Next Article in Journal
Outer Membrane Vesicles Formed by Clinical Proteus mirabilis Strains May Be Incorporated into the Outer Membrane of Other P. mirabilis Cells and Demonstrate Lytic Properties
Next Article in Special Issue
Studies of the Sorption-Desorption of Pesticides from Cellulose-Based Derivative Nanocomposite Hydrogels
Previous Article in Journal
GC-MS and PCA Analysis of Fatty Acid Profile in Various Ilex Species
Previous Article in Special Issue
Composite Hydrogels with Rapid Self-Healing, Stretchable, Moldable and Antibacterial Properties Based on PVA/ε-Poly-l-lysine/Hyaluronic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development of Slow-Release Fertilizers with Function of Water Retention Using Eco-Friendly Starch Hydrogels

1
Institute of Chemistry, Henan Academy of Sciences, Zhengzhou 450002, China
2
High & New Technology Research Center of Henan Academy of Sciences, Zhengzhou 450002, China
3
School of Food Science and Engineering, South China University of Technology, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(20), 4835; https://doi.org/10.3390/molecules29204835
Submission received: 13 September 2024 / Revised: 5 October 2024 / Accepted: 9 October 2024 / Published: 12 October 2024
(This article belongs to the Special Issue Hydrogels: Preparation, Characterization, and Applications)

Abstract

:
Over the past two decades, the development and commercialization of slow-release fertilizers (SRFs) have significantly advanced, with the primary aim of mitigating environmental issues associated with excessive fertilizer use. A range of methodologies, including chemical and physical reactions, incorporation into carriers with porous and layered structures, and coating techniques, have been explored and refined. On the other hand, global challenges such as drought and desertification further underscore the need for SRFs that not only control nutrient release but also improve soil moisture retention. This paper reviews the development and application of eco-friendly starch hydrogels as fertilizer carriers and water retention for SRFs, particularly starch-based superabsorbent polymers (SAPs) produced through grafting copolymerization with acrylamide. This review explores both scientific issues, such as the microstructures and releasing mechanisms of SAPs, and technical development, involving copolymerization technologies, multi-initialization processes, methods of loading fertilizer into hydrogel, etc. Starch, as both a biodegradable and renewable carbohydrate polymer, offers distinct advantages due to its excellent chemical stability and high reactivity. The fabrication techniques of SAPs have been developed from traditional batch polymerization in aqueous solutions to more efficient, solvent-free reactive extrusion. The benefits of SRFs based on SAPs encompass enhanced soil aeration, the prevention of soil deterioration, the minimization of water evaporation, environmental pollution control, reduction in plant mortality, and prolonged nutrient retention within soil. In this review, we summarize the current progress, identify limitations in existing technologies, and propose future research directions to further enhance the performance of starch-based SRFs.

1. Introduction of Slow-Release Fertilizers

Data from the International Fertilizer Association (IFA) reveal a steady increase in global fertilizer consumption, reflecting the increasing reliance on chemical inputs to enhance crop yields and meet the demands of a growing global population. From 2019 to 2023, fertilizer use increased by an average of 1.3% annually [1], underscoring the agricultural sector’s continued dependence on these resources. Nitrogen fertilizers dominate the market due to their critical role in promoting plant growth by providing essential nutrients. However, their efficiency is undermined by significant losses—up to 70% of conventional fertilizers are wasted. This inefficiency has far-reaching environmental implications, contributing to pollution through decomposition, leaching, and ammonium volatilization. These processes not only deplete the quality of water resources but also contaminate soil and air, leading to a range of environmental issues such as eutrophication, soil acidification, and the emission of greenhouse gases [2,3].
To address the problems caused by overusing fertilizers, slow-release fertilizers (SRFs) have emerged as a promising solution. These fertilizers are designed to release nutrients gradually over an extended period, thereby enhancing nutrient uptake by plants and reducing the environmental impact associated with conventional fertilizers [4,5]. SRFs include a variety of products, such as controlled-release fertilizers, enhanced-efficiency fertilizers, and smart-release fertilizers, among others. Despite differences in nomenclatures, all these products should be classified as SRFs, because their primary function is to decelerate the release rate, without the ability to modulate performance based on specific crops or environmental conditions. Evaluation standards from the EU, ISO, and GB (China) also categorize these fertilizers as SRFs. Technically, there are two main strategies SRFs use to supply nutrients to plants: one involves slowing down nutrient availability in soil or reducing their solubility in water [6,7,8].
Various slow-release methodologies, including physical or chemical reactions, loading into carriers, and coating techniques, have been established [9,10,11]. Among these, coating is particularly favored for developing SRFs due to its simplicity, efficiency, and precise control over coating thickness, thereby modulating the release rate [12]. A wide range of materials have been explored as coating agents for SRFs, including petrochemical-derived polymers and biodegradable compounds like synthetic polylactic acid (PLA) and natural materials such as cellulose, lignin, starch, chitosan, and alginate [13,14,15,16]. Given concerns about potential environmental impacts from non-biodegradable materials, systems derived from biodegradable polymers, particularly those from renewable sources, have garnered significant interest. Additionally, to address challenges such as drought and desertification in agriculture, incorporating superabsorbent polymers (SAPs) into polymer-coated SRF products offers benefits such as enhanced water retention in soil and reduced soil pollution [17,18,19]. Recently, the use of eco-friendly hydrogels or SAPs as fertilizer carriers for developing SRFs has attracted growing attention [20,21]. SRFs based on SAPs offer numerous advantages, including improved soil aeration, the mitigation of soil degradation, reduced carbon emissions, and enhanced nutrient retention in soil.
While many reviews have covered SRF development from various perspectives and stages [9,22,23,24,25,26,27,28,29,30,31,32], this review focuses specifically on the use of eco-friendly hydrogels as fertilizer carriers for SRFs, emphasizing their water retention capabilities and unique release mechanisms. It addresses both scientific and technical developments, summarizing current achievements, identifying limitations in existing technologies, and proposing future research directions in this area.

2. Development of Starch-Based Hydrogels

Hydrogels, characterized by their unique capacity to absorb and retain large amounts of water, play a crucial role in various scientific and industrial applications. These materials are highly versatile, offering a range of properties that can be tailored to specific needs. Hydrogels consist of a network of hydrophilic polymers that are crosslinked, allowing them to maintain a solid structure despite their high water content [33]. This crosslinking is critical, as it prevents the polymer chains from dissolving in water, enabling the hydrogel to absorb water without losing its structural integrity. Both synthetic and natural polymers are used in hydrogel composition. Synthetic polymers like polyacrylic acid (PAA), polyvinyl alcohol (PVA), and polyacrylamide (PAM) are commonly used due to their controllable properties and ease of synthesis [34,35,36]. Natural polymers, such as proteins, polysaccharides, and nucleic acids, offer advantages like being bioresorbable and non-toxic, making them ideal for biomedical applications. Polysaccharides, in particular, are favored for their robustness and cost-effectiveness, with starch, alginate, chitosan, and hyaluronic acid being prominent examples. Manuel and Jennifer [37] recently conducted a comprehensive review of hydrogels, encompassing their properties, classification, synthesis mechanisms, and applications across various industries, with an emphasis on starch and cellulose as copolymer components in hydrogel formulation.
Starch, a polysaccharide composed of amylose and amylopectin (Figure 1) [38], has high hydrophilicity, making it an attractive precursor for hydrogel synthesis. Starch-based hydrogels have garnered significant attention in agriculture due to their inherent properties such as biodegradability and eco-friendliness. However, native starch-based hydrogels face limitations, including inadequate water absorption capacity, poor mechanical stability, and limited salt tolerance [39]. To address these challenges, researchers have explored various modification techniques to enhance the properties of starch-based hydrogels [40,41,42]. One of the most effective methods is grafting, in which hydrophilic vinyl monomers such as acrylic acid (AA) and acrylamide (AM) are covalently attached to the starch backbone [43,44,45,46]. This grafting process introduces additional hydrophilic groups into the starch matrix, significantly boosting the hydrogel’s water absorption capacity and mechanical strength. Grafting is typically initiated through free radical polymerization, using chemical initiators like ammonium cerium nitrate, persulfate, and redox systems [47,48]. Crosslinking is also a critical step in hydrogel synthesis, as it determines the structural integrity and swelling behavior of the hydrogel. Various crosslinking methods have been employed, including irradiation with ultraviolet (UV) light, microwave irradiation, or gamma (γ) radiation, and chemical crosslinking agents [49,50].
Starch-based superabsorbent polymers (SAPs), known for their exceptional water absorption capability, are often synthesized via graft polymerization onto starch substrates. These SAPs can absorb several hundred to thousands of times their weight in water, making them ideal for applications such as agriculture, where they help retain water in soil and improve plant growth conditions. Additionally, SAPs can encapsulate nutrients through physicochemical interactions, allowing for controlled nutrient release, which is beneficial in slow-release fertilizer (SRF) formulations. Recent studies by Supare and Mhanwar [49] have extensively reviewed the role of SAPs in agriculture, focusing on starch-derived SAPs, crosslinking, and graft copolymerization methods used in their synthesis. This chapter delves into the methods of synthesizing starch-based SAPs under various process conditions.

2.1. Grafting Copolymerization with Starches

Starch-grafted PAM and PAA are among the most effective superabsorbent materials due to their superior water absorption capabilities. These materials are synthesized through graft polymerization, which typically takes place in an aqueous medium using batch processing techniques. This process enables the production of starch-based SAPs that are highly efficient in retaining water and can be used in various fields, ranging from agriculture to personal care products [51,52,53]. A schematic diagram of the process for preparing a hydrogel through starch graft polymerization in aqueous solution is shown in Figure 2. The starch is first gelatinized at a certain temperature, which destroys the crystalline structure of the starch and promotes its uniform dissolution. This step is essential, as it enhances the accessibility of starch molecules to the chemical reactions involved in grafting. Subsequently, a chemical initiator, such as a redox system, is then introduced to generate free radicals required to initiate the grafting reaction. Monomers and crosslinkers are subsequently added to the solution, and the grafting polymerization process proceeds until the desired degree of polymerization is achieved. The resulting starch-based SAPs are isolated from the reaction mixture through filtration or centrifugation, followed by washing and drying to remove any residual monomers, initiators, and by-products. The final product is a highly absorbent material capable of retaining large amounts of water, making it suitable for a wide range of applications [54,55].
Mahmoodi-Babolan et al. [56] synthesized starch-based SAPs via solution polymerization by incorporating AA and AM. The material’s performance was improved by grafting catecholamine functional groups onto its pore surfaces through dopamine oxidation polymerization. The resultant bimodal mesoporous adsorbent demonstrated a swelling ratio of 5914.66%. SAPs have also shown high efficacy as adsorbents for pollutant removal, particularly dyes, with a maximum adsorption capacity of 2276 mg/g for methylene blue.
Various starches from different resources with different microstructures have been evaluated to be used as raw materials. Zhang et al. [57] developed three types of rice starch-based superabsorbent materials (SPAMs). The findings indicated that there was a minimal occurrence of side reactions, resulting in an exceptionally high acrylic grafting efficiency of 93–94%. Substantial amylose content enhances polymer chain interactions and structural integrity through hydrogen bonding. The structured rigidity of amylopectin’s branched chains in glutinous rice starch contributes to the SPAM’s favorable water absorption properties. Siyamak et al. [58] reported on the grafting of acrylamide monomers through free radical copolymerization onto four distinct starches as potential ammonium sorbents. The findings indicated that there was no significant difference in the average monomer conversion and graft efficiency among the different types of starch. Zou et al. [59] performed an extensive investigation into the impact of the amylose–amylopectin ratio in corn-derived starches on the grafting reactions and the resultant properties of starch-based SAPs. Their results demonstrated that, under consistent reaction conditions, both the grafting ratio and efficiency increased with higher amylose content, correlating with the water absorption capacity. The high molecular weight and branched structure of amylopectin reduce polymer chain mobility, thereby increasing viscosity and potentially affecting the grafting reactions and performance attributes of starch-based SAPs. Bao et al. [60,61] investigated the relationship between molecular structures, viscoelastic properties, graft polymerization, and hydrogel microstructures using cornstarch models with varying amylose–amylopectin ratios. They found that an increase in amylose content resulted in a higher viscoelastic modulus in starch melts, which was associated with a reduced degree of micro-mixing at lower rheological dynamic rates. Although monomer conversion remained nearly constant with increasing amylose content, the grafting efficiency decreased. This decrease was attributed to the higher tendency of high-amylose starch to form crosslinks with grafted components.
In order to enhance the performance of hydrogels, particularly in terms of gel strength and water absorption capacity, researchers have been exploring the synthesis and application of composite hydrogels. One of the primary strategies in enhancing gel strength involves the introduction of reinforcing agents into the hydrogel matrix. These agents can include various types of fibers, nanoparticles, and other polymers that can increase the structural integrity of the hydrogel [62,63,64]. Li et al. [65] synthesized a starch-g-PAA–attapulgite superabsorbent composite by incorporating attapulgite micropowder into an aqueous reaction medium. The optimal synthesis conditions, which included 10 w% attapulgite, resulted in a composite with a water absorption capacity of 1077 g H2O/g sample in distilled water and 61 g H2O/g sample in a 0.9 w% NaCl solution. This biodegradable, cost-effective, and environmentally friendly superabsorbent composite demonstrates superior water absorbency and water retention capabilities under load, making it particularly suitable for agricultural and horticultural applications.

2.2. Radiation In Situ Polymerization

Radiation technology has been extensively employed in the modification of starch, with numerous radiation techniques being utilized to produce starch-based hydrogels [66,67,68,69]. Radiations are typically classified into ionizing and nonionizing categories based on their ability to ionize materials. Ionizing radiation has the capability to ionize the medium it passes through, either through direct or indirect interaction with the material. In contrast, nonionizing radiation, due to its low ionizing potential, is incapable of ionizing materials [70]. Among ionizing radiation techniques, gamma radiation and electron beams are extensively utilized for hydrogel synthesis. Nonionizing radiation has been explored for the microwave-assisted synthesis of hydrogels [71]. As depicted in Figure 3, the radiation-induced ionization process of starch generally involves two stages. First, covalent bonds are disrupted, leading to the formation of free radicals. These radicals then initiate chemical reactions among molecules at varying concentrations [72].
Gamma irradiation has proven to be particularly effective in forming three-dimensional polymeric networks in superabsorbent materials, showcasing high efficiency [73]. This method offers several advantages over conventional hydrogel fabrication techniques. Traditional techniques often require toxic initiators and crosslinking agents, which pose risks to both human health and the environment [74]. In contrast, gamma irradiation eliminates the need for these chemicals, making the process safer and more environmentally friendly. Additionally, it allows for reactions under milder conditions, reducing energy consumption and operational costs [75]. The formation of minimal by-products further enhances the sustainability of the process by reducing waste generation and simplifies the purification steps. The irradiation process is highly controllable, allowing for precise adjustments in radiation dose to achieve the desired degree of polymerization and crosslinking [76,77]. This flexibility enables the production of starch hydrogels with tailored properties, such as absorption capacity, mechanical strength, and biodegradability [78]. Similar to gamma ray irradiation, electron beam crosslinking represents a clean and safe technique that requires no external initiators or crosslinking agents. Electron beam irradiation surpasses gamma irradiation techniques in hydrogel production, primarily due to its faster processing speed and precise beam control [79].
Starch-based hydrogels were synthesized at ambient temperature using gamma radiation [80,81,82], as shown in Figure 4. This study reveals that, within a certain range, increasing radiation dosage leads to a higher gel fraction. Higher doses generate more persistent free radicals, which enhances the probability of homopolymerization and the formation of an entangled network. As the radiation dose escalates, so does the concentration of free radicals, leading to a higher degree of conversion and crosslinking. However, at higher doses, intramolecular free radical recombination becomes more dominant than intermolecular recombination, particularly at sufficiently high absorbed doses, which can inhibit further crosslinking and initiate the degradation of the polymer network. When superabsorbent polymers (SAPs) were synthesized via the gamma irradiation of aqueous mixtures containing carboxymethylcellulose (CMC) and starch, the incorporation of starch significantly increased both the gel fraction and water absorption capacity at relatively low irradiation doses of 20 kGy. However, excessive starch content hindered gel formation, reducing the gel fraction. Starch is inherently prone to radiation-induced degradation, but in the presence of CMC, radicals formed on starch molecules interact with those on the CMC chains, promoting crosslinking rather than degradation [83].

2.3. Application of Reactive Extrusion Technology

Traditional methods of producing starch-based SAPs often rely on batch processes that are solvent-intensive and inefficient in terms of energy consumption and waste generation. The shift towards reactive extrusion (REX) represents a significant advancement, offering a more sustainable and efficient approach to SAP manufacturing [84,85,86,87,88]. As shown in Figure 5, REX is a continuous process that uses extruders as chemical reactors to facilitate the grafting and polymerization of starch-based materials [58]. This method is particularly advantageous for handling high-viscosity polymers, which are often difficult to process using conventional batch techniques [89]. By reducing solvent use, REX minimizes environmental impact while improving the safety and efficiency of the production process.
One of the primary advantages of REX is its operational flexibility. Various processing parameters, such as temperature and pressure, can be precisely controlled, allowing manufacturers to tailor the polymer’s properties to specific requirements [90,91,92,93,94]. Additionally, REX offers precise control over the residence time of reactants within the extruder, optimizing reaction kinetics and enhancing product performance. The continuous nature of REX leads to higher throughput rates compared to batch processes, significantly reducing production time and associated costs. This increased efficiency also enables manufacturers to respond more quickly to market demands [92]. Furthermore, the reduced need for solvents and the associated decrease in waste treatment contribute to a more sustainable manufacturing approach. Siyamak et al. [95] demonstrated the use of REX to synthesize starch-g-PAM copolymers, utilizing a screw configuration with various elements. The solvent-free graft copolymerization of starch was performed at high solid concentrations, with the reaction reaching completion within 5 min. This process resulted in an average monomer conversion rate of 80% and a grafting efficiency of approximately 74%. Another key advantage of REX is its modular design, which allows for easy modifications and optimization of the system for different types of starch-based materials and desired product characteristics [96]. This adaptability is crucial for ongoing research and development, where continuous improvements are being made to enhance SAP performance and expand their applications.
Jiang et al. [97] developed a REX system (illustrated in Figure 6) for the production of starch-g-PAM hydrogels, utilizing a dual-initiation mechanism. The REX system, equipped with a twin-screw co-extruder, a solid feeder, and four injection ports, facilitated the separate addition of an acrylamide solution, initiator-1, initiator-2, and a saponifying agent. The dual-initiation approach was found to enhance the homogeneity of the hydrogel’s network microstructure and augment its gel strength.

2.4. Characterization Techniques for Hydrogel Network Structure

Rheological Characterization of Hydrogels. Rheology is the science of studying the flow and deformation behavior of materials, aiming to investigate the response of materials under external forces [98,99]. It covers both the hydrodynamic theory of fluids and the elasticity theory of elastomers. For simple Newtonian fluids and elastomers, their rheological properties can be described by simple continuity equations such as Hooke’s law and Newton’s law of viscosity. However, the rheological properties of complex viscoelastic materials like hydrogels are much more intricate [100]. Hydrogels exhibit both viscous and elastic properties, and their behavior is influenced by stress, strain, strain rate, and even deformation history, displaying notable nonlinear characteristics [101,102]. As a typical complex viscoelastic material, the rheological characterization of starch-based hydrogels is crucial for understanding the relationship between their microstructure and macroscopic mechanical properties. Dynamic shear testing using a rotational rheometer, known for its operational simplicity and clear demonstration of viscoelastic characteristics, is the preferred technique for evaluating the viscoelastic behavior of starch-based hydrogels. This method applies periodic sinusoidal displacements and torque, measuring the response torque or strain and its phase difference from the applied action, effectively revealing the material’s viscoelastic characteristics, including storage modulus (G′), loss modulus (G″), and complex viscosity (η*). These rheological parameters not only help in understanding the interactions between starch molecular chains and the influence of crosslinking density on material performance but also guide the optimization of formulations and processing conditions to ensure product quality and meet specific application needs. (1) Strain Sweep: The linear viscoelastic region (LVR) of hydrogels can be identified through strain sweeps at constant frequency. In the low strain range, viscoelastic moduli remain independent of strain but exceed the critical strain, and the gel structure begins to break down. Chemically crosslinked gels typically have larger critical strain values than physical gels, with the LVR depending on chain interactions and crosslinking density. (2) Frequency Sweep: Frequency sweep is essential for probing the molecular structure of gel networks. By analyzing viscoelastic moduli versus frequency in a log–log plot, the relaxation behavior of gel chains is revealed. Physical crosslinking systems formed by hydrogen bonds, double helices, and ionic bonds show varied gel strengths and relaxation spectra. In highly crosslinked gels, the storage modulus (G′) is significantly higher than the loss modulus (G″) over the entire frequency range, indicating elastic behavior. (3) Time Sweep: Time sweep at constant amplitude and frequency monitors the real-time changes in the hydrogel network. It can reveal molecular structure changes and the kinetics of gel formation, with viscoelastic moduli increasing significantly during the process. (4) Temperature Sweep: Temperature sweeps involve heating or cooling gels while keeping strain and frequency constant. The intersection of G′ and G″ indicates the sol–gel transition temperature, providing insights into gel thermal stability and crosslinking strength.
Dynamic mechanical thermal analysis (DMTA) of hydrogels. DMTA is a key technique used to study the mechanical properties of materials under dynamic stress as a function of temperature or frequency [103,104]. By applying a periodic sinusoidal load to polymer-based composites and precisely recording their response, it is possible to observe a phase lag between the deformation and the applied load due to the viscoelastic nature of the material. As the temperature increases, molecular segments within the material are activated, leading to significant changes in the stiffness of the matrix. These changes can be reflected through parameters such as the storage modulus (E′), loss modulus (E″), and the loss factor (tan δ = E″/E′), which provide valuable insights for understanding and predicting the material’s performance in practical applications, particularly in assessing its viscoelasticity and energy dissipation capabilities [105].
Microstructural parameters, such as molecular weight, molecular orientation, crystallinity, the degree of crosslinking, and the nature of copolymers, significantly influence the macroscopic mechanical properties of materials. DMTA determines the glass transition temperature (Tg) and secondary relaxation transitions by analyzing how E′, E″, and tan δ change with temperature. These transitions reflect changes in the material’s internal structure. For example, the glass transition temperature is the point at which a polymer transitions from a glassy state to a rubbery state, associated with the increased mobility of molecular chains. This transition directly illustrates the relationship between the material’s microstructure and its macroscopic mechanical properties. In reality, the glass transition does not occur at a single, specific temperature but over a range of temperatures where the material’s characteristics change. Various methods are used to determine Tg, including the peak of the tan δ curve, the peak of the E″ curve, and the inflection point of the E’ curve. Each method has its advantages and disadvantages, and they may yield slightly different results. Typically, the temperature corresponding to the peak of the tan δ curve is taken as the Tg of the material.

3. Loading Fertilizer into Hydrogels

3.1. Coating by Starch

Among the various methods for creating SRFs, the coated and matrix types have garnered considerable attention [106]. Coated-type SRFs, in particular, have been extensively researched for their capacity to modulate nutrient release through the manipulation of coating materials [11,107,108]. These SRFs are typically prepared by applying a layer of inert material onto solid fertilizers to reduce their dissolution rate. The effectiveness of this approach heavily relies on the intrinsic physical attributes of the coating material, including its hydrophobic/hydrophilic nature and porosity. Starch granules have emerged as a promising coating material due to their ability to enhance the adhesion on fertilizer surfaces. For instance, Khan et al. [109] demonstrated a process where starch was incorporated into NPK fertilizers as a binder, ensuring the uniform coating of the NPK granules with starch particles. However, native starch has limitations such as poor film-forming ability, low water resistance, and susceptibility to damage in humid environments. To overcome these challenges, starch is often chemically modified through processes like esterification, crosslinking, and graft copolymerization to enhance its water resistance [110,111,112,113]. Modified starch as a coating material has been extensively reviewed [38], and this chapter specifically delves into the novel application of starch-based hydrogel coatings in SRFs.
Qiao et al. [114] introduced an innovative double-coated SRF design, utilizing ethyl cellulose (EC) as the inner layer and a starch-based SAP as the outer layer, with urea as the core, accounting for 55% w/w (Figure 7). The release of nutrients from the double-coated SRF occurs in three distinct phases: (1) the absorption of water by the starch-based SAP and its subsequent permeation through the EC layer, (2) the dissolution of the urea core’s nutrients by water, and (3) nutrient delivery into soil through the dual-layered system. The release dynamics of the fertilizer are primarily influenced by the properties of the starch–SAP layer. This double-coating technique significantly augments the controlled release of nutrients and improves the overall efficiency and sustainability of fertilizer application. Similarly, Lü et al. [115] developed another dual-coated SRF that incorporated starch acetate and a combination of carboxymethyl starch and xanthan gum. By carefully selecting and modifying coating materials, researchers cancustomize the release profiles of SRFs to meet specific agricultural requirements,, ultimately leading to more effective and environmentally friendly farming practices.
Polyvinyl alcohol (PVA), a biodegradable polymer, can form a strong and environmentally friendly coating matrix when combined with starch. The enhancement in interactions and crosslinking within PVA–starch blends not only augments the strength and density of the resulting films but also impedes the swelling of the starch structure by reducing accessible regions, thereby increasing their resistance to dissolution. Zafar et al. [111] used starch and PVA coatings, acrylic acid, citric acid, and maleic acid as crosslinking agents and prepared a new type of coated urea SRF through a granulator/fluidized bed coater. The incorporation of clay nanoparticles, such as bentonite, further augments the mechanical properties of the coating film, making it more durable and effective in controlling the release of nutrients. Sarkar et al. [116] demonstrated a novel approach to encapsulate diammonium phosphate (DAP) using a blend of wheat starch, PVA, and bentonite clay, aiming to optimize the release kinetics and improve the overall performance of the fertilizer. The resulting coated DAP particles exhibited controlled release characteristics, with the release rate influenced by the concentration of bentonite clay. Higher concentrations of bentonite resulted in slower release rates, as the clay nanoparticles increased the film’s barrier properties and reduced permeability. This controlled release mechanism is beneficial for optimizing nutrient availability for plants and minimizing environmental impact by reducing nutrient leaching. This approach holds promise for the development of sustainable and efficient slow-release fertilizers, contributing to improved agricultural practices and environmental sustainability.

3.2. Swelling and Absorption Equilibrium Method

When hydrogels are immersed in a fertilizer solution, they absorb nutrients and swell as they take up liquid. This process continues until the hydrogels reach an equilibrium state, where they have absorbed as much fertilizer solution as they can hold. These nutrient-loaded hydrogels can then be applied to soil, where they gradually release the absorbed fertilizers over time as the liquid is released (Figure 8). This method allows for the controlled and sustained release of fertilizers, improving nutrient uptake by plants and reducing the risk of nutrient leaching into the environment.
Leόn et al. [117] explored the graft copolymerization of starch with itaconic acid to produce SAPs. The findings indicated that increasing the fertilizer concentration from 0.5 g/L to 10 g/L enhanced urea absorption but reduced the adsorption of KNO3 and NH4NO3. The release behavior of fertilizers also varied, with higher fertilizer concentrations leading to slower release rates. Similarly, Perez et al. [118] studied macrospheres fabricated from chitosan and chitosan–starch blends. The dry polymer matrices were immersed in a fertilizer-enriched solution for 4 h at ambient temperature, then dried at 40 °C for 48 h. Analyses revealed that the bead structure significantly influenced swelling behavior, which in turn affected their fertilizer loading capacity and release kinetics. Although the swelling and absorption equilibrium method is simple, it has certain limitations. Drying the loaded hydrogels requires substantial energy, and the fertilizer loading capacity is restricted by the hydrogel’s swelling capability. Furthermore, the swelling and absorption capacity of all SAPs is significantly reduced in ionic solutions, such as those found in various fertilizers. As a result, the concentration of fertilizer within SAPs is typically modest.

3.3. In Situ Polymerization Technology

The in situ polymerization of starch is a method of directly initiating the polymerization of monomers in the presence of starch, which enables the formation of polymers on the surface or inside starch granules, thus improving the physicochemical properties of starch or giving it new functions [119,120]. Salimi et al. [121] fabricated a novel slow-release urea fertilizer using an in situ polymerization process. This method integrated acrylic monomers with starch in the presence of urea, achieving a homogeneous blend. The innovation was further enhanced by incorporating natural char nanoparticles as nano-fillers, which uniformly dispersed within the polymer matrix. This uniform distribution significantly improved the interfacial interactions between the polymer and the fillers, effectively retarding nitrogen diffusion and reducing the release rate in both aqueous and soil conditions.
Chen et al. [82] developed an approach for synthesizing SRFs through the in situ radiation-induced polymerization of monolithic hydrogels. This method involves embedding urea within a starch-based matrix and grafting it with polyacrylamide. The analysis demonstrated distinct microstructures in hydrogels prepared under different conditions (Figure 9). Low irradiation intensities preserved some honeycomb structures with weaker fibers, but higher intensities resulted in the disappearance of these structures, replaced by larger holes and a denser, more homogeneous crosslinked network. Higher concentrations led to porous structures, whereas lower concentrations produced less dense networks and thinner cell walls. The presence of porous or cottony structures was primarily attributed to the grafted starch matrix, emphasizing the significant influence of irradiation intensity, concentration, and the AM–starch ratio on the hydrogel microstructure. An increase in radiation intensity and concentration enhanced grafting efficiency and monomer conversion. Enhanced gel strength, associated with higher radiation levels, AM content, and concentration, was correlated with a reduced urea release rate. For example, the release rates of samples U-1~U-7 were only about 26.2%, 14.9%, 20.2%, 11.6%, 33.5%, 9.5%, and 30.2% in 2 h, respectively. After 42 h, the release rate of urea remained above 80%, and the difference was not significant.

3.4. Reactive Extrusion Technology

REX is a continuous process that involves the mixing and chemical modification of polymers in the molten state within an extruder. The extruder is a long, cylindrical machine equipped with a screw that rotates and conveys the polymer forward while mixing it with other components [122]. The REX process for loading fertilizer into hydrogels typically involves the following steps: (1) Preparation of hydrogel precursors. The first step is the preparation of hydrogel precursors, which are usually hydrophilic polymers such as PVA, PAA, PAM, starch, cellulose, etc. (2) Incorporation of fertilizer. The fertilizer, which can be in the form of solid particles or a liquid solution, is introduced into the extruder along with the hydrogel precursors. The extruder is designed to provide a controlled environment for the mixing and encapsulation of the fertilizer within the polymer matrix. (3) Crosslinking reaction. As the mixture moves along the extruder, a crosslinking reaction is initiated, typically through the addition of a crosslinking agent or by using heat or radiation. This reaction leads to the formation of the hydrogel network, with the fertilizer particles or molecules being trapped within the polymer matrix. (4) Formation of hydrogel beads. The hydrogel–fertilizer mixture is then extruded through a die, typically in the form of small beads or pellets. These beads are rapidly cooled to solidify the hydrogel structure and stabilize the encapsulated fertilizer.
REX research encompasses a broad spectrum of disciplines, such as chemical reaction engineering, rheology, polymer processing, and mechanics, involving numerous intricate reaction processes. Given the complexity of these processes, the utilization of reactive extrusion in the production of starch-based SRFs remains relatively constrained. A modified Hakke internal mixer was employed as a reactor to synthesize SAPs, mimicking the operation of a twin-screw extruder [60,61,123,124,125]; Under a high-viscosity system, the fertilizer loading rate can be increased to over 37 w% [126]. The REX process, while efficient for many polymer processing applications, presents a unique challenge when it comes to the preservation of the crosslinked structure of hydrogels. The crosslinking in hydrogels is a critical factor that determines their mechanical strength, swelling behavior, and degradation rate. During the reaction extrusion process, which involves the extrusion of a polymer melt through a die, the high shear forces and thermal conditions can potentially disrupt the delicate crosslinked structure of the hydrogel. This disruption can lead to a decrease in the mechanical integrity of the hydrogel, affecting its performance in practical applications. To circumvent the potential degradation of the crosslinked structure during the reaction extrusion process, an alternative approach known as post-crosslinking can be employed. Post-crosslinking involves the formation of crosslinks in a polymer matrix after the initial processing steps. This method allows for the preparation of hydrogels with controlled crosslinking density and distribution, which is crucial for optimizing the properties of the hydrogel for specific applications.

3.5. Compounding with Cellulose

Cellulose, a polymeric compound abundantly found in plant cell walls, is the most plentiful and sustainably sourced material on Earth. It possesses a rich variety of functional groups, exceptional mechanical properties, and a high degree of chemical adaptability, making it ideal for modification [127,128]. Furthermore, the capability to render cellulose into fibrous forms at the micro- or nanometer scale has positioned it as an increasingly promising contender for the creation of hydrogels in recent years. Li et al. [129] reviewed various modification methods of cellulose and the application of stimuli-responsive hydrogels in slow-release fertilizers. Manuel et al. [37] conducted a review on hydrogels, delving into their properties, classifications, synthesis mechanisms, and diverse applications across various industries. In their article, starch and cellulose were highlighted as copolymers, showcasing their integral roles in the development and functionality of hydrogels.
Cellulose-based composites have garnered significant attention in recent years due to their potential as reinforcing fillers in biopolymer matrices derived from starch [64,130,131,132] (Figure 10). Starch composite cellulose hydrogels are both biodegradable and renewable polymers, and these hydrogels can slowly release nutrients to plants, providing a controlled and sustained supply of essential elements for plant growth. Bora et al. [133] utilized wastepaper powder as a modifier in a biodegradable hydrogel composite composed of starch, itaconic acid, and acrylic acid. The incorporation of an optimal quantity of the modifier increased the hydrogel’s swelling capacity from 503 g/g to 647 g/g. The NPK-laden hydrogel demonstrated effective sustained-release characteristics, with 98% of nitrogen, 81% of phosphorus, and 95% of potassium being released over 20 days.

3.6. Other Hydrogel Systems

Interpenetrating polymer network (IPN) hydrogels based on starch represent a promising class of materials that combine the biodegradability and biocompatibility of starch with the mechanical properties and functional versatility of synthetic polymers. These hydrogels are formed by the simultaneous or sequential polymerization of starch with one or more synthetic polymers, resulting in a network structure where both components are interwoven. This interpenetration enhances the overall performance of the hydrogel, rendering it appropriate for SRFs. Vudjung et al. explored the synthesis of crosslinked natural rubber and cassava starch hydrogels through the IPN method [134]. This approach aimed to enhance mechanical strength and compatibility by integrating multiple network polymers. These hydrogels demonstrated high water swelling capabilities and favorable biodegradation properties. However, the rigid nature of the final product precluded its application as a coating on urea bead surfaces. To mitigate this issue, they sought to innovate it by developing a novel coating membrane from pre-vulcanized natural rubber (NR) and starch (St), employing the IPN method with sulfur and glutaraldehyde as crosslinking agents [135]. The release mechanism of urea in both aqueous and soil environments was characterized as non-Fickian diffusion, suggesting transport through a porous matrix. The efficacy of encapsulated urea beads in corn and basil cultivation was significantly superior to that of native urea beads.
Apart from cellulose, starch can also form complexes with other polysaccharides to construct an SRF system. Pimsen et al. [136] focused on the creation of SRFs utilizing a nano-zeolite (NZ) composite incorporated into chitosan (CS)–sago starch (ST)-based biopolymers. The biopolymer nanocomposite was synthesized through an ionotropic gelation method, employing sodium tripolyphosphate as the crosslinking agent. The swelling capacity of the biopolymer nanocomposite notably increased with higher molecular weights of CS and increased crosslinking durations. The NZ-CS/ST nanocomposite exhibited a release of 64.00% phosphorus and 41.93% urea by the 14th day. Majeed et al. [137] investigated the impact of lignin concentration on the release and biodegradability of urea-modified tapioca starch-based enhanced-efficiency fertilizers. They found that incorporating lignin at levels of 5%, 10%, 15%, and 20% into the starch matrix decreased nutrient release in moist soil. Furthermore, an increase in lignin content was associated with enhanced biodegradability.
The integration of inorganic fillers into starch-based hydrogel SRFs represents a significant advancement in agricultural technology, offering multifaceted benefits that enhance the overall efficacy and sustainability of fertilizer applications. This integration is particularly crucial in modern agriculture, where the need for efficient, environmentally friendly, and cost-effective fertilization methods is paramount. These fillers serve as potent reinforcing agents, significantly bolstering the hydrogel’s resilience against mechanical stress, including wear and compression [138]. Furthermore, they enhance the hydrogel’s capacity to retain water, ensuring a robust performance in various applications. Lu et al. [139] studied a novel slow-release fertilizer characterized by high water retention, designated as HS-BCF, which was synthesized by integrating hydrotalcite and starch into biochar-based compound fertilizers (BCFs). The findings indicate that the addition of hydrotalcite and starch to BCFs enhanced the soil water retention capacity by 5–10%. Over 30 days, the cumulative leaching amounts of nitrogen, phosphorus, and potassium from HS-BCF in soil were, at most, 49.4%, 13.3%, and 87.4% of those from BCFs, respectively. Furthermore, hydrotalcite was found to bind with phosphorus in HS-BCF, thereby improving the longevity of phosphorus in the fertilizer. Wei et al. [140] reported on a novel slow-release and water retention fertilizer through the free radical copolymerization of potato starch, acrylic acid, acrylamide, and maleic anhydride-modified β-cyclodextrin. This synthesis process was enhanced by incorporating acid-treated halloysite nanotubes, which were engineered to expand their internal cavities for optimal urea pre-loading. The addition of these halloysite nanotubes improved the fertilizer’s release profile, effectively regulating the cumulative release rate of urea from the slow-release formulation.
Starch-based hydrogels SRFs have been extensively studied, with numerous research findings presented in Table 1.

4. Release Mechanisms and Kinetics

4.1. Release Mechanisms

The release of nutrients from SRF hydrogels is a complex, multi-stage process involving various mechanisms. A comprehensive understanding of these mechanisms is essential for optimizing the design and application of hydrogels in agriculture. Nutrient release typically follows three distinct phases, the lag phase, the constant release phase, and the decay period, each contributing uniquely to the overall release dynamics [150]. During the lag phase, the hydrogel absorbs water from the surrounding environment, initiating the swelling process. This initial stage is characterized by minimal nutrient release, as the hydrogel primarily undergoes hydration and expansion. Following this is the constant release stage, during which the hydrogel reaches an equilibrium state, and nutrients are steadily released. This stage is driven by the diffusion of nutrients through the polymer matrix and the continued hydration and expansion of polymer chains. The decay phase occurs when the release rate of nutrients begins to decline, typically due to the depletion of available nutrients within the hydrogel or changes in environmental conditions.
Release mechanisms involve multiple processes, including diffusion through the polymer and the hydration, expansion, and dissolution of polymer chains [151,152,153]. Diffusion through the polymer matrix is critical, enabling nutrients to diffuse from higher to lower concentration regions. This process is influenced by the polymer network, nutrient molecule size, and the crosslinking degree. The hydration and expansion of polymer chains also play an essential role. As the hydrogel absorbs water, the polymer swells, creating additional space within the polymer matrix, which facilitates nutrient movement and release into the environment. The degree of swelling depends on factors like polymer hydrophilicity, crosslinking, and environmental conditions such as ionic strength and pH. Polymers with higher hydrophilicity and lower crosslinking density tend to exhibit greater swelling, enhancing nutrient release.
Various kinetic models have been developed to understand how swelling impacts nutrient release [154,155]. Swelling behavior occurs in two stages: an initial rapid swelling phase, followed by a gradual and slower swelling phase. In the initial phase, water rapidly penetrates the hydrogel’s porous matrix, driven by the polymer’s hydrophilic nature. In the gradual swelling phase, water penetrates the pores induced by the relaxation of polymer chains. This process is slower and more protracted, reflecting the gradual uptake of water and the continued expansion of the polymer matrix. Polymers with higher hydrophilicity and lower crosslinking tend to exhibit more rapid swelling, while higher crosslinking density and lower hydrophilicity result in slower, more gradual swelling.
Environmental factors, such as temperature, pH, and ionic strength, significantly affect swelling and, consequently, nutrient release. For instance, higher temperatures enhance the diffusion of water into the hydrogel matrix, speeding up swelling and nutrient delivery. Changes in pH can affect the ionization of functional groups, altering the swelling behavior of the hydrogel. Ionic strength can affect osmotic pressure, which drives water absorption [156,157]. Dudu et al. [144] synthesized a novel superabsorbent based on N,N-dimethylacrylamide, maleic acid (MA), and starch (St), designated as DMSt1, and evaluated its performance as a controlled-release fertilizer in lettuce cultivation. They subjected DMSt1 to surface modifications using HCl and NaOH, resulting in negatively and positively charged variants, DMSt2 and DMSt3, respectively. The latter exhibited the highest maximum swelling rate of 37.38% in deionized water. The acid and base treatments altered the anionic and cationic properties of DMSt1, leading to distinct swelling behaviors at varying pH levels. The hydrogels derived from DMSt1, DMSt2, and DMSt3 demonstrated remarkable water uptake capacities, with equilibrium swelling values of 7163.7% at pH 10, 15708% at pH 10, and 27838% at pH 8, respectively. de Lima et al. [158] developed a hydrogel through the copolymerization of cassava gum and polyacrylamide. The hydrogel exhibited the maximum swelling at pH 7, absorbing approximately 270 times its initial mass. At pHs 4 and 2, the swelling capacity decreased to 150 and 250 times, respectively. This reduction in swelling at lower pH values is attributed to the protonation of carboxylate anions, which diminishes repulsive forces and impedes swelling. In more alkaline environments, the swelling percentage also decreased, reaching 100 at pH 10. This phenomenon is explained by the “charge shielding effect”, which weakens electrostatic interactions, as excess Na+ ions hinder the formation of simple COO bonds, a consequence of the hydrolysis process. Al Rohily et al. [159] investigated the swelling properties of fertilizer materials derived from phosphorylated alginate grafted with polyacrylamide (PAM). They found that the degree of crosslinking and the presence of pendant ionic groups within the polymer matrix significantly influence the swelling behavior. In these hydrogels, ionic charges play a key role, as the repulsion between adjacent fixed charges causes network expansion and subsequent swelling. Shang et al. [154] used the Korsmeyer–Peppas and pseudo-second-order kinetic models to analyze the swelling characteristics of temperature-responsive hydrogels, concluding that water transport is primarily governed by Fickian diffusion and polymer chain relaxation.
Overall, the release of nutrients from SRF hydrogels involves a combination of diffusion, swelling and deswelling, chemical degradation, mechanical stress, ion exchange, and temperature effects. These mechanisms work synergistically to provide a controlled and sustained release of nutrients to plants, enhancing nutrient use efficiency and promoting plant growth. A clear understanding of these mechanisms and their interaction with environmental factors is essential for designing effective hydrogel-based fertilizers.

4.2. Kinetic Models

The development of release kinetic models for fertilizers plays a critical role in agricultural science, particularly optimizing nutrient delivery to plants. These models are essential for understanding and predicting the behavior of fertilizers under various environmental conditions, aiding in maximizing crop yields while minimizing environmental impact. Quantitative analysis is vital for developing these models because nutrient release is dynamic—fertilizers do not release nutrients at a constant rate. Instead, the release rate can vary significantly depending on factors such as temperature, moisture, and soil biological activity. Quantitative analysis allows for the measurement of these variables over time, providing data that can be used to construct mathematical models that accurately predict nutrient release patterns. For instance, the type of fertilizer, whether a conventional, slow-release, or controlled-release formulation, significantly impacts the release kinetics. Given the variability in nutrient release mechanisms among various types of SRFs and the intricate interplay of factors such as composition, soil moisture, and temperature, the release mechanisms of SRFs cannot be simply characterized. Typically, four primary mathematical models are employed to elucidate these mechanisms: the zero-order, first-order, Korsmeyer–Peppas, and Higuchi kinetic models can be represented by the following equations (Equations (1)–(4), respectively):
Zero-order model: Mt/M = k0t
First-order model: Mt/M = 1 − e(−k1t)
Korsmeyer–Peppas model: Mt/M = ktn
Higuchi model: Mt/M = kHt1/2
Specifically, the parameter Mt/M denotes the percentage of fertilizer released from the hydrogel at a specific time t. The constants k0, k1, kH, and k represent the release constants for the zero-order, first-order, Higuchi, and Korsmeyer–Peppas models, respectively. The release mechanisms are classified according to the diffusion index (n) as follows: (1) Fickian diffusion mechanism: n < 0.43; (2) non-Fickian diffusion mechanism: 0.43 < n < 0.85; and (3) Case-II transport mechanism: n > 0.85 [129].
Indeed, the slow-release model of the SRF hydrogel exhibits variation across different stages. Shaghaleh et al. [160] developed an aminated cellulose nanofiber (A-CNF) fertilizer hydrogel by encapsulating ammonium nitrate (AN) within the A-CNF matrix. During the initial phase, the hydrogel demonstrated rapid AN release within the first 72 h, conforming to a first-order kinetic model (k1 = 0.068–0.0575). This phase involved the swelling of A-CNF when exposed to buffer or soil media, where pH levels influenced AN diffusion. In the subsequent phase, spanning 144 to 504 h, the hydrogel followed a zero-order kinetic mechanism, demonstrating a controlled release with a lower AN release rate (k0 = 0.0023–0.0027). The final release phase, occurring between 720 and 1540 h, involved the degradation of the hydrogel’s polymeric network, particularly in soil conditions. This degradation resulted in the breakdown of A-CNFs and the formation of oligomeric fragments, enlarging the pore structure and enhancing AN release, which was previously restricted within the dense hydrogel matrix. This stage followed the Higuchi model, with an increased release rate (kH = 0.0281–0.0268). Notably, soil properties significantly moderated the AN release rates across all stages and pH levels, compared to buffer media, leading to a more consistent and stable AN release profile.
Water permeation and the subsequent chain relaxation in hydrogels are pivotal in determining swelling dynamics and nutrient release kinetics. Tanan et al. [141] developed biodegradable semi-IPN hydrogels composed of cassava starch (CSt)-g-PAA, natural rubber (NR), and PVA. Urea release in both aqueous and soil environments conformed to the Korsmeyer–Peppas model. When NR content was ≤50%, the behavior was non-Fickian, with both water diffusion and polymer chain relaxation governing the swelling process. When NR content exceeded 50%, diffusion was pseudo-Fickian (n < 0.5), indicating that hydrophobicity rather than chain relaxation primarily controlled water penetration, leading to a lower equilibrium swelling ratio (Seq). These findings show that Seq and n values can be adjusted by altering the NR/PVA ratio to meet specific application needs. Chamorro et al. [161] prepared cassava starch hydrogels in an aqueous environment. Upon swelling, the polymer matrix undergoes macromolecular relaxation, transitioning to a rubbery state that typically allows for solute diffusion into the surrounding aqueous phase. However, if the rate of water penetration is significantly lower than the polymer chain relaxation rate, the diffusion is characterized as “less Fickian” with an “n” value less than 0.5, although it remains within the Fickian classification. The release kinetics of potassium followed a less Fickian diffusion pattern, suggesting that its release is predominantly governed by the water intake rate, which promotes swift diffusion through the hydrogel’s rubbery matrix. Nitrogen release, on the other hand, was influenced by the hydrogel’s high fluidity, which facilitated water permeation into the biopolymer network. The release kinetic models mentioned above are the most widely used in the study of hydrogel SRFs, but due to the complexity of the fertilizer release process, other kinetic models have been studied and applied, such as the Peppas–Sahlin, Hixson–Crowell, and Ritger–Peppas models [150,158,162,163].
In conclusion, release kinetic models are indispensable tools in agricultural science. They provide a quantitative framework for understanding and predicting fertilizer behavior, which is crucial for optimizing nutrient delivery to plants and ensuring sustainable agricultural practices. By combining quantitative analysis with mathematical and geometrical models, researchers can develop more effective and environmentally friendly fertilizer formulations, ultimately leading to healthier plants and higher crop yields

5. Conclusions and Future Directions

In the pursuit of sustainable agricultural practices, the development of controlled-release fertilizers (CRFs) and smart fertilizers has garnered significant attention. These innovative fertilizers are designed to optimize nutrient delivery to crops in a manner that closely aligns with their growth requirements, thereby enhancing plant health and productivity while minimizing environmental impacts. The use of eco-friendly starch hydrogels in the formulation of these fertilizers not only aids in slow-release mechanisms but also contributes to improved water retention in soil, a critical factor in arid and semi-arid regions where water scarcity is a pressing issue. Starch hydrogels, derived from renewable plant sources, offer a biodegradable and environmentally friendly alternative to traditional synthetic polymers used in fertilizer encapsulation. These hydrogels can be engineered to have varying degrees of permeability, allowing for precise control over the release of nutrients into soil. The incorporation of starch hydrogels into fertilizer formulations can significantly reduce the initial burst effect commonly associated with conventional fertilizers, where a large amount of nutrients is released quickly, often leading to nutrient leaching and runoff.
Figure 11 provides a schematic depiction of plant growth in relation to the release of fertilizer, highlighting the discrepancy between the conventional rates of fertilizer release and the nutrient requirements of plants [164,165]. Corn, for instance, undergoes distinct growth phases, sowing, seedling, tasseling, and grain filling, each with varying nutrient demands. Traditional fertilizers like urea, due to their high solubility, release nutrients rapidly, often at times when the plant requires them the least. This mismatch can lead to inefficiencies in nutrient utilization and potential environmental harm.
Smart fertilizers, on the other hand, are designed to release nutrients in a controlled manner that mimics the natural uptake patterns of plants. By adjusting the composition and structure of starch hydrogels, researchers can modulate the release kinetics of nutrients, ensuring that they are available to the plant when needed. This approach not only enhances crop yields but also reduces the overall amount of fertilizer required, making agriculture more sustainable and cost-effective. The potential of smart fertilizers extends beyond traditional soil-based agriculture to soilless cultures, including hydroponics and aeroponics. In these systems, precise control over nutrient delivery is crucial for maintaining optimal plant growth conditions. Starch hydrogels can be adapted for use in nutrient reservoirs, providing a steady and controlled release of nutrients directly to plant roots. This application can lead to more efficient nutrient use and reduced waste, further supporting the sustainability of soilless cultivation methods.
Despite promising developments in the field of smart fertilizers, several challenges remain. The development of truly responsive fertilizers that can adapt to changing environmental conditions and plant needs is still in its infancy. Future research should focus on integrating advanced sensing technologies with fertilizer formulations to create fertilizers that can dynamically adjust their release rates based on real-time data on soil conditions, weather patterns, and plant health. Moreover, the commercialization of these advanced fertilizers requires a careful consideration of cost-effectiveness and scalability. While the benefits of using starch hydrogels and other eco-friendly materials are clear, their adoption will depend on their economic viability compared to traditional fertilizers.
The development of slow-releasing fertilizers with water retention capabilities using eco-friendly starch hydrogels represents a significant step forward in sustainable agriculture. By aligning nutrient release with plant growth needs and enhancing soil moisture retention, these fertilizers can contribute to more efficient and environmentally friendly farming practices. As research continues to advance, the potential for these smart fertilizers to revolutionize agriculture and support global food security becomes increasingly evident.

Funding

This research was funded by National Natural Science Foundation of China (No. 22178124), High-level Talent Research Start-up Project Funding of Henan Academy of Sciences (No. 232018005, 231818017), Joint Fund of Henan Province Science and Technology R&D Program (No. 225200810010, 230618025), Key Scientific and Technological Project of Henan Province (No. 232102230101), and Fundamental Research Fund of Henan Academy of Sciences (No. 230618025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Association, I.F. Fertilizer Outlook 2019–2023. In Proceedings of the 87th IFA Annual Conference, Montreal, QC, Canada, 11–13 June 2019. [Google Scholar]
  2. Naz, M.Y.; Sulaiman, S.A. Slow release coating remedy for nitrogen loss from conventional urea: A review. J. Control. Release 2016, 225, 109–120. [Google Scholar] [CrossRef] [PubMed]
  3. Govindasamy, P.; Muthusamy, S.K.; Bagavathiannan, M.; Mowrer, J.; Jagannadham, P.T.; Maity, A.; Halli, H.M.; Sujayananad, G.K.; Vadivel, R.; Das, T.K.; et al. Nitrogen use efficiency—A key to enhance crop productivity under a changing climate. Front. Plant Sci. 2023, 14, 1121073. [Google Scholar] [CrossRef] [PubMed]
  4. Gil-Ortiz, R.; Naranjo, M.A.; Ruiz-Navarro, A.; Atares, S.; Garcia, C.; Zotarelli, L.; San Bautista, A.; Vicente, O. Enhanced Agronomic Efficiency Using a New Controlled-Released, Polymeric-Coated Nitrogen Fertilizer in Rice. Plants 2020, 9, 1183. [Google Scholar] [CrossRef]
  5. Messiga, A.J.; Dyck, K.; Ronda, K.; van Baar, K.; Haak, D.; Yu, S.; Dorais, M. Nutrients Leaching in Response to Long-Term Fertigation and Broadcast Nitrogen in Blueberry Production. Plants 2020, 9, 1530. [Google Scholar] [CrossRef]
  6. Li, S.; Chen, G. Contemporary strategies for enhancing nitrogen retention and mitigating nitrous oxide emission in agricultural soils: Present and future. Environ. Dev. Sustain. 2019, 22, 2703–2741. [Google Scholar] [CrossRef]
  7. Li, T.; Gao, B.; Tong, Z.; Yang, Y.; Li, Y. Chitosan and graphene oxide nanocomposites as coatings for controlled-release fertilizer. Water Air Soil Pollut. 2019, 230, 1–9. [Google Scholar] [CrossRef]
  8. Wei, X.; Bao, X.; Yu, L.; Liu, H.; Lu, K.; Chen, L.; Bai, L.; Zhou, X.; Li, Z.; Li, W. Correlation Between Gel Strength of Starch-Based Hydrogel and Slow Release Behavior of Its Embedded Urea. J. Polym. Environ. 2020, 28, 863–870. [Google Scholar] [CrossRef]
  9. Duan, Q.; Jiang, S.; Chen, F.; Li, Z.; Ma, L.; Song, Y.; Yu, X.; Chen, Y.; Liu, H.; Yu, L. Fabrication, evaluation methodologies and models of slow-release fertilizers: A review. Ind. Crops Prod. 2023, 192, 116075. [Google Scholar] [CrossRef]
  10. Fu, J.; Wang, C.; Chen, X.; Huang, Z.; Chen, D. Classification research and types of slow controlled release fertilizers (SRFs) used—A review. Commun. Soil Sci. Plant Anal. 2018, 49, 2219–2230. [Google Scholar] [CrossRef]
  11. Kassem, I.; Ablouh, E.-H.; El Bouchtaoui, F.-Z.; Jaouahar, M.; El Achaby, M. Polymer coated slow/controlled release granular fertilizers: Fundamentals and research trends. Prog. Mater. Sci. 2024, 144, 101269. [Google Scholar] [CrossRef]
  12. Beig, B.; Niazi, M.B.K.; Jahan, Z.; Hussain, A.; Zia, M.H.; Mehran, M.T. Coating materials for slow release of nitrogen from urea fertilizer: A review. J. Plant Nutr. 2020, 43, 1510–1533. [Google Scholar] [CrossRef]
  13. El Assimi, T.; Blažic, R.; Vidović, E.; Raihane, M.; El Meziane, A.; Baouab, M.H.V.; Khouloud, M.; Beniazza, R.; Kricheldorf, H.; Lahcini, M. Polylactide/cellulose acetate biocomposites as potential coating membranes for controlled and slow nutrients release from water-soluble fertilizers. Prog. Org. Coat. 2021, 156, 106255. [Google Scholar] [CrossRef]
  14. Li, H.; Wang, N.; Zhang, L.; Wei, Y.; Zhang, L.; Ma, Y.; Ruso, J.M.; Liu, Z. Engineering and slow-release properties of lignin-based double-layer coated fertilizer. Polym. Adv. Technol. 2023, 34, 2029–2043. [Google Scholar] [CrossRef]
  15. Jyothi, A.N.; Pillai, S.S.; Aravind, M.; Salim, S.A.; Kuzhivilayil, S.J. Cassava starch-graft-poly (acrylonitrile)-coated urea fertilizer with sustained release and water retention properties. Adv. Polym. Technol. 2018, 37, 2687–2694. [Google Scholar] [CrossRef]
  16. Meng, W.; Zhang, X.; Zhang, Y.; Zhang, X.; Zhu, W.; Huang, H.; Han, X.; Liu, Y.; Xu, C. Poly (vinyl alcohol)/sodium alginate polymer membranes as eco-friendly and biodegradable coatings for slow release fertilizers. J. Sci. Food Agric. 2023, 103, 3592–3601. [Google Scholar] [CrossRef]
  17. Andry, H.; Yamamoto, T.; Irie, T.; Moritani, S.; Inoue, M.; Fujiyama, H. Water retention, hydraulic conductivity of hydrophilic polymers in sandy soil as affected by temperature and water quality. J. Hydrol. 2009, 373, 177–183. [Google Scholar] [CrossRef]
  18. Burke, D.R.; Akay, G.; Bilsborrow, P.E. Development of novel polymeric materials for agroprocess intensification. J. Appl. Polym. Sci. 2010, 118, 3292–3299. [Google Scholar] [CrossRef]
  19. Snyder, C.S.; Bruulsema, T.W.; Jensen, T.L.; Fixen, P.E. Review of greenhouse gas emissions from crop production systems and fertilizer management effects. Agric. Ecosyst. Environ. 2009, 133, 247–266. [Google Scholar] [CrossRef]
  20. Channab, B.-E.; El Idrissi, A.; Zahouily, M.; Essamlali, Y.; White, J.C. Starch-based controlled release fertilizers: A review. Int. J. Biol. Macromol. 2023, 238, 124075. [Google Scholar] [CrossRef]
  21. Matmin, J.; Ibrahim, S.I.; Mohd Hatta, M.H.; Ricky Marzuki, R.; Jumbri, K.; Nik Malek, N.A.N. Starch-Derived Superabsorbent Polymer in Remediation of Solid Waste Sludge Based on Water–Polymer Interaction. Polymers 2023, 15, 1471. [Google Scholar] [CrossRef]
  22. Azeem, B.; Elboughdiri, N.; KuShaari, K.; Jamoussi, B.; Ghernaout, D.; Ghareba, S.; Raza, S.; Gasmi, A. Valorization of almond shells’ lignocellulosic microparticles for controlled release urea production: Interactive effect of process parameters on longevity and kinetics of nutrient release. J. Coat. Technol. Res. 2021, 19, 643–660. [Google Scholar] [CrossRef]
  23. Azeem, B.; KuShaari, K.; Man, Z.B.; Basit, A.; Thanh, T.H. Review on materials & methods to produce controlled release coated urea fertilizer. J. Control. Release 2014, 181, 11–21. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, J.; Fan, X.; Zhang, L.; Chen, X.; Sun, S.; Sun, R.C. Research Progress in Lignin-Based Slow/Controlled Release Fertilizer. ChemSusChem 2020, 13, 4356–4366. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, J.; Lu, S.; Zhang, Z.; Zhao, X.; Li, X.; Ning, P.; Liu, M. Environmentally friendly fertilizers: A review of materials used and their effects on the environment. Sci. Total Environ. 2018, 613–614, 829–839. [Google Scholar] [CrossRef] [PubMed]
  26. Fertahi, S.; Ilsouk, M.; Zeroual, Y.; Oukarroum, A.; Barakat, A. Recent trends in organic coating based on biopolymers and biomass for controlled and slow release fertilizers. J. Control. Release 2021, 330, 341–361. [Google Scholar] [CrossRef]
  27. Irfan, S.A.; Razali, R.; KuShaari, K.; Mansor, N.; Azeem, B.; Ford Versypt, A.N. A review of mathematical modeling and simulation of controlled-release fertilizers. J. Control. Release 2018, 271, 45–54. [Google Scholar] [CrossRef]
  28. Jennett, T.S.; Zheng, Y. Component characterization and predictive modeling for green roof substrates optimized to adsorb P and improve runoff quality: A review. Environ. Pollut. 2018, 237, 988–999. [Google Scholar] [CrossRef]
  29. Lawrencia, D.; Wong, S.K.; Low, D.Y.S.; Goh, B.H.; Goh, J.K.; Ruktanonchai, U.R.; Soottitantawat, A.; Lee, L.H.; Tang, S.Y. Controlled release fertilizers: A review on coating materials and mechanism of release. Plants 2021, 10, 238. [Google Scholar] [CrossRef]
  30. Vejan, P.; Khadiran, T.; Abdullah, R.; Ahmad, N. Controlled release fertilizer: A review on developments, applications and potential in agriculture. J. Control. Release 2021, 339, 321–334. [Google Scholar] [CrossRef]
  31. Wesołowska, M.; Rymarczyk, J.; Góra, R.; Baranowski, P.; Sławiński, C.; Klimczyk, M.; Supryn, G.; Schimmelpfennig, L. New slow-release fertilizers—Economic, legal and practical aspects: A Review. Int. Agrophys. 2021, 35, 11–24. [Google Scholar] [CrossRef]
  32. Yuan, S.; Cheng, L.; Tan, Z. Characteristics and preparation of oil-coated fertilizers: A review. J. Control. Release 2022, 345, 675–684. [Google Scholar] [CrossRef] [PubMed]
  33. Madduma-Bandarage, U.S.K.; Madihally, S.V. Synthetic hydrogels: Synthesis, novel trends, and applications. J. Appl. Polym. Sci. 2021, 138, 50376. [Google Scholar] [CrossRef]
  34. Liu, Y.; Zhu, Y.; Mu, B.; Wang, Y.; Quan, Z.; Wang, A. Synthesis, characterization, and swelling behaviors of sodium carboxymethyl cellulose-g-poly(acrylic acid)/semi-coke superabsorbent. Polym. Bull. 2022, 79, 935–953. [Google Scholar] [CrossRef]
  35. Yang, Y.; Liang, Z.; Zhang, R.; Zhou, S.; Yang, H.; Chen, Y.; Zhang, J.; Yin, H.; Yu, D. Research Advances in Superabsorbent Polymers. Polymers 2024, 16, 501. [Google Scholar] [CrossRef] [PubMed]
  36. Ma, X.; Wen, G. Development history and synthesis of super-absorbent polymers: A review. J. Polym. Res. 2020, 27, 136. [Google Scholar] [CrossRef]
  37. Manuel, M.; Jennifer, A. A review on starch and cellulose-enhanced superabsorbent hydrogel. J. Chem. Rev. 2023, 5, 183–203. [Google Scholar] [CrossRef]
  38. Salimi, M.; Channab, B.-E.; El Idrissi, A.; Zahouily, M.; Motamedi, E. A comprehensive review on starch: Structure, modification, and applications in slow/controlled-release fertilizers in agriculture. Carbohydr. Polym. 2023, 322, 121326. [Google Scholar] [CrossRef]
  39. Qamruzzaman, M.; Ahmed, F.; Mondal, M.I.H. An Overview on Starch-Based Sustainable Hydrogels: Potential Applications and Aspects. J. Polym. Environ. 2022, 30, 19–50. [Google Scholar] [CrossRef]
  40. Meng, R.; Wu, Z.; Xie, H.-Q.; Xu, G.-X.; Cheng, J.-S.; Zhang, B. Preparation, characterization, and encapsulation capability of the hydrogel cross-linked by esterified tapioca starch. Int. J. Biol. Macromol. 2020, 155, 1–5. [Google Scholar] [CrossRef]
  41. Haq, F.; Yu, H.; Wang, L.; Teng, L.; Haroon, M.; Khan, R.U.; Mehmood, S.; Bilal Ul, A.; Ullah, R.S.; Khan, A.; et al. Advances in chemical modifications of starches and their applications. Carbohydr. Res. 2019, 476, 12–35. [Google Scholar] [CrossRef]
  42. Masina, N.; Choonara, Y.E.; Kumar, P.; du Toit, L.C.; Govender, M.; Indermun, S.; Pillay, V. A review of the chemical modification techniques of starch. Carbohydr. Polym. 2017, 157, 1226–1236. [Google Scholar] [CrossRef] [PubMed]
  43. Dragan, E.S.; Apopei, D.F. Synthesis and swelling behavior of pH-sensitive semi-interpenetrating polymer network composite hydrogels based on native and modified potatoes starch as potential sorbent for cationic dyes. Chem. Eng. J. 2011, 178, 252–263. [Google Scholar] [CrossRef]
  44. Kiatkamjornwong, S.; Mongkolsawat, K.; Sonsuk, M. Synthesis and property characterization of cassava starch grafted poly [acrylamide-co-(maleic acid)] superabsorbent via γ-irradiation. Polymer 2002, 43, 3915–3924. [Google Scholar] [CrossRef]
  45. Czarnecka, E.; Nowaczyk, J. Semi-Natural Superabsorbents Based on Starch-g-poly(acrylic acid): Modification, Synthesis and Application. Polymers 2020, 12, 1794. [Google Scholar] [CrossRef] [PubMed]
  46. Czarnecka, E.; Nowaczyk, J. Synthesis and Characterization Superabsorbent Polymers Made of Starch, Acrylic Acid, Acrylamide, Poly(Vinyl Alcohol), 2-Hydroxyethyl Methacrylate, 2-Acrylamido-2-methylpropane Sulfonic Acid. Int. J. Mol. Sci. 2021, 22, 4325. [Google Scholar] [CrossRef]
  47. Noordergraaf, I.-W.; Witono, J.R.; Heeres, H.J. Grafting Starch with Acrylic Acid and Fenton’s Initiator: The Selectivity Challenge. Polymers 2024, 16, 255. [Google Scholar] [CrossRef]
  48. Aniagor, C.O.; Afifi, M.A.; Hashem, A. Rapid and efficient uptake of aqueous lead pollutant using starch-based superabsorbent hydrogel. Polym. Bull. 2022, 79, 6373–6388. [Google Scholar] [CrossRef]
  49. Supare, K.; Mahanwar, P.A. Starch-derived superabsorbent polymers in agriculture applications: An overview. Polym. Bull. 2022, 79, 5795–5824. [Google Scholar] [CrossRef]
  50. Khan, M.A.; Bhattacharia, S.K.; Kader, M.A.; Bahari, K. Preparation and characterization of ultra violet (UV) radiation cured bio-degradable films of sago starch/PVA blend. Carbohydr. Polym. 2006, 63, 500–506. [Google Scholar] [CrossRef]
  51. Parvathy, P.C.; Jyothi, A.N. Synthesis, characterization and swelling behaviour of superabsorbent polymers from cassava starch-graft-poly(acrylamide). Starch Starke 2012, 64, 207–218. [Google Scholar] [CrossRef]
  52. Amiri, F.; Kabiri, K.; Bouhendi, H.; Abdollahi, H.; Najafi, V.; Karami, Z. High gel-strength hybrid hydrogels based on modified starch through surface cross-linking technique. Polym. Bull. 2019, 76, 4047–4068. [Google Scholar] [CrossRef]
  53. Ma, D.; Zhu, B.; Cao, B.; Wang, J.; Zhang, J. The Microstructure and Swelling Properties of Poly Acrylic Acid-Acrylamide Grafted Starch Hydrogels. J. Macromol. Sci. Part B 2016, 55, 1124–1133. [Google Scholar] [CrossRef]
  54. Zamani-Babgohari, F.; Irannejad, A.; Kalantari Pour, M.; Khayati, G.R. Synthesis of carboxymethyl starch co (polyacrylamide/ polyacrylic acid) hydrogel for removing methylene blue dye from aqueous solution. Int. J. Biol. Macromol. 2024, 269, 132053. [Google Scholar] [CrossRef] [PubMed]
  55. Bazoubandi, B.; Soares, J.B.P. Amylopectin-graft-polyacrylamide for the flocculation and dewatering of oil sands tailings. Miner. Eng. 2020, 148, 106196. [Google Scholar] [CrossRef]
  56. Mahmoodi-Babolan, N.; Nematollahzadeh, A.; Heydari, A.; Merikhy, A. Bioinspired catecholamine/starch composites as superadsorbent for the environmental remediation. Int. J. Biol. Macromol. 2019, 125, 690–699. [Google Scholar] [CrossRef]
  57. Zhang, Q.; Wang, Z.; Zhang, C.; Aluko, R.E.; Yuan, J.; Ju, X.; He, R. Structural and functional characterization of rice starch-based superabsorbent polymer materials. Int. J. Biol. Macromol. 2020, 153, 1291–1298. [Google Scholar] [CrossRef]
  58. Siyamak, S.; Laycock, B.; Luckman, P. Synthesis of starch graft-copolymers via reactive extrusion and their ammonium sorption properties. Chem. Eng. J. 2020, 398, 124291. [Google Scholar] [CrossRef]
  59. Zou, W.; Yu, L.; Liu, X.; Chen, L.; Zhang, X.; Qiao, D.; Zhang, R. Effects of amylose/amylopectin ratio on starch-based superabsorbent polymers. Carbohydr. Polym. 2012, 87, 1583–1588. [Google Scholar] [CrossRef]
  60. Bao, X.; Yu, L.; Simon, G.P.; Shen, S.; Xie, F.; Liu, H.; Chen, L.; Zhong, L. Rheokinetics of graft copolymerization of acrylamide in concentrated starch and rheological behaviors and microstructures of reaction products. Carbohydr. Polym. 2018, 192, 1–9. [Google Scholar] [CrossRef]
  61. Bao, X.; Yu, L.; Shen, S.; Simon, G.P.; Liu, H.; Chen, L. How rheological behaviors of concentrated starch affect graft copolymerization of acrylamide and resultant hydrogel. Carbohydr. Polym. 2019, 219, 395–404. [Google Scholar] [CrossRef]
  62. Sethi, S.; Saruchi; Medha; Thakur, S.; Kaith, B.S.; Sharma, N.; Ansar, S.; Pandey, S.; Kuma, V. Biopolymer starch-gelatin embedded with silver nanoparticle–based hydrogel composites for antibacterial application. Biomass Convers. Biorefinery 2022, 12, 5363–5384. [Google Scholar] [CrossRef]
  63. Motamedi, E.; Motesharezedeh, B.; Shirinfekr, A.; Samar, S.M. Synthesis and swelling behavior of environmentally friendly starch-based superabsorbent hydrogels reinforced with natural char nano/micro particles. J. Environ. Chem. Eng. 2020, 8, 103583. [Google Scholar] [CrossRef]
  64. Olad, A.; Doustdar, F.; Gharekhani, H. Fabrication and characterization of a starch-based superabsorbent hydrogel composite reinforced with cellulose nanocrystals from potato peel waste. Colloids Surf. A 2020, 601, 124962. [Google Scholar] [CrossRef]
  65. Li, A.; Zhang, J.; Wang, A. Utilization of starch and clay for the preparation of superabsorbent composite. Bioresour. Technol. 2007, 98, 327–332. [Google Scholar] [CrossRef]
  66. Mahmoud, G.A.; Abdel-Aal, S.E.; Badway, N.A.; Abo Farha, S.A.; Alshafei, E.A. Radiation synthesis and characterization of starch-based hydrogels for removal of acid dye. Starch Stärke 2014, 66, 400–408. [Google Scholar] [CrossRef]
  67. Relleve, L.S.; Aranilla, C.T.; Barba, B.J.D.; Gallardo, A.K.R.; Cruz, V.R.C.; Ledesma, C.R.M.; Nagasawa, N.; Abad, L.V. Radiation-synthesized polysaccharides/polyacrylate super water absorbents and their biodegradabilities. Radiat. Phys. Chem. 2020, 170, 108618. [Google Scholar] [CrossRef]
  68. Yoshii, F.; Zhao, L.; Wach, R.A.; Nagasawa, N.; Mitomo, H.; Kume, T. Hydrogels of polysaccharide derivatives crosslinked with irradiation at paste-like condition. Nucl. Instrum. Methods Phys. Res. Sect. B 2003, 208, 320–324. [Google Scholar] [CrossRef]
  69. Zhai, M.; Yoshii, F.; Kume, T.; Hashim, K. Syntheses of PVA/starch grafted hydrogels by irradiation. Carbohydr. Polym. 2002, 50, 295–303. [Google Scholar] [CrossRef]
  70. Saif, M.J.; Naveed, M.; Asif, H.M.; Akhtar, R. Irradiation applications for polymer nano-composites: A state-of-the-art review. J. Ind. Eng. Chem. 2018, 60, 218–236. [Google Scholar] [CrossRef]
  71. Goganian, A.M.; Hamishehkar, H.; Arsalani, N.; Khiabani, H.K. Microwave-Promoted Synthesis of Smart Superporous Hydrogel for the Development of Gastroretentive Drug Delivery System. Adv. Polym. Technol. 2015, 34, 21490. [Google Scholar] [CrossRef]
  72. Naikwadi, A.T.; Sharma, B.K.; Bhatt, K.D.; Mahanwar, P.A. Gamma Radiation Processed Polymeric Materials for High Performance Applications: A Review. Front. Chem. 2022, 10, 837111. [Google Scholar] [CrossRef] [PubMed]
  73. Abd El-Rehim, H.; Hegazy, E.-S.A.; Diaa, D. Radiation synthesis of eco-friendly water reducing sulfonated starch/acrylic acid hydrogel designed for cement industry. Radiat. Phys. Chem. 2013, 85, 139–146. [Google Scholar] [CrossRef]
  74. Sapuła, P.; Bialik-Wąs, K.; Malarz, K. Are Natural Compounds a Promising Alternative to Synthetic Cross-Linking Agents in the Preparation of Hydrogels? Pharmaceutics 2023, 15, 253. [Google Scholar] [CrossRef]
  75. Braşoveanu, M.; Nemţanu, M.R.; Duţă, D. Electron-beam processed corn starch: Evaluation of physicochemical and structural properties and technical-economic aspects of the processing. Braz. J. Chem. Eng. 2013, 30, 847–856. [Google Scholar] [CrossRef]
  76. Bashir, K.; Aggarwal, M. Physicochemical, structural and functional properties of native and irradiated starch: A review. J. Food Sci. Technol. 2019, 56, 513–523. [Google Scholar] [CrossRef]
  77. Sánchez-Moreno, V.E.; Sandoval-Pauker, C.; Aldas, M.; Ciobotă, V.; Luna, M.; Vargas Jentzsch, P.; Muñoz Bisesti, F. Synthesis of inulin hydrogels by electron beam irradiation: Physical, vibrational spectroscopic and thermal characterization and arsenic removal as a possible application. J. Polym. Res. 2020, 27, 184. [Google Scholar] [CrossRef]
  78. More, A.P.; Chapekar, S. Irradiation assisted synthesis of hydrogel: A Review. Polym. Bull. 2024, 81, 5839–5908. [Google Scholar] [CrossRef]
  79. Moghaddam, R.H.; Dadfarnia, S.; Shabani, A.M.H.; Moghaddam, Z.H.; Tavakol, M. Electron beam irradiation synthesis of porous and non-porous pectin based hydrogels for a tetracycline drug delivery system. Mater. Sci. Eng. C 2019, 102, 391–404. [Google Scholar] [CrossRef]
  80. Abd El-Mohdy, H.L.; Hegazy, E.A.; El-Nesr, E.M.; El-Wahab, M.A. Synthesis, characterization and properties of radiation-induced Starch/(EG-co-MAA) hydrogels. Arab. J. Chem. 2016, 9, S1627–S1635. [Google Scholar] [CrossRef]
  81. Bhuyan, M.M.; Chandra Dafader, N.; Hara, K.; Okabe, H.; Hidaka, Y.; Rahman, M.M.; Mizanur Rahman Khan, M.; Rahman, N. Synthesis of Potato Starch-Acrylic-Acid Hydrogels by Gamma Radiation and Their Application in Dye Adsorption. Int. J. Polym. Sci. 2016, 2016, 9867859. [Google Scholar] [CrossRef]
  82. Chen, F.; Miao, C.; Duan, Q.; Jiang, S.; Liu, H.; Ma, L.; Li, Z.; Bao, X.; Lan, B.; Chen, L.; et al. Developing slow release fertilizer through in-situ radiation-synthesis of urea-embedded starch-based hydrogels. Ind. Crops Prod. 2023, 191, 115971. [Google Scholar] [CrossRef]
  83. Fekete, T.; Borsa, J.; Takács, E.; Wojnárovits, L. Synthesis of carboxymethylcellulose/starch superabsorbent hydrogels by gamma-irradiation. Chem. Cent. J. 2017, 11, 46. [Google Scholar] [CrossRef] [PubMed]
  84. Finkenstadt, V.L.; Willett, J.L. Reactive Extrusion of Starch-Polyacrylamide Graft Copolymers: Effects of Monomer/Starch Ratio and Moisture Content. Macromol. Chem. Phys. 2005, 206, 1648–1652. [Google Scholar] [CrossRef]
  85. Willett, J.L.; Finkenstadt, V.L. Initiator effects in reactive extrusion of starch-polyacrylamide graft copolymers. J. Appl. Polym. Sci. 2006, 99, 52–58. [Google Scholar] [CrossRef]
  86. Willett, J.L.; Finkenstadt, V.L. Reactive Extrusion of Starch–Polyacrylamide Graft Copolymers Using Various Starches. J. Polym. Environ. 2006, 14, 125–129. [Google Scholar] [CrossRef]
  87. Willett, J.L.; Finkenstadt, V.L. Comparison of Cationic and Unmodified Starches in Reactive Extrusion of Starch–Polyacrylamide Graft Copolymers. J. Appl. Polym. Sci. 2009, 17, 248–253. [Google Scholar] [CrossRef]
  88. Willett, J.L.; Finkenstadt, V.L. Starch-poly (acrylamide-co-2-acrylamido-2-methylpropanesulfonic acid) graft copolymers prepared by reactive extrusion. J. Appl. Polym. Sci. 2015, 132, 42405. [Google Scholar] [CrossRef]
  89. Li, T.-T.; Feng, L.-F.; Gu, X.-P.; Zhang, C.-L.; Wang, P.; Hu, G.-H. Intensification of Polymerization Processes by Reactive Extrusion. Ind. Eng. Chem. Res. 2021, 60, 2791–2806. [Google Scholar] [CrossRef]
  90. Liu, H.; Xie, F.; Yu, L.; Chen, L.; Li, L. Thermal processing of starch-based polymers. Prog. Polym. Sci. 2009, 34, 1348–1368. [Google Scholar] [CrossRef]
  91. McGauran, T.; Harris, M.; Dunne, N.; Smyth, B.M.; Cunningham, E. Development and optimisation of extruded bio-based polymers from poultry feathers. Eur. Polym. J. 2021, 158, 110678. [Google Scholar] [CrossRef]
  92. Moad, G. Chemical modification of starch by reactive extrusion. Prog. Polym. Sci. 2011, 36, 218–237. [Google Scholar] [CrossRef]
  93. Xie, F.; Yu, L.; Liu, H.; Chen, L. Starch Modification Using Reactive Extrusion. Starch Starke 2006, 58, 131–139. [Google Scholar] [CrossRef]
  94. Xu, E.; Campanella, O.H.; Ye, X.; Jin, Z.; Liu, D.; BeMiller, J.N. Advances in conversion of natural biopolymers: A reactive extrusion (REX)–enzyme-combined strategy for starch/protein-based food processing. Trends Food Sci. Technol. 2020, 99, 167–180. [Google Scholar] [CrossRef]
  95. Siyamak, S.; Laycock, B.; Luckman, P. Synthesis of starch graft-copolymers via reactive extrusion: Process development and structural analysis. Carbohydr. Polym. 2020, 227, 115066. [Google Scholar] [CrossRef]
  96. Zhuang, Y.; Saadatkhah, N.; Morgani, M.S.; Xu, T.; Martin, C.; Patience, G.S.; Ajji, A. Experimental methods in chemical engineering: Reactive extrusion. Can. J. Chem. Eng. 2023, 101, 59–77. [Google Scholar] [CrossRef]
  97. Jiang, T.; Chen, F.; Duan, Q.; Bao, X.; Jiang, S.; Liu, H.; Chen, L.; Yu, L. Designing and application of reactive extrusion with twice initiations for graft copolymerization of acrylamide on starch. Eur. Polym. J. 2022, 165, 111008. [Google Scholar] [CrossRef]
  98. McKenna, G.B. Deformation and flow of matter: Interrogating the physics of materials using rheological methods. J. Rheol. 2012, 56, 113–158. [Google Scholar] [CrossRef]
  99. Serrero, A.; Trombotto, S.; Cassagnau, P.; Bayon, Y.; Gravagna, P.; Montanari, S.; David, L. Polysaccharide Gels Based on Chitosan and Modified Starch: Structural Characterization and Linear Viscoelastic Behavior. Biomacromolecules 2010, 11, 1534–1543. [Google Scholar] [CrossRef]
  100. Stojkov, G.; Niyazov, Z.; Picchioni, F.; Bose, R.K. Relationship between Structure and Rheology of Hydrogels for Various Applications. Gels 2021, 7, 255. [Google Scholar] [CrossRef]
  101. Narita, T.; Mayumi, K.; Ducouret, G.; Hébraud, P. Viscoelastic properties of poly (vinyl alcohol) hydrogels having permanent and transient cross-links studied by microrheology, classical rheometry, and dynamic light scattering. Macromolecules 2013, 46, 4174–4183. [Google Scholar] [CrossRef]
  102. Kavanagh, G.M.; Ross-Murphy, S.B. Rheological characterisation of polymer gels. Prog. Polym. Sci. 1998, 23, 533–562. [Google Scholar] [CrossRef]
  103. Wetton, R.E.; Marsh, R.D.L.; Van-de-Velde, J.G. Theory and application of dynamic mechanical thermal analysis. Thermochim. Acta 1991, 175, 1–11. [Google Scholar] [CrossRef]
  104. Sha, H.; Zhang, X.; Harrison, I.R. A dynamic mechanical thermal analysis (DMTA) study of polyethylenes. Thermochim. Acta 1991, 192, 233–242. [Google Scholar] [CrossRef]
  105. Badia, J.-D.; Santonja-Blasco, L.; Martínez-Felipe, A.; Ribes-Greus, A. Dynamic Mechanical Thermal Analysis of Polymer Blends. In Characterization of Polymer Blends; Wiley: Hoboken, NJ, USA, 2014; pp. 365–392. [Google Scholar]
  106. Ramli, R.A. Slow release fertilizer hydrogels: A review. Polym. Chem. 2019, 10, 6073–6090. [Google Scholar] [CrossRef]
  107. Kalita, A.; Elayarajan, M.; Janaki, P.; Suganya, S.; Sankari, A.; Parameswari, E. Organo-monomers coated slow-release fertilizers: Current understanding and future prospects. Int. J. Biol. Macromol. 2024, 274, 133320. [Google Scholar] [CrossRef]
  108. Wang, X.; Yang, Y.; Zhong, S.; Meng, Q.; Li, Y.; Wang, J.; Gao, Y.; Cui, X. Advances in controlled-release fertilizer encapsulated by organic-inorganic composite membranes. Particuology 2024, 84, 236–248. [Google Scholar] [CrossRef]
  109. Ahmed Khan, T.; Zakaria, M.E.T.; Kim, H.J.; Ghazali, S.; Jamari, S.S. Carbonaceous microsphere-based superabsorbent polymer as filler for coating of NPK fertilizer: Fabrication, properties, swelling, and nitrogen release characteristics. J. Appl. Polym. Sci. 2020, 137, 48396. [Google Scholar] [CrossRef]
  110. Chen, L.; Xie, Z.; Zhuang, X.; Chen, X.; Jing, X. Controlled release of urea encapsulated by starch-g-poly(l-lactide). Carbohydr. Polym. 2008, 72, 342–348. [Google Scholar] [CrossRef]
  111. Zafar, N.; Niazi, M.B.K.; Sher, F.; Khalid, U.; Jahan, Z.; Shah, G.A.; Zia, M. Starch and polyvinyl alcohol encapsulated biodegradable nanocomposites for environment friendly slow release of urea fertilizer. Chem. Eng. J. Adv. 2021, 7, 100123. [Google Scholar] [CrossRef]
  112. Boetje, L.; Lan, X.; Silvianti, F.; van Dijken, J.; Polhuis, M.; Loos, K. A more efficient synthesis and properties of saturated and unsaturated starch esters. Carbohydr. Polym. 2022, 292, 119649. [Google Scholar] [CrossRef]
  113. Salimi, M.; Motamedi, E.; Safari, M.; Motesharezadeh, B. Synthesis of urea slow-release fertilizer using a novel starch-g-poly (styrene-co-butylacrylate) nanocomposite latex and its impact on a model crop production in greenhouse. J. Cleaner Prod. 2021, 322, 129082. [Google Scholar] [CrossRef]
  114. Qiao, D.; Liu, H.; Yu, L.; Bao, X.; Simon, G.P.; Petinakis, E.; Chen, L. Preparation and characterization of slow-release fertilizer encapsulated by starch-based superabsorbent polymer. Carbohydr. Polym. 2016, 147, 146–154. [Google Scholar] [CrossRef] [PubMed]
  115. Lü, S.; Gao, C.; Wang, X.; Xu, X.; Bai, X.; Gao, N.; Feng, C.; Wei, Y.; Wu, L.; Liu, M. Synthesis of a starch derivative and its application in fertilizer for slow nutrient release and water-holding. RSC Adv. 2014, 4, 51208–51214. [Google Scholar] [CrossRef]
  116. Sarkar, A.; Biswas, D.R.; Datta, S.C.; Dwivedi, B.S.; Bhattacharyya, R.; Kumar, R.; Bandyopadhyay, K.K.; Saha, M.; Chawla, G.; Saha, J.K.; et al. Preparation of novel biodegradable starch/poly(vinyl alcohol)/bentonite grafted polymeric films for fertilizer encapsulation. Carbohydr. Polym. 2021, 259, 117679. [Google Scholar] [CrossRef]
  117. León, O.; Soto, D.; González, J.; Piña, C.; Muñoz-Bonilla, A.; Fernandez-García, M. Environmentally Friendly Fertilizers Based on Starch Superabsorbents. Materials 2019, 12, 3493. [Google Scholar] [CrossRef]
  118. Perez, J.J.; Francois, N.J. Chitosan-starch beads prepared by ionotropic gelation as potential matrices for controlled release of fertilizers. Carbohydr. Polym. 2016, 148, 134–142. [Google Scholar] [CrossRef]
  119. Ramírez-Hernández, A.; Aparicio-Saguilán, A.; Mata-Mata, J.L.; González-García, G.; Hernández-Mendoza, H.; Gutiérrez-Fuentes, A.; Báez-García, E. Chemical modification of banana starch by the in situ polymerization of ϵ-caprolactone in one step. Starch Starke 2017, 69, 1600197. [Google Scholar] [CrossRef]
  120. Yao, Y.; Shen, Y.; Hu, C.; Wu, H. Superabsorbent fabric: In situ polymerisation of macroporous starch–sodium alginate–polyacrylate hydrogel on fibre surface. Cellulose 2023, 30, 7113–7128. [Google Scholar] [CrossRef]
  121. Salimi, M.; Motamedi, E.; Motesharezedeh, B.; Hosseini, H.M.; Alikhani, H.A. Starch-g-poly(acrylic acid-co-acrylamide) composites reinforced with natural char nanoparticles toward environmentally benign slow-release urea fertilizers. J. Environ. Chem. Eng. 2020, 8, 103765. [Google Scholar] [CrossRef]
  122. Bolt, R.R.; Leitch, J.A.; Jones, A.C.; Nicholson, W.I.; Browne, D.L. Continuous flow mechanochemistry: Reactive extrusion as an enabling technology in organic synthesis. Chem. Soc. Rev. 2022, 51, 4243–4260. [Google Scholar] [CrossRef]
  123. Dong, Q.; Xian, B.; Hong, L.; Xing, L.; Ling, C.; Long, Y.; Xiao, Z.; Pei, C. Preparation of cassava starch-based superabsorbent polymer using a twin-roll mixer as reactor. Chin. J. Polym. Sci. 2014, 32, 1348–1356. [Google Scholar] [CrossRef]
  124. Qiao, D.; Ma, H.; Yu, L.; Liu, H.; Zou, W.; Chen, L.; Chen, P. Synthesis and characteristics of graft copolymerization of starch-G-PAM using A twin-roll mixer as reactor for cornstarch with different amylose/amylopectin ratios. Int. Polym. Proc. 2014, 29, 252–259. [Google Scholar] [CrossRef]
  125. Qiao, D.; Tu, W.; Wang, Z.; Yu, L.; Zhang, B.; Bao, X.; Jiang, F.; Lin, Q. Influence of crosslinker amount on the microstructure and properties of starch-based superabsorbent polymers by one-step preparation at high starch concentration. Int. J. Biol. Macromol. 2019, 129, 679–685. [Google Scholar] [CrossRef] [PubMed]
  126. Xiao, X.; Yu, L.; Xie, F.; Bao, X.; Liu, H.; Ji, Z.; Chen, L. One-step method to prepare starch-based superabsorbent polymer for slow release of fertilizer. Chem. Eng. J. 2017, 309, 607–616. [Google Scholar] [CrossRef]
  127. Bhaladhare, S.; Das, D. Cellulose: A fascinating biopolymer for hydrogel synthesis. J. Mater. Chem. B 2022, 10, 1923–1945. [Google Scholar] [CrossRef]
  128. Tan, X.; Lv, Q.; Dong, G.; Zhang, Z.; Chai, D.-f.; Zhao, M.; Zhang, W.; Li, J. Insight into the preparation and improved properties of cellulose citrate ester hydrogel slow-release fertilizer. Ind. Crops Prod. 2024, 222, 119517. [Google Scholar] [CrossRef]
  129. Li, Z.; Zhang, M. Progress in the preparation of stimulus-responsive cellulose hydrogels and their application in slow-release fertilizers. Polymers 2023, 15, 3643. [Google Scholar] [CrossRef]
  130. González, K.; Guaresti, O.; Palomares, T.; Alonso-Varona, A.; Eceiza, A.; Gabilondo, N. The role of cellulose nanocrystals in biocompatible starch-based clicked nanocomposite hydrogels. Int. J. Biol. Macromol. 2020, 143, 265–272. [Google Scholar] [CrossRef]
  131. Moharrami, P.; Motamedi, E. Application of cellulose nanocrystals prepared from agricultural wastes for synthesis of starch-based hydrogel nanocomposites: Efficient and selective nanoadsorbent for removal of cationic dyes from water. Bioresour. Technol. 2020, 313, 123661. [Google Scholar] [CrossRef]
  132. Bahadoran Baghbadorani, N.; Behzad, T.; Karimi Darvanjooghi, M.H.; Etesami, N. Modelling of water absorption kinetics and biocompatibility study of synthesized cellulose nanofiber-assisted starch-graft-poly(acrylic acid) hydrogel nanocomposites. Cellulose 2020, 27, 9927–9945. [Google Scholar] [CrossRef]
  133. Bora, A.; Sarmah, D.; Karak, N. Cellulosic wastepaper modified starch/itaconic acid/acrylic acid-based biodegradable hydrogel as a sustain release of NPK fertilizer vehicle for agricultural applications. Int. J. Biol. Macromol. 2023, 253, 126555. [Google Scholar] [CrossRef] [PubMed]
  134. Vudjung, C.; Saengsuwan, S. Synthesis and properties of biodegradable hydrogels based on cross-linked natural rubber and cassava starch. J. Elastomers Plast. 2017, 49, 574–594. [Google Scholar] [CrossRef]
  135. Vudjung, C.; Saengsuwan, S. Biodegradable IPN hydrogels based on pre-vulcanized natural rubber and cassava starch as coating membrane for environment-friendly slow-release urea fertilizer. J. Polym. Environ. 2018, 26, 3967–3980. [Google Scholar] [CrossRef]
  136. Pimsen, R.; Porrawatkul, P.; Nuengmatcha, P.; Ramasoot, S.; Chanthai, S. Efficiency enhancement of slow release of fertilizer using nanozeolite–chitosan/sago starch-based biopolymer composite. J. Coat. Technol. Res. 2021, 18, 1321–1332. [Google Scholar] [CrossRef]
  137. Majeed, Z.; Ramli, N.K.; Binti Mansor, N.; Man, Z. Lignin loading effect on biodegradability and nitrogen release properties of urea modified tapioca starch in wet soil. Key Eng. Mater. 2014, 594, 798–802. [Google Scholar] [CrossRef]
  138. Abdul Khalil, H.P.S.; Chong, E.W.N.; Owolabi, F.A.T.; Asniza, M.; Tye, Y.Y.; Rizal, S.; Nurul Fazita, M.R.; Mohamad Haafiz, M.K.; Nurmiati, Z.; Paridah, M.T. Enhancement of basic properties of polysaccharide-based composites with organic and inorganic fillers: A review. J. Appl. Polym. Sci. 2019, 136, 47251. [Google Scholar] [CrossRef]
  139. Lu, J.; Li, Y.; Cai, Y.; Jiang, P.; Yu, B. Co-incorporation of hydrotalcite and starch into biochar-based fertilizers for the synthesis of slow-release fertilizers with improved water retention. Biochar 2023, 5, 44. [Google Scholar] [CrossRef]
  140. Wei, H.; Wang, H.; Chu, H.; Li, J. Preparation and characterization of slow-release and water-retention fertilizer based on starch and halloysite. Int. J. Biol. Macromol. 2019, 133, 1210–1218. [Google Scholar] [CrossRef]
  141. Tanan, W.; Panichpakdee, J.; Suwanakood, P.; Saengsuwan, S. Biodegradable hydrogels of cassava starch-g-polyacrylic acid/natural rubber/polyvinyl alcohol as environmentally friendly and highly efficient coating material for slow-release urea fertilizers. J. Ind. Eng. Chem. 2021, 101, 237–252. [Google Scholar] [CrossRef]
  142. Jin, S.; Wang, Y.; He, J.; Yang, Y.; Yu, X.; Yue, G. Preparation and properties of a degradable interpenetrating polymer networks based on starch with water retention, amelioration of soil, and slow release of nitrogen and phosphorus fertilizer. J. Appl. Polym. Sci. 2013, 128, 407–415. [Google Scholar] [CrossRef]
  143. León, O.; Soto, D.; Muñoz-Bonilla, A.; Fernández-García, M. Amylose Modified Starches as Superabsorbent Systems for Release of Potassium Fertilizers. J. Polym. Environ. 2022, 30, 2314–2328. [Google Scholar] [CrossRef]
  144. Dudu, T.E.; Alpaslan, D.; Aktas, N. Synthesis of controlled release hydrogels from dimethylacrylamide/maleic acid/starch and its application in lettuce cultivation. J. Polym. Res. 2022, 29, 524. [Google Scholar] [CrossRef]
  145. Li, Y.; Ma, Y.; Chang, F.; Zhu, H.; Tian, C.; Jia, F.; Ke, Y.; Dai, J. Slow-Release Urea Fertilizer with Water Retention and Photosensitivity Properties Based on Sodium Alginate/Carboxymethyl Starch Sodium/Polydopamine. Processes 2024, 12, 842. [Google Scholar] [CrossRef]
  146. Zhao, C.; Tian, H.; Zhang, Q.; Liu, Z.; Zhang, M.; Wang, J. Preparation of urea-containing starch-castor oil superabsorbent polyurethane coated urea and investigation of controlled nitrogen release. Carbohydr. Polym. 2021, 253, 117240. [Google Scholar] [CrossRef] [PubMed]
  147. Lv, Q.; Xiao, T.; Dong, G.; Tan, X.; Zhang, Z.; Zhao, M.; Zhu, M.; Li, J.; Zhang, W. Preparation and characterization of starch carbamate modified natural sodium alginate composite hydrogel blend formulation and its application for slow-release fertilizer. Int. J. Biol. Macromol. 2024, 278, 134713. [Google Scholar] [CrossRef]
  148. Mu, Z.; Zhang, W.; Chai, D.-f.; Lv, Q.; Tan, X.; Yuan, R.; Dong, G. Preparation and characterization of slow-release urea fertilizer encapsulated by a blend of starch derivative and polyvinyl alcohol with desirable biodegradability and availability. Int. J. Biol. Macromol. 2024, 271, 132693. [Google Scholar] [CrossRef]
  149. Ee Huey, C.; Zaireen Nisa Yahya, W.; Mansor, N. Allicin incorporation as urease inhibitor in a chitosan/starch based biopolymer for fertilizer application. Mater. Today Proc. 2019, 16, 2187–2196. [Google Scholar] [CrossRef]
  150. Lakshani, N.; Wijerathne, H.S.; Sandaruwan, C.; Kottegoda, N.; Karunarathne, V. Release Kinetic Models and Release Mechanisms of Controlled-Release and Slow-Release Fertilizers. ACS Agric. Sci. Technol. 2023, 3, 939–956. [Google Scholar] [CrossRef]
  151. Priya, E.; Sarkar, S.; Maji, P.K. A review on slow-release fertilizer: Nutrient release mechanism and agricultural sustainability. J. Environ. Chem. Eng. 2024, 12, 113211. [Google Scholar] [CrossRef]
  152. Firmanda, A.; Fahma, F.; Syamsu, K.; Mahardika, M.; Suryanegara, L.; Munif, A.; Gozan, M.; Wood, K.; Hidayat, R.; Yulia, D. Biopolymer-based slow/controlled-release fertilizer (SRF/CRF): Nutrient release mechanism and agricultural sustainability. J. Environ. Chem. Eng. 2024, 12, 112177. [Google Scholar] [CrossRef]
  153. Lin, C.-C.; Metters, A.T. Hydrogels in controlled release formulations: Network design and mathematical modeling. Adv. Drug Deliv. Rev. 2006, 58, 1379–1408. [Google Scholar] [CrossRef] [PubMed]
  154. Shang, H.; Yang, X.; Liu, H. Temperature-responsive hydrogel prepared from carboxymethyl cellulose-stabilized N-vinylcaprolactam with potential for fertilizer delivery. Carbohydr. Polym. 2023, 313, 120875. [Google Scholar] [CrossRef] [PubMed]
  155. Dai, H.; Zhang, Y.; Ma, L.; Zhang, H.; Huang, H. Synthesis and response of pineapple peel carboxymethyl cellulose-g-poly (acrylic acid-co-acrylamide)/graphene oxide hydrogels. Carbohydr. Polym. 2019, 215, 366–376. [Google Scholar] [CrossRef] [PubMed]
  156. Damiri, F.; Salave, S.; Vitore, J.; Bachra, Y.; Jadhav, R.; Kommineni, N.; Karouach, F.; Paiva-Santos, A.C.; Varma, R.S.; Berrada, M. Properties and valuable applications of superabsorbent polymers: A comprehensive review. Polym. Bull. 2024, 81, 6671–6701. [Google Scholar] [CrossRef]
  157. Zhang, W.; Wang, P.; Liu, S.; Chen, J.; Chen, R.; He, X.; Ma, G.; Lei, Z. Factors affecting the properties of superabsorbent polymer hydrogels and methods to improve their performance: A review. J. Mater. Sci. 2021, 56, 16223–16242. [Google Scholar] [CrossRef]
  158. de Lima, I.S.; Sousa, H.R.; Silva, A.S.; de Oliveira, L.H.; Muniz, E.C.; Osajima, J.A.; Silva-Filho, E.C. Superabsorbent hydrogel from cassava gum (Manihot esculenta) to release water and macronutrients. Ind. Crops Prod. 2024, 219, 119045. [Google Scholar] [CrossRef]
  159. Al Rohily, K.; El-Hamshary, H.; Ghoneim, A.; Modaihsh, A. Controlled Release of Phosphorus from Superabsorbent Phosphate-Bound Alginate-Graft-Polyacrylamide: Resistance to Soil Cations and Release Mechanism. ACS Omega 2020, 5, 32919–32929. [Google Scholar] [CrossRef]
  160. Shaghaleh, H.; Alhaj Hamoud, Y.; Xu, X.; Wang, S.; Liu, H. A pH-responsive/sustained release nitrogen fertilizer hydrogel based on aminated cellulose nanofiber/cationic copolymer for application in irrigated neutral soils. J. Cleaner Prod. 2022, 368, 133098. [Google Scholar] [CrossRef]
  161. Chamorro, A.F.; Palencia, M.; Arrieta, Á.A. Development of High-Efficiency Fertilizer by Hydrogels Obtained from Cassava Starch and Citric Acid for Slow Release of Ammonium and Potassium. Gels 2024, 10, 434. [Google Scholar] [CrossRef]
  162. Sempeho, S.I.; Kim, H.T.; Mubofu, E.; Hilonga, A. Meticulous Overview on the Controlled Release Fertilizers. Adv. Chem. 2014, 2014, 363071. [Google Scholar] [CrossRef]
  163. Vo, P.T.; Nguyen, H.T.; Trinh, H.T.; Nguyen, V.M.; Le, A.-T.; Tran, H.Q.; Nguyen, T.T.T. The nitrogen slow-release fertilizer based on urea incorporating chitosan and poly(vinyl alcohol) blend. Environ. Technol. Innov. 2021, 22, 101528. [Google Scholar] [CrossRef]
  164. Pogorzelski, D.; Filho, J.F.L.; Matias, P.C.; Santos, W.O.; Vergütz, L.; Melo, L.C.A. Biochar as composite of phosphate fertilizer: Characterization and agronomic effectiveness. Sci. Total Environ. 2020, 743, 140604. [Google Scholar] [CrossRef] [PubMed]
  165. Jiang, S.; Duan, Q.; Ma, L.; Song, Y.; Xie, H.; Liu, H.; Chen, L.; Yu, L. Preparation and characterization of slow-release fertilizer through coating acrylate epoxidized soybean oil. Environ. Technol. Innov. 2024, 34, 103626. [Google Scholar] [CrossRef]
Figure 1. Chemical composition of starch featuring units of amylose and amylopectin.
Figure 1. Chemical composition of starch featuring units of amylose and amylopectin.
Molecules 29 04835 g001
Figure 2. Schematic diagram of process of preparing hydrogel by starch graft polymerization in aqueous solution.
Figure 2. Schematic diagram of process of preparing hydrogel by starch graft polymerization in aqueous solution.
Molecules 29 04835 g002
Figure 3. Schematic diagram of hydrogels prepared by starch irradiation graft polymerization.
Figure 3. Schematic diagram of hydrogels prepared by starch irradiation graft polymerization.
Molecules 29 04835 g003
Figure 4. Schematic representation of procedure and principle of polyacrylamide grafted starch using γ-ray [82].
Figure 4. Schematic representation of procedure and principle of polyacrylamide grafted starch using γ-ray [82].
Molecules 29 04835 g004
Figure 5. Schematic diagram of reactive extrusion to prepare starch-based materials.
Figure 5. Schematic diagram of reactive extrusion to prepare starch-based materials.
Molecules 29 04835 g005
Figure 6. Schematic representation of dual-stage and dual-initiator reactive extrusion systems [97].
Figure 6. Schematic representation of dual-stage and dual-initiator reactive extrusion systems [97].
Molecules 29 04835 g006
Figure 7. A schematic representation of the relationship between starch-based SAP characteristics and the slow-release behaviors of the double-coated fertilizer [114].
Figure 7. A schematic representation of the relationship between starch-based SAP characteristics and the slow-release behaviors of the double-coated fertilizer [114].
Molecules 29 04835 g007
Figure 8. Schematic diagram of hydrogel absorption swelling and fertilizer release process.
Figure 8. Schematic diagram of hydrogel absorption swelling and fertilizer release process.
Molecules 29 04835 g008
Figure 9. SEM images of radiation-induced in situ synthesis of starch-based hydrogels under diverse conditions. S-1–S-4, S-8: AM–starch = 1:2, water–starch = 10:1, irradiation doses are 3 kGy, 6 kGy, 9 kGy, 12 kGy, and 24 kGy, respectively. S-9: AM–starch = 1:2, water–starch = 6:1; S-15: AM–starch = 1:2, water–starch = 15:1; S-16: AM–starch = 2:1, water–starch = 10:1; S-17: AM–starch = 2:2, water–starch = 10:1; S-20: AM–starch = 2:6, water–starch = 10:1, irradiation dose is similarly 12 kGy [82].
Figure 9. SEM images of radiation-induced in situ synthesis of starch-based hydrogels under diverse conditions. S-1–S-4, S-8: AM–starch = 1:2, water–starch = 10:1, irradiation doses are 3 kGy, 6 kGy, 9 kGy, 12 kGy, and 24 kGy, respectively. S-9: AM–starch = 1:2, water–starch = 6:1; S-15: AM–starch = 1:2, water–starch = 15:1; S-16: AM–starch = 2:1, water–starch = 10:1; S-17: AM–starch = 2:2, water–starch = 10:1; S-20: AM–starch = 2:6, water–starch = 10:1, irradiation dose is similarly 12 kGy [82].
Molecules 29 04835 g009
Figure 10. Illustrative scheme of structure of cellulose nanocrystals compounded with starch-based hydrogel [130].
Figure 10. Illustrative scheme of structure of cellulose nanocrystals compounded with starch-based hydrogel [130].
Molecules 29 04835 g010
Figure 11. Schematic representation of plant growth vs. released fertilizer for different fertilizers: A: raw fertilizer; B: slow-release fertilizer; C: controlled or smart fertilizer [9].
Figure 11. Schematic representation of plant growth vs. released fertilizer for different fertilizers: A: raw fertilizer; B: slow-release fertilizer; C: controlled or smart fertilizer [9].
Molecules 29 04835 g011
Table 1. Summary of starch-based SRF hydrogels.
Table 1. Summary of starch-based SRF hydrogels.
Starch/AdditivesFertilizerModificationPerformanceReferences
Cassava starch–PAA/natural rubber/PVAUreaCSt-g-PAA networks and linear NR/PVA blends In water: hydrogel wax coated urea (BHWCU) <60% within 24 h, in 7 days, 89.1%, 68.6%, and 41.5% in BHWCU/2:8, BHWCU/6:4, and BHWCU/9:1 formulations. In soil: hydrogel wax coated urea <68% in 20 days[141]
Cassava starch-g-poly(acrylonitrile)UreaCrosslinking of cassava starch and acrylonitrile using MBA applied as a coatingUncoated urea released 100% after 28 days in soil, coated urea 70% in 108 days[15]
Corn starch–Poly (AA-co-AM)Diammonium phosphateCrosslinking of starch, AA, and AM using MBA60% N release during 30 days in soil[142]
High-amylose maize starch-g-poly(sodium acid maleate-disodium maleate)PotassiumHigh-amylose maize starch graft copolymerization with disodium maleate, using KMnO4-NaHSO3 redox system, and covalently crosslinked with MBAThe release percentages of KNO3 and KH2PO4 were between 92 and 67% and 89 and 61% in 320 h[143]
Poly (N, N-dimethylacrylamide –maleic acid)–starchUreaCrosslinking of starch, N, N-dimethylacrylamide, and maleic acid using MBAUrea released from hydrogels DMSt1, DMSt2, and DMSt3 after 14,000 min was found as 80.2%, 45.5%, and 67.7% in well water and 81.5%, 46.6%, and 68.9%
in tap water
[144]
Sodium alginate–carboxymethyl starch sodium–polydopamineUreaSodium alginate and carboxymethyl starch sodium were compounded, and polydopamine (PDA) film was formed on its surface by self-polymerization In soil: >25 days[145]
Corn starch–castor oil superabsorbent–polyurethaneUreaStarch–castor oil SAP–polyurethane-coated urea in smooth rotating drumNitrogen-controlled release period of 60–150 days in water[146]
Starch carbamate–sodium alginate–SRFUreaStarch carbamate and sodium alginate through cationic ion crosslinkingIn water: 61.6% within 10 h and almost completely release >16 h. In soil: 58.5% of urea released within 25 days and exceeded 50 days for complete release [147]
Starch phosphate carbamate–PVA–stearic acidUreaStearic acid as inner coating layer and starch phosphate carbamate–PVA film crosslinked with citric acid as external layerIn water: 50.3% within 15 h, nearly complete release over 25 h. In soil: 46.6% was released within 20 d, extending to approximately 30 d[148]
Chitosan–starchUreaCrosslinking of starch and chitosan using glutaraldehyde 63.71% N release during seven days in soil[149]
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

Song, Y.; Ma, L.; Duan, Q.; Xie, H.; Dong, X.; Zhang, H.; Yu, L. Development of Slow-Release Fertilizers with Function of Water Retention Using Eco-Friendly Starch Hydrogels. Molecules 2024, 29, 4835. https://doi.org/10.3390/molecules29204835

AMA Style

Song Y, Ma L, Duan Q, Xie H, Dong X, Zhang H, Yu L. Development of Slow-Release Fertilizers with Function of Water Retention Using Eco-Friendly Starch Hydrogels. Molecules. 2024; 29(20):4835. https://doi.org/10.3390/molecules29204835

Chicago/Turabian Style

Song, Yue, Litao Ma, Qingfei Duan, Huifang Xie, Xinyi Dong, Huaran Zhang, and Long Yu. 2024. "Development of Slow-Release Fertilizers with Function of Water Retention Using Eco-Friendly Starch Hydrogels" Molecules 29, no. 20: 4835. https://doi.org/10.3390/molecules29204835

Article Metrics

Back to TopTop