biodiesel Production from High Free Fatty Acid Feedstock; University of Malaya: Kuala Lumpur, Malaysia, 2015. [Google Scholar]
  130. Jatoi, A.S.; Hashmi, Z.; Anjum, A.; Bhatti, Z.A.; Siyal, S.H.; Mazari, S.; Akhter, F.; Mubarak, N.; Iqbal, A. Overview of bioelectrochemical approaches for sulfur reduction: Current and future perspectives. Biomass Convers. Biorefinery 2021, 13, 12333–12348. [Google Scholar] [CrossRef]
  131. Modiri, S.; Sharafi, H.; Alidoust, L.; Hajfarajollah, H.; Haghighi, O.; Azarivand, A.; Zamanzadeh, Z.; Zahiri, H.S.; Vali, H.; Noghabi, K.A. Lipid production and mixotrophic growth features of cyanobacterial strains isolated from various aquatic sites. Microbiology 2015, 161 Pt 3, 662–673. [Google Scholar] [CrossRef]
  132. Etim, A.O.; Jisieike, C.F.; Ibrahim, T.H.; Betiku, E. Biodiesel and its properties. In Production of Biodiesel from Non-Edible Sources; Elsevier: Amsterdam, The Netherlands, 2022; pp. 39–79. [Google Scholar]
  133. Atadashi, I.; Aroua, M.; Aziz, A.A. Biodiesel separation and purification: A review. Renew. Energy 2011, 36, 437–443. [Google Scholar] [CrossRef]
  134. Avagyan, A.B.; Singh, B. Biodiesel: Feedstocks, Technologies, Economics and Barriers. Assessment of Environmental Impact in Producing and Using Chains; Springer: Singapore, 2019. [Google Scholar]
  135. Babadi, A.A.; Rahmati, S.; Fakhlaei, R.; Barati, B.; Wang, S.; Doherty, W.; Ostrikov, K.K. Emerging technologies for biodiesel production: Processes, challenges, and opportunities. Biomass Bioenergy 2022, 163, 106521. [Google Scholar] [CrossRef]
  136. Kawale, H.D.; Kishore, N. Production of hydrocarbons from a green algae (Oscillatoria) with exploration of its fuel characteristics over different reaction atmospheres. Energy 2019, 178, 344–355. [Google Scholar] [CrossRef]
  137. Kuppusamy, S.; Maddela, N.R.; Megharaj, M.; Venkateswarlu, K.; Kuppusamy, S.; Maddela, N.R.; Megharaj, M.; Venkateswarlu, K. An overview of total petroleum hydrocarbons. In Total Petroleum Hydrocarbons: Environmental Fate, Toxicity, and Remediation; Springer: Cham, Switzerland, 2020; pp. 1–27. [Google Scholar]
  138. Farrokh, P.; Sheikhpour, M.; Kasaeian, A.; Asadi, H.; Bavandi, R. Cyanobacteria as an eco-friendly resource for biofuel production: A critical review. Biotechnol. Prog. 2019, 35, e2835. [Google Scholar] [CrossRef] [PubMed]
  139. Koller, M.; Salerno, A.; Dias, M.; Reiterer, A.; Braunegg, G. Modern biotechnological polymer synthesis: A review. Food Technol. Biotechnol. 2010, 48, 255–269. [Google Scholar]
  140. Serrano-Ruiz, J.C.; Dumesic, J.A. Catalytic routes for the conversion of biomass into liquid hydrocarbon transportation fuels. Energy Environ. Sci. 2011, 4, 83–99. [Google Scholar] [CrossRef]
  141. Günay, M.E.; Türker, L.; Tapan, N.A. Significant parameters and technological advancements in biodiesel production systems. Fuel 2019, 250, 27–41. [Google Scholar] [CrossRef]
  142. Ndaya, D.M. Biodiesel Production From Candlenut And Calodendrum Capense Seeds: Process Design And Technological Assessment. Ph.D. Thesis, University of Nairobi, Nairobi, Kenya, 2013. [Google Scholar]
  143. Wang, J.; Singer, S.D.; Souto, B.A.; Asomaning, J.; Ullah, A.; Bressler, D.C.; Chen, G. Current progress in lipid-based biofuels: Feedstocks and production technologies. Bioresour. Technol. 2022, 351, 127020. [Google Scholar] [CrossRef] [PubMed]
  144. Pasha, M.K.; Dai, L.; Liu, D.; Guo, M.; Du, W. An overview to process design, simulation and sustainability evaluation of biodiesel production. Biotechnol. Biofuels 2021, 14, 129. [Google Scholar] [CrossRef] [PubMed]
  145. Knight, R.A. Coordinated Response and Regulation of Carotenogenesis in Thermosynechococcus elongatus (BP-1): Implications for Commercial Application. Ph.D. Thesis, The University of Texas at Austin, Austin, TX, USA, 2014. [Google Scholar]
  146. Piloto-Rodríguez, R.; Sánchez-Borroto, Y.; Melo-Espinosa, E.A.; Verhelst, S. Assessment of diesel engine performance when fueled with biodiesel from algae and microalgae: An overview. Renew. Sustain. Energy Rev. 2017, 69, 833–842. [Google Scholar] [CrossRef]
  147. Sitther, V.; Tabatabai, B.; Fathabad, S.G.; Gichuki, S.; Chen, H.; Arumanayagam, A.C.S. Cyanobacteria as a biofuel source: Advances and applications. In Advances in Cyanobacterial Biology; Academic Press: Cambridge, MA, USA, 2020; pp. 269–289. [Google Scholar]
  148. Elgarahy, A.M.; Eloffy, M.; Hammad, A.; Saber, A.N.; El-Sherif, D.M.; Mohsen, A.; Abouzid, M.; Elwakeel, K.Z. Hydrogen production from wastewater, storage, economy, governance and applications: A review. Environ. Chem. Lett. 2022, 20, 3453–3504. [Google Scholar] [CrossRef]
  149. Hassan, H.; Ansari, F.A.; Ingle, K.N.; Singh, K.; Bux, F. Commercial products and environmental benefits of algal diversity. In Biodiversity and Bioeconomy; Elsevier: Amsterdam, The Netherlands, 2024; pp. 475–502. [Google Scholar]
  150. Francisco, E.C.; Neves, D.B.; Jacob-Lopes, E.; Franco, T.T. Microalgae as feedstock for biodiesel production: Carbon dioxide sequestration, lipid production and biofuel quality. J. Chem. Technol. Biotechnol. 2010, 85, 395–403. [Google Scholar] [CrossRef]
  151. Zhang, A.; Carroll, A.L.; Atsumi, S. Carbon recycling by cyanobacteria: Improving CO2 fixation through chemical production. FEMS Microbiol. Lett. 2017, 364, fnx165. [Google Scholar] [CrossRef]
  152. Arias, D.M.; Uggetti, E.; García, J. Assessing the potential of soil cyanobacteria for simultaneous wastewater treatment and carbohydrate-enriched biomass production. Algal Res. 2020, 51, 102042. [Google Scholar] [CrossRef]
  153. Xu, H.; Ou, L.; Li, Y.; Hawkins, T.R.; Wang, M. Life cycle greenhouse gas emissions of biodiesel and renewable diesel production in the United States. Environ. Sci. Technol. 2022, 56, 7512–7521. [Google Scholar] [CrossRef] [PubMed]
  154. Mathimani, T.; Baldinelli, A.; Rajendran, K.; Prabakar, D.; Matheswaran, M.; van Leeuwen, R.P.; Pugazhendhi, A. Review on cultivation and thermochemical conversion of microalgae to fuels and chemicals: Process evaluation and knowledge gaps. J. Clean. Prod. 2019, 208, 1053–1064. [Google Scholar] [CrossRef]
  155. Wang, H.; Peng, X.; Zhang, H.; Yang, S.; Li, H. Microorganisms-promoted biodiesel production from biomass: A review. Energy Convers. Manag. X 2021, 12, 100137. [Google Scholar] [CrossRef]
  156. Yang, X.; Mukherjee, S.; O’Carroll, T.; Hou, Y.; Singh, M.R.; Gauthier, J.A.; Wu, G. Achievements, Challenges, and Perspectives on Nitrogen Electrochemistry for Carbon-Neutral Energy Technologies. Angew. Chem. 2023, 135, e202215938. [Google Scholar] [CrossRef]
  157. Borah, D.; Rout, J.; Nooruddin, T. Phycoremediation and water reuse in bioenergy production from algae and cyanobacteria in relevance to sustainable development goals. In Water, the Environment and the Sustainable Development Goals; Elsevier: Amsterdam, The Netherlands, 2024; pp. 375–406. [Google Scholar]
  158. Gomes Gradíssimo, D.; Pereira Xavier, L.; Valadares Santos, A. Cyanobacterial polyhydroxyalkanoates: A sustainable alternative in circular economy. Molecules 2020, 25, 4331. [Google Scholar] [CrossRef] [PubMed]
  159. Jiang, N.; Wang, M.; Song, L.; Yu, D.; Zhou, S.; Li, Y.; Li, H.; Han, X. Polyhydroxybutyrate production by recombinant Escherichia coli based on genes related to synthesis pathway of PHB from Massilia sp. UMI-21. Microb. Cell Factories 2023, 22, 129. [Google Scholar] [CrossRef] [PubMed]
  160. Zhila, N.O.; Sapozhnikova, K.Y.; Kiselev, E.G.; Shishatskaya, E.I.; Volova, T.G. Biosynthesis of Poly (3-hydroxybutyrate-co-4-hydroxybutyrate) from Different 4-Hydroxybutyrate Precursors by New Wild-Type Strain Cupriavidus necator IBP/SFU-1. Processes 2023, 11, 1423. [Google Scholar] [CrossRef]
  161. Turco, R.; Santagata, G.; Corrado, I.; Pezzella, C.; Di Serio, M. In vivo and post-synthesis strategies to enhance the properties of PHB-based materials: A review. Front. Bioeng. Biotechnol. 2021, 8, 619266. [Google Scholar] [CrossRef]
  162. Zheng, L.Z.; Li, Z.; Tian, H.-L.; Li, M.; Chen, G.-Q. Molecular cloning and functional analysis of (R)-3-hydroxyacyl-acyl carrier protein: Coenzyme A transacylase from Pseudomonas mendocina LZ. FEMS Microbiol. Lett. 2005, 252, 299–307. [Google Scholar] [CrossRef]
  163. Kassab, E.; Mehlmer, N.; Brueck, T. GFP Scaffold-Based Engineering for the Production of Unbranched Very Long Chain Fatty Acids in Escherichia coli With Oleic Acid and Cerulenin Supplementation. Front. Bioeng. Biotechnol. 2019, 7, 408. [Google Scholar] [CrossRef]
  164. Mendhulkar, V.D.; Shetye, L.A. Synthesis of biodegradable polymer polyhydroxyalkanoate (PHA) in cyanobacteria Synechococcus elongates under mixotrophic nitrogen-and phosphate-mediated stress conditions. Ind. Biotechnol. 2017, 13, 85–93. [Google Scholar] [CrossRef]
  165. Riaz, S.; Rhee, K.Y.; Park, S.J. Polyhydroxyalkanoates (PHAs): Biopolymers for biofuel and biorefineries. Polymers 2021, 13, 253. [Google Scholar] [CrossRef] [PubMed]
  166. Zhou, W.; Bergsma, S.; Colpa, D.I.; Euverink, G.-J.W.; Krooneman, J. Polyhydroxyalkanoates (PHAs) synthesis and degradation by microbes and applications towards a circular economy. J. Environ. Manag. 2023, 341, 118033. [Google Scholar] [CrossRef]
  167. Kamravamanesh, D.; Lackner, M.; Herwig, C. Bioprocess engineering aspects of sustainable polyhydroxyalkanoate production in cyanobacteria. Bioengineering 2018, 5, 111. [Google Scholar] [CrossRef]
  168. Troschl, C.; Meixner, K.; Drosg, B. Cyanobacterial PHA production—Review of recent advances and a summary of three years’ working experience running a pilot plant. Bioengineering 2017, 4, 26. [Google Scholar] [CrossRef] [PubMed]
  169. Karan, H.; Funk, C.; Grabert, M.; Oey, M.; Hankamer, B. Green bioplastics as part of a circular bioeconomy. Trends Plant Sci. 2019, 24, 237–249. [Google Scholar] [CrossRef] [PubMed]
  170. Price, S.; Kuzhiumparambil, U.; Pernice, M.; Ralph, P.J. Cyanobacterial polyhydroxybutyrate for sustainable bioplastic production: Critical review and perspectives. J. Environ. Chem. Eng. 2020, 8, 104007. [Google Scholar] [CrossRef]
  171. Yadav, B.; Pandey, A.; Kumar, L.R.; Tyagi, R.D. Bioconversion of waste (water)/residues to bioplastics-A circular bioeconomy approach. Bioresour. Technol. 2020, 298, 122584. [Google Scholar] [CrossRef] [PubMed]
  172. Campbell, A.M.; Paradise, C.J. Photosynthesis; Momentum Press: New York, NY, USA, 2016. [Google Scholar]
  173. Tiwari, S.; Patel, A.; Pandey, N.; Singh, G.; Pandey, A.; Prasad, S.M. Photosynthesis and Energy Flow in Cyanobacteria. In Ecophysiology and Biochemistry of Cyanobacteria; Springer: Berlin/Heidelberg, Germany, 2022; pp. 49–64. [Google Scholar]
  174. Razeghifard, R. Natural and Artificial Photosynthesis: Solar Power as an Energy Source; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  175. Norena-Caro, D.; Benton, M.G. Cyanobacteria as photoautotrophic biofactories of high-value chemicals. J. CO2 Util. 2018, 28, 335–366. [Google Scholar] [CrossRef]
  176. Gayathri, R.; Mahboob, S.; Govindarajan, M.; Al-Ghanim, K.A.; Ahmed, Z.; Al-Mulhm, N.; Vodovnik, M.; Vijayalakshmi, S. A review on biological carbon sequestration: A sustainable solution for a cleaner air environment, less pollution and lower health risks. J. King Saud Univ.-Sci. 2021, 33, 101282. [Google Scholar] [CrossRef]
  177. Tuit, C.; Waterbury, J.; Ravizza, G. Diel variation of molybdenum and iron in marine diazotrophic cyanobacteria. Limnol. Oceanogr. 2004, 49, 978–990. [Google Scholar] [CrossRef]
  178. Khatami, K.; Perez-Zabaleta, M.; Owusu-Agyeman, I.; Cetecioglu, Z. Waste to bioplastics: How close are we to sustainable polyhydroxyalkanoates production? Waste Manag. 2021, 119, 374–388. [Google Scholar] [CrossRef]
  179. Koller, M. A review on established and emerging fermentation schemes for microbial production of polyhydroxyalkanoate (PHA) biopolyesters. Fermentation 2018, 4, 30. [Google Scholar] [CrossRef]
  180. Singh, A.K.; Mallick, N. Advances in cyanobacterial polyhydroxyalkanoates production. FEMS Microbiol. Lett. 2017, 364, fnx189. [Google Scholar] [CrossRef]
  181. Lau, N.-S.; Matsui, M.; Abdullah, A.A.-A. Cyanobacteria: Photoautotrophic microbial factories for the sustainable synthesis of industrial products. BioMed Res. Int. 2015, 2015, 754934. [Google Scholar] [CrossRef]
  182. Metsoviti, M.N.; Papapolymerou, G.; Karapanagiotidis, I.T.; Katsoulas, N. Effect of light intensity and quality on growth rate and composition of Chlorella vulgaris. Plants 2019, 9, 31. [Google Scholar] [CrossRef] [PubMed]
  183. Chen, M.-Y.; Teng, W.-K.; Zhao, L.; Hu, C.-X.; Zhou, Y.-K.; Han, B.-P.; Song, L.-R.; Shu, W.-S. Comparative genomics reveals insights into cyanobacterial evolution and habitat adaptation. ISME J. 2021, 15, 211–227. [Google Scholar] [CrossRef]
  184. Bugnicourt, E.; Cinelli, P.; Lazzeri, A.; Alvarez, V.A. Polyhydroxyalkanoate (PHA): Review of synthesis, characteristics, processing and potential applications in packaging. Express Polym. Lett. 2014, 8, 791–808. [Google Scholar] [CrossRef]
  185. Rujnić-Sokele, M.; Pilipović, A. Challenges and opportunities of biodegradable plastics: A mini review. Waste Manag. Res. 2017, 35, 132–140. [Google Scholar] [CrossRef] [PubMed]
  186. Singh, J.S.; Kumar, A.; Rai, A.N.; Singh, D.P. Cyanobacteria: A precious bio-resource in agriculture, ecosystem, and environmental sustainability. Front. Microbiol. 2016, 7, 529. [Google Scholar] [CrossRef] [PubMed]
  187. Das, M.; Maiti, S.K. Recent progress and challenges in cyanobacterial autotrophic production of polyhydroxybutyrate (PHB), a bioplastic. J. Environ. Chem. Eng. 2021, 9, 105379. [Google Scholar]
  188. Pandey, A.; Adama, N.; Adjallé, K.; Blais, J.-F. Sustainable applications of polyhydroxyalkanoates in various fields: A critical review. Int. J. Biol. Macromol. 2022, 221, 1184–1201. [Google Scholar] [CrossRef]
  189. Mukherjee, A.; Koller, M. Microbial polyHydroxyAlkanoate (PHA) biopolymers—Intrinsically natural. Bioengineering 2023, 10, 855. [Google Scholar] [CrossRef]
  190. Law, K.L.; Narayan, R. Reducing environmental plastic pollution by designing polymer materials for managed end-of-life. Nat. Rev. Mater. 2022, 7, 104–116. [Google Scholar] [CrossRef]
  191. Koller, M. “Bioplastics from microalgae”—Polyhydroxyalkanoate production by cyanobacteria. In Handbook of Microalgae-Based Processes and Products; Academic Press: Cambridge, MA, USA, 2020; pp. 597–645. [Google Scholar]
  192. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The role of renewable energy in the global energy transformation. Energy Strategy Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  193. Adeleye, A.T.; Odoh, C.K.; Enudi, O.C.; Banjoko, O.O.; Osiboye, O.O.; Odediran, E.T.; Louis, H. Sustainable synthesis and applications of polyhydroxyalkanoates (PHAs) from biomass. Process Biochem. 2020, 96, 174–193. [Google Scholar] [CrossRef]
  194. Farrelly, D.J.; Everard, C.D.; Fagan, C.C.; McDonnell, K.P. Carbon sequestration and the role of biological carbon mitigation: A review. Renew. Sustain. Energy Rev. 2013, 21, 712–727. [Google Scholar] [CrossRef]
  195. Muradov, N.; Muradov, N. Industrial utilization of CO2: A win–win solution. In Liberating Energy from Carbon: Introduction to Decarbonization; Springer: New York, NY, USA, 2014; pp. 325–383. [Google Scholar]
  196. Ali, S.; Isha; Chang, Y. -C. Ecotoxicological Impact of Bioplastics Biodegradation: A Comprehensive Review. Processes 2023, 11, 3445. [Google Scholar] [CrossRef]
  197. Obruča, S.; Dvořák, P.; Sedláček, P.; Koller, M.; Sedlář, K.; Pernicová, I.; Šafránek, D. Polyhydroxyalkanoates synthesis by halophiles and thermophiles: Towards sustainable production of microbial bioplastics. Biotechnol. Adv. 2022, 58, 107906. [Google Scholar] [CrossRef]
  198. Nikolaivits, E.; Pantelic, B.; Azeem, M.; Taxeidis, G.; Babu, R.; Topakas, E.; Brennan Fournet, M.; Nikodinovic-Runic, J. Progressing plastics circularity: A review of mechano-biocatalytic approaches for waste plastic (re) valorization. Front. Bioeng. Biotechnol. 2021, 9, 696040. [Google Scholar] [CrossRef] [PubMed]
  199. Rasul, I.; Azeem, F.; Siddique, M.H.; Muzammil, S.; Rasul, A.; Munawar, A.; Afzal, M.; Ali, M.A.; Nadeem, H. Algae biotechnology: A green light for engineered algae. In Algae Based Polymers, Blends, and Composites; Elsevier: Amsterdam, The Netherlands, 2017; pp. 301–334. [Google Scholar]
  200. Iles, A.; Martin, A.; Rosen, C.M. Undoing Chemical Industry Lock-ins: Polyvinyl Chloride and Green Chemistry. In Ethics of Chemistry: From Poison Gas to Climate Engineering; World Scientific: Singapore, 2021; pp. 279–316. [Google Scholar]
  201. Meereboer, K.W.; Misra, M.; Mohanty, A.K. Review of recent advances in the biodegradability of polyhydroxyalkanoate (PHA) bioplastics and their composites. Green Chem. 2020, 22, 5519–5558. [Google Scholar] [CrossRef]
  202. Agboola, O.; Babatunde, D.E.; Fayomi, O.S.I.; Sadiku, E.R.; Popoola, P.; Moropeng, L.; Yahaya, A.; Mamudu, O.A. A review on the impact of mining operation: Monitoring, assessment and management. Results Eng. 2020, 8, 100181. [Google Scholar] [CrossRef]
  203. Blaser, K.K. The Triple Bottom Line of Sustainable Fashion. Master’s Thesis, Ohio University, Athens, OH, USA, 2017. [Google Scholar]
  204. Jacob-Lopes, E.; Zepka, L.Q.; Queiroz, M.I. Cyanobacteria and carbon sequestration. In Cyanobacteria: An Economic Perspective; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 65–71. [Google Scholar]
  205. Andrady, A.L. Plastics and Environmental Sustainability; John Wiley & Sons: Hoboken, NJ, USA, 2015. [Google Scholar]
  206. Ajao, V.O. Biomaterials from Renewable Sources: Biosurfactants and Biopolymers. Master’s Thesis, Universidade do Algarve, Faro, Portugal, 2015. [Google Scholar]
  207. Mah, A. Plastic Unlimited: How Corporations Are Fuelling the Ecological Crisis and What We Can Do about It; John Wiley & Sons: Hoboken, NJ, USA, 2022. [Google Scholar]
  208. Kandpal, V.; Jaswal, A.; Santibanez Gonzalez, E.D.; Agarwal, N. Circular Economy Principles: Shifting Towards Sustainable Prosperity. In Sustainable Energy Transition: Circular Economy and Sustainable Financing for Environmental, Social and Governance (ESG) Practices; Springer: Berlin/Heidelberg, Germany, 2024; pp. 125–165. [Google Scholar]
  209. Kourmentza, C.; Plácido, J.; Venetsaneas, N.; Burniol-Figols, A.; Varrone, C.; Gavala, H.N.; Reis, M.A. Recent advances and challenges towards sustainable polyhydroxyalkanoate (PHA) production. Bioengineering 2017, 4, 55. [Google Scholar] [CrossRef]
  210. Poltronieri, P.; Kumar, P. Polyhydroxyalkanoates (PHAs) in industrial applications. Handb. Ecomater. 2017, 4, 2843–2872. [Google Scholar]
  211. Raza, Z.A.; Abid, S.; Banat, I.M. Polyhydroxyalkanoates: Characteristics, production, recent developments and applications. Int. Biodeterior. Biodegrad. 2018, 126, 45–56. [Google Scholar] [CrossRef]
  212. Phung Hai, T.A.; Tessman, M.; Neelakantan, N.; Samoylov, A.A.; Ito, Y.; Rajput, B.S.; Pourahmady, N.; Burkart, M.D. Renewable polyurethanes from sustainable biological precursors. Biomacromolecules 2021, 22, 1770–1794. [Google Scholar] [CrossRef] [PubMed]
  213. Vu, H.P.; Nguyen, L.N.; Zdarta, J.; Nga, T.T.; Nghiem, L.D. Blue-green algae in surface water: Problems and opportunities. Curr. Pollut. Rep. 2020, 6, 105–122. [Google Scholar] [CrossRef]
  214. Dick, G.J.; Grim, S.L.; Klatt, J.M. Controls on O2 production in cyanobacterial mats and implications for Earth’s oxygenation. Annu. Rev. Earth Planet. Sci. 2018, 46, 123–147. [Google Scholar] [CrossRef]
  215. Xu, Z.; Jiang, Y.; Zhou, G. Response and adaptation of photosynthesis, respiration, and antioxidant systems to elevated CO2 with environmental stress in plants. Front. Plant Sci. 2015, 6, 701. [Google Scholar] [CrossRef]
  216. Singh, S.K.; Sundaram, S.; Kishor, K.; Singh, S.K.; Sundaram, S.; Kishor, K. Photosynthetic microorganism-based CO2 mitigation system: Integrated approaches for global sustainability. Photosynthetic Microorganisms: Mechanism For Carbon Concentration; Springer: Cham, Switzerland, 2014; pp. 83–123. [Google Scholar]
  217. Paul, S.; Parvez, S.S.; Goswami, A.; Banik, A. Exopolysaccharides from agriculturally important microorganisms: Conferring soil nutrient status and plant health. Int. J. Biol. Macromol. 2024, 262, 129954. [Google Scholar] [CrossRef]
  218. Kang, Y.-H.; Park, C.-S.; Han, M.-S. Pseudomonas aeruginosa UCBPP-PA14 a useful bacterium capable of lysing Microcystis aeruginosa cells and degrading microcystins. J. Appl. Phycol. 2012, 24, 1517–1525. [Google Scholar] [CrossRef]
  219. Monib, A.W.; Niazi, P.; Barai, S.M.; Sawicka, B.; Baseer, A.Q.; Nikpay, A.; Fahmawi, S.M.S.; Singh, D.; Alikhail, M.; Thea, B. Nitrogen Cycling Dynamics: Investigating Volatilization and its Interplay with N2 Fixation. J. Res. Appl. Sci. Biotechnol. 2024, 3, 17–31. [Google Scholar] [CrossRef]
  220. Gholamhosseinian, A.; Sepehr, A.; Asgari Lajayer, B.; Delangiz, N.; Astatkie, T. Biological soil crusts to keep soil alive, rehabilitate degraded soil, and develop soil habitats. In Microbial Polymers: Applications and Ecological Perspectives; Springer: Singapore, 2021; pp. 289–309. [Google Scholar]
  221. Mona, S.; Kumar, V.; Deepak, B.; Kaushik, A. Cyanobacteria: The eco-friendly tool for the treatment of industrial wastewaters. In Bioremediation of Industrial Waste for Environmental Safety: Volume II: Biological Agents and Methods for Industrial Waste Management; Springer: Singapore, 2020; pp. 389–413. [Google Scholar]
  222. Sadvakasova, A.K.; Kossalbayev, B.D.; Zayadan, B.K.; Kirbayeva, D.K.; Alwasel, S.; Allakhverdiev, S.I. Potential of cyanobacteria in the conversion of wastewater to biofuels. World J. Microbiol. Biotechnol. 2021, 37, 140. [Google Scholar] [CrossRef]
  223. Ullah, H.; Nagelkerken, I.; Goldenberg, S.U.; Fordham, D.A. Climate change could drive marine food web collapse through altered trophic flows and cyanobacterial proliferation. PLoS Biol. 2018, 16, e2003446. [Google Scholar] [CrossRef] [PubMed]
  224. Wicker, R.J. Bioprospecting for Nordic Photosynthetic Consortia for Wastewater Treatment, Carbon Capture, and Value Creation. Ph.D. Thesis, Lappeenranta-Lahti University of Technology LUT, Lappeenranta, Finland, 2023. [Google Scholar]
  225. Boonbangkeng, D.; Thiemsorn, W.; Ruangrit, K.; Pekkoh, J. Promoting the simultaneous removal of Microcystis bloom and microcystin-RR by Bacillus sp. AK3 immobilized on floating porous glass pellets. J. Appl. Phycol. 2022, 34, 1513–1525. [Google Scholar] [CrossRef]
  226. Krishnan, A.; Zhang, Y.-Q.; Mou, X. Isolation and characterization of microcystin-degrading bacteria from Lake Erie. Bull. Environ. Contam. Toxicol. 2018, 101, 617–623. [Google Scholar] [CrossRef] [PubMed]
  227. Shen, R.; Chen, Z.; Dong, X.; Shen, H.; Su, P.; Mao, L.; Zhang, W. Biodegradation kinetics of microcystins-LR crude extract by Lysinibacillus boronitolerans strain CQ5. Ann. Microbiol. 2019, 69, 1259–1266. [Google Scholar] [CrossRef]
  228. Lawton, L.; Welgamage, A.; Manage, P.; Edwards, C. Novel bacterial strains for the removal of microcystins from drinking water. Water Sci. Technol. 2011, 63, 1137–1142. [Google Scholar] [CrossRef]
  229. Manage, P.M.; Edwards, C.; Singh, B.K.; Lawton, L.A. Isolation and identification of novel microcystin-degrading bacteria. Appl. Environ. Microbiol. 2009, 75, 6924–6928. [Google Scholar] [CrossRef] [PubMed]
  230. Benegas, G.R.S.; Bernal, S.P.F.; de Oliveira, V.M.; Passarini, M.R.Z. Antimicrobial activity against Microcystis aeruginosa and degradation of microcystin-LR by bacteria isolated from Antarctica. Environ. Sci. Pollut. Res. 2021, 28, 52381–52391. [Google Scholar] [CrossRef]
  231. Nybom, S.M.; Salminen, S.J.; Meriluoto, J.A. Removal of microcystin-LR by strains of metabolically active probiotic bacteria. FEMS Microbiol. Lett. 2007, 270, 27–33. [Google Scholar] [CrossRef]
  232. Jia, Y.; Du, J.; Song, F.; Zhao, G.; Tian, X. A fungus capable of degrading microcystin-LR in the algal culture of Microcystis aeruginosa PCC7806. Appl. Biochem. Biotechnol. 2012, 166, 987–996. [Google Scholar] [CrossRef]
  233. Jia, Y.; Wang, C.; Zhao, G.; Guo, P.; Tian, X. The possibility of using cyanobacterial bloom materials as a medium for white rot fungi. Lett. Appl. Microbiol. 2012, 54, 96–101. [Google Scholar] [CrossRef]
  234. Bawazeer, S.; Rauf, A.; Nawaz, T.; Khalil, A.A.; Javed, M.S.; Muhammad, N.; Shah, M.A. Punica granatum peel extracts mediated the green synthesis of gold nanoparticles and their detailed in vivo biological activities. Green Process. Synth. 2021, 10, 882–892. [Google Scholar] [CrossRef]
  235. Esterhuizen-Londt, M.; Hertel, S.; Pflugmacher, S. Uptake and biotransformation of pure commercial microcystin-LR versus microcystin-LR from a natural cyanobacterial bloom extract in the aquatic fungus Mucor hiemalis. Biotechnol. Lett. 2017, 39, 1537–1545. [Google Scholar] [CrossRef] [PubMed]
  236. Mohamed, Z.A.; Alamri, S.; Hashem, M.; Mostafa, Y. Growth inhibition of Microcystis aeruginosa and adsorption of microcystin toxin by the yeast Aureobasidium pullulans, with no effect on microalgae. Environ. Sci. Pollut. Res. 2020, 27, 38038–38046. [Google Scholar] [CrossRef] [PubMed]
  237. Valeria, A.M.; Ricardo, E.J.; Stephan, P.; Alberto, W.D. Degradation of microcystin-RR by Sphingomonas sp. CBA4 isolated from San Roque reservoir (Córdoba–Argentina). Biodegradation 2006, 17, 447–455. [Google Scholar] [CrossRef] [PubMed]
  238. Jones, G.J.; Bourne, D.G.; Blakeley, R.L.; Doelle, H. Degradation of the cyanobacterial hepatotoxin microcystin by aquatic bacteria. Nat. Toxins 1994, 2, 228–235. [Google Scholar] [CrossRef]
  239. Tsuji, K.; Asakawa, M.; Anzai, Y.; Sumino, T.; Harada, K.-i. Degradation of microcystins using immobilized microorganism isolated in an eutrophic lake. Chemosphere 2006, 65, 117–124. [Google Scholar] [CrossRef] [PubMed]
  240. Imanishi, S.; Kato, H.; Mizuno, M.; Tsuji, K.; Harada, K.-i. Bacterial degradation of microcystins and nodularin. Chem. Res. Toxicol. 2005, 18, 591–598. [Google Scholar] [CrossRef]
  241. Somdee, T.; Thunders, M.; Ruck, J.; Lys, I.; Allison, M.; Page, R. Degradation of [Dha7] MC-LR by a Microcystin Degrading Bacterium Isolated from Lake Rotoiti, New Zealand. Int. Sch. Res. Not. 2013, 2013, 596429. [Google Scholar] [CrossRef]
  242. Ishii, H.; Nishijima, M.; Abe, T. Characterization of degradation process of cyanobacterial hepatotoxins by a gram-negative aerobic bacterium. Water Res. 2004, 38, 2667–2676. [Google Scholar] [CrossRef]
  243. Park, H.D.; Sasaki, Y.; Maruyama, T.; Yanagisawa, E.; Hiraishi, A.; Kato, K. Degradation of the cyanobacterial hepatotoxin microcystin by a new bacterium isolated from a hypertrophic lake. Environ. Toxicol. Int. J. 2001, 16, 337–343. [Google Scholar] [CrossRef]
  244. Zhang, J.; Wei, J.; Massey, I.Y.; Peng, T.; Yang, F. Immobilization of Microbes for Biodegradation of Microcystins: A Mini Review. Toxins 2022, 14, 573. [Google Scholar] [CrossRef]
  245. Zeng, Y.-H.; Cai, Z.-H.; Zhu, J.-M.; Du, X.-P.; Zhou, J. Two hierarchical LuxR-LuxI type quorum sensing systems in Novosphingobium activate microcystin degradation through transcriptional regulation of the mlr pathway. Water Res. 2020, 183, 116092. [Google Scholar] [CrossRef] [PubMed]
  246. Phujomjai, Y.; Somdee, A.; Somdee, T. Biodegradation of microcystin [Dha7] MC-LR by a novel microcystin-degrading bacterium in an internal airlift loop bioreactor. Water Sci. Technol. 2016, 73, 267–274. [Google Scholar] [CrossRef] [PubMed]
  247. Ho, L.; Hoefel, D.; Saint, C.P.; Newcombe, G. Isolation and identification of a novel microcystin-degrading bacterium from a biological sand filter. Water Res. 2007, 41, 4685–4695. [Google Scholar] [CrossRef]
  248. Jing, W.; Sui, G.; Liu, S. Characteristics of a microcystin-LR biodegrading bacterial isolate: Ochrobactrum sp. FDT5. Bull. Environ. Contam. Toxicol. 2014, 92, 119–122. [Google Scholar] [CrossRef]
  249. Zhang, M.; Yan, H.; Pan, G. Microbial degradation of microcystin-LR by Ralstonia solanacearum. Environ. Technol. 2011, 32, 1779–1787. [Google Scholar] [CrossRef]
  250. Yang, F.; Zhou, Y.; Sun, R.; Wei, H.; Li, Y.; Yin, L.; Pu, Y. Biodegradation of microcystin-LR and-RR by a novel microcystin-degrading bacterium isolated from Lake Taihu. Biodegradation 2014, 25, 447–457. [Google Scholar] [CrossRef]
  251. Mou, X.; Lu, X.; Jacob, J.; Sun, S.; Heath, R. Metagenomic identification of bacterioplankton taxa and pathways involved in microcystin degradation in Lake Erie. PLoS ONE 2013, 8, e61890. [Google Scholar] [CrossRef] [PubMed]
  252. Hu, L.B.; Yang, J.D.; Zhou, W.; Yin, Y.F.; Chen, J.; Shi, Z.Q. Isolation of a Methylobacillus sp. that degrades microcystin toxins associated with cyanobacteria. New Biotechnol. 2009, 26, 205–211. [Google Scholar] [CrossRef] [PubMed]
  253. You, D.-J.; Chen, X.-G.; Xiang, H.-Y.; Ouyang, L.; Yang, B. Isolation, identification and characterization of a microcystin-degrading bacterium Paucibacter sp. strain CH. Huan Jing Ke Xue=Huanjing Kexue 2014, 35, 313–318. [Google Scholar]
  254. Santos, A.A.; Soldatou, S.; de Magalhães, V.F.; Azevedo, S.M.; Camacho-Muñoz, D.; Lawton, L.A.; Edwards, C. Degradation of multiple peptides by Microcystin-Degrader Paucibacter toxinivorans (2C20). Toxins 2021, 13, 265. [Google Scholar] [CrossRef] [PubMed]
  255. Chen, J.; Hu, L.B.; Zhou, W.; Yan, S.H.; Yang, J.D.; Xue, Y.F.; Shi, Z.Q. Degradation of microcystin-LR and RR by a Stenotrophomonas sp. strain EMS isolated from Lake Taihu, China. Int. J. Mol. Sci. 2010, 11, 896–911. [Google Scholar] [CrossRef]
  256. Yang, F.; Massey, I.Y.; Guo, J.; Yang, S.; Pu, Y.; Zeng, W.; Tan, H. Microcystin-LR degradation utilizing a novel effective indigenous bacterial community YFMCD1 from Lake Taihu. J. Toxicol. Environ. Health Part A 2018, 81, 184–193. [Google Scholar] [CrossRef]
  257. Yang, F.; Zhou, Y.; Yin, L.; Zhu, G.; Liang, G.; Pu, Y. Microcystin-degrading activity of an indigenous bacterial strain Stenotrophomonas acidaminiphila MC-LTH2 isolated from Lake Taihu. PLoS ONE 2014, 9, e86216. [Google Scholar] [CrossRef]
  258. Idroos, F.S.; De Silva, B.; Manage, P.M. Biodegradation of microcystin analogues by Stenotrophomonas maltophilia isolated from Beira Lake Sri Lanka. J. Natl. Sci. Found. Sri Lanka 2017, 45, 91–99. [Google Scholar] [CrossRef]
  259. Eleuterio, L.; Batista, J.R. Biodegradation studies and sequencing of microcystin-LR degrading bacteria isolated from a drinking water biofilter and a fresh water lake. Toxicon 2010, 55, 1434–1442. [Google Scholar] [CrossRef]
  260. Crettaz-Minaglia, M.; Fallico, M.; Aranda, O.; Juarez, I.; Pezzoni, M.; Costa, C.; Andrinolo, D.; Giannuzzi, L. Effect of temperature on microcystin-LR removal and lysis activity on Microcystis aeruginosa (cyanobacteria) by an indigenous bacterium belonging to the genus Achromobacter. Environ. Sci. Pollut. Res. 2020, 27, 44427–44439. [Google Scholar] [CrossRef] [PubMed]
  261. Huang, F.; Feng, H.; Li, X.; Yi, X.; Guo, J.; Clara, T.; Yang, F. Anaerobic degradation of microcystin-LR by an indigenous bacterial Enterobacter sp. YF3. J. Toxicol. Environ. Health Part A 2019, 82, 1120–1128. [Google Scholar] [CrossRef]
  262. Alamri, S.A. Biodegradation of microcystin-RR by Bacillus flexus isolated from a Saudi freshwater lake. Saudi J. Biol. Sci. 2012, 19, 435–440. [Google Scholar] [CrossRef]
  263. Massey, I.Y.; Yang, F. A mini review on microcystins and bacterial degradation. Toxins 2020, 12, 268. [Google Scholar] [CrossRef]
  264. Eloka-Eboka, A.C.; Bwapwa, J.K.; Maroa, S. Biomass for CO2 sequestration. Encycl. Renew. Sustain. Mater. 2019, 1, 4252. [Google Scholar]
  265. Sarma, M.K.; Kaushik, S.; Goswami, P. Cyanobacteria: A metabolic power house for harvesting solar energy to produce bio-electricity and biofuels. Biomass Bioenergy 2016, 90, 187–201. [Google Scholar] [CrossRef]
  266. Zhang, S.; Liu, Z. Advances in the biological fixation of carbon dioxide by microalgae. J. Chem. Technol. Biotechnol. 2021, 96, 1475–1495. [Google Scholar] [CrossRef]
  267. Paerl, H.W.; Paul, V.J. Climate change: Links to global expansion of harmful cyanobacteria. Water Res. 2012, 46, 1349–1363. [Google Scholar] [CrossRef]
  268. Anu, K.; Sneha, V.; Busheera, P.; Muhammed, J.; Augustine, A. Mangroves in environmental engineering: Harnessing the multifunctional potential of Nature’s coastal architects for sustainable ecosystem management. Results Eng. 2024, 21, 101765. [Google Scholar]
  269. Goodchild-Michelman, I.M.; Church, G.M.; Schubert, M.G.; Tang, T.-C. Light and carbon: Synthetic biology toward new cyanobacteria-based living biomaterials. Mater. Today Bio 2023, 19, 100583. [Google Scholar] [CrossRef]
  270. Vasile, N.S.; Cordara, A.; Usai, G.; Re, A. Computational analysis of dynamic light exposure of unicellular algal cells in a flat-panel photobioreactor to support light-induced CO2 bioprocess development. Front. Microbiol. 2021, 12, 639482. [Google Scholar] [CrossRef]
  271. Ashokkumar, V.; Venkatkarthick, R.; Jayashree, S.; Chuetor, S.; Dharmaraj, S.; Kumar, G.; Chen, W.-H.; Ngamcharussrivichai, C. Recent advances in lignocellulosic biomass for biofuels and value-added bioproducts-A critical review. Bioresour. Technol. 2022, 344, 126195. [Google Scholar] [CrossRef] [PubMed]
  272. Yue, X.-L.; Xu, L.; Cui, L.; Fu, G.-Y.; Xu, X.-W. Metagenome-based analysis of carbon-fixing microorganisms and their carbon-fixing pathways in deep-sea sediments of the southwestern Indian Ocean. Mar. Genom. 2023, 70, 101045. [Google Scholar] [CrossRef] [PubMed]
  273. Hamilton, D.P.; Salmaso, N.; Paerl, H.W. Mitigating harmful cyanobacterial blooms: Strategies for control of nitrogen and phosphorus loads. Aquat. Ecol. 2016, 50, 351–366. [Google Scholar] [CrossRef]
  274. Singh, P.K.; Kumar, A.; Singh, V.K.; Shrivistava, A.K. Advances in Cyanobacterial Biology; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar]
  275. Boucot, A.; Gray, J. A critique of Phanerozoic climatic models involving changes in the CO2 content of the atmosphere. Earth-Sci. Rev. 2001, 56, 1–159. [Google Scholar] [CrossRef]
  276. Mani, A.; Budd, T.; Maine, E. Emissions-Intensive and Trade-Exposed Industries: Technological Innovation and Climate Policy Solutions to Achieve Net-Zero Emissions by 2050. RSC Sustain. 2024, 2, 903–927. [Google Scholar] [CrossRef]
  277. Mantzouki, E.; Visser, P.M.; Bormans, M.; Ibelings, B.W. Understanding the key ecological traits of cyanobacteria as a basis for their management and control in changing lakes. Aquat. Ecol. 2016, 50, 333–350. [Google Scholar] [CrossRef]
  278. Olivares, J.; Bedmar, E.J.; Sanjuán, J. Biological nitrogen fixation in the context of global change. Mol. Plant-Microbe Interact. 2013, 26, 486–494. [Google Scholar] [CrossRef]
  279. Timsina, J. Preface to” Fertilizer Application on Crop Yield”. Fertilizer Application on Crop Yield; MDPI: Basel, Switzerland, 2019. [Google Scholar]
  280. Alvarez, A.L.; Weyers, S.L.; Gardner, R.D. Cyanobacteria-based soil amendments in the soil-plant system: Effects of inoculations on soil nutrient and microbial dynamics under spring wheat growth. Algal Res. 2024, 77, 103326. [Google Scholar] [CrossRef]
  281. Maestre, F.T.; Bowker, M.A.; Cantón, Y.; Castillo-Monroy, A.P.; Cortina, J.; Escolar, C.; Escudero, A.; Lázaro, R.; Martínez, I. Ecology and functional roles of biological soil crusts in semi-arid ecosystems of Spain. J. Arid Environ. 2011, 75, 1282–1291. [Google Scholar] [CrossRef]
  282. Scarlat, N.; Dallemand, J.-F. Future role of bioenergy. In The Role of Bioenergy in the Bioeconomy; Elsevier: Amsterdam, The Netherlands, 2019; pp. 435–547. [Google Scholar]
  283. Rony, Z.I.; Rasul, M.; Jahirul, M.; Mofijur, M. Harnessing marine biomass for sustainable fuel production through pyrolysis to support United Nations’ Sustainable Development Goals. Fuel 2024, 358, 130099. [Google Scholar] [CrossRef]
  284. Gregucci, D.; Nazir, F.; Calabretta, M.M.; Michelini, E. Illuminating Progress: The Contribution of Bioluminescence to Sustainable Development Goal 6—Clean Water and Sanitation—Of the United Nations 2030 Agenda. Sensors 2023, 23, 7244. [Google Scholar] [CrossRef]
  285. Lane, C.E. Bioplastic Production in Cyanobacteria and Consensus Degenerate PCR Probe Design. Ph.D. Thesis, Louisiana State University and Agricultural & Mechanical College, Baton Rouge, LA, USA, 2015. [Google Scholar]
  286. Tan, C.; Xu, P.; Tao, F. Carbon-negative synthetic biology: Challenges and emerging trends of cyanobacterial technology. Trends Biotechnol. 2022, 40, 1488–1502. [Google Scholar] [CrossRef]
  287. Koller, M. Production, properties, and processing of microbial polyhydroxyalkanoate (PHA) biopolyesters. In Microbial and Natural Macromolecules; Elsevier: Amsterdam, The Netherlands, 2021; pp. 3–55. [Google Scholar]
  288. Reisoglu, Ş.; Aydin, S. Microalgae as a Promising Candidate for Fighting Climate Change and Biodiversity Loss; IntechOpen: London, UK, 2023. [Google Scholar]
  289. Zhang, J.; Gao, D.; Li, Q.; Zhao, Y.; Li, L.; Lin, H.; Bi, Q.; Zhao, Y. Biodegradation of polyethylene microplastic particles by the fungus Aspergillus flavus from the guts of wax moth Galleria mellonella. Sci. Total Environ. 2020, 704, 135931. [Google Scholar] [CrossRef] [PubMed]
  290. Sánchez, C. Fungal potential for the degradation of petroleum-based polymers: An overview of macro-and microplastics biodegradation. Biotechnol. Adv. 2020, 40, 107501. [Google Scholar] [CrossRef] [PubMed]
  291. El-Shafei, H.A.; Abd El-Nasser, N.H.; Kansoh, A.L.; Ali, A.M. Biodegradation of disposable polyethylene by fungi and Streptomyces species. Polym. Degrad. Stab. 1998, 62, 361–365. [Google Scholar] [CrossRef]
  292. Delacuvellerie, A.; Cyriaque, V.; Gobert, S.; Benali, S.; Wattiez, R. The plastisphere in marine ecosystem hosts potential specific microbial degraders including Alcanivorax borkumensis as a key player for the low-density polyethylene degradation. J. Hazard. Mater. 2019, 380, 120899. [Google Scholar] [CrossRef]
  293. Harshvardhan, K.; Jha, B. Biodegradation of low-density polyethylene by marine bacteria from pelagic waters, Arabian Sea, India. Mar. Pollut. Bull. 2013, 77, 100–106. [Google Scholar] [CrossRef]
  294. Kunlere, I.O.; Fagade, O.E.; Nwadike, B.I. Biodegradation of low density polyethylene (LDPE) by certain indigenous bacteria and fungi. Int. J. Environ. Stud. 2019, 76, 428–440. [Google Scholar] [CrossRef]
  295. Park, S.Y.; Kim, C.G. Biodegradation of micro-polyethylene particles by bacterial colonization of a mixed microbial consortium isolated from a landfill site. Chemosphere 2019, 222, 527–533. [Google Scholar] [CrossRef]
  296. Skariyachan, S.; Patil, A.A.; Shankar, A.; Manjunath, M.; Bachappanavar, N.; Kiran, S. Enhanced polymer degradation of polyethylene and polypropylene by novel thermophilic consortia of Brevibacillus sps. and Aneurinibacillus sp. screened from waste management landfills and sewage treatment plants. Polym. Degrad. Stab. 2018, 149, 52–68. [Google Scholar] [CrossRef]
  297. Muhonja, C.N.; Makonde, H.; Magoma, G.; Imbuga, M. Biodegradability of polyethylene by bacteria and fungi from Dandora dumpsite Nairobi-Kenya. PLoS ONE 2018, 13, e0198446. [Google Scholar] [CrossRef]
  298. Giacomucci, L.; Raddadi, N.; Soccio, M.; Lotti, N.; Fava, F. Polyvinyl chloride biodegradation by Pseudomonas citronellolis and Bacillus flexus. New Biotechnol. 2019, 52, 35–41. [Google Scholar] [CrossRef]
  299. Giacomucci, L.; Raddadi, N.; Soccio, M.; Lotti, N.; Fava, F. Biodegradation of polyvinyl chloride plastic films by enriched anaerobic marine consortia. Mar. Environ. Res. 2020, 158, 104949. [Google Scholar] [CrossRef]
  300. Kumari, A.; Chaudhary, D.R.; Jha, B. Destabilization of polyethylene and polyvinylchloride structure by marine bacterial strain. Environ. Sci. Pollut. Res. 2019, 26, 1507–1516. [Google Scholar] [CrossRef] [PubMed]
  301. Zhang, N.; Ding, M.; Yuan, Y. Current advances in biodegradation of polyolefins. Microorganisms 2022, 10, 1537. [Google Scholar] [CrossRef] [PubMed]
  302. Auta, H.S.; Emenike, C.U.; Jayanthi, B.; Fauziah, S.H. Growth kinetics and biodeterioration of polypropylene microplastics by Bacillus sp. and Rhodococcus sp. isolated from mangrove sediment. Mar. Pollut. Bull. 2018, 127, 15–21. [Google Scholar] [CrossRef]
  303. Aravinthan, A.; Arkatkar, A.; Juwarkar, A.A.; Doble, M. Synergistic growth of Bacillus and Pseudomonas and its degradation potential on pretreated polypropylene. Prep. Biochem. Biotechnol. 2016, 46, 109–115. [Google Scholar] [CrossRef]
  304. Habib, S.; Iruthayam, A.; Abd Shukor, M.Y.; Alias, S.A.; Smykla, J.; Yasid, N.A. Biodeterioration of untreated polypropylene microplastic particles by Antarctic bacteria. Polymers 2020, 12, 2616. [Google Scholar] [CrossRef]
  305. Sekhar, V.C.; Nampoothiri, K.M.; Mohan, A.J.; Nair, N.R.; Bhaskar, T.; Pandey, A. Microbial degradation of high impact polystyrene (HIPS), an e-plastic with decabromodiphenyl oxide and antimony trioxide. J. Hazard. Mater. 2016, 318, 347–354. [Google Scholar] [CrossRef]
  306. Mohan, A.J.; Sekhar, V.C.; Bhaskar, T.; Nampoothiri, K.M. Microbial assisted high impact polystyrene (HIPS) degradation. Bioresour. Technol. 2016, 213, 204–207. [Google Scholar] [CrossRef] [PubMed]
  307. Gupta, A.; Thakur, I.S. Study of optimization of wastewater contaminant removal along with extracellular polymeric substances (EPS) production by a thermotolerant Bacillus sp. ISTVK1 isolated from heat shocked sewage sludge. Bioresour. Technol. 2016, 213, 21–30. [Google Scholar] [CrossRef] [PubMed]
  308. Wang, Z.; Xin, X.; Shi, X.; Zhang, Y. A polystyrene-degrading Acinetobacter bacterium isolated from the larvae of Tribolium castaneum. Sci. Total Environ. 2020, 726, 138564. [Google Scholar] [CrossRef]
  309. Urbanek, A.K.; Rybak, J.; Wróbel, M.; Leluk, K.; Mirończuk, A.M. A comprehensive assessment of microbiome diversity in Tenebrio molitor fed with polystyrene waste. Environ. Pollut. 2020, 262, 114281. [Google Scholar] [CrossRef]
  310. Yang, Y.; Yang, J.; Wu, W.-M.; Zhao, J.; Song, Y.; Gao, L.; Yang, R.; Jiang, L. Biodegradation and mineralization of polystyrene by plastic-eating mealworms: Part 2. Role of gut microorganisms. Environ. Sci. Technol. 2015, 49, 12087–12093. [Google Scholar] [CrossRef] [PubMed]
  311. Kim, H.R.; Lee, H.M.; Yu, H.C.; Jeon, E.; Lee, S.; Li, J.; Kim, D.-H. Biodegradation of polystyrene by Pseudomonas sp. isolated from the gut of superworms (larvae of Zophobas atratus). Environ. Sci. Technol. 2020, 54, 6987–6996. [Google Scholar] [CrossRef]
  312. Shah, Z.; Gulzar, M.; Hasan, F.; Shah, A.A. Degradation of polyester polyurethane by an indigenously developed consortium of Pseudomonas and Bacillus species isolated from soil. Polym. Degrad. Stab. 2016, 134, 349–356. [Google Scholar] [CrossRef]
  313. Rafiemanzelat, F.; Jafari, M.; Emtiazi, G. Study of biological degradation of new poly (ether-urethane-urea) s containing cyclopeptide moiety and PEG by Bacillus amyloliquefaciens isolated from soil. Appl. Biochem. Biotechnol. 2015, 177, 842–860. [Google Scholar] [CrossRef] [PubMed]
  314. Ekanayaka, A.H.; Tibpromma, S.; Dai, D.; Xu, R.; Suwannarach, N.; Stephenson, S.L.; Dao, C.; Karunarathna, S.C. A review of the fungi that degrade plastic. J. Fungi 2022, 8, 772. [Google Scholar] [CrossRef] [PubMed]
  315. Lin, Z.; Jin, T.; Xu, X.; Yin, X.; Zhang, D.; Geng, M.; Pang, C.; Luo, G.; Xiong, L.; Peng, J. Screening and degradation characteristics of plastic-degrading microorganisms in film-mulched vegetable soil. Int. Biodeterior. Biodegrad. 2024, 186, 105686. [Google Scholar] [CrossRef]
  316. Ojha, N.; Pradhan, N.; Singh, S.; Barla, A.; Shrivastava, A.; Khatua, P.; Rai, V.; Bose, S. Evaluation of HDPE and LDPE degradation by fungus, implemented by statistical optimization. Sci. Rep. 2017, 7, 39515. [Google Scholar] [CrossRef] [PubMed]
  317. Vivi, V.K.; Martins-Franchetti, S.M.; Attili-Angelis, D. Biodegradation of PCL and PVC: Chaetomium globosum (ATCC 16021) activity. Folia Microbiol. 2019, 64, 1–7. [Google Scholar] [CrossRef]
  318. Kudzin, M.H.; Piwowarska, D.; Festinger, N.; Chruściel, J.J. Risks Associated with the Presence of Polyvinyl Chloride in the Environment and Methods for Its Disposal and Utilization. Materials 2023, 17, 173. [Google Scholar] [CrossRef]
  319. Shimpi, N.; Borane, M.; Mishra, S.; Kadam, M.; Sonawane, S. Biodegradation of isotactic polypropylene (iPP)/poly (lactic acid)(PLA) and iPP/PLA/nano calcium carbonates using phanerochaete chrysosporium. Adv. Polym. Technol. 2018, 37, 522–530. [Google Scholar] [CrossRef]
  320. Paço, A.; Duarte, K.; da Costa, J.P.; Santos, P.S.; Pereira, R.; Pereira, M.; Freitas, A.C.; Duarte, A.C.; Rocha-Santos, T.A. Biodegradation of polyethylene microplastics by the marine fungus Zalerion maritimum. Sci. Total Environ. 2017, 586, 10–15. [Google Scholar] [CrossRef] [PubMed]
  321. Awasthi, S.; Srivastava, N.; Singh, T.; Tiwary, D.; Mishra, P.K. Biodegradation of thermally treated low density polyethylene by fungus Rhizopus oryzae NS 5. 3 Biotech 2017, 7, 73. [Google Scholar] [CrossRef] [PubMed]
  322. Anand, U.; Dey, S.; Bontempi, E.; Ducoli, S.; Vethaak, A.D.; Dey, A.; Federici, S. Biotechnological methods to remove microplastics: A review. Environ. Chem. Lett. 2023, 21, 1787–1810. [Google Scholar] [CrossRef] [PubMed]
  323. Fahad, S.; Saud, S.; Zhou, R. Biochar-Assisted Remediation of Contaminated Soils Under Changing Climate; Elsevier: Amsterdam, The Netherlands, 2024. [Google Scholar]
  324. Sheik, S.; Chandrashekar, K.; Swaroop, K.; Somashekarappa, H. Biodegradation of gamma irradiated low density polyethylene and polypropylene by endophytic fungi. Int. Biodeterior. Biodegrad. 2015, 105, 21–29. [Google Scholar] [CrossRef]
  325. Khan, S.; Nadir, S.; Shah, Z.U.; Shah, A.A.; Karunarathna, S.C.; Xu, J.; Khan, A.; Munir, S.; Hasan, F. Biodegradation of polyester polyurethane by Aspergillus tubingensis. Environ. Pollut. 2017, 225, 469–480. [Google Scholar] [CrossRef] [PubMed]
  326. Su, T.; Zhang, T.; Liu, P.; Bian, J.; Zheng, Y.; Yuan, Y.; Li, Q.; Liang, Q.; Qi, Q. Biodegradation of polyurethane by the microbial consortia enriched from landfill. Appl. Microbiol. Biotechnol. 2023, 107, 1983–1995. [Google Scholar] [CrossRef] [PubMed]
  327. Choudhary, S.; Tripathi, S.; Poluri, K.M. Microalgal-based bioenergy: Strategies, prospects, and sustainability. Energy Fuels 2022, 36, 14584–14612. [Google Scholar] [CrossRef]
  328. Larkum, A.W. Light-harvesting in cyanobacteria and eukaryotic algae: An overview. In Photosynthesis in Algae: Biochemical and Physiological Mechanisms; Springer: Cham, Switzerland, 2020; pp. 207–260. [Google Scholar]
  329. Sachse, M. Regulation of the Calvin Cycle in the Diatom Phaeodactylum tricornutum. Ph.D. Thesis, Universität Konstanz, Konstanz, Germany, 2013. [Google Scholar]
  330. Perera-Castro, A.V.; Flexas, J. Recent advances in understanding and improving photosynthesis. Fac. Rev. 2020, 9, 5. [Google Scholar] [CrossRef]
  331. Peck, B. The Chemistry of Photosynthesis. 2022, Bibliotex.
  332. Sanfilippo, J.E.; Garczarek, L.; Partensky, F.; Kehoe, D.M. Chromatic acclimation in cyanobacteria: A diverse and widespread process for optimizing photosynthesis. Annu. Rev. Microbiol. 2019, 73, 407–433. [Google Scholar] [CrossRef]
  333. Kirst, H.; Formighieri, C.; Melis, A. Maximizing photosynthetic efficiency and culture productivity in cyanobacteria upon minimizing the phycobilisome light-harvesting antenna size. Biochim. Biophys. Acta (BBA)-Bioenerg. 2014, 1837, 1653–1664. [Google Scholar] [CrossRef]
  334. Khetkorn, W.; Khanna, N.; Incharoensakdi, A.; Lindblad, P. Metabolic and genetic engineering of cyanobacteria for enhanced hydrogen production. Biofuels 2013, 4, 535–561. [Google Scholar] [CrossRef]
  335. KhokharVoytas, A.; Shahbaz, M.; Maqsood, M.F.; Zulfiqar, U.; Naz, N.; Iqbal, U.Z.; Sara, M.; Aqeel, M.; Khalid, N.; Noman, A. Genetic modification strategies for enhancing plant resilience to abiotic stresses in the context of climate change. Funct. Integr. Genom. 2023, 23, 283. [Google Scholar] [CrossRef] [PubMed]
  336. Kitchener, R.L.; Grunden, A.M. Methods for enhancing cyanobacterial stress tolerance to enable improved production of biofuels and industrially relevant chemicals. Appl. Microbiol. Biotechnol. 2018, 102, 1617–1628. [Google Scholar] [CrossRef] [PubMed]
  337. Nwoba, E.G.; Parlevliet, D.A.; Laird, D.W.; Alameh, K.; Moheimani, N.R. Light management technologies for increasing algal photobioreactor efficiency. Algal Res. 2019, 39, 101433. [Google Scholar] [CrossRef]
  338. Orr, D.J.; Pereira, A.M.; da Fonseca Pereira, P.; Pereira-Lima, Í.A.; Zsögön, A.; Araújo, W.L. Engineering photosynthesis: Progress and perspectives. F1000Research 2017, 6, 1891. [Google Scholar] [CrossRef] [PubMed]
  339. Xie, Y.; Khoo, K.S.; Chew, K.W.; Devadas, V.V.; Phang, S.J.; Lim, H.R.; Rajendran, S.; Show, P.L. Advancement of renewable energy technologies via artificial and microalgae photosynthesis. Bioresour. Technol. 2022, 363, 127830. [Google Scholar] [CrossRef] [PubMed]
  340. Mittler, R.; Blumwald, E. Genetic engineering for modern agriculture: Challenges and perspectives. Annu. Rev. Plant Biol. 2010, 61, 443–462. [Google Scholar] [CrossRef]
  341. Stephens, S.; Mahadevan, R.; Allen, D.G. Engineering photosynthetic bioprocesses for sustainable chemical production: A review. Front. Bioeng. Biotechnol. 2021, 8, 610723. [Google Scholar] [CrossRef]
  342. Jingura, R.M.; Matengaifa, R.; Musademba, D.; Musiyiwa, K. Characterisation of land types and agro-ecological conditions for production of Jatropha as a feedstock for biofuels in Zimbabwe. Biomass Bioenergy 2011, 35, 2080–2086. [Google Scholar] [CrossRef]
  343. Knoot, C.J.; Ungerer, J.; Wangikar, P.P.; Pakrasi, H.B. Cyanobacteria: Promising biocatalysts for sustainable chemical production. J. Biol. Chem. 2018, 293, 5044–5052. [Google Scholar] [CrossRef]
  344. Abo-Shady, A.M.; Osman, M.E.-A.H.; Gaafar, R.M.; Ismail, G.A.; El-Nagar, M.M. Cyanobacteria as a Valuable Natural Resource for Improved Agriculture, Environment, and Plant Protection. Water Air Soil Pollut. 2023, 234, 313. [Google Scholar] [CrossRef]
  345. Dange, P.; Gawas, S.; Pandit, S.; Mekuto, L.; Gupta, P.K.; Shanmugam, P.; Patil, R.; Banerjee, S. Trends in photobioreactor technology for microalgal biomass production along with wastewater treatment: Bottlenecks and breakthroughs. In An Integration of Phycoremediation Processes in Wastewater Treatment; Elsevier: Amsterdam, The Netherlands, 2022; pp. 135–154. [Google Scholar]
  346. Fabris, M.; Abbriano, R.M.; Pernice, M.; Sutherland, D.L.; Commault, A.S.; Hall, C.C.; Labeeuw, L.; McCauley, J.I.; Kuzhiuparambil, U.; Ray, P. Emerging technologies in algal biotechnology: Toward the establishment of a sustainable, algae-based bioeconomy. Front. Plant Sci. 2020, 11, 279. [Google Scholar] [CrossRef] [PubMed]
  347. Rajvanshi, M.; Gautam, K.; Manjre, S.; Krishna Kumar, G.R.; Kumar, C.; Govindachary, S.; Dasgupta, S. Stoichiometrically balanced nutrient management using a newly designed nutrient medium for large scale cultivation of Cyanobacterium aponinum. J. Appl. Phycol. 2019, 31, 2779–2789. [Google Scholar] [CrossRef]
  348. do Vale Barreto Figueiredo, M.; do Espírito Santo Mergulhão, A.C.; Sobral, J.K.; de Andrade Lira Junior, M.; de Araújo, A.S.F. Biological nitrogen fixation: Importance, associated diversity, and estimates. In Plant Microbe Symbiosis: Fundamentals and Advances; Springer: Berlin/Heidelberg, Germany, 2013; pp. 267–289. [Google Scholar]
  349. Cai, T.; Park, S.Y.; Li, Y. Nutrient recovery from wastewater streams by microalgae: Status and prospects. Renew. Sustain. Energy Rev. 2013, 19, 360–369. [Google Scholar] [CrossRef]
  350. Mishra, V.; Mudgal, N.; Rawat, D.; Poria, P.; Mukherjee, P.; Sharma, U.; Kumria, P.; Pani, B.; Singh, M.; Yadav, A. Integrating microalgae into textile wastewater treatment processes: Advancements and opportunities. J. Water Process Eng. 2023, 55, 104128. [Google Scholar] [CrossRef]
  351. Yadav, S.; Rai, S.; Rai, R.; Shankar, A.; Singh, S.; Rai, L. Cyanobacteria: Role in agriculture, environmental sustainability, biotechnological potential and agroecological impact. In Plant-Microbe Interactions in Agro-Ecological Perspectives: Volume 2: Microbial Interactions and Agro-Ecological Impacts; Springer: Singapore, 2017; pp. 257–277. [Google Scholar]
  352. Sakarika, M.; Sventzouri, E.; Pispas, K.; Ali, S.; Kornaros, M. Production of biopolymers from microalgae and cyanobacteria. In Algal Systems for Resource Recovery from Waste and Wastewater; IWA Publishing: London, UK, 2023; p. 207. [Google Scholar]
  353. Alvarez, A.L.; Weyers, S.L.; Goemann, H.M.; Peyton, B.M.; Gardner, R.D. Microalgae, soil and plants: A critical review of microalgae as renewable resources for agriculture. Algal Res. 2021, 54, 102200. [Google Scholar] [CrossRef]
  354. Yap, X.Y.; Gew, L.T.; Khalid, M.; Yow, Y.-Y. Algae-Based Bioplastic for Packaging: A Decade of Development and Challenges (2010–2020). J. Polym. Environ. 2023, 31, 833–851. [Google Scholar] [CrossRef]
  355. Arias, D.M.; Ortíz-Sánchez, E.; Okoye, P.U.; Rodríguez-Rangel, H.; Ortega, A.B.; Longoria, A.; Domínguez-Espíndola, R.; Sebastian, P. A review on cyanobacteria cultivation for carbohydrate-based biofuels: Cultivation aspects, polysaccharides accumulation strategies, and biofuels production scenarios. Sci. Total Environ. 2021, 794, 148636. [Google Scholar] [CrossRef]
  356. Kadri, M.S.; Singhania, R.R.; Anisha, G.S.; Gohil, N.; Singh, V.; Patel, A.K.; Patel, A.K. Microalgal lutein: Advancements in production, extraction, market potential, and applications. Bioresour. Technol. 2023, 389, 129808. [Google Scholar] [CrossRef]
  357. Rosenboom, J.-G.; Langer, R.; Traverso, G. Bioplastics for a circular economy. Nat. Rev. Mater. 2022, 7, 117–137. [Google Scholar] [CrossRef]
  358. Costa, A.; Encarnação, T.; Tavares, R.; Todo Bom, T.; Mateus, A. Bioplastics: Innovation for Green Transition. Polymers 2023, 15, 517. [Google Scholar] [CrossRef] [PubMed]
  359. Kollmen, J.; Strieth, D. The beneficial effects of cyanobacterial co-culture on plant growth. Life 2022, 12, 223. [Google Scholar] [CrossRef]
  360. Govindasamy, R.; Gayathiri, E.; Sankar, S.; Venkidasamy, B.; Prakash, P.; Rekha, K.; Savaner, V.; Pari, A.; Thirumalaivasan, N.; Thiruvengadam, M. Emerging Trends of Nanotechnology and Genetic Engineering in Cyanobacteria to Optimize Production for Future Applications. Life 2022, 12, 2013. [Google Scholar] [CrossRef] [PubMed]
  361. Shuba, E.S.; Kifle, D. Microalgae to biofuels:‘Promising’alternative and renewable energy, review. Renew. Sustain. Energy Rev. 2018, 81, 743–755. [Google Scholar] [CrossRef]
  362. Abinandan, S.; Subashchandrabose, S.R.; Venkateswarlu, K.; Megharaj, M. Nutrient removal and biomass production: Advances in microalgal biotechnology for wastewater treatment. Crit. Rev. Biotechnol. 2018, 38, 1244–1260. [Google Scholar] [CrossRef] [PubMed]
  363. Singh, B.J.; Chakraborty, A.; Sehgal, R. A systematic review of industrial wastewater management: Evaluating challenges and enablers. J. Environ. Manag. 2023, 348, 119230. [Google Scholar] [CrossRef] [PubMed]
  364. Abbott, M.L.; Fisher, M.T. The Art of Scalability: Scalable Web Architecture, Processes, and Organizations for the Modern Enterprise; Addison-Wesley Professional: Boston, MA, USA, 2015. [Google Scholar]
  365. Rawat, I.; Kumar, R.R.; Mutanda, T.; Bux, F. Biodiesel from microalgae: A critical evaluation from laboratory to large scale production. Appl. Energy 2013, 103, 444–467. [Google Scholar] [CrossRef]
  366. Bajpai, V.K.; Shukla, S.; Kang, S.-M.; Hwang, S.K.; Song, X.; Huh, Y.S.; Han, Y.-K. Developments of cyanobacteria for nano-marine drugs: Relevance of nanoformulations in cancer therapies. Mar. Drugs 2018, 16, 179. [Google Scholar] [CrossRef] [PubMed]
  367. Liu, M.; Wu, J.; Zhang, S.; Liang, J. Cyanobacterial blooms management: A modified optimization model for interdisciplinary research. Ecol. Model. 2023, 484, 110480. [Google Scholar] [CrossRef]
  368. Kumar, R.; Lalnundiki, V.; Shelare, S.D.; Abhishek, G.J.; Sharma, S.; Sharma, D.; Kumar, A.; Abbas, M. An investigation of the environmental implications of bioplastics: Recent advancements on the development of environmentally friendly bioplastics solutions. Environ. Res. 2023, 244, 117707. [Google Scholar] [CrossRef]
  369. Bello, I.A. Development of Soy Formulate and Bioprocessing for Biological Ammonia Production via Hyperammonia Bacteria Fermentation. Ph.D. Thesis, North Dakota State University, Fargo, ND, USA, 2023. [Google Scholar]
  370. Nobles, D.R., Jr. Cellulose in the Cyanobacteria. Ph.D. Thesis, The University of Texas at Austin, Austin, TX, USA, 2006. [Google Scholar]
  371. Huang, F.; Li, M.; Siffalovic, P.; Cao, G.; Tian, J. From scalable solution fabrication of perovskite films towards commercialization of solar cells. Energy Environ. Sci. 2019, 12, 518–549. [Google Scholar] [CrossRef]
  372. Bernard, O. Hurdles and challenges for modelling and control of microalgae for CO2 mitigation and biofuel production. J. Process Control 2011, 21, 1378–1389. [Google Scholar] [CrossRef]
  373. Pruvost, J.; Le Borgne, F.; Artu, A.; Cornet, J.-F.; Legrand, J. Industrial photobioreactors and scale-up concepts. In Advances in Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2016; Volume 48, pp. 257–310. [Google Scholar]
  374. Borowitzka, M.A.; Vonshak, A. Scaling up microalgal cultures to commercial scale. Eur. J. Phycol. 2017, 52, 407–418. [Google Scholar] [CrossRef]
  375. Li, F.; Liu, X.; Zhang, X.; Zhao, D.; Liu, H.; Zhou, C.; Wang, R. Urban ecological infrastructure: An integrated network for ecosystem services and sustainable urban systems. J. Clean. Prod. 2017, 163, S12–S18. [Google Scholar] [CrossRef]
  376. Bosma, R.H.; Verdegem, M.C. Sustainable aquaculture in ponds: Principles, practices and limits. Livest. Sci. 2011, 139, 58–68. [Google Scholar] [CrossRef]
  377. Page, M.A.; MacAllister, B.A.; Urban, A.; MacAllister, I.E.; White, C.S.; Pokrzywinski, K.L.; Martinez-Guerra, E.; Grasso, C.; Kennedy, A.J.; Thomas, C.C. Harmful Algal Bloom Interception, Treatment, and Transformation System,” HABITATS”: Pilot Research Study Phase I-Summer 2019; US Army Engineer Research and Development Center, Construction Engineering: Vicksburg, MS, USA, 2020. [Google Scholar]
  378. Kayode-Ajala, O. Establishing cyber resilience in developing countries: An exploratory investigation into institutional, legal, financial, and social challenges. Int. J. Sustain. Infrastruct. Cities Soc. 2023, 8, 1–10. [Google Scholar]
  379. Bade, S.O.; Tomomewo, O.S.; Meenakshisundaram, A.; Ferron, P.; Oni, B.A. Economic, social, and regulatory challenges of green hydrogen production and utilization in the US: A review. Int. J. Hydrog. Energy 2023, 49, 314–335. [Google Scholar] [CrossRef]
  380. Price, S. Developing Next Generation Algae Bioplastic Technology. Ph.D. Thesis, University of Technology Sydney, Sydney, Australia, 2022. [Google Scholar]
  381. Dixon, C.; Wilken, L.R. Green microalgae biomolecule separations and recovery. Bioresour. Bioprocess. 2018, 5, 14. [Google Scholar] [CrossRef]
  382. Noble, R.D.; Agrawal, R. Separations research needs for the 21st century. Ind. Eng. Chem. Res. 2005, 44, 2887–2892. [Google Scholar] [CrossRef]
  383. Poumadère, M.; Bertoldo, R.; Samadi, J. Public perceptions and governance of controversial technologies to tackle climate change: Nuclear power, carbon capture and storage, wind, and geoengineering. Wiley Interdiscip. Rev. Clim. Chang. 2011, 2, 712–727. [Google Scholar] [CrossRef]
  384. McGinn, M.M. Inclusive Outreach and Public Engagement Guide; Race and Social Justice Initiative: Seattle, DC, USA, 2009. [Google Scholar]
  385. Kamravamanesh, D.; Kiesenhofer, D.; Fluch, S.; Lackner, M.; Herwig, C. Scale-up challenges and requirement of technology-transfer for cyanobacterial poly (3-hydroxybutyrate) production in industrial scale. Int. J. Biobased Plast. 2019, 1, 60–71. [Google Scholar] [CrossRef]
  386. Bansal, S.; Mahendiratta, S.; Kumar, S.; Sarma, P.; Prakash, A.; Medhi, B. Collaborative research in modern era: Need and challenges. Indian J. Pharmacol. 2019, 51, 137. [Google Scholar] [PubMed]
  387. Gupta, V.; Ratha, S.K.; Sood, A.; Chaudhary, V.; Prasanna, R. New insights into the biodiversity and applications of cyanobacteria (blue-green algae)—Prospects and challenges. Algal Res. 2013, 2, 79–97. [Google Scholar] [CrossRef]
  388. Baunach, M.; Guljamow, A.; Miguel-Gordo, M.; Dittmann, E. Harnessing the potential: Advances in cyanobacterial natural product research and biotechnology. Nat. Prod. Rep. 2024, 41, 347–369. [Google Scholar] [CrossRef]
  389. Johnson, T.J.; Jahandideh, A.; Johnson, M.D.; Fields, K.H.; Richardson, J.W.; Muthukumarappan, K.; Cao, Y.; Gu, Z.; Halfmann, C.; Zhou, R. Producing next-generation biofuels from filamentous cyanobacteria: An economic feasibility analysis. Algal Res. 2016, 20, 218–228. [Google Scholar] [CrossRef]
  390. Majidian, P.; Tabatabaei, M.; Zeinolabedini, M.; Naghshbandi, M.P.; Chisti, Y. Metabolic engineering of microorganisms for biofuel production. Renew. Sustain. Energy Rev. 2018, 82, 3863–3885. [Google Scholar] [CrossRef]
  391. Yadav, I.; Rautela, A.; Kumar, S. Approaches in the photosynthetic production of sustainable fuels by cyanobacteria using tools of synthetic biology. World J. Microbiol. Biotechnol. 2021, 37, 1–17. [Google Scholar]
  392. Tanvir, R.U.; Zhang, J.; Canter, T.; Chen, D.; Lu, J.; Hu, Z. Harnessing solar energy using phototrophic microorganisms: A sustainable pathway to bioenergy, biomaterials, and environmental solutions. Renew. Sustain. Energy Rev. 2021, 146, 111181. [Google Scholar] [CrossRef] [PubMed]
  393. Tiwari, A.; Pandey, A. Cyanobacterial hydrogen production—A step towards clean environment. Int. J. Hydrogen Energy 2012, 37, 139–150. [Google Scholar] [CrossRef]
  394. Ahmad, A.; Khan, S.; Chhabra, T.; Tariq, S.; Javed, M.S.; Li, H.; Naqvi, S.R.; Rajendran, S.; Luque, R.; Ahmad, I. Synergic impact of renewable resources and advanced technologies for green hydrogen production: Trends and perspectives. Int. J. Hydrogen Energy 2024, 67, 788–806. [Google Scholar] [CrossRef]
  395. Zhao, J.; Li, X.F.; Ren, Y.P.; Wang, X.H.; Jian, C. Electricity generation from Taihu Lake cyanobacteria by sediment microbial fuel cells. J. Chem. Technol. Biotechnol. 2012, 87, 1567–1573. [Google Scholar] [CrossRef]
  396. Indrama, T.; Tiwari, O.; Bandyopadhyay, T.K.; Mondal, A.; Bhunia, B. Cyanobacteria: A biocatalyst in microbial fuel cell for sustainable electricity generation. In Waste to Sustainable Energy; CRC Press: Boca Raton, FL, USA, 2019; pp. 125–140. [Google Scholar]
  397. Hallenbeck, P.C.; Grogger, M.; Veverka, D. Recent advances in microbial electrocatalysis. Electrocatalysis 2014, 5, 319–329. [Google Scholar] [CrossRef]
  398. Kumar, M.; Sundaram, S.; Gnansounou, E.; Larroche, C.; Thakur, I.S. Carbon dioxide capture, storage and production of biofuel and biomaterials by bacteria: A review. Bioresour. Technol. 2018, 247, 1059–1068. [Google Scholar] [CrossRef] [PubMed]
  399. López-Pacheco, I.Y.; Rodas-Zuluaga, L.I.; Fuentes-Tristan, S.; Castillo-Zacarías, C.; Sosa-Hernández, J.E.; Barceló, D.; Iqbal, H.M.; Parra-Saldívar, R. Phycocapture of CO2 as an option to reduce greenhouse gases in cities: Carbon sinks in urban spaces. J. CO2 Util. 2021, 53, 101704. [Google Scholar] [CrossRef]
  400. Lin, W.-R.; Tan, S.-I.; Hsiang, C.-C.; Sung, P.-K.; Ng, I.-S. Challenges and opportunity of recent genome editing and multi-omics in cyanobacteria and microalgae for biorefinery. Bioresour. Technol. 2019, 291, 121932. [Google Scholar] [CrossRef]
  401. Parmar, A.; Singh, N.K.; Pandey, A.; Gnansounou, E.; Madamwar, D. Cyanobacteria and microalgae: A positive prospect for biofuels. Bioresour. Technol. 2011, 102, 10163–10172. [Google Scholar] [CrossRef]
  402. Prabha, S.; Vijay, A.K.; Paul, R.R.; George, B. Cyanobacterial biorefinery: Towards economic feasibility through the maximum valorization of biomass. Sci. Total Environ. 2022, 814, 152795. [Google Scholar] [CrossRef]
  403. Quintana, N.; Van der Kooy, F.; Van de Rhee, M.D.; Voshol, G.P.; Verpoorte, R. Renewable energy from Cyanobacteria: Energy production optimization by metabolic pathway engineering. Appl. Microbiol. Biotechnol. 2011, 91, 471–490. [Google Scholar] [CrossRef]
  404. Kashyap, A.K.; Dubey, S.K.; Jain, B.P. Cyanobacterial diversity concerning the extreme environment and their bioprospecting. In Cyanobacterial Lifestyle and Its Applications in Biotechnology; Elsevier: Amsterdam, The Netherlands, 2022; pp. 1–22. [Google Scholar]
  405. Meshram, B.G.; Chaugule, B.B. An introduction to cyanobacteria: Diversity and potential applications. In The Role of Photosynthetic Microbes in Agriculture and Industry; Nova Science Publishers: Hauppauge, NY, USA, 2018; p. 1. [Google Scholar]
  406. Chakdar, H.; Pabbi, S. Cyanobacterial phycobilins: Production, purification, and regulation. In Frontier Discoveries and Innovations in Interdisciplinary Microbiology; Springer: New Delhi, India, 2016; pp. 45–69. [Google Scholar]
  407. Hamilton, T.L.; Bryant, D.A.; Macalady, J.L. The role of biology in planetary evolution: Cyanobacterial primary production in low-oxygen Proterozoic oceans. Environ. Microbiol. 2016, 18, 325–340. [Google Scholar] [CrossRef]
  408. Bandyopadhyay, A.; Elvitigala, T.; Welsh, E.; Stöckel, J.; Liberton, M.; Min, H.; Sherman, L.A.; Pakrasi, H.B. Novel metabolic attributes of the genus Cyanothece, comprising a group of unicellular nitrogen-fixing cyanobacteria. MBio 2011, 2. [Google Scholar] [CrossRef]
  409. Zhang, Y.; Whalen, J.K.; Cai, C.; Shan, K.; Zhou, H. Harmful cyanobacteria-diatom/dinoflagellate blooms and their cyanotoxins in freshwaters: A nonnegligible chronic health and ecological hazard. Water Res. 2023, 233, 119807. [Google Scholar] [CrossRef] [PubMed]
  410. Willis, A.; Woodhouse, J.N. Defining cyanobacterial species: Diversity and description through genomics. Crit. Rev. Plant Sci. 2020, 39, 101–124. [Google Scholar] [CrossRef]
  411. Sadvakasova, A.K.; Kossalbayev, B.D.; Bauenova, M.O.; Balouch, H.; Leong, Y.K.; Zayadan, B.K.; Huang, Z.; Alharby, H.F.; Tomo, T.; Chang, J.-S. Microalgae as a key tool in achieving carbon neutrality for bioproduct production. Algal Res. 2023, 72, 103096. [Google Scholar] [CrossRef]
  412. Brunschwig, C.; Moussavou, W.; Blin, J. Use of bioethanol for biodiesel production. Prog. Energy Combust. Sci. 2012, 38, 283–301. [Google Scholar] [CrossRef]
  413. Moravvej, Z.; Makarem, M.A.; Rahimpour, M.R. The fourth generation of biofuel. In Second and Third Generation of Feedstocks; Elsevier: Amsterdam, The Netherlands, 2019; pp. 557–597. [Google Scholar]
  414. Lin, B.B. Resilience in agriculture through crop diversification: Adaptive management for environmental change. BioScience 2011, 61, 183–193. [Google Scholar] [CrossRef]
  415. Novoveská, L.; Nielsen, S.L.; Eroldoğan, O.T.; Haznedaroglu, B.Z.; Rinkevich, B.; Fazi, S.; Robbens, J.; Vasquez, M.; Einarsson, H. Overview and challenges of large-scale cultivation of photosynthetic microalgae and cyanobacteria. Mar. Drugs 2023, 21, 445. [Google Scholar] [CrossRef]
  416. Das, M.; Maiti, S.K. Current knowledge on cyanobacterial biobutanol production: Advances, challenges, and prospects. Rev. Environ. Sci. Bio/Technol. 2022, 21, 483–516. [Google Scholar] [CrossRef]
  417. Carroll, A.L.; Case, A.E.; Zhang, A.; Atsumi, S. Metabolic engineering tools in model cyanobacteria. Metab. Eng. 2018, 50, 47–56. [Google Scholar] [CrossRef]
  418. Singhal, A.; Prashant. Biofuels: Perspective for sustainable development and climate change mitigation. In Biotechnology for Biofuels: A Sustainable Green Energy Solution; Springer: Singapore, 2020; pp. 1–22. [Google Scholar]
  419. Mastropetros, S.G.; Pispas, K.; Zagklis, D.; Ali, S.S.; Kornaros, M. Biopolymers production from microalgae and cyanobacteria cultivated in wastewater: Recent advances. Biotechnol. Adv. 2022, 60, 107999. [Google Scholar] [CrossRef]
  420. Shah, S.; Kumar, A. Polyhydroxyalkanoates: Advances in the synthesis of sustainable bio-plastics. Eur. J. Environ. Sci. 2020, 10, 76–88. [Google Scholar] [CrossRef]
  421. Umar, A. The Screening, Fabrication and Production of Microalgae Biocomposites for Carbon Capture and Utilisation. Ph.D. Thesis, Newcastle University, Newcastle upon Tyne, UK, 2019. [Google Scholar]
  422. Kovacic, Z.; Strand, R.; Völker, T. The Circular Economy in Europe: Critical Perspectives on Policies and Imaginaries; Taylor & Francis: Abingdon, UK, 2020. [Google Scholar]
  423. Caldwell, G.S.; In-na, P.; Hart, R.; Sharp, E.; Stefanova, A.; Pickersgill, M.; Walker, M.; Unthank, M.; Perry, J.; Lee, J.G. Immobilising microalgae and cyanobacteria as biocomposites: New opportunities to intensify algae biotechnology and bioprocessing. Energies 2021, 14, 2566. [Google Scholar] [CrossRef]
  424. Singh, S.; Tiwari, B.S. Biosynthesis of high-value amino acids by synthetic biology. In Current Developments in Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2019; pp. 257–294. [Google Scholar]
  425. Guzmán-Lagunes, F.; Wongsirichot, P.; Winterburn, J.; Guerrero Sanchez, C.; Montiel, C. Polyhydroxyalkanoates Production: A Challenge for the Plastic Industry. Ind. Eng. Chem. Res. 2023, 62, 18133–18158. [Google Scholar] [CrossRef]
  426. Pan, Y.; Zhang, H.; Zhang, B.; Gong, F.; Feng, J.; Huang, H.; Vanka, S.; Fan, R.; Cao, Q.; Shen, M. Renewable formate from sunlight, biomass and carbon dioxide in a photoelectrochemical cell. Nat. Commun. 2023, 14, 1013. [Google Scholar] [CrossRef]
  427. Mitra, R.; Xu, T.; Xiang, H.; Han, J. Current developments on polyhydroxyalkanoates synthesis by using halophiles as a promising cell factory. Microb. Cell Factories 2020, 19, 86. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Engineering photosynthesis for increased carbon assimilation.
Figure 1. Engineering photosynthesis for increased carbon assimilation.
Fuels 05 00023 g001
Figure 2. Life-cycle assessment of biodiesel production from microalgae.
Figure 2. Life-cycle assessment of biodiesel production from microalgae.
Fuels 05 00023 g002
Figure 3. Innovative synthesis pathways for polyhydroxyalkanoates.
Figure 3. Innovative synthesis pathways for polyhydroxyalkanoates.
Fuels 05 00023 g003
Figure 4. Photosynthetic cyanobacteria transformation and energy harvesting for ALA synthesis.
Figure 4. Photosynthetic cyanobacteria transformation and energy harvesting for ALA synthesis.
Fuels 05 00023 g004
Table 2. Degradation of nitroaromatic compounds and their chlorinated derivatives by different microorganisms.
Table 2. Degradation of nitroaromatic compounds and their chlorinated derivatives by different microorganisms.
Common SynonymsStrainsInitiate EnzymesRing-Cleavage Substrates
anilineP. pseudoalcaligenes JS45NR o-Aminopheno
2-nitrochlorobenzeneP. putida OCNB-1NR1-Chloro-2,3-dihydroxybenzene
4-Chloro-1-nitrobenzeneP. putida ZWL73NR(4-CAP)
2-hydroxynitrobenzeneP. putida B2, One-component
Monooxygenases
pyrocatechol
3-Nitrophenol (m-)P. putida B2NR1,2,4-Benzenetriol
p-nitrophenolPseudomonas sp. WBC-3One-component
Monooxygenases
benzene-1,4-diol
2-4-DNPR. erythropolis HL24-1
R. erythropolis HL24-2
(anionic ς-complex)
2-6-DNPC. necator JMP1343-Nitroquinolin-4-ol
Picric acidNocardioides sp. CB22-2
R. erythropolis HLPM-1
(anionic ς-complex)
nitrofunginBurkholderia sp. strain SJ98
Burkholderia sp. RKJ 800
Arthrobacter sp. SJCon
One-component
Monooxygenases
2-Chlorohydroquinone
C6H5ClO2
4C2NPExiguobacterium sp. PMANR2-Amino-1-hydroxybenzene
2C5NPC. pinatubonensis JMP134NR2-Aminohydroquinone hydrobromide
2,6-Dihalo-4-nitrophenolCupriavidus sp. CNP-82-dioxygenase 6-Quinolinol
2-nitro-benzoic acidP. fluorescens KU-7NR3-hydroxyanthranilic acid
3-nitro-benzoic acidPseudomonas sp. JS51
Comamonas sp. JS46
ADO, 2-aminoethanethiol dioxygenase;3,4-Dihydroxybenzoic acid
4-nitro-benzoic acidC. acidovoransNBA-10
Pseudomonas sp. 4NT
Pseudomonas putida TW3
Ralstonia sp. SJ98
NR3,4-Dihydroxybenzoic acid
2C4NPAcinetobacter sp.RKJ12(MMO)pyrocatechol
2-4-DNP (Dinitrophenol), 4C2NP (4-Chloro-2-nitrophenol), 2C5NP (2-Chloro-5-nitrophenol), NR (Nitroreductases), 4-Chloro 2-Amino Phenol (4-CAP).
Table 4. The remarkable plastic-degrading bacteria and fungi.
Table 4. The remarkable plastic-degrading bacteria and fungi.
Plastic Type of (Bacteria)Strain SpeciesEfficiency of DegradationTimeInitiation of the Strain
Top of Form
Refs.
PolyethyleneStreptomyces spp.
Streptomyces caatingaensis
Streptomyces cacaoi
Streptomyces cadmiisoli
Streptomyces caelestis
Streptomyces caeni
Streptomyces caeruleatus
Streptomyces caespitosus
27.5% 6 weeksNile Delta Lagoons[291]
Alcanivorax borkumensis3.6% ± 0.36%3 monthsSeawater samples from northern Corsica (Gulf of Calvi, Mediterranean)[292]
Kocuria palustris M161% ± 0.063%1 month[293]
P. aeruginosa,
P. fluorescens,
P. putida, P cepacia,
P stutzeri,
P. maltophilia, and P putrefaciens.
3.765%2 monthsScrapyard in the north of Ibadan[294]
Bacillus sp.
B. cereus
B. clausii
B. polyfermenticus SCD
B. pumilus
16.7%2 monthsWaste Site in Incheon, South Korea[295]
Brevibacillus sp. LDPE: 54.21% ± 7%
HDPE: 46.4% ± 5%
4 monthsAn abandoned dump in Karnataka, India[296]
Bacillus anthracis34.76% ± 4.04%4 monthsDandora Dumpsite landfill[297]
Brevibacillus borstelensis strain B2, 2 (MG645267)22.23% ± 2.20%4 monthsDandora Dumpsite landfill[297]
Polyvinyl chloride (PVC)Pseudomonas citronellolis18.56% ± 0.02%1 month[298]
Erysipelothrix rhusiopathiae sp. 11.4% ± 0.6%6 monthsoceanic[299]
Bacillus sp. AIIW20.29%3 monthsoceanic[300]
Polypropylene
(pp)
Stenotrophomonas panacihumi PA3-220.9% ± 1.30% (molecular weight 10 300)
16.8% ± 1.70% (molecular weight 19 700)
3 monthsOpen-air solid wasteyard in South Korea[301]
Nocardia, Rhodococcus, Amycolata,. Amycolatopsis, Gordona and Pseudoamycolata4.4%
6.5%
1 monthSediment samples from mangroves in Matang, Peninsular Malaysia[302]
Bacillus and Pseudomonas1.93% ± 0.18%1 year[303]
Aneurinibacillus sp. 53% ± 2% (PP band)
46.2% ± 3% (PP board)
4 monthsWastewater Treatment Plant and Transfer Station Site[296]
Pseudomonas azotoformans, P. stutzeri, Bacillus subtilis, B. flexus2.6%1 yearEarth for dumping plastic waste[304]
Polystyrene (PS)Enterobacter sp., 15.4%1 monthtransfer station[305]
Pseudomonas sp., Bacillus sp.25%1 monthExploring Soil Composition Near Plastic Waste Dump Site[306]
Pseudomonas aeruginosaBacillus subtilis exhibits the highest degradation efficiency.
Top of Form
terrestrial sample[307]
Acinetobacter sp.12.16%2 monthsTribolium castaneum Intestine of a larva[308]
Pseudomonas aeruginosaTribolium castaneum Intestine of a larva[309]
Exiguobacterium sp. strain YT27.5% ± 0.5%1 monthLarval intestine[310]
Pseudomonas aeruginosa strain DSM 50071Zophobas atratus[311]
polyurethane (PUR/PU)Bacillus subtilis MZA-75, Enhanced Degradation Potential through Co-cultivation1 month[312]
Bacillus amyloliquefaciens35–45%4 weeks[313]
Plastic type of (Fungi)Strain name Evaluation of Degradation Efficiency through Weight Loss AnalysisIncubation timeOrigin of Microbial Strain[314,315]
PE (HDPE)
High-density polyethylene
Aspergillus flavus PEDX33.6125% ± 1.18%1 monthIntestinal contents of the Galleria mellonella[289]
Penicillium oxalicum NS4 (KU559906)20.18%1 monthPlastic wasteyardnear Mohanpur campus[316]
48.53%2 months
45.54%3 months
Penicillium chrysogenum NS10 (KU559907)14.05%1 monthPlastic wasteyard near Mohanpur campus[316]
44.04%2 months
56.548%3 months
Polyvinyl chloride (PVC)Chaetomium globosum (ATCC 16021)“Incipient PVC Absorption by Spores and Hyphae”
Top of Form
1 monthBought standard trunks[317]
Mucor rouxii“Onset of Degradation in Glycerin and Urea Modified PVC”
Top of Form
1 monthPolymer recycling site[318]
Phanerocheate chrysosporium14%1 monthPlastic waste landfill[319]
PE (LDPE)
Polyethylene
Aspergillus oryzae strain A5, 1 (MG779508)34.5% ± 4.55%4 monthsDandola landfill soil[297]
Zalerion maritimum57.4% ± 2.5%14 days[320]
Rhizopus oryzae NS58.6% ± 4%1 monthLaboratory isolate accession number KT160362[321]
Penicillium oxalicum NS4 (KU559906)26.72%1 monthPlastic wasteyard near Mohanpur campus[316]
28.70%1 month
30.60%3 months
Penicillium chrysogenum NS10 (KU559907)30.32%1 monthPlastic wasteyard near Mohanpur campus[316]
36.73%2 months
44.36%3 months
PolypropylenePhanerochaete chrysosporium5.5% (iPP/PLA/nCaCO3 Nanocomposite)1 monthSoil from the Indian campus[319]
Aspergillus niger0.68% (PP)1 month
2.40% (PP/PET/thermoplastic starch blend)1 month [322]
Engyodontium album MTP091 (F2)9.50% (UV pretreatment)I yearPlastic trunk[323]
Lasiodiplodia theobromae1.5%3 monthsPlant Endophytes inhabiting Psychotria flavida[324]
PUR (polyurethane)Aspergillus tubingensis80%60 daysWaste disposal site located in Islamabad, Pakistan.[325]
Chaetomium globosum20–18%3 months[326]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nawaz, T.; Gu, L.; Hu, Z.; Fahad, S.; Saud, S.; Zhou, R. Advancements in Synthetic Biology for Enhancing Cyanobacterial Capabilities in Sustainable Plastic Production: A Green Horizon Perspective. Fuels 2024, 5, 394-438. https://doi.org/10.3390/fuels5030023

AMA Style

Nawaz T, Gu L, Hu Z, Fahad S, Saud S, Zhou R. Advancements in Synthetic Biology for Enhancing Cyanobacterial Capabilities in Sustainable Plastic Production: A Green Horizon Perspective. Fuels. 2024; 5(3):394-438. https://doi.org/10.3390/fuels5030023

Chicago/Turabian Style

Nawaz, Taufiq, Liping Gu, Zhong Hu, Shah Fahad, Shah Saud, and Ruanbao Zhou. 2024. "Advancements in Synthetic Biology for Enhancing Cyanobacterial Capabilities in Sustainable Plastic Production: A Green Horizon Perspective" Fuels 5, no. 3: 394-438. https://doi.org/10.3390/fuels5030023

Article Metrics

Back to TopTop