Carbon Footprint Reduction and Climate Change Mitigation: Comparison
Please note this is a comparison between Version 2 by Mona Zou and Version 1 by Nikolay Lobus.

Since the Industrial Revolution, human economic activity and the global development of society in general have been heavily dependent on the exploitation of natural resources. The use of fossil fuels, deforestation, the drainage of wetlands, the transformation of coastal marine ecosystems, unsustainable land use, and many other unbalanced processes of human activity have led to an increase both in the anthropogenic emissions of climate-active gases and in their concentration in the atmosphere. It is believed that over the past ~150 years these phenomena have contributed to an increase in the global average temperature in the near-surface layer of the atmosphere by ~1 °C. Currently, the most pressing tasks facing states and scientific and civil societies are to reduce anthropogenic CO2 emissions and to limit the global air temperature increase. In this regard, there is an urgent need to change existing production systems in order to reduce greenhouse gas emissions and to sequester them.

  • renewable energy sources
  • carbon sequestration
  • carbon sink
  • CCUS technologies
  • green carbon
  • blue carbon

1. Physicochemical Methods for the Capturing, Separating, Storing, and Using of Anthropogenic CO2

The capture-and-storage technologies (CCS) and capture-and-use technologies (CCU) of carbon (in the form of CO2) are key techniques for reducing anthropogenic greenhouse gas emissions and combating climate change [105][1]. The International Energy Agency (IEA) predicts that reducing emissions cannot be achieved by improving the efficiency of renewable energy use and adjusting the energy mix alone. Without CCS and CCU, the total cost of reducing carbon dioxide emissions will increase by 70% by 2050 [105,106][1][2]. The use of CCS and CCU impacts directly on the cost of energy generated and the rate of economic viability (cost to benefit) of such technologies. As the world continues to rely heavily on fossil energy sources, the need for an efficient method of carbon capture, storage, and/or use is critical to achieving carbon neutrality [107][3].

1.1. CO2 Capture Technologies

CO2 capture is the basis of the CCS technology concept, which was first developed in 1977. Depending on the configuration of fossil fuel power plants, the partial pressure of CO2, and the pressure of the gas stream, there are three approaches by which CO2 may be captured and sequestered. They are classified as pre-combustion, post-combustion, and oxy-combustion carbon capture technologies [105][1].
Technologies for separating CO2 from flue gases, generated by the large-scale combustion of fossil fuels, include physical absorption, chemical absorption, membrane separation, etc. [108][4]. Due to the large volume of flue gases and their low concentrations of CO2, the chemical absorption method is the most suitable technology for separating and capturing CO2 from other combustion gases [105][1]. This method has certain advantages: the process is easy to control and there is no need to strongly modify the power generation system. However, the current focus in gas separation and CO2 capture after combustion is on the searching for effective absorbers and optimizing the process to reduce energy costs [109][5]. A low CO2 concentration in the resulting flue gas mixture is the main reason for a high energy consumption when separating gases after fuel combustion [105][1]. Typically, the CO2 concentration in the exhaust gas of coal-fired power plants is 10−15%, and that of natural gas-fired power plants is even lower (<3−5%), while the volume of these exhaust gases is large. Technologically low concentrations of CO2 in flue gases predetermine the high final costs of the generated energy [1][6].
The method used to separate CO2 before combustion is called pre-combustion. This technology is a promising method for reducing our carbon footprint [110][7]. During the reforming process, fuel is pre-gasified into synthesis gas (mainly composed of H2 and CO). Then, CO in the syngas is converted into CO2 and hydrogen and, afterward, CO2 is separated from the H2. Since CO2 separation occurs before the fuel combustion process, and the fuel gas is not diluted, the CO2 concentration in the synthesis gas is more than 30% [111][8]. According to calculations, capturing 90% of CO2 before combustion reduces the net energy efficiency of the energy production process much less than capturing CO2 after combustion [112][9]. However, the further development of advanced technologies for coal gasification and gas turbines running on hydrogen-enriched gas is necessary for fuel pre-combustion [105][1].
Carbon dioxide capture in the oxy-combustion process is based on technologies for burning fossil fuels in pure O2, in which nitrogen-free flue gases containing only CO2 and H2O are formed. The condensation of flue gases contributes to the production of pure CO2 (concentration of ~95%) and NOx gas impurities. The bottleneck in this method is the need to increase the energy efficiency of the fuel combustion system in pure oxygen through the development of advanced and inexpensive technologies for obtaining O2 from the atmosphere [1,111][6][8].

1.2. CO2 Extraction Technologies

The development of technologies for separating carbon dioxide from the flue/fuel gas stream before its transportation is one of the cutting-edge tasks associated with the implementation of a program to reduce anthropogenic CO2 emissions into the atmosphere. Advanced CO2 recovery techniques have already been developed yet, although technologies such as wet scrubbers, dry regenerable sorbents, membranes, cryogenics, pressure–temperature swing adsorption, and other up-to-date approaches have been proposed. Table 1 provides a comparison of various CO2 extraction technologies [113][10].
Table 1. Comparison of technologies for separating CO2 from the flue gas stream.
Technologies Advantages Flaws Source
Absorption
  • High absorption efficiency (>90%)
  • The absorption efficiency depends on the CO2 concentration in the flue gases
[114][11]
  • Sorbents may be regenerated by heating and/or depressurization
  • A significant amount of heat is required to regenerate the absorbent
  • Most developed CO2 separation process
  • It is necessary to understand the environmental impacts associated with sorbent degradation
Adsorption
  • The process is reversible and the absorbent may be recycled
2 volumes may only be achieved if carbon dioxide utilization processes are combined with renewable energy sources [128][25].
The use of CO2 as an alternative carbon feedstock opens up new opportunities for the producing of fuels and valuable materials or chemicals with a high added value, complementing fossil fuel-based products, and then completely replacing them in the long term. The conversion of carbon into chemicals is an important pathway for CO2 utilization, representing great potential for its sequestration [129][26]. By capturing and utilizing CO2, it is possible to produce various chemicals such as urea, formic acid, salicylic acid, organic carbonates (e.g., acyclic carbonate), cyclic carbonates (e.g., ethylene carbonate), polycarbonates, and fine chemicals such as biotin, etc. [130][27].
Carbon dioxide may be converted to produce such fuels as methane, methanol, and synthesis gas (syngas). Dry methane reforming and hydrogenating are considered the main methods of converting CO2 into fuel. The technical, economic, and environmental performance of the hydrocarbon synthesis process using CO2 has been recently improved [130][27]. The hydrogenation process involves using CO2 instead of CO to produce methanol. The conventional methanol production process is based on the conversion of syngas obtained from natural gas. Methanol is a liquid petrochemical used as an energy carrier in the transportation sector, as a feedstock and solvent, and for the production of other chemicals (e.g., acetic acid, formaldehyde, methylamines) and fuel additives [129][26]. It is a particularly valuable chemical because it may be produced by the low-temperature reaction of CO2 with hydrogen and it is easy to store and transport [131][28].
Mineral carbonation is another method for capturing, storing, and/or using CO2. This process includes the reaction of CO2 with natural minerals or industrial wastes containing metal ions with the subsequent formation of inorganic carbonates [130][27]. Since there is no need to purify the gas to remove impurities (NOxand SOx) formed during fuel combustion, there is an obvious advantage to this method of CO2 collecting. Nitrogen and sulfur oxides do not affect the carbonization reaction; therefore, this allows us to reduce the costs of the process of CO2 capturing and purifying. The resulting mineral carbonates are widely used. In the construction industry, they are fillers and additives in building mixtures or compounds in the production of carbonate blocks, replacing Portland cement-based concrete blocks characterized by a negative carbon footprint. Wastes from the steel or cement industries (i.e., rich in calcium and magnesium oxides) may also be used as an alkali to form carbonates in the presence of CO2. This is a promising technology with a potential CO2 sequestration capacity of up to 3.3 Gt CO2 per year, which could represent 5–12% of its total emissions by 2100. Mineral carbonates, such as hydrotalcite, may be used as catalysts in chemical reactions, for example, in polyester transesterification. In general, the mineral carbonization process is considered not only a method for obtaining high value-added products, but also a method for CO2 storage in geological formations. One ton of CO2 may be absorbed by approximately 1.6–3.7 tons of rock [132][29]. Along with the production of mineral carbonates, the production of organic carbonates is also of great industrial importance. Both linear and cyclic carbonates are generally non-toxic compounds that are widely used for the synthesis of important chemicals, including monomers, polymers, surfactants, plasticizers, and as fuel additives. Aromatic polycarbonates, which do not contain phosgene or aliphatic polycarbonates (such as polypropylene carbonate, polyethylene carbonate, polylimonene carbonate, and polyurethanes), are made from CO2 [133][30].

2. Biological Methods for Capturing, Storing, and Using CO2

Searching for the environmentally friendly technologies that make it possible to ensure the necessary level of economic growth without creating additional risks for the environment is one of the key areas of technological development worldwide [134,135][31][32]. Among the wide range of living organisms tested during the inventing and developing of innovative biotechnologies, green plants are the most popular and promising object of research, since they are widely used in various areas of human economic activity [136][33]. Currently, methods for the biological sequestration of climate-active gases are being actively developed simultaneously to physical and chemical approaches to capturing and storing CO2. Biological carbon fixation technologies use the photosynthesis of green plants, inhabiting both aquatic and terrestrial ecosystems, to convert CO2 into organic matter, ensuring the C–O balance in the atmosphere [137,138][34][35]. The technologies focused on these strategies for the biological assimilation of CO2 include blue carbon sequestration and green carbon sequestration [19,139][36][37].

2.1. Technologies for CO2 Sequestration by Terrestrial Ecosystems (“Green Carbon”)

Under natural conditions, terrestrial ecosystems are sinks for atmospheric CO2, influencing significantly the global carbon cycle [140][38]. A part of the carbon fixed by plants is converted into stable soil organic carbon (SOC) through forming organomineral complexes. Another part of the carbon contributes to the formation of soil inorganic carbon (SIC) due to the formation of carbonates/bicarbonates of calcium, magnesium, potassium, and sodium. These two systems provide soil carbon storage [141][39]. The gross primary productivity (GPP) of terrestrial ecosystems represents the annual flow of carbon between the atmosphere and the land surface. However, only about 8% of the GPP remains in the ecosystem as net primary production (NPP). The rest is lost to the atmosphere through plant and microbial respiration and heterotrophic nutrition. The main processes associated with terrestrial carbon sequestration include the retention of fixed carbon as NPP and the formation of SOC and SIC [142][40]. CO2 sequestration by terrestrial ecosystems may be increased through technologies aimed at increasing the carbon content of biomass and soil, namely conservation agriculture, agroforestry, biochar applications, and forest and wetland restoration (Table 2).
Table 2. Approaches to increasing the soil carbon content.
Sequestration Strategy Potential Increase in VOC Stocks Advantages Flaws Source
Resource-saving rural farming up to 1.01 t C ha−1 year−1
  • Increasing biodiversity
  • Improving soil nutrition
  • Increasing water retention
  • Various data on effects on SOC
  • Various data on the impact on crop yields
[143][41]
Agroforestry up to 5.3 Gt C year−1 (global)
  • Increase in aboveground biomass
  • Increasing carbon input into soil
  • The level of SOC sequestration depends on the climate, soil type, management practices, age and type of organic feedstock
[144,145][42][43]
Reforestation 0.75–5.80 Gt C year−1 depending on the price of land, region, and time
  • Increase in aboveground biomass
  • Increasing carbon input into soil
  • Different impacts on SOC stocks depending on the soil type, climate, tree species
[19,146][36][44]
  • Requires a high-temperature adsorbent
[115][12]
  • High adsorption efficiency (>85%) is achievable
  • CO2 desorption requires a lot of energy
Chemical loop combustion
  • CO2 is the main combustion product that does not mix with N2, thereby avoiding energy-intensive air separation
  • The process is still in development and there is no experience in its large-scale operation
[116][13]
Membrane separation
  • The process has been adopted for the separation of other gases
  • Operational problems include low flows and clogging
[114][11]
  • High separation efficiency is achievable (>80%)
Cryogenic distillation
  • Mature technology
  • Only suitable for very high CO2 concentrations in flue gases (>90% by volume)
[117][14]
  • Used for many years in industry for CO2 recovery
  • Should be carried out at very low temperatures
  • The process is very energy intensive

1.3. CO2 Transportation and Storage Technologies

After CO2 is separated from the remaining components of the flue gas, it must be transported to a storage location or to facilities for its further use. Whatever the final fate of CO2, a reliable, safe, and cost-effective transport system is a key feature of any CCS or CCU project. Depending on the volumes involved, a variety of vehicles may be used, from tank trucks to offshore vessels and pipelines. Pipelines are considered the most viable method for transporting large volumes of CO2 over long distances overland. For commercial-scale CCS projects, it is necessary to develop an extensive network of these pipelines [118,119][15][16]. In order to optimize the mass/volume ratio, CO2 is transported as a dense phase in either a liquid or a supercritical state. This requires maintaining certain temperature and pressure conditions. These technologies are quite mature and are actively used in various types of carbon dioxide transportation. However, impurities in the gases and water vapor in CO2 capture pose a serious problem, since their presence may change the boundaries of the range of pressures and temperatures providing a stable single-phase state to pure carbon oxide [113][10]. Once captured, the high-CO2 stream may be transported for long-term storage or industrial reuse to produce high value-added products.
Technologies for storing CO2 after its extraction from flue gases, generated by the combustion of hydrocarbon fossil fuels, are critical technologies for implementing CCS projects and achieving carbon neutrality targets [120,121][17][18]. Compressed carbon dioxide may be stored in geological formations, such as deep saline aquifers that have no other practical use, and oil, gas, or coal reservoirs. Storing CO2 in porous geological media is a promising method for reducing anthropogenic greenhouse gas emissions [122][19]. A typical geological repository may contain dozens of millions of tons of CO2 captured through a variety of physical and chemical methods. Typically, three different geological formations are considered for CO2 storage: depleted (or nearly depleted) oil and gas reservoirs, stranded coal seams, and saline aquifers. The CO2 storage potential may be as high as 400–10,000 GT for deep-saline aquifers, about 920 GT for depleted oil and gas fields, and >15 GT for undeveloped coal seams [123][20]. However, suitable geological sites for CO2 storage must be carefully selected. The general requirements for the geological storage of CO2 include the specific tectonic setting and geology of the basin, its geothermal regime, hydrology, hydrocarbon potential, maturity of the basin, porosity, the thickness and permeability of the reservoir rock, the presence of cap rock with a good sealing ability, and stable seismic conditions [122,124][19][21]. However, technologies for CO2 storage in geological reservoirs have certain risks that must be taken into account when choosing the final pool for carbon dioxide injection. Injecting CO2 into saline aquifers lowers the pH of the brine and dissolves iron carbonates and oxyhydrates. The dissolution of carbonates and minerals weakens the surrounding rocks and may create cracks, through which CO2 will leak, so the chemical equilibrium of geological formations may change, creating mobile toxic trace elements and organic compounds [121,125][18][22].
Deep ocean storage technologies are another option for CO2 immobilization. However, they are not technically mature, as they are at the stage of laboratory development and causing active discussion and criticism in the scientific community. Their essence lies in the fact that CO2 liquefies at depths of more than 3 km and sinks to the bottom of the sea due to its higher density than the surrounding seawater. Mathematical models suggest that CO2 introduced in this way may provide the permanent geological storage of CO2 even under large geomechanical disturbances [112][9]. However, this approach is more controversial than other geological storage methods. The direct release of large amounts of CO2 into the ocean may affect the chemistry of seawater, which may lead to catastrophic consequences for marine ecosystems. Comparatively less research has been conducted in this area, especially regarding the impact of CO2 on marine ecosystems [126][23]. Although the Intergovernmental Panel on Climate Change has recognized the potential of ocean CO2 storage, it has also noted the risks that this technology may pose [112][9].
Since CCS technologies involve the long-term immobilization of CO2, monitoring storage locations with highly qualified personnel and appropriate infrastructure becomes necessary. According to an IPCC report, the monitoring of disposal sites will require a slight increase in energy consumption, but this increases significantly the cost of storage, so monitoring costs may be similar to the costs of transporting CO2 to conservation sites [112,121][9][18].

1.4. CO2 Utilization Technologies

Undoubtedly, underground CO2 storage technologies in geological reservoirs are the fastest and largest-scale solution aimed at reducing anthropogenic greenhouse gas emissions to achieve carbon neutrality in the world economy [127][24]. However, it should be taken into account that the current volume of the secondary industrial use of CO2 (>200 Mt per year) is clearly insufficient compared to the global CO2 production (37,000 Mt per year). The conversion of large CO
Wetland restoration
0.35–1.10 t C ha
−1 year−1 depending on the landscape and depth
  • Higher SOC content compared to cultivated wetlands
  • Varying mitigation potential depending on natural and anthropogenic factors
[147][45]
Biochar can offset up to 12% of annual net anthropogenic CO2 emissions
  • Initial carbon retention of 50%
  • Acts as both a source and sink for soil carbon
  • Various data on effects on soil quality
  • Different effects on priming vary depending on soil type
  • Differential impacts on SOC stocks depending on biochar type and age
[19,148][36][46]
The Food and Agriculture Organization (FAO) defines conservation agriculture (CA) as a system of farming that promotes minimal soil disturbance (by using no-till or low-till practices) and crop diversification, and thus maintains permanent soil cover [143][41]. Because CA is often considered synonymous with low or no tillage, this issue demands assessing the potential of CA methods to improve the SOC and its yield due to data variability. Because three-part CA is often studied separately or applied in research or practice, the actual impact of CA on POC is unclear. The amount of carbon input plays a decisive role in increasing the SOC due to its GPP. The type of crop, intensity, and duration of cultivation predetermine the carbon input. Using deep-rooted plant species with high above- and below-ground biomass and increasing the number of harvests per year may thus increase the carbon storage capacity of CA systems [149][47]. In addition to the benefits associated with SOC, CA improves farm economics, the planting schedule’s flexibility, weed control, soil protection and fertilization, the efficiency of nutrient usage, and water usage and retention. This has accelerated the implementation of CA worldwide with the average rate of global expansion amounting to 10.5 million hectares of arable land per year [150][48].
Agroforestry is alternative climate change mitigation strategy. The system integrates trees into the agricultural landscape. The conversion of land use from forests and grasslands to plantations significantly reduces SOC stocks. Agroforestry may restore up to 35% of lost forest carbon reserves [151][49]. Initially, the carbon pool of agricultural biomass was considered insignificant compared to SOC. However, agroforestry makes a significant contribution to carbon storage in terms of agricultural biomass. In particular, trees have a positive effect on SOC through their deeper deposition and reduced decomposition capacity [152][50]. The changes in soil carbon stocks resulting from agroforestry vary with tree species, tree population density, and climate, but generally appear to fit the range between agricultural and forest carbon stocks [153][51]. All studies conducted to date in the field of agroforestry have shown an increase in SOC storage compared to agricultural monocultures, in which SOC storage rates were influenced by climate zones, the management methods used, the age of the system, the soil type, and the type of organic raw materials [154][52]. Agroforestry systems may store ~5.3 Gt of carbon over 944 million hectares worldwide, with the greatest potential in the tropics and subtropics [144][42].
The idea of forest restoration is central to climate change mitigation because it offers an effective alternative to other, more costly measures involving renewable energy or industrial CO2 sequestration. Since the 1990s, afforestation has been widely adopted throughout the world, increasing the area of artificial planted forests by approximately 1.05 × 108 ha. However, achieving mitigation targets through afforestation depends on the area and carbon sequestration potential of each individual forest stand [112][9]. From a long-term perspective, the mitigation potential of climate change through afforestation varies widely from 1.5 to 4.9 Gt CO2 year−1 by 2050 and from 1.1 to 5.8 Gt CO2 year−1 by 2100 [155][53]. However, the effects of forest restoration on soil carbon dynamics are not well understood yet and a variety of assumptions have been made so far. SOC stocks may increase, decrease, or remain unchanged after afforestation under the influence of factors such as tree species, land use history, plant age, climate and soil type, etc. [156][54].
Currently, about 50% of the world’s wetlands have been lost, due to either agriculture, industry, or urbanization. The loss of this species-rich habitat both has a major impact on biodiversity decline and increases CO2 emissions. Wetlands and their associated soils represent a large soil carbon reservoir [147][45]. The drainage and anthropogenic transformation of these landscapes transform them from a major carbon sink into a major source of CO2. Restoring wetlands by banning their development and replenishing their SOC is believed to be a potential approach for mitigating climate change [19][36]. However, assessing the potential of their restoration to meet decarbonization goals is problematic due to the many variables affecting wetland dynamics [157,158][55][56]. When analyzing the costs of wetland restoration in relation to the carbon sequestered, it appears that restoration is more cost-effective in coastal areas (such as mangroves) compared to inland wetlands. In the latter case, conservation rather than restoration is recommended [159][57].
Biochar is a biomass-derived solid material derived mainly through pyrolysis, a thermochemical process completed under high-temperature oxygen-deficit conditions. Syngas and bio-oil are formed in addition to biochar as a result of this process. Biochar is a stable product with a half-life of several hundred to several thousand years, which has emerged as a promising solution for soil carbon sequestration [160,161][58][59]. Compared to burning or decomposing crop residues, which retain only ~3% and 10−20% carbon, respectively, biochar retains about 50% of its original content [162][60]. It is estimated that biochar may offset up to 12% of annual net anthropogenic CO2 emissions [163][61]. The biochar industry and market have grown around the world, realizing its potential for carbon capture and agricultural production. However, recent studies report contrasting results on the effects of biochar on soil quality, nutrient availability, and soil carbon mineralization [19][36]. Together, soil properties (intrinsic SOC content, pH, soil type) and biochar properties (pyrolysis temperature, source type, application rate, age, C:N ratio) jointly influence the effectiveness of the biochar addition to soil in terms of soil content [164][62]. Long-term field experiments with biochar are needed to assess the climate change mitigation potential of biochar [165][63]. Currently, various factors affecting the SOC uptake from biochar applications make it difficult to assess its decarbonization potential, so the economic feasibility of implementing large-scale biochar application is unclear [165][63]. The average increase in SOC content with biochar use is estimated to be 45.8%, but with large regional differences. The global potential of biochar for climate change mitigation using existing volumes of plant pyrolysis is estimated at 6.6 Mt C year−1 [148][46].

2.2. Technologies for CO2 Sequestration by Aquatic Ecosystems (“Blue Carbon”)

The term “blue carbon” was coined in 2009 and was initially used as a metaphor to draw attention to coastal ecosystems and their role in CO2 sequestering and storing. This metaphor subsequently evolved into a strategy for mitigating and adapting to climate change through the conservation and restoration of biodiversity in coastal marine ecosystems [166][64]. The blue carbon concept places particular emphasis on areas with rich vegetation, which primarily include seagrass thickets, tidal marshes, and mangroves. Blue carbon ecosystems have a high potential for converting CO2 into plant biomass. The average carbon sequestration potential is estimated at 24.0 ± 3.2 Mt C year−1 for mangroves, 13.4 ± 1.4 Mt C year−1 for salt marshes, and 43.9 ± 12.1 Mt C year−1 for seagrass thickets at the global scale [167][65]. The low oxygen concentration in their soil reduces the mineralization of organic matter in the plants’ biomass, and improves their accumulation and preservation of organomineral detritus. That is why coastal biotopes, belonging to blue carbon ecosystems, are hotbeds of CO2 sequestration [168][66].
Numerous studies of mangrove, tidal marsh, and seagrass ecosystems reveal their potential for an integrated approach to mitigate climate change. This led many countries to include measures for the restoration and conservation of coastal ecosystems in their national protocols for implementing the Paris Agreement goals [169][67]. However, the practice of destructive land use in coastal ecosystems turns these water areas from CO2 accumulators into its sources. The disruption of ecosystem networks contributes to the mobilization of previously accumulated carbon and its release into the atmosphere. According to experts, disturbed mangroves and seagrass areas account for 18% and 29% of CO2 emissions in all tropical coastal ecosystems, respectively [170][68]. Protecting existing blue carbon ecosystems could prevent the release of 304 Mt of inorganic carbon per year, with large-scale restoration potentially eliminating an additional 841 Mt per year by 2030, equivalent to 3% of annual global greenhouse gas emissions [171][69].
The potential of other marine ecosystems as CO2 sequestration sites is the subject of ongoing debate. Calcifying organisms release CO2 during the calcification process, making coral reefs and oyster banks a source of CO2 rather than a sink. However, the environment-forming role of coral reefs have a beneficial effect on the accumulation of carbonate sediments, the development of sea meadows and mangroves; in turn, the latter prevent high-turbidity water from entering coral reefs. Therefore, coral reefs are interdependent with mangroves and seagrass, which enhances their inorganic carbon sequestration potential [172][70].
Pelagic ecosystems of the world ocean, especially macro- and microalgae, are considered another important link in the sequestration of CO2 from the atmosphere [134,139,173,174][31][37][71][72]. The net primary production of phytoplankton in the world ocean is ~60 × 103 Mt of organic carbon per year, which corresponds to ~200 × 103 Mt of CO2 captured by primary producers from the atmosphere [175][73]. However, only a small part of the assimilated carbon reaches the bottom and is buried in sediments (~250 Mt), equivalent to ~900–950 Mt CO2. At the same time, the bottom sediments of the world ocean are one of the main reservoirs of carbon storage in the biosphere on a geological time scale [176,177,178][74][75][76].
Freshwater ecosystems are of no small importance. Despite the significant difficulty in estimating the total production of freshwater macro- and microalgae, it is believed that ~70–72 Mt of organic carbon per year accumulates in river sediments, which is equivalent to ~250 Mt CO2 per annum [177][75].
The development of blue carbon programs may undoubtedly have a significant impact on our efforts to reduce and sequester CO2 emissions. However, some uncertainties, such as the consequences of climate change for aquatic ecosystems, the degree of greenhouse gas emissions as a result of the anthropogenic transformation of aquatic landscapes, the role of freshwater and marine pelagic ecosystems in the global carbon cycle, etc., require further fundamental research involving large investments [179][77].

3. Technologies for Biological CO2 Sequestration and the Production of Products with a High Added Value

The biotechnologies based on the use of the photosynthetic activity of microalgae and cyanobacteria have received widespread development in solving applied problems related to CO2 capture and its recycling [134,180,181][31][78][79]. Historically, microalgae have been widely used as an alternative source of raw materials for the production of renewable green energy, so CO2 sequestration by these means was initially established as one of the potential directions for sustainable energy development research in many countries around the world [182,183][80][81]. The uniqueness of microalgae is that they may capture CO2 from various sources, including CO2 from the atmosphere, from flue gases, and even in the form of soluble carbonates, and then process it into high-carbon organic biomass [138][35]. Historically, researchers have studied microalgae both for the discovery and development of alternative energy resources and for the development of technologies protecting the environment from air pollution and allowing the sequestration of climate-active gases, primarily CO2 [184,185][82][83].
The efficiency of microalgae at assimilating CO2 using solar energy is 10–50 times higher than that of terrestrial plants [186,187][84][85]. In terms of CO2, 1.0 kg of algae biomass may assimilate ~1.83 kg of CO2, which makes it possible to rear microalgae near thermal power plants or any other sources of greenhouse gases [181,182,188][79][80][86]. Currently, microalgae are actively used to obtain a wide range of biologically active components, such as proteins (including the production of amino acids), fats (including polyunsaturated fatty acids), carbohydrates (including starch and fiber), carotenoids, pigments, vitamins, and biologically active forms of major and trace elements [189,190,191][87][88][89]. This allows us to consider them alternative and industrially promising sources that ensure the sustainable production of many commercial products with a high added value [135,173,189][32][71][87].
Along with the physical and chemical methods of CO2 sequestration, biological methods are dynamically developing [10][90]. The latter have some advantages. Firstly, their use makes it possible to reduce the volume of anthropogenic CO2 emissions and to produce, simultaneously, a wide range of biologically active compounds that are highly commercially attractive for investment [185,192][83][91]. Secondly, the cultivation of microalgae may be carried out using wastewater to provide them with nutrients. This promotes a combined approach for solving the important environmental problems associated not only with the reduction of anthropogenic CO2 emissions, but also with the development of water purification technologies [193][92]. Third, microalgae may survive and adapt to various extreme conditions. Their cultivation may be carried out on lands and in reservoirs, where the soil, water quality, or climate are not suitable for growing conventional crops or aquaculture, so they do not demand arable land or a source of clean water. This reduces their competition with agricultural crops for fertile land, increasing food crop areas [194,195,196][93][94][95]. This approach helps to redistribute income in dry regions and to create new jobs for the local population [197][96].
The main problem associated with the use of biological CO2 sequestration is the high temperatures of the flue gases and the presence of CO, NOx, SOx, and certain amounts of other impurities in the fossil fuel used [10,181][79][90]. The process of CO2 sequestration requires detailed knowledge of the component composition of flue gases and cell biology. The main factors influencing this process may be temperature, pH, SOx and NOx, light, microalgae strain, culture density, critical CO2 concentration, CO2 mass transfer, and O2 accumulation. Growing algae also requires selecting the species and strain for cultivation and developing a suitable photobioreactor [138,198][35][97].
Despite the promise of using biological methods of CO2 sequestration that use microalgae, there are still a large number of unresolved issues associated with them: as a rule, they concern assessing the economic profitability of the technological chains [181,183][79][81]. Currently, active work is underway, the main result of which is the development of a set of biotechnological approaches associated with the sequestration of CO2 by microalgae and the production of biologically active components with a high added value [10][90]. Economic profitability, in this case, is achieved by localizing the entire biotechnological chain (sequestration → production of bioproducts → primary processing) to one location [134,181,185,199][31][79][83][98].
Despite all the difficulties associated with the development of innovative technologies for biological CO2 sequestration, it is believed that this will be an economically feasible, environmentally friendly, and sustainable technology for CO2 fixing and obtaining the biologically active components produced by microalgae in the long term [2,181,195,200][79][94][99][100].

References

  1. Wilberforce, T.; Olabi, A.G.; Sayed, E.T.; Elsaid, K.; Abdelkareem, M.A. Progress in carbon capture technologies. Sci. Total Environ. 2021, 761, 143203.
  2. IEA. UNIDO Technology Roadmap: Carbon Capture and Storage in Industrial Applications. Available online: https://www.iea.org/reports/technology-roadmap-carbon-capture-and-storage-2013 (accessed on 1 August 2023).
  3. Zhang, Z.; Pan, S.-Y.; Li, H.; Cai, J.; Olabi, A.G.; Anthony, E.J.; Manovic, V. Recent advances in carbon dioxide utilization. Renew. Sustain. Energy Rev. 2020, 125, 109799.
  4. Chao, C.; Deng, Y.; Dewil, R.; Baeyens, J.; Fan, X. Post-combustion carbon capture. Renew. Sustain. Energy Rev. 2021, 138, 110490.
  5. Zhang, X.; Singh, B.; He, X.; Gundersen, T.; Deng, L.; Zhang, S. Post-combustion carbon capture technologies: Energetic analysis and life cycle assessment. Int. J. Greenh. Gas Control 2014, 27, 289–298.
  6. Wang, F.; Harindintwali, J.D.; Yuan, Z.; Wang, M.; Wang, F.; Li, S.; Yin, Z.; Huang, L.; Fu, Y.; Li, L.; et al. Technologies and perspectives for achieving carbon neutrality. Innovation 2021, 2, 100180.
  7. Jansen, D.; Gazzani, M.; Manzolini, G.; van Dijk, E.; Carbo, M. Pre-combustion CO2 capture. Int. J. Greenh. Gas Control 2015, 40, 167–187.
  8. Nazir, S.M.; Cloete, J.H.; Cloete, S.; Amini, S. Efficient hydrogen production with CO2 capture using gas switching reforming. Energy 2019, 185, 372–385.
  9. Metz, B.; Davidson, O.; De Coninck, H.C.; Loos, M.; Meyer, L. IPCC Special Report on Carbon Dioxide Capture and Storage; Cambridge University Press: Cambridge, UK, 2005.
  10. Leung, D.Y.C.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443.
  11. Aaron, D.; Tsouris, C. Separation of CO2 from Flue Gas: A Review. Sep. Sci. Technol. 2005, 40, 321–348.
  12. Clausse, M.; Merel, J.; Meunier, F. Numerical parametric study on CO2 capture by indirect thermal swing adsorption. Int. J. Greenh. Gas Control 2011, 5, 1206–1213.
  13. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gayan, P.; de Diego, L.F. Progress in Chemical-Looping Combustion and Reforming technologies. Prog. Energy Combust. Sci. 2012, 38, 215–282.
  14. Tuinier, M.J.; van Sint Annaland, M.; Kramer, G.J.; Kuipers, J.A.M. Cryogenic CO2 capture using dynamically operated packed beds. Chem. Eng. Sci. 2010, 65, 114–119.
  15. Witkowski, A.; Majkut, M.; Rulik, S. Analysis of pipeline transportation systems for carbon dioxide sequestration. Arch. Thermodyn. 2014, 35, 117–140.
  16. Aminu, M.D.; Nabavi, S.A.; Rochelle, C.A.; Manovic, V. A review of developments in carbon dioxide storage. Appl. Energy 2017, 208, 1389–1419.
  17. Rackley, S.A. Carbon Capture and Storage, 2nd ed.; Elsevier: Oxford, UK, 2017; ISBN 978-0-12-812041-5.
  18. Michaelides, E.E. Thermodynamic analysis and power requirements of CO2 capture, transportation, and storage in the ocean. Energy 2021, 230, 120804.
  19. Ali, M.; Jha, N.K.; Pal, N.; Keshavarz, A.; Hoteit, H.; Sarmadivaleh, M. Recent advances in carbon dioxide geological storage, experimental procedures, influencing parameters, and future outlook. Earth-Sci. Rev. 2022, 225, 103895.
  20. IEA Greenhouse Gas Programme. Improvement in Power Generation with Post-Combustion Capture of CO2. Available online: https://ieaghg.org/docs/General_Docs/Reports/PH4-33%20post%20combustion.pdf (accessed on 15 November 2004).
  21. Baines, S.J.; Worden, R.H. Geological Storage of Carbon Dioxide; Geological Society London Special Publications: Bath, UK, 2004; Volume 233, pp. 1–6.
  22. Kharaka, Y.K.; Cole, D.R.; Hovorka, S.D.; Gunter, W.D.; Knauss, K.G.; Freifeld, B.M. Gas-water-rock interactions in Frio Formation following CO2 injection: Implications for the storage of greenhouse gases in sedimentary basins. Geology 2006, 34, 577.
  23. Seibel, B.A.; Walsh, P.J. Potential Impacts of CO2 Injection on Deep-Sea Biota. Science 2001, 294, 319–320.
  24. Vasudevan, S.; Aggarwal, S.; Farooq, S.; Karimi, I.A.; Quah, M.C.G. Technoenergetic and Economic Analysis of CO2 Conversion. In An Economy Based on Carbon Dioxide and Water; Aresta, M., Karimi, I., Kawi, S., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 413–430.
  25. Nocito, F.; Dibenedetto, A. Atmospheric CO2 mitigation technologies: Carbon capture utilization and storage. Curr. Opin. Green Sustain. Chem. 2020, 21, 34–43.
  26. Chauvy, R.; Meunier, N.; Thomas, D.; De Weireld, G. Selecting emerging CO2 utilization products for short- to mid-term deployment. Appl. Energy 2019, 236, 662–680.
  27. Ghiat, I.; Al-Ansari, T. A review of carbon capture and utilisation as a CO2 abatement opportunity within the EWF nexus. J. CO2 Util. 2021, 45, 101432.
  28. Abdelaziz, O.Y.; Hosny, W.M.; Gadalla, M.A.; Ashour, F.H.; Ashour, I.A.; Hulteberg, C.P. Novel process technologies for conversion of carbon dioxide from industrial flue gas streams into methanol. J. CO2 Util. 2017, 21, 52–63.
  29. Alper, E.; Yuksel Orhan, O. CO2 utilization: Developments in conversion processes. Petroleum 2017, 3, 109–126.
  30. Grignard, B.; Gennen, S.; Jérôme, C.; Kleij, A.W.; Detrembleur, C. Advances in the use of CO2 as a renewable feedstock for the synthesis of polymers. Chem. Soc. Rev. 2019, 48, 4466–4514.
  31. Lobus, N.V. Biogeochemical Role of Algae in Aquatic Ecosystems: Basic Research and Applied Biotechnology. J. Mar. Sci. Eng. 2022, 10, 1846.
  32. Orejuela-Escobar, L.; Gualle, A.; Ochoa-Herrera, V.; Philippidis, G.P. Prospects of Microalgae for Biomaterial Production and Environmental Applications at Biorefineries. Sustainability 2021, 13, 3063.
  33. Ścieszka, S.; Klewicka, E. Algae in food: A general review. Crit. Rev. Food Sci. Nutr. 2019, 59, 3538–3547.
  34. da Rosa, G.M.; de Morais, M.G.; Costa, J.A.V. Green alga cultivation with monoethanolamine: Evaluation of CO2 fixation and macromolecule production. Bioresour. Technol. 2018, 261, 206–212.
  35. Kumar, K.; Dasgupta, C.N.; Nayak, B.; Lindblad, P.; Das, D. Development of suitable photobioreactors for CO2 sequestration addressing global warming using green algae and cyanobacteria. Bioresour. Technol. 2011, 102, 4945–4953.
  36. Nayak, N.; Mehrotra, R.; Mehrotra, S. Carbon biosequestration strategies: A review. Carbon Capture Sci. Technol. 2022, 4, 100065.
  37. Lobus, N.V.; Kulikovskiy, M.S. The Co-Evolution Aspects of the Biogeochemical Role of Phytoplankton in Aquatic Ecosystems: A Review. Biology 2023, 12, 92.
  38. Keenan, T.F.; Williams, C.A. The Terrestrial Carbon Sink. Annu. Rev. Environ. Resour. 2018, 43, 219–243.
  39. Lal, R. Sequestration of atmospheric CO2 in global carbon pools. Energy Environ. Sci. 2008, 1, 86.
  40. Lal, R.; Smith, P.; Jungkunst, H.F.; Mitsch, W.J.; Lehmann, J.; Nair, P.K.R.; McBratney, A.B.; de Moraes Sá, J.C.; Schneider, J.; Zinn, Y.L.; et al. The carbon sequestration potential of terrestrial ecosystems. J. Soil Water Conserv. 2018, 73, 145A–152A.
  41. Pittelkow, C.M.; Liang, X.; Linquist, B.A.; van Groenigen, K.J.; Lee, J.; Lundy, M.E.; van Gestel, N.; Six, J.; Venterea, R.T.; van Kessel, C. Productivity limits and potentials of the principles of conservation agriculture. Nature 2015, 517, 365–368.
  42. Shi, L.; Feng, W.; Xu, J.; Kuzyakov, Y. Agroforestry systems: Meta-analysis of soil carbon stocks, sequestration processes, and future potentials. Land Degrad. Dev. 2018, 29, 3886–3897.
  43. Tschora, H.; Cherubini, F. Co-benefits and trade-offs of agroforestry for climate change mitigation and other sustainability goals in West Africa. Glob. Ecol. Conserv. 2020, 22, e00919.
  44. Zomer, R.J.; Neufeldt, H.; Xu, J.; Ahrends, A.; Bossio, D.; Trabucco, A.; van Noordwijk, M.; Wang, M. Global Tree Cover and Biomass Carbon on Agricultural Land: The contribution of agroforestry to global and national carbon budgets. Sci. Rep. 2016, 6, 29987.
  45. Tangen, B.A.; Bansal, S. Soil organic carbon stocks and sequestration rates of inland, freshwater wetlands: Sources of variability and uncertainty. Sci. Total Environ. 2020, 749, 141444.
  46. Han, M.; Zhao, Q.; Li, W.; Ciais, P.; Wang, Y.; Goll, D.S.; Zhu, L.; Zhao, Z.; Wang, J.; Wei, Y.; et al. Global soil organic carbon changes and economic revenues with biochar application. GCB Bioenergy 2022, 14, 364–377.
  47. Palm, C.; Blanco-Canqui, H.; DeClerck, F.; Gatere, L.; Grace, P. Conservation agriculture and ecosystem services: An overview. Agric. Ecosyst. Environ. 2014, 187, 87–105.
  48. Kassam, A.; Friedrich, T.; Derpsch, R. Global spread of Conservation Agriculture. Int. J. Environ. Stud. 2019, 76, 29–51.
  49. Abbas, F.; Hammad, H.M.; Fahad, S.; Cerdà, A.; Rizwan, M.; Farhad, W.; Ehsan, S.; Bakhat, H.F. Agroforestry: A sustainable environmental practice for carbon sequestration under the climate change scenarios—A review. Environ. Sci. Pollut. Res. 2017, 24, 11177–11191.
  50. De Stefano, A.; Jacobson, M.G. Soil carbon sequestration in agroforestry systems: A meta-analysis. Agrofor. Syst. 2017, 92, 285–299.
  51. Chatterjee, N.; Nair, P.K.R.; Chakraborty, S.; Nair, V.D. Changes in soil carbon stocks across the Forest-Agroforest-Agriculture/Pasture continuum in various agroecological regions: A meta-analysis. Agric. Ecosyst. Environ. 2018, 266, 55–67.
  52. Rakotovao, N.H.; Rasoarinaivo, A.R.; Razafimbelo, T.; Blanchart, E.; Albrecht, A. Organic inputs in agroforestry systems improve soil organic carbon storage in Itasy, Madagascar. Reg. Environ. Chang. 2022, 22, 6.
  53. Doelman, J.C.; Stehfest, E.; Vuuren, D.P.; Tabeau, A.; Hof, A.F.; Braakhekke, M.C.; Gernaat, D.E.H.J.; Berg, M.; Zeist, W.; Daioglou, V.; et al. Afforestation for climate change mitigation: Potentials, risks and trade-offs. Glob. Chang. Biol. 2020, 26, 1576–1591.
  54. Hong, S.; Yin, G.; Piao, S.; Dybzinski, R.; Cong, N.; Li, X.; Wang, K.; Peñuelas, J.; Zeng, H.; Chen, A. Divergent responses of soil organic carbon to afforestation. Nat. Sustain. 2020, 3, 694–700.
  55. Tangen, B.A.; Finocchiaro, R.G.; Gleason, R.A. Effects of land use on greenhouse gas fluxes and soil properties of wetland catchments in the Prairie Pothole Region of North America. Sci. Total Environ. 2015, 533, 391–409.
  56. Tangen, B.A.; Bansal, S. Hydrologic Lag Effects on Wetland Greenhouse Gas Fluxes. Atmosphere 2019, 10, 269.
  57. Taillardat, P.; Thompson, B.S.; Garneau, M.; Trottier, K.; Friess, D.A. Climate change mitigation potential of wetlands and the cost-effectiveness of their restoration. Interface Focus 2020, 10, 20190129.
  58. Hussain, M.; Farooq, M.; Nawaz, A.; Al-Sadi, A.M.; Solaiman, Z.M.; Alghamdi, S.S.; Ammara, U.; Ok, Y.S.; Siddique, K.H.M. Biochar for crop production: Potential benefits and risks. J. Soils Sediments 2017, 17, 685–716.
  59. Li, S.; Tasnady, D. Biochar for Soil Carbon Sequestration: Current Knowledge, Mechanisms, and Future Perspectives. J. Carbon Res. 2023, 9, 67.
  60. El-Naggar, A.; Awad, Y.M.; Tang, X.-Y.; Liu, C.; Niazi, N.K.; Jien, S.-H.; Tsang, D.C.W.; Song, H.; Ok, Y.S.; Lee, S.S. Biochar influences soil carbon pools and facilitates interactions with soil: A field investigation. Land Degrad. Dev. 2018, 29, 2162–2171.
  61. Woolf, D.; Amonette, J.E.; Street-Perrott, F.A.; Lehmann, J.; Joseph, S. Sustainable biochar to mitigate global climate change. Nat. Commun. 2010, 1, 56.
  62. Xu, H.; Cai, A.; Wu, D.; Liang, G.; Xiao, J.; Xu, M.; Colinet, G.; Zhang, W. Effects of biochar application on crop productivity, soil carbon sequestration, and global warming potential controlled by biochar C:N ratio and soil pH: A global meta-analysis. Soil Tillage Res. 2021, 213, 105125.
  63. Gross, A.; Bromm, T.; Glaser, B. Soil Organic Carbon Sequestration after Biochar Application: A Global Meta-Analysis. Agronomy 2021, 11, 2474.
  64. Wylie, L.; Sutton-Grier, A.E.; Moore, A. Keys to successful blue carbon projects: Lessons learned from global case studies. Mar. Policy 2016, 65, 76–84.
  65. Bertram, C.; Quaas, M.; Reusch, T.B.H.; Vafeidis, A.T.; Wolff, C.; Rickels, W. The blue carbon wealth of nations. Nat. Clim. Chang. 2021, 11, 704–709.
  66. Geraldi, N.R.; Ortega, A.; Serrano, O.; Macreadie, P.I.; Lovelock, C.E.; Krause-Jensen, D.; Kennedy, H.; Lavery, P.S.; Pace, M.L.; Kaal, J.; et al. Fingerprinting Blue Carbon: Rationale and Tools to Determine the Source of Organic Carbon in Marine Depositional Environments. Front. Mar. Sci. 2019, 6, 236.
  67. Herr, D.; Landis, E. Coastal Blue Carbon Ecosystems: Opportunities for Nationally Determined Contributions; IUCN: Gland, Switzerland, 2016.
  68. Alongi, D.M. Global Significance of Mangrove Blue Carbon in Climate Change Mitigation. Science 2020, 2, 67.
  69. Macreadie, P.I.; Costa, M.D.P.; Atwood, T.B.; Friess, D.A.; Kelleway, J.J.; Kennedy, H.; Lovelock, C.E.; Serrano, O.; Duarte, C.M. Blue carbon as a natural climate solution. Nat. Rev. Earth Environ. 2021, 2, 826–839.
  70. Kuwae, T.; Hori, M. (Eds.) Blue Carbon in Shallow Coastal Ecosystems; Springer: Gateway East, Singapore, 2019; ISBN 978-981-13-1294-6.
  71. Lobus, N.V.V.; Udalov, A.A.A. Chemical composition of brown algae Laminaria digitata (Hudson) J.V. Lamouroux, 1813 and Fucus distichus (Linnaeus, 1767) from the bays of the Novaya Zemlya Archipelago (the Kara Sea). Russ. J. Mar. Biol. 2021, 47, 407–412.
  72. Lobus, N.V.; Glushchenko, A.M.; Osadchiev, A.A.; Maltsev, Y.I.; Kapustin, D.A.; Konovalova, O.P.; Kulikovskiy, M.S.; Krylov, I.N.; Drozdova, A.N. Production of Fluorescent Dissolved Organic Matter by Microalgae Strains from the Ob and Yenisei Gulfs (Siberia). Plants 2022, 11, 3361.
  73. Romankevich, E.A.; Vetrov, A.A. Carbon in the World Ocean; GEOS: Moscow, Russia, 2021; ISBN 978-5-89118-835-8.
  74. Romankevich, E. Geochemistry of Organic Matter in the Ocean; Springer: Berlin, Germany, 1984.
  75. Romankevich, E.A.; Vetrov, A.A. Masses of carbon in the Earth’s hydrosphere. Geochem. Int. 2013, 51, 431–455.
  76. Lasareva, E.V.; Parfenova, A.M.; Romankevich, E.A.; Lobus, N.V.; Drozdova, A.N. Organic Matter and Mineral Interactions Modulate Flocculation Across Arctic River Mixing Zones. J. Geophys. Res. Biogeosci. 2019, 124, 1651–1664.
  77. Macreadie, P.I.; Anton, A.; Raven, J.A.; Beaumont, N.; Connolly, R.M.; Friess, D.A.; Kelleway, J.J.; Kennedy, H.; Kuwae, T.; Lavery, P.S.; et al. The future of Blue Carbon science. Nat. Commun. 2019, 10, 3998.
  78. Gabrielyan, D.A.; Sinetova, M.A.; Gabrielyan, A.K.; Bobrovnikova, L.A.; Bedbenov, V.S.; Starikov, A.Y.; Zorina, A.A.; Gabel, B.V.; Los, D.A. Laboratory System for Intensive Cultivation of Microalgae and Cyanobacteria. Russ. J. Plant Physiol. 2023, 70, 20.
  79. Xu, P.; Li, J.; Qian, J.; Wang, B.; Liu, J.; Xu, R.; Chen, P.; Zhou, W. Recent advances in CO2 fixation by microalgae and its potential contribution to carbon neutrality. Chemosphere 2023, 319, 137987.
  80. Chisti, Y. Biodiesel from microalgae. Biotechnol. Adv. 2007, 25, 294–306.
  81. Zhang, S.; Liu, Z. Advances in the biological fixation of carbon dioxide by microalgae. J. Chem. Technol. Biotechnol. 2021, 96, 1475–1495.
  82. Morales, M.; Sánchez, L.; Revah, S. The impact of environmental factors on carbon dioxide fixation by microalgae. FEMS Microbiol. Lett. 2018, 365, fnx262.
  83. Li, G.; Xiao, W.; Yang, T.; Lyu, T. Optimization and Process Effect for Microalgae Carbon Dioxide Fixation Technology Applications Based on Carbon Capture: A Comprehensive Review. J. Carbon Res. 2023, 9, 35.
  84. Bohutskyi, P.; Bouwer, E. Biogas Production from Algae and Cyanobacteria Through Anaerobic Digestion: A Review, Analysis, and Research Needs. In Advanced Biofuels and Bioproducts; Springer: New York, NY, USA, 2013; pp. 873–975.
  85. Rossi, F.; Olguín, E.J.; Diels, L.; De Philippis, R. Microbial fixation of CO2 in water bodies and in drylands to combat climate change, soil loss and desertification. New Biotechnol. 2015, 32, 109–120.
  86. Gabrielyan, D.A.; Gabel, B.V.; Sinetova, M.A.; Gabrielian, A.K.; Markelova, A.G.; Shcherbakova, N.V.; Los, D.A. Optimization of CO2 Supply for the Intensive Cultivation of Chlorella sorokiniana IPPAS C-1 in the Laboratory and Pilot-Scale Flat-Panel Photobioreactors. Life 2022, 12, 1469.
  87. Bux, F.; Chisti, Y. (Eds.) Algae Biotechnology; Green Energy and Technology; Springer International Publishing: Cham, Switzerland, 2016; ISBN 978-3-319-12333-2.
  88. Sinetova, M.A.; Sidorov, R.A.; Starikov, A.Y.; Voronkov, A.S.; Medvedeva, A.S.; Krivova, Z.V.; Pakholkova, M.S.; Bachin, D.V.; Bedbenov, V.S.; Gabrielyan, D.A.; et al. Assessment of the Biotechnological Potential of Cyanobacterial and Microalgal Strains from IPPAS Culture Collection. Appl. Biochem. Microbiol. 2020, 56, 794–808.
  89. Lobus, N.V.; Kulikovskiy, M.S.; Maltsev, Y.I. Multi-Element Composition of Diatom Chaetoceros spp. from Natural Phytoplankton Assemblages of the Russian Arctic Seas. Biology 2021, 10, 1009.
  90. Zhang, L.; Ling, J.; Lin, M. Carbon neutrality: A comprehensive bibliometric analysis. Environ. Sci. Pollut. Res. 2023, 30, 45498–45514.
  91. Tsai, D.D.-W.; Chen, P.H.; Ramaraj, R. The potential of carbon dioxide capture and sequestration with algae. Ecol. Eng. 2017, 98, 17–23.
  92. Posten, C.; Walter, C. (Eds.) Microalgal Biotechnology: Potential and Production; De Gruyter: Berlin, Germany, 2012; ISBN 978-3-11-022501-3.
  93. Chew, K.W.; Yap, J.Y.; Show, P.L.; Suan, N.H.; Juan, J.C.; Ling, T.C.; Lee, D.-J.; Chang, J.-S. Microalgae biorefinery: High value products perspectives. Bioresour. Technol. 2017, 229, 53–62.
  94. Xu, X.; Gu, X.; Wang, Z.; Shatner, W.; Wang, Z. Progress, challenges and solutions of research on photosynthetic carbon sequestration efficiency of microalgae. Renew. Sustain. Energy Rev. 2019, 110, 65–82.
  95. Khoo, K.S.; Chew, K.W.; Yew, G.Y.; Leong, W.H.; Chai, Y.H.; Show, P.L.; Chen, W.-H. Recent advances in downstream processing of microalgae lipid recovery for biofuel production. Bioresour. Technol. 2020, 304, 122996.
  96. Trentacoste, E.M.; Martinez, A.M.; Zenk, T. The place of algae in agriculture: Policies for algal biomass production. Photosynth. Res. 2015, 123, 305–315.
  97. Gabrielyan, D.A.; Sinetova, M.A.; Gabel, B.V.; Gabrielian, A.K.; Markelova, A.G.; Rodionova, M.V.; Bedbenov, V.S.; Shcherbakova, N.V.; Los, D.A. Cultivation of Chlorella sorokiniana IPPAS C-1 in Flat-Panel Photobioreactors: From a Laboratory to a Pilot Scale. Life 2022, 12, 1309.
  98. Ding, G.T.; Mohd Yasin, N.H.; Takriff, M.S.; Kamarudin, K.F.; Salihon, J.; Yaakob, Z.; Mohd Hakimi, N.I.N. Phycoremediation of palm oil mill effluent (POME) and CO2 fixation by locally isolated microalgae: Chlorella sorokiniana UKM2, Coelastrella sp. UKM4 and Chlorella pyrenoidosa UKM7. J. Water Process Eng. 2020, 35, 101202.
  99. Yang, S.; Yang, D.; Shi, W.; Deng, C.; Chen, C.; Feng, S. Global evaluation of carbon neutrality and peak carbon dioxide emissions: Current challenges and future outlook. Environ. Sci. Pollut. Res. 2022, 30, 81725–81744.
  100. Zhao, B.; Su, Y. Process effect of microalgal-carbon dioxide fixation and biomass production: A review. Renew. Sustain. Energy Rev. 2014, 31, 121–132.
More
Video Production Service