Dimethylolurea as a novel slow-release nitrogen source for nitrogen

Jun 3, 2019 - Dimethylolurea as a novel slow-release nitrogen source for nitrogen leaching mitigation and crop production ... PDF (777 KB) · Go to Vol...
0 downloads 0 Views 1MB Size
Article Cite This: J. Agric. Food Chem. 2019, 67, 7616−7625

pubs.acs.org/JAFC

Dimethylolurea as a Novel Slow-Release Nitrogen Source for Nitrogen Leaching Mitigation and Crop Production Jinhui Yang,*,† Tai Liu,†,‡ Hongbin Liu,‡ Limei Zhai,‡ Man Wang,† Yonggang Du,† Yanxue Chen,† Cheng Yang,† Huining Xiao,§ and Hongyuan Wang*,‡ †

School of Materials Science and Engineering, Shijiazhuang Tiedao University, Shijiazhuang, Hebei Province 050043, China Key Laboratory of Non-point Source Pollution Control, Ministry of Agriculture, Institute of Agricultural Resources and Regional Planning, Chinese Academy of Agricultural Sciences, Beijing 100081, China § Department of Chemical Engineering, University of New Brunswick, Fredericton, NB E3B 5A3 Canada Downloaded via GUILFORD COLG on July 20, 2019 at 06:58:22 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Rapid hydrolysis of urea results in further fertilization frequency and excessive nitrogen (N) input. A modified urea, dimethylolurea (DMU), was synthesized in this study. The structure of the sample was characterized by Fourier transform infrared and nuclear magnetic resonance analysis, manifesting the formation of DMU. N release investigation confirmed that DMU enabling provided a gradual N supply. The N leaching experiment indicated that increasing the applied DMU significantly reduced the NH4+-N, NO3−-N, and total N leaching, compared with urea application alone. The application effect on maize and wheat was evaluated. The results revealed that singly applied DMU with 100% or 80% N input, irrespective of the amount, promoted crop yield and agronomic characteristic and N use efficiency (NUE) of maize and wheat, beyond urea with two split applications at the recommended rate. Thus, the potential availability of DMU was proven; this could be widely used in agricultural fields as a slow-release fertilizer. KEYWORDS: dimethylolurea, slow release, N leaching, yield, N use efficiency



INTRODUCTION The increasing world population requires higher food production, where fertilizer is one of the most important elements.1,2 Therefore, large quantities of chemical fertilizers, particularly in relation to the nitrogen (N) element, have been used in agriculture. At present, the major and traditional N fertilizer is urea in China. However, high N loss via volatilization, leaching, and greenhouse gases causes large economic losses and a series of adverse environmental problems due to the rapid hydrolytic characteristic.3 Meanwhile, the imbalance between the rates of N release from urea and the rates of N uptake by plant roots result in lower N use efficiency (NUE). In order to provide enough nutrients for the demands of plants, a split application of N fertilizer and a labor intensive fertilization method have been applied over the past decades.4 However, the lack of fertilizer application machines and the shortage of agricultural workers in China make it more and more difficult.5 The use of slow-release fertilizers (SRFs), which can release nutrients more slowly than commonly used fertilizers, is a promising approach to mitigate the aforementioned problem. SRFs supply N that aims to be better synchronized with crop demand, facilitating better N uptake and utilization by crop plants. Furthermore, more of the potential benefits from SRFs are decreasing the fertilizer loss rate, improving NUE, lowering application frequency, and minimizing negative environmental effects associated with over dosage, including reduced N volatilization loss as well as N leaching.6,7 Furthermore, SRFs can be sown with seeds at the same time, utilizing agricultural mechanization to reduce labor input and frequency of fertilization.8 Comprehensively, conventional SRFs have been © 2019 American Chemical Society

classified into three major categories according to the mechanism of slow release: organic compounds, water-soluble fertilizers with physical barriers, and inorganic compounds with low solubility.9 Organic compounds mainly consist of natural organic compounds (animal manure, sewage sludge, etc.) and synthetic organic compounds with low solubility, which generally include condensation products from urea and other materials. In the second major category, hydrophobic polymers, actually appearing as a physical barrier of the coated SRFs, encumber fertilizer dissolution. Inorganic compounds with low solubility, in terms of N and P fertilizer, fundamentally denote the metal ammonium phosphates, e.g. KNH 4 PO 4 and MgNH4PO4. With the recent technology boom, diversity, preparation materials, and methods of SRFs have developed rapidly. Cheng et al.10 prepared a multifunctional SRF by means of chemical synthesis with neutralized acrylic acid, urea, potassium persulfate, and N,N′-methylenebis (acrylamide), which possesses three-dimensional structure with excellent water conservation and slow N release properties. To reduce the high cost of polymer synthesis and improve the sustainability, biobased polymers derived from renewable biomass (castor oil, palm oil, lignin, feather, etc.), in place of nonsustainable petrochemicals, have served as coating materials of SRFs.11−14 Nevertheless, undesirable nutrient release characteristics caused by the inadequate hydrophobicity of biobased coating materials Received: Revised: Accepted: Published: 7616

March 4, 2019 May 25, 2019 June 3, 2019 June 3, 2019 DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry

this disadvantage. Numerous studies have presented the superiority of this N fertilizer mixture on more plants such as maize,36 wheat,37 and Brassica campestris L,38 expanding the application range of UF. As an intermediate in the synthesis of UF, dimethylolurea (DMU) is regarded as a simple compound or one kind of modified urea according to its specific molecular structure similar to that of urea,39 which has not served as an N fertilizer in agricultural fields. The less synthetic procedure provides DMU with a lower material cost than traditional UF fertilizer. Theoretically, it is thought that DMU has a longer release period of N in soil due to its more complex molecular structure compared to with urea. In addition, lower leaching loss of mineral N caused by lower availability of NH4+, in particular to NO3−-N, was decreased considerably under the use of SRFs, since the stimulation of nitrifying bacteria population depends on the release rate of NH4+.40,41 The application of DMU may be another viable approach to mitigate leaching losses of N. Therefore, it is imperative to examine the slow-release behavior of DMU as well as its performance on N leaching, and investigate its feasibility in crop cultivation experiments. Based on the above background information, the primary objectives of this present work were to (i) synthesize the target product of DMU; (ii) determine the N release rate of N fertilizer mixture including DMU and urea with seven concentrations; (iii) explore leaching N loss in the presence of DMU and urea with four different formulations in layered columns filled with agricultural soil; (iv) investigate the effect of three N application levels of DMU applied as a single basal fertilization and urea applied as a split fertilization on maize yield in soil-based pot; and the effects of two N application levels of DMU applied as a single basal fertilization and urea applied as a split fertilization on wheat yield under field conditions, respectively.

generates a potential limitation for their commercial applications.9 It is thus necessary to develop novel and ecofriendly biobased coating materials with superhydrophobic surfaces. Cottonseed oil and pig fat have been used as feedstock to prepare biopolymer coatings; the resulting biopolymer coating has increased surface roughness as well as decreased surface energy.15,16 Once the surface is immersed into water, a threephase solid−vapor−liquid interface occurs with an accompanying air membrane; as a result, water slowly permeated into the inner fertilizer core in vapor instead of the liquid phase and consequently reduced the nutrient release rate. For coated SRFs, nutrient release rate is based on the thickness of the coating layer; however, batch leaching of the fertilizers depends on highly water-soluble species after the breakup of the encapsulation.17 Melt mixing incorporated with the recrystallization and extrusion process was confirmed as a novel and better approach to produce SRFs by completely blending fertilizer with various matrixes, for instant, bentonite, clay, and montmorillonite, entrapping the hydrophilic filler (fertilizer) inside the hydrophobic materials (matrix).18−20 The mechanochemical reaction of various calcined layered double hydroxides of other inorganic compounds (Mg2Al-CO3 and Mg2Fe-CO3) and K2HPO4 was conducted by Borges et al.,21 where a large amount of P was immobilized, achieving the controlled release over a long period. In addition, used as a support material for delivering plant nutrients, biochar has excellent adsorption capacity due to coarse and highly porous microstructure, and attracted much attention.22 Biochar participated in production of SRFs in diverse manners: constituent of coating materials,23 matrix of chemical fertilizer,21 alternatives of N fertilizer stored in superabsorbent material,24 etc. In order to acquire SRFs with more advantages, the combination of production method, matrix, and fertilizer type in different formulations were employed to improve the application effect.25 Recent work by Wen et al.,26−28 who attempted to substitute microwave irradiation for an external heat source, showed that introduction of bentonite in matrixes derived from various materials in concert with an N source (NH4+-loaded biochar and urea). The products obtained not only effectively increased swelling ability, gel strength, mechanical properties, and thermal stability but also facilitated the growth of plants along with biodegradable behavior. Ureaformaldehyde (UF), one of the slow-release degradable N fertilizers (known as methylene urea as well), has been demonstrated to prolong availability of N throughout the growing season.29,30 As the condensation product of urea and formaldehyde, UF consisting of numerous polymers with various chain lengths (methylenediurea, dimethylenetriurea, trimethylenetetraurea, etc.) is hydrolyzed to NH4+ slowly by a limited number of microbial enzyme.31 To date, UF has been used widely in golf greens, lawns, greenhouses, and public parks.32 UF has been proven to increase NUE of plants and reduce volatilization as well as leaching losses of N in these fields.33 However, the N release rate of UF depends on the activity index (AI) referring to the proportion of N (faster N release rate with lower AI).34 Meanwhile, UF does not respond directly to the nutrient demand of plants and releases N at the same rate regardless of whether a plant is demanding more nutrients or none at all.35 As a result, N release from UF, especially these products with higher AI, has been often too slow and not synchronized with the N uptake by several plant species.33 In recent years, more research has been focused increasingly on the incorporation of urea and UF to diminish



MATERIALS AND METHODS

Experimental Site and Materials. All experiments in the present study were conducted at the Chinese Academy of Agricultural Sciences, Changping County, Beijing, China. The study site has a temperate and monsoonal climate. The soil in the study site is aquic cinnamon with light loam texture and widely distributed in the North China Plain. Its basic physical and chemical properties were bulk density = 1.3 g cm−3, pH = 8.3, soil organic carbon (SOC) = 7.9 g kg−1, total nitrogen (TN) = 0.8 g kg−1, total potassium (TK) = 16.8 g kg−1, total phosphorus (TP) = 0.87 g kg−1, cation exchange capacity (CEC) = 17.4 cmol kg−1, and electrical conductivity (EC) = 183.0 cmol kg−1. The soil used in leaching and potted maize experiments was collected from the plow layer, passed through a 2 mm sieve to remove stones, plant roots as well as residues, and then mixed into a homogeneous sample prior to use. Urea (46.4% N), used as N fertilizer, was purchased from PetroChina Limited Company (Ningxia, China). P fertilizer was super phosphate (12.0% P2O5, Jiuhuashan Limited Company. Anhui, China); K fertilizer was potassium chloride (60.0% K2O, China Fertilizer Limited Company. Beijing, China). The cultivars in the present study were “Jingdan 28” for maize and “Zhongmai 175” for wheat. DMU was prepared in our laboratory. And the urea (46.0% N), involved in synthesis of DMU, was supplied by Shijiazhuang Baipo Zhengyuan Chemical Fertilizer Limited Company (Shijiazhuang, China). Formaldehyde solution (37∼40% HCHO) was collected from Tianjin Fengchuan Chemical Reagent Technologies Limited Company (Tianjin, China). Preparation of DMU. A 300 mL portion of formaldehyde was poured into a 1000 mL three-necked flask equipped with mechanical stirring, and then, 150 g of urea was added into the formaldehyde solution under gentle stirring at 45 °C in a water bath. As stirring continued, the pH of the solution was controlled at 7.5∼8.5 with a certain amount of sodium hydroxide, while a circling line was 7617

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry incorporated in case of volatilization of formaldehyde. After 4 h, the reaction was completed. And then, the water phase obtained was filtrated after it was cooled to room temperature. The production was purified with 95% ethanol and dried to constant weight in a vacuum oven at 60 °C. Finally, the DMU was obtained and stored for further use. The preparation process of DMU is depicted in Scheme 1:

soil leachate at different layers in the profile, experimentally simulating four soil depths of 10, 20, 30, and 40 cm along with the soil profile, respectively. A simple water container was used to supply deionized water into the column to simulate leaching conditions. A layer of filter paper was placed in the upper pad to prevent leaching water from disturbing the surface soil. A 5 cm layer of quartz sand and paper−nylon mesh was placed at the bottom of each soil column to further prevent the underlying soil erosion.44 The holes are well plugged when not in use. Before the N leaching experiment, 4.4 kg prepared soil was packed carefully into different column sections, corresponding to the original soil bulk density (about 1.3 g cm−3) of each layer in the field. In addition, mixtures of N fertilizer were mixed with the top 10 cm soil in the column physically and thoroughly. The soil on the edge of the soil column wall was compacted to eliminate edge effects. The N fertilizer application rate was 0.15 g N kg−1 soil, equal to 250 kg N ha−1. Five different N fertilizer formulations with three replicates were arranged in a random design: (1) CK (unfertilized), (2) D0U1 (urea-N only), (3) D1U3 (DMU-N:urea-N = 1:3), (4) D1U1 (DMU-N:urea-N = 1:1), and (5) D1U0 (DMU-N), respectively. The leaching experiment was conducted at 25 ± 2 °C in an artificial greenhouse controlled with a relative humidity of 65%. During the leaching period, approximately 100 mL of distilled water was added every 3 days from four openings along each soil column at 10, 20, 30, and 40 cm, respectively. All the sampling pots were sealed when not sampled. The leachate samples at different depths in the soil column profile were collected through the sampling ports, respectively, and the total volume of leachate was measured. Leachate was sampled at an interval of 10 days. The leachate samples were filtered through a 0.45 mm disposable pore-size filters (Whatman, Clifton NJ, USA) and analyzed for NO3−-N, NH4+-N, concentration by means of a flow injector auto analyzer (Auto Analyzer 3, High Resolution Digital Colorimeter, Germany) and total N following the description in N release experiment. The soil in all layers was sampled at the end of the experiment. The residual soil inorganic N (NO3−-N and NH4+-N) were measured using a flow injector auto analyzer (Auto Analyzer 3, High Resolution Digital Colorimeter, Germany) after extraction with 0.01 mol L−1 CaCl2 for 30 min at a soil to solution mass ratio of 1:10. Residual TN was detected by a Kjeldahl analyzer (KDY-9830, China) after H2SO4 digestion. Potted Maize Experiment. The experiments were performed with five treatments (Table 1), including PK, NuPK, NdPK1, NdPK2, and NdPK3. The height of the pots was 0.44 m with a round area of approximately 0.07 m2. The pots were placed randomly with three replications for each treatment. The N source in NdPK1, NdPK2, and NdPK3 was DMU, whereas urea was applied in NuPK. Fertilizers required were calculated according to the investigation of local N management (150 kg N ha−1) and the mass of soil used before application. A 150 mg P2O5 kg−1 soil as superphosphate, 100 mg K2O kg−1 soil as potassium chloride, and controlled TN at rates of 0, 100, 200, 160, and 120 mg N kg−1 soil (referred to as PK, NuPK, NdPK1, NdPK2, and NdPK3, respectively) were mixed with 8 kg soil and then

Scheme 1. Synthesis Route of DMU

Structure Characterization of DMU. FT-IR Analysis. The spectroscopic characterization of the DMU was determined by a FTIR spectrometer (Nicolet iS10, Thermofisher). The sample DMU was completely dried, ground to fine powder, mixed thoroughly with KBr, and pressed forming KBr tablets, successively. The FT-IR spectrum of the sample was recorded in the wavenumber range of 3500∼500 cm−1. NMR Analysis. The solid-state 1H NMR and 13C NMR spectra were recorded at 25 °C on a 900 MHz AVANACIII NMR spectrometer equipped with 3.2 mm triple resonance MAS probe (Bruker BioSpin, Rheinstetten, Germany), using D2O as a solvent and tetramethylsilane (TMS) as an internal standard. N Release Experiment. The nitrogen release experiment was carried out according to the procedure reported by Duan et al.42 Seven treatments with same the dosage of N (2.3 g N L−1) and different N sources formulations, i.e., 100% urea-N (D0), 50% DMU-N and 50% urea-N (D50), 60% DMU-N and 40% urea-N (D60), 70% DMU-N and 30% urea-N (D70), 80% DMU-N and 20% urea-N (D80), 90% DMUN and 10% urea-N (D90), and 100% DMU-N (D100), were designed with three replications. Sealed plastic cups filled with N fertilizer mixed with 100 mL of deionized water were incubated in a sterilized incubated chamber (25 °C). Solutions in plastic cups were filtered through disposable filters of 0.45 mm pore-size to determine the TN concentration in the filtrates using an elemental analyzer (vario PYRO cube, Germany). After that, solid residues on filters were washed in another plastic cup with 100 mL of deionized water for further incubation. Samples were collected at an interval of 2 days (first week), 7 days (2nd−7th weeks), and 14 days (8th−10th weeks) for the duration of the experiment as described above. Nitrogen release rate (%) is presented according to

cumulative N content released × 100% added N content (1) Leaching Experiment. The multilayer soil column device was constructed based on the previous method43 as shown in Figure S1. Each column with 10 cm inner diameter, and 52 cm length, consisted of four separated sections, which were well coated with vaseline, joined, and sealed throughout the experiment. There were four sampling holes and a small tap drilled in the middle sidewall of each section to extract nitrogen release rate =

Table 1. Experimental Design of the Potted Maize and Field Wheat treatments maize

wheat

N fertilizer type

PK NuPK

urea

NdPK1 NdPK2 NdPK3

dimethylolurea dimethylolurea dimethylolurea

control UNstd DNstd DNcsv

N fertilizer application rate no N fertilizer applied broadcast at planting (50%) and elongation stages (50%) broadcast at planting (100%) broadcast at planting (100%) broadcast at planting (100%)

urea dimethylolurea dimethylolurea

N supplied (mg kg−1)

P2O5 supplied (mg kg−1)

K2O supplied (mg kg−1)

0 200

150 150

100 100

200 160 120

150 150 150 (kg ha−1)

(kg ha−1)

100 100 100 (kg ha−1)

0 150 150 120

112.5 112.5 112.5 112.5

75 75 75 75

no N fertilizer applied broadcast prior to sowing (50%) and elongation stages (50%) broadcast prior to sowing (100%) broadcast prior to sowing (100%) 7618

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry added into the corresponding pot at basal stage before sowing. In NuPK, another 100 mg N kg−1 soil in the form of urea was applied at elongation stages. Six maize seeds were sowed, and the three best strains were selected and retained in each pot. Additional field management was carried out following the local agricultural practice. The maize in the pot was harvested after three months and kept separately for each pot. The maize plants, heights of which were recorded with a ruler, were separated to obtain grain, stem, and leaf samples. Fresh samples were oven-dried at 105 °C for 1 h and then kept at 80 °C until a constant weight was reached to determine the dry matter accumulation of organ and plant. Grain and straw yields were multiplied by the grain and straw N/P/K concentration, respectively, to calculate plant N/P/K accumulation (N/P/K uptake), as described by Salvagiotti et al.45 The dried organ samples were milled and sifted through a 0.5 mm screen. After wet digestion with H2SO4−H2O2,46 the TN, TP, and TK of the plants were determined using a Kjeldahl analyzer (KDY-9830, China), a continuous-flow analyzer (Auto Analyzer 3, High Resolution Digital Colorimeter, Germany), and an atomic absorption spectrometer (ZCA-1000, China), respectively. Nitrogen use efficiency-NUE47 caused by N fertilizer addition can be written as

NUE =

NUnf − NUck × 100% NAR

Figure 1. FT-IR spectrum of DMU.

spectrum, indicating the complete participation of urea in reaction. The strong absorption peak at 3339.69 cm−1 is N−H in −NH−, which is caused by the adjacent carbonyl coupling effect. The peaks at 2960.53 and 3010.67 cm−1 are −CH2−, and the peak at 1389.88 cm−1 denotes stretching vibration of C−H in −NHCH2−, demonstrating the existence of DMU. NMR Spectrum. The 1H NMR spectrum of sample in D2O is shown in Figure S2. It can be seen that 4.65 ppm is assigned to the solvent residual peak of D2O, and 4.52 ppm is the signal of −CH2−. −OH and −NH− disappeared in the spectrum because the active −H was contained. Figure S3 is the spectrum of 13C NMR of the synthesized sample. From the spectrum, 159.49 and 63.88 ppm are ascribed to the carbonyl group and −CH2−, respectively. Consistency with the standard 13C NMR spectrum of DMU (Figure S4), acquired in Spectral Database for Organic Compounds SDBS (Japan), fully confirms that the sample was indeed the target compound. N Release Characteristics. The N cumulative release curve of different formulations presented a variable release trend (Figure 2). D100 showed nearly 14.8% and 51.7% N release rate after 1 day and 70 days incubation, respectively, which suggests its slow release capacity. Gradual release behavior of DMU can

(2)

Where NUnf and NUck represent N uptake of plant in treatments with N fertilizer addition and N uptake of plants in treatment with zero nitrogen application, respectively; NAR is N application rate. 2.5. Field Wheat Experiment. The field experiments were conducted for six months and 20 days and then harvested, comprising four treatments of N fertilizer as depicted in Table 1 (control, no nitrogen fertilizer; UNstd, standard urea-N fertilization rate, 150 kg N ha−1; DNstd, standard DMU-N fertilization rate, 150 kg N ha−1; DNcsv, conservation DMU-N fertilization rate, 120 kg N ha−1). Each treatment was replicated three times. For the experiment, 12 rectangular plots (length 2 m; width 3 m) were arranged evenly in two parallel rows (each row with 6 plots) perpendicular adjacent plots were separated by concrete walls (depth 100 cm; width 10 cm) with 20 cm above the ground. In all treatments, P fertilizer was applied at 112.5 kg P2O5 ha−1 and K fertilizer at 75.0 kg K2O ha−1. As a basal fertilization, both of these two fertilizers were broadcast and then incorporated into 0−20 cm soil layers before sowing wheat. DMU was applied as basal N fertilizers, whereas urea in UNstd was added in two applications as both the basal and supplementary N fertilizer, which was applied to the surface by hand and then immediately incorporated into the ploughed layer with irrigation water. Seeds of wheat were sown at the rate of 180 kg ha−1. All treatments were subjected to identical management in order to minimize baseline heterogeneities among field plots. Ten ears were collected randomly from each plot and saturated grain number per ear was determined by hand. The ear of plant was measured with a ruler with three replicates. The 1000-grain weight was determined using an automatic seed-counting system (Tuopu Co., Ltd., Zhejiang, China). Grain yield, N accumulation, and NUE were determined in terms of the methods mentioned above. Data Statistical Analysis. The results in N release, N leaching, potted maize, and field wheat experiments were expressed means and standard deviations. Analysis of variance (one-way ANOVA) with means compared by least significant difference (LSD) calculations at P = 0.05 was performed to identify statistically significant differences among different treatments, using the SPSS 19.0 (Inc., Chicago, IL, USA) software package for Windows. All figures were processed by Origin 8.0 (Origin Lab Corporation, USA). Any differences between the mean values at P < 0.05 were considered statistically significant.



RESULTS Characterization. FT-IR Analysis. Figure 1 shows the FT-IR spectrum of DMU. The absorption peak at 3200−3400 and 998.95 cm−1 are attributed to the stretching and bending vibration of −OH, respectively. The characteristic absorption peaks of N−H at 3434 and 3344 cm−1 in urea disappear in the

Figure 2. Temporal change of nitrogen release rate subjected to various N treatments. (D100, 100% dimethylolurea-N; D90, 90% dimethylolurea-N and 10% urea-N; D80, 80% dimethylolurea-N and 20% urea-N; D70, 70% dimethylolurea-N and 30% urea-N; D60, 60% dimethylolurea-N and 40% urea-N; D50, 50% dimethylolurea-N and 50% urea-N; D0, 100% urea-N, respectively). Values are means ± 1 SD (n = 3). 7619

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry

Figure 3. Cumulative mass of (a) NO3−-N, (b) NH4−-N, and (c) TN in the leachates from all four layers of the soil columns subjected to various N treatments. (CK unfertilized, D0U1 urea-N, D1U3 dimethylolurea-N:urea-N = 1:3, D1U1 dimethylolurea-N:urea-N = 1:1, and D1U0 dimethylolurea-N, respectively). Values are means ± 1 SD (n = 3). Significant differences are indicated by different letters (P < 0.05).

Figure 4. Residual soil (a) NO3−-N and (b) NH4−-N content in all four layers of the soil columns after incubation subjected to various N treatments. (CK unfertilized, D0U1 urea-N, D1U3 dimethylolurea-N:urea-N = 1:3, D1U1 dimethylolurea-N:urea-N = 1:1, and D1U0 dimethylolurea-N, respectively). Values are means ± 1 SD (n = 3). Significant differences are indicated by different letters (P < 0.05).

Table 2. TN of Residual Soil in Four Different Layers TN (g kg−1)b a

layers (cm) treatments

CK

D0U1

D1U3

D1U1

D1U0

0−10 10−20 20−30 30−40

1.28 ± 0.05 1.29 ± 0.02 1.28 ± 0.01 1.27 ± 0.01

1.34 ± 0.01 1.29 ± 0.05 1.29 ± 0.06 1.27 ± 0.01

1.35 ± 0.01 1.32 ± 0.03 1.31 ± 0.03 1.30 ± 0.03

1.37 ± 0.01 1.32 ± 0.04 1.32 ± 0.02 1.30 ± 0.02

1.38 ± 0.04 1.34 ± 0.07 1.33 ± 0.04 1.31 ± 0.06

a

Treatments were CK (unfertilized), D0U1 (urea-N), D1U3 (dimethylolurea-N:urea-N = 1:3), D1U1 (dimethylolurea-N:urea-N = 1:1), and D1U0 (dimethylolurea-N). The N application rate in D0U1, D1U3, D1U1, and D1U0 was 0.15 g kg−1 soil. bValues are presented as means ± standard deviation with n = 3.

to various N treatments are shown in Figure 3. NO3−-N dominated TN leaching losses in all treatments, expressed as the magnitude of leached N in form of NO3−-N was obviously higher than that of NH4+-N. The application of DMU altered the amount of leached NH4+-N, NO3−-N, and TN, which was decreased with increasing DMU-N rate in N source. In detail, the cumulative amount of leached NH4+-N, NO3−-N, and TN in these four layers was reduced by 15.8%, 4.1%, and 3.6% for D1U3, 14.5%, 12.1%, and 11.3% for D1U1, 20.1%, 17.4%, and 20.9% for D1U0 significantly (P < 0.05), when compared with D0U1 (urea only). The continuous addition water modified the profile of vertical NH4+-N, NO3−-N, and total N distribution in the soil. After

be identified clearly with the increase of cumulative N content up to the 70th day, and it is emphasized that augmentation of DMU content extended the N release stage. In comparison, the pattern of temporal release from the N source with 100% DMU was almost slow and linear (51.7% at the 70th day), whereas in D0 containing pure urea as the N source, release had stopped at the second sampling (3 days after incubation), ended with a negligible N release stage (4−70 day). In summary, addition of DMU delayed the N release process, deriving from the steeper N release curve in corresponding treatment with increasing N content in form of DMU (Figure 2). Nitrogen Leaching. The cumulative masses of NH4+-N, NO3−-N, and TN in the leachates from all four layers subjected 7620

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry

Table 3. Plant Height (PH), Straw Dry Weight (SDW), Grain Dry Weight (GDW) per Maize Plant under Different Treatments treatments

a

PHc,d

incrementc,d

cm 101.3 ± 7.8 c 108.0 ± 3.9 b 113.0 ± 7.4 a 114.6 ± 8.2 a 112.1 ± 5.3 a

PK NuPK NdPK1 NdPK2 NdPK3

SDWc,d

incrementb

GDWc,d

incrementb

%

g

4.6 6.1 3.8

27.2 ± 1.4 d 36.5 ± 5.9 c 46.6 ± 3.2 a 42.7 ± 5.4 b 42.2 ± 4.1 b

%

g

%

27.7 17.0 15.6

78.4 ± 2.3 e 135.9 ± 10.2 c 165.4 ± 12.4 a 151.4 ± 16.9 b 124.8 ± 11.0 d

19.5 11.4 −0.82

Treatments were PK, no nitrogen fertilizer application; NuPK, 200 mg kg−1 urea-N application rate; NdPK1, 200 mg kg−1dimethylolurea-N application rate; NdPK2, 160 mg kg−1 dimethylolurea-N application rate; NdPK3, 120 mg kg−1 dimethylolurea-N application rate, respectively. b Increase percentages are computed compared with NuPK. cValues are presented as means ± standard deviation with n = 3. dDifferent letters in a single row indicate significant difference between the treatments at P < 0.05 (Duncan’s multiple range test). a

Table 4. Total N/P/K Uptake (TNU, TPU, and TKU) and Nitrogen Use Efficiency (NUE) of Maize Plant under Different Treatments TNUc,d treatments

a

PK NuPK NdPK1 NdPK2 NdPK3

−1

mg hill

283.1 ± 6.9 d 677.0 ± 61.1 c 828.4 ± 29.0 a 737.3 ± 39.2 b 614.8 ± 40.1 d

NUEc,d

TPUc,d

incrementb

−1

%

mg hill

28.6 ± 2.6 d 39.5 ± 2.8 c 41.1 ± 3.4 a 40.3 ± 4.1 b

%

45.2 ± 4.6 d 64.1 ± 8.3 c 84.3 ± 5.1 a 75.2 ± 6.4 b 72.4 ± 7.8 c

31.5 17.3 12.9

TKUc,d

incrementb

−1

mg hill

351.9 ± 21.3 e 513.7 ± 29.1 d 699.6 ± 3.1 a 655.3 ± 30.1 b 638.4 ± 35.5 c

%

36.2 27.6 24.3

Treatments were PK, no nitrogen fertilizer application; NuPK, 200 mg kg−1 urea-N application rate; NdPK1, 200 mg kg−1 dimethylolurea-N application rate; NdPK2, 160 mg kg−1dimethylolurea-N application rate; NdPK3, 120 mg kg−1 dimethylolurea-N application rate, respectively. b Increase percentages are computed compared with NuPK. cValues are presented as means ± standard deviation with n = 3. dDifferent letters in a single row indicate significant difference between the treatments at P < 0.05 (Duncan’s multiple range test). a

Table 5. Saturated Grain Number per Ear (SGNPE), Ear Length (EL), 1000-Grain Weight (TGW), and Grain Yield (GY) of Field Wheat under Different Treatments ELc,d

TGWc,d

GYc,d

incrementb

treatmentsa

SGNPEc,d

cm

g

kg ha−1

%

control UNstd DNstd DNcsv

26.1 ± 1.5 c 28.7 ± 0.3 b 31.7 ± 1.2 a 31.4 ± 1.0 a

7.4 ± 0.5 d 7.7 ± 0.4 c 8.7 ± 0.7 a 8.5 ± 1.0 b

41.5 ± 1.4 c 43.8 ± 1.4 b 44.3 ± 0.6 a 43.9 ± 1.2 b

2363.3 ± 100.5 d 3895.0 ± 221.6 c 4425.5 ± 176.9 a 4275.0 ± 284.6 b

13.6 9.76

Treatments were control, no nitrogen fertilizer; UNstd, standard urea-N fertilization rate, 150 kg N ha−1; DNstd, standard dimethylolurea-N fertilization rate, 150 kg N ha−1; DNcsv, conservation dimethylolurea-N fertilization rate, 120 kg N ha−1. bIncrease percentages are computed compared with UNstd. cValues are presented as means ± standard deviation with n = 3. dDifferent letters in a single row indicate significant difference between the treatments at P < 0.05 (Duncan’s multiple range test). a

incubation, the contents of NH4+-N and NO3−-N were concentrated in the 20−40 cm soil layers (Figure 4), but the opposite trend was found in residual soil total N, which decreased with increasing depth (Table 2). The highest NH4+-N and NO3−-N content occurred in D1U0, followed in descending order by D1U1, D1U3, and D0U1. Meanwhile, the TN content of D1U3, D1U1, and D1U0 were higher than that of D0U1 over all soil layers. TN content was elevated with increasing rates of DMU application. These results indicated that better results in mitigated N leaching, especially for NO3−-N, was achieved when larger urea proportion was replaced by DMU. Agronomic Characteristics, Yield, and Nutrition Uptake in Maize. The yields of potted maize are expressed in centimeters of plant height (PH), grams of straw dry weight (SDW), and grams of grain dry weight (GDW) per plant, respectively (Table 3). N fertilizer application had significant interactions on the three indexes above (P < 0.05). NuPK, containing the same N dosage with NuPK, attained the highest SDW (46.6 ± 3.2 g) and GDW (165.4 ± 12.4 g) of maize, higher than those in NuPK for 27.7% and 19.5%. NdPK1, 20% N

application lower than NuPK, also increased PH, SDW, and GDW significantly (P < 0.05). However, there is a negative result on GDW of NdPK3, at 60% N rates of NuPK, which elevated PH and SDW significantly when compared with NuPK (P < 0.05). As shown in Table 4, increasing N rates under DMU treatment, except NdPK3, increased N/P/K uptake significantly over NuPK (P < 0.05). Compared with NuPK, NdPK3 (with 80% N application rate) still showed a significant increase in P/K uptake (12.9% and 24.3%, respectively). There is a converse result of N uptake in NdPK3, the highest NUE (41.1%) was observed due to relatively lower N import among all treatments with either urea or DMU (P < 0.05). Agronomic Characteristic, Yield, and Nutrition Uptake in Wheat. The effects on plant yield, including saturated grain number per ear (SGNPE), ear length (EL), 1000-grain weight (TGW), and grain yield (GW), were varied with type and level of N fertilization (Table 5). The application of DMU in DNstd increased wheat yield significantly compared with wheat plots with two-split applications of urea (UNstd) (P < 0.05), with 7621

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry

Table 6. Dry Matter Weight (DMW), Total N Uptake (TNU) of Straw and Grain Nitrogen Use Efficiency (NUE) of Plants under Different Treatments DMWb,c straw treatments control UNstd DNstd DNcsv

a

kg ha

−1

3595.0 ± 246.1 d 5080.0 ± 123.5 b 5758.3 ± 197.8 c 5591.7 ± 364.2 a

TNUb,c grain −1

kg ha

2363.3 ± 100.5 e 3895.0 ± 221.6 d 4425.5 ± 176.9 a 4275.0 ± 284.6 b

straw −1

g kg

21.4 ± 1.2 b 23.4 ± 2.8 a 22.1 ± 0.6 a 23.0 ± 2.1 a

grain −1

g kg

1.4 ± 0.3 d 1.5 ± 0.1 c 1.9 ± 0.3 a 1.6 ± 0.2 b

total kg ha

−1

83.2 ± 5.3 d 124.7 ± 6.4 c 135.9 ± 2.7 a 129.6 ± 7.9 b

NUEb,c % 34.6 ± 1.8 c 41.8 ± 4.1 a 37.8 ± 2.6 b

a Treatments were control, no nitrogen fertilizer; UNstd, standard urea-N fertilization rate, 150 kg N ha−1; DNstd, standard dimethylolurea-N fertilization rate, 150 kg N ha−1; DNcsv, conservation dimethylolurea-N fertilization rate, 120 kg N ha−1. bValues are presented as means ± standard deviation with n = 3. cDifferent letters in a single row indicate significant difference between the treatments at P < 0.05 (Duncan’s multiple range test).

overall mean values of 31.7, 8.7 cm, 44.3 g, and 4425.5 kg ha−1 for SGNPE, EL, TGW, and GW, respectively. These parameters except for TGW from DNcsv, at 80% N fertilization rate (120 kg N ha−1), were elevated by 9.4%, 10.4%, and 9.8% over those in the UNstd treatment, respectively (P < 0.05). No significant differences in TGW between UNstd and DNcsv were observed. The DMU application at 150 kg N ha−1, produced significantly higher dry matter weight (DMW), TNU and NUE than UNstd (P < 0.05) (Table 6). This further confirms that DMU is definitely superior to widely use in wheat field at the same N input. The TNU and NUE in UNstd, where the urea was broadcast in two stages, were lower than those obtained in DNcsv under 120 kg N ha−1 rate, for 4.9 kg ha−1 and 3.2%, respectively.

due to the driving force of water within soil.55 As reported by previous authors, SRF potentially can reduce N leaching because of the nature of the N release pattern over a longer period of time compared with traditional N source products such as urea.56 Figure 2 has provided the evidence regarding the slow-release behavior of DMU. In addition, treatments with DMU incorporation showed significantly lower NH4+-N, NO3−-N, and TN leaching loss than D0U1, implying that the application of DMU could reduce N leaching loss. Residual NH4+-N and NO3−-N, though concentrated in deeper layers, was found to be higher in the presence of DMU with increasing percentage. Conversely, TN after incubation gave a positive response to the dosage of DMU with respect to the same layer, particular in upland soils. These results were presumed to favor its slow release capacity as well. Effect of DMU Application on Crop Production. Numerous studies have demonstrated clearly that crop yield was undisputedly affected by N rates,57 N sources,58 and N management.59 When considering large-scale implementation, the performance of DMU in agricultural fields, a more reliable indicator, must be considered simultaneously in addition to slow and gradual release behavior of N.60 In other words, it is necessary to confirm whether the DMU application was sufficient to meet the crop needs. DMU basally applied irrespective of N level, facilitating potted maize and field wheat growth effectively when compared with urea with two split adoption, and the benefits varied from DMU application rate to the harvest or agronomic parameters, in spite of singly applied DMU with 60% N application rate (NdPK3) yielding a lower GDW. Increases of DMU rates were accompanied by increasing SDW, GDW, TNU, TPU, and TKU of maize (Table 2). In this study, NdPK2 with reduced N rate (160 mg kg−1) by one-fifth achieved the increasingly higher yield than NuPK (200 mg kg−1) with two split N fertilizer applications (P < 0.05). A similar impact was observed in GY and TNU of field wheat (Tables 5 and 6). Conceivably, split application of N urea in NuPK was more costly in labor input than one-time fertilization carried out in NdPK1, NdPK2, NdPK3 (maize), DNstd, and DNcsv (wheat). NUE in DMU treatments was boosted, compared with the treatments containing pure urea as the N source. Rapid transform of urea provided excessive N availability at early stages, which was not consistent with peak N uptake stage, resulting in limited supply of N to the plants and consequently decreased crop yield. Finally, the excessive N chemical fertilizer input became a common practice aiming at promoting crop yield. Meanwhile, the slow hydrolysis process from DMU, caused relatively higher NUE as well as productivity because of reductive N leaching and successive N supplement in latter



DISCUSSION Effect of DMU on N Release Characteristics. Synchronized fertilizer input with the crop needs is of great importance in crop production.48 Fertilizer can synchronize nutrient release with the crop growth, meanwhile, maximizing crop yield and minimizing environmental impact with decreased N fertilization rate.49 The results indicated that DMU does present a slow and gradual release behavior and the increase of the DMU ratio in synchronized fertilizer could reduce the N release rate. The N release rate in combination with urea and DMU (CUD) can be adjusted by changing the ratio of DMU to suit the N requirements of different crops, which have diverse N uptake peak stage, accumulative N demand and growth period. In addition, CUD in varying proportions adjusts and controls N release rate (Figure 2). Thus, CUD with various ratios can be adopted as effective alternatives of controlled release urea (CRU), designed to enhance crop yields by synchronizing N supply with plant demand rationally. Studies on the traits of CRU performance on crop yield and soil N supply were conducted in various crop systems: cotton,50 rice,51 corn,52 and potato53 as well as soil nitrate leaching. As a result, it has been proven that the release rate curves of CRU were synchronized with N requirements of crop. The application of CUD in crop, which would be an advantage from the agricultural production point of view, is worthy of development and research in the future. Effect of DMU Application on N Leaching. In China, excessive N chemical fertilizer input for increasing crops yield and inappropriate application methods lead to high N losses through leaching.49 However, N leaching is affected simultaneously by the soil, environmental, and management conditions.54 NO3−-N was the major form of N leaching in the present study, similar to the effect of that heavy rainfall and excessive irrigation resulting in NO3−-N leaching in croplands 7622

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Journal of Agricultural and Food Chemistry



growth period. In agreement with other SRFs, DMU possibly alleviated the risk of N losses through ammonia (NH3) volatilization, nitrification, and denitrification, which depends on the level of N released. Further attempts are worthy of investigation for NH3 and green house gas emission once DMU is applied. The use of DMU indicated better economic and environmental effects than the use of traditional urea, so it is reasonable to conclude that DMU has real potential as a novel slow release N supplier in agriculture fields. Analysis of N Slow Release Mechanism and Biodegradable Behavior. In the N release experiment, N release rate was obtained by determining the total N dissolved in water, which was mainly related to the solubility. As a result, N slow release mechanism of DMU is attributed to its slight solubility (Figure 2). Meanwhile, Figure S5 suggests −H is substituted by −CH2OH in DMU compared with urea. The retardant release of N from DMU in soil, in addition to the dissolution characteristic, was assigned to the better stability of C−N bond, resulting from strengthened electron cloud density caused by the electron-donating group of −CH2OH possibly. Jahns et al.31 found that UF of different chain lengths can be hydrolyzed to NH4+, formaldehyde, urea, and carbon dioxide by a bacterium called Ochrobactrum anthropi. Except for two −OH groups, the rest of molecular skeleton of DMU constitute the constitutional unit of repeated portion in UF (Figure S5). Therefore, it is believable that DMU was biodegradable, in spite of the lack of direct evidence in the lab results in the current work. Economic Evaluation of SRFs. The production cost is considered as a major concern limiting the wide application of SRFs. Therefore, the economic evaluation of MDU is necessary to conduct. As shown in Table S1, the production cost of DMU was approximately $386.4−412.5 ton−1, increasing by 25.1− 38.9% compared with that of urea. However, N slow-release behavior was observed in DMU synthesized in our laboratory in comparison with urea. On the other hand, the increasing proportion of DMU-N, at the same N dosage, generated less N loss in leaching process, which is beneficial to environment protection and resource conservation. The application of DMU only with single fertilization, even at a lower N application rate, significantly facilitates crop growth, alleviating the labor investment of farmer in the field of plant cultivation. Therefore, the advantages of DMU above are superior to its production cost.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Phone number: +86 1733293-1675 (J.Y.). *E-mail: [email protected]. Phone number: +86 1860028-2261 (H.W.). ORCID

Jinhui Yang: 0000-0003-1955-3134 Huining Xiao: 0000-0003-3500-2308 Funding

This research was supported via the National Key Research and Development Program of China [2016YFD0800500, 2018YFD0200200]; the Key Laboratory of Nonpoint Source Pollution Control, Ministry of Agriculture, Beijing [1610132018031, 1610132018025]. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS We are grateful to all staff and partners participating in the study. ABBREVIATIONS USED SRFs, slow-release fertilizers; UF, urea formaldehyde; AI, activity index of UF, denoting the available content of UF in fertilizer; DMU, dimethylolurea, the intermediate to synthesize UF, its molecular structure is presented in Figure S5; CUD, combination of urea and dimethylolurea



REFERENCES

(1) Bhardwaj, D.; Sharma, M.; Sharma, P.; Tomar, R. Synthesis and surfactant modification of clinoptilolite and montmorillonite for the removal of nitrate and preparation of slow release nitrogen fertilizer. J. Hazard. Mater. 2012, 227, 292−300. (2) Wen, P.; Han, Y.; Wu, Z.; He, Y.; Ye, B. C.; Wang, J. Rapid synthesis of a corncob-based semi-interpenetrating polymer network slow-release nitrogen fertilizer by microwave irradiation to control water and nutrient losses. Arabian J. Chem. 2017, 10, 922−934. (3) Nardi, P.; Neri, U.; Matteo, G. D.; Trinchera, A.; Napoli, R.; Farina, R.; Subbarao, G. V.; Benedetti, A. Nitrogen release from slowrelease fertilizers in soils with different microbial activities. Pedosphere 2018, 28, 332−340. (4) Cui, Z. L.; Zhang, H. Y.; Chen, X. P.; Zhang, C. C.; Ma, W. Q.; Huang, C. D.; Zhang, W. F.; Mi, G. H.; Miao, Y. X.; Li, X. L.; et al. Pursuing sustainable productivity with millions of smallholder farmers. Nature 2018, 555, 363−366. (5) Tang, Y.; Li, X.; Shen, W. Z.; Duan, Z. Q. Effect of the slow-release nitrogen fertilizer oxamide on ammonia volatilization and nitrogen use efficiency in paddy soil. Agronomy 2018, 8, 53. (6) Chalk, P. M.; Craswell, E. C.; Polidoro, J. C.; Chen, D. L. Fate of efficiency of 15N-labelled slow-and controlled-release fertilizers. Nutr. Cycling Agroecosyst. 2015, 102, 167−178. (7) Solihin; Zhang, Q.; Tongamp, W.; Saito, F. Mechanochemical route for synthesizing KMgPO4 and NH4MgPO4 for application as slow-release fertilizers. Ind. Eng. Chem. Res. 2010, 49, 2213−2216. (8) Liu, J. L.; Yang, Y. C.; Gao, B.; Li, Y. C.; Xie, J. Z. Bio-based elastic polyurethane for controlled-release urea fertilizer: Fabrication, properties, swelling and nitrogen release characteristics. J. Cleaner Prod. 2019, 209, 528−537. (9) Azeem, B.; KuShaari, K.; Man, Z. B.; Basit, A.; Thanh, T. H. Review on materials & methods to produce controlled release coated urea fertilizer. J. Controlled Release 2014, 181, 11−21. (10) Cheng, D. D.; Liu, Y.; Yang, G. T.; Zhang, A. P. Water-and Fertilizer-Integrated Hydrogel Derived from the Polymerization of Acrylic Acid and Urea as a Slow-Release N Fertilizer and Water

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jafc.9b01432. Figure S1: Structural diagram of multilayer soil column device in leaching experiment. Figure S4: Standard 13C NMR spectrum of DMU acquired in the Spectral Database for Organic Compounds SDBS (Japan). Figure S5: the molecular structural formulas of urea, DMU and UF. Table S1: the production cost and corresponding calculating process of DMU (PDF) 1

H NMR spectrum of DMU synthesized (PDF)

13

Article

C NMR spectrum of DMU synthesized (PDF) 7623

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry Retention in Agriculture. J. Agric. Food Chem. 2018, 66 (23), 5762− 5769. (11) Liu, L.; Shen, T. L.; Yang, Y. C.; Gao, B.; Li, Y. C. C.; Xie, J. Z.; Tang, Y. F.; Zhang, S. G.; Wang, Z. H.; Chen, J. Q. Bio-based Large Tablet Controlled-Release Urea: Synthesis, Characterization, and Controlled-Released Mechanisms. J. Agric. Food Chem. 2018, 66 (43), 11265−11272. (12) Liu, J. L.; Yang, Y. C.; Gao, B.; Li, Y. C. C.; Xie, J. Z. Bio-based elastic polyurethane for controlled-release urea fertilizer: fabrication, properties, swelling and nitrogen release characteristics. J. Cleaner Prod. 2019, 209, 528−537. (13) Pang, W. C.; Hou, D. J.; Wang, H.; Sai, S.; Wang, B. R.; Ke, J. W.; Wu, G. B.; Li, Q.; Holtzapple, M. T. Preparation of microcapsules of slow-release NPK compound fertilizer and the release characteristics. J. Braz. Chem. Soc. 2018, 29 (11), 2397−2404. (14) Kong, W. J.; Li, Q.; Li, X. D.; Su, Y.; Yue, Q. Y.; Gao, B. Y. A biodegradable biomass-based polymeric composite for slow release and water retention. J. Environ. Manage. 2019, 230, 190−198. (15) Xie, J. Z.; Yang, Y. C.; Gao, B.; Wan, Y. S.; Li, Y. Y. C.; Xu, J.; Zhao, Q. H. Biomimetic Superhydrophobic Biobased PolyurethaneCoated Fertilizer with Atmosphere “Outerwear. ACS Appl. Mater. Interfaces 2017, 9 (18), 15868−15879. (16) Xie, J. Z.; Yang, Y. C.; Gao, B.; Wan, Y. S.; Li, Y. Y. C.; Cheng, D. D.; Xiao, T. Q.; Li, K.; Fu, Y. N.; Xu, J.; Zhao, Q. H.; Zhang, Y. F.; Tang, Y. F.; Yao, Y. Y.; Wang, Z. H.; Liu, L. Magnetic-Sensitive Nanoparticle Self-Assembled Superhydrophobic Biopolymer-Coated Slow-Release Fertilizer: Fabrication, Enhanced Performance and Mechanism. ACS Nano 2019, 13 (3), 3320−3333. (17) Noppakundilograt, S.; Pheatcharat, N.; Kiatkamjornwong, S. Multilayer-coated NPK compound fertilizer hydrogel with controlled nutrient release and water absorbency. J. Appl. Polym. Sci. 2015, 132 (2), 1−11. (18) Baldanza, V. A. R.; Souza, F. G.; Filho, S. T.; Franco, H. A.; Oliveira, G. E.; Caetano, R. M. J.; Hernandez, J. A. R.; Leite, S. G. F.; Sousa, A. M. F.; Silva, A. L. N. Controlled-release fertilizer based on poly(butylene succinate)/urea/clay and its effect on lettuce growth. J. Appl. Polym. Sci. 2018, 135, 46858. (19) Daitx, T. S.; Giovanela, M.; Carli, L. N.; Mauler, R. S. Biodegradable polymer/clay systems for highly controlled release of NPK fertilizer. Polym. Adv. Technol. 2019, 30 (3), 631−639. (20) Giroto, A. S.; Guimaraes, G. G.; Colnago, L. A.; Klamczynski, A.; Glenn, G.; Ribeiro, C. Controlled release of Nitrogen using UreaMelamine-Starch Composites. J. Cleaner Prod. 2019, 217, 448−455. (21) Borges, R.; Wypych, F.; Petit, E.; Forano, C.; Prevot, V. Potential Sustainable Slow-Release Fertilizers Obtained by Mechanochemical Activation of MgAl and MgFe Layered Double Hydroxides and K2HPO4. Nanomaterials 2019, 9 (2), 183. (22) Gwenzi, W.; Nyambishi, T. J.; Chaukura, N.; Mapope, N. Synthesis and nutrient release patterns of a biochar-based N-P-K slowrelease fertilizer. Int. J. Environ. Sci. Technol. 2018, 15 (2), 405−414. (23) Chen, S. L.; Yang, M.; Ba, C.; Yu, S. S.; Jiang, Y. F.; Zou, H. T.; Zhang, Y. L. Preparation and characterization of slow-release fertilizer encapsulated by biochar-based waterborne copolymers. Sci. Total Environ. 2018, 615, 431−437. (24) Baki, M.; Abedi-Koupai, J. Preparation and characterization of a superabsorbent slow-release fertilizer with sodium alginate and biochar. J. Appl. Polym. Sci. 2018, 135 (10), 45966. (25) Wen, P.; Wu, Z. S.; He, Y. H.; Han, Y. J.; Tong, Y. B. Characterization of p(AA-co-AM)/bent/urea and its swelling and slow release behavior in a simulative soil environment. J. Appl. Polym. Sci. 2016, 133 (12), 1. (26) Wen, P.; Wu, Z.; Han, Y.; Cravotto, G.; Wang, J.; Ye, B. C. Microwave-assisted synthesis of a novel biochar-based slow-release nitrogen fertilizer with enhanced water-retention capacity. ACS Sustainable Chem. Eng. 2017, 5, 7374. (27) Wen, P.; Wu, Z. S.; He, Y. H.; Ye, B. C.; Han, Y. J.; Wang, J.; Guan, X. Y. Microwave-assisted synthesis of a semi-IPN slow-release nitrogen fertilizer with water absorbency from cotton stalks. ACS Sustainable Chem. Eng. 2016, 4 (12), 6572−6579.

(28) Wen, P.; Wu, Z. S.; He, Y. H.; Ye, B. C.; Han, Y. J.; Guan, X. Y.; Wang, J. Microwave-assisted one-step synthesis and characterization of a slow release nitrogen fertilizer with inorganic and organic composites. RSC Adv. 2016, 6 (44), 37337−37346. (29) Koivunen, M. E.; Morisseau, C.; Horwath, W. R.; Hammock, B. D. Isolation of a strain of Agrobacterium tumefaciens (Rhizobium radiobacter) utilizing methylene urea (ureaformaldehyde) as nitrogen source. Can. J. Microbiol. 2004, 50, 167−174. (30) Steiner, C.; Glaser, B.; Geraldes Teixeira, W.; Lehmann, J.; Blum, W. E. H.; Zech, W. Nitrogen retention and plant uptake on a highly weathered central amazonian ferralsol amended with compost and charcoal. J. Plant Nutr. Soil Sci. 2008, 171, 893−899. (31) Jahns, T.; Kaltwasser, H. Mechanism of Microbial Degradation of Slow-Release Fertilizers. J. Polym. Environ. 2000, 8, 11−16. (32) Yang, Z. H.; Shi, Y. L.; Sun, Y. H.; Wang, L. L.; Guan, F. The study on biodegradation of methylene urea by activated sludge. Polym. Degrad. Stab. 2016, 128, 107−114. (33) Aarnio, T.; Mccullough, K.; Trofymow, J. A. Fate of urea and ureaformaldehyde nitrogen in a one-year laboratory incubation with douglas-fir forest floor. Soil Biol. Biochem. 1996, 28, 1407−1415. (34) Koivunen, M. E.; Horwath, W. R. Methylene urea as a slowrelease nitrogen source for processing tomatoes. Nutr. Cycling Agroecosyst. 2005, 71, 177−190. (35) Azeem, B.; Kushaari, K. Z.; Man, Z. B.; Basit, A.; Thanh, T. H. Review on materials & methods to produce controlled release coated urea fertilizer. J. Controlled Release 2014, 181, 11−21. (36) Ni, L.; Bai, Y. L.; Yang, L. P.; Lu, Y. L.; Wang, L.; Zhou, L. P. The effect of urea-formaldehyde fertilizer under different components by summer maize. Sci. Agric. Sin. 2016, 49, 3370−3379. (37) Yang, J. F.; Huang, S. H.; Xing, S. L.; Yuan, W. H.; Liu, F.; Yan, X. F.; Jia, L. L.; Liu, X. T. Study on effect of urea formaldehyde slowrelease fertilizer on winter wheat. J. Hebei Agric. Sci. 2017, 21, 47−52. (38) Chang, H. B.; Liu, Q. Y.; Xie, L. H.; Yu, C. B. Study on appropriate proportion of slow release nitrogen fertilizer in rape. Hubei Agric. Sci. 2016, 55, 6096. (39) Zhang, B.; Lli, Y. Q.; Gao, J. F.; Wang, H.; Zhu, F. T.; Ma, C. Synthesis and characterization of dimethylolurea and its condensation polymer. J. North Univ. China 2008, 29, 550−553. (40) Paramashivam, D.; Clough, T. J.; Carlton, A.; Gough, K.; Dickinson, N. M.; Horswell, J.; Sherlock, R. R.; Clucas, L.; Robinson, B. The effect of lignite on nitrogen mobility in a low-fertility soil amended with biosolids and urea. Sci. Total Environ. 2016, 543, 601−608. (41) Rose, M. T.; Perkins, E. L.; Saha, B. K.; Tang, E. C. W.; Cavagnaro, T. R.; Jackson, W. R.; Hapgood, K.; Hoadley, A.; Patti, A. F. A slow release nitrogen fertiliser produced by simultaneous granulation and superheated steam drying of urea with brown coal. Chem. Biol. Technol. Agric. 2016, 3, 10. (42) Duan, L. L.; Fan, B.; Wang, Y.; Liu, G.; Wang, L. Study of Ureaform Slow-Release Fertilizers by Infrared Spectrometry and Nutrient Release Characteristics. J. Chem. Fertilizer Ind. 2010, 37, 16−22. (43) Xu, N.; Tan, G. C.; Wang, H. Y.; Gai, X. P. Effect of biochar additions to soil on nitrogen leaching, microbial biomass and bacterial community structure. Eur. J. Soil Biol. 2016, 74, 1−8. (44) Ding, Y.; Liu, Y. X.; Wu, W. X.; Shi, D. Z.; Yang, M.; Zhong, Z. K. Evaluation of biochar effects on nitrogen retention and leaching in multi-layered soil columns. Water, Air, Soil Pollut. 2010, 213, 47−55. (45) Salvagiotti, F.; Castellarin, J. M.; Miralles, D. J.; Pedrol, H. M. Sulfur fertilization improves nitrogen use efficiency in wheat by increasing nitrogen uptake. Field Crop. Res. 2009, 113, 170−177. (46) Wei, T. X.; Zhang, X. J.; Zhu, J. Z. The nutrient accumulation pattern and cycling in natural secondary forests in North China. A case study from the Caijiachuan watershed, Shanxi Province. Phyton-int. J. Exp. Bot. 2014, 88, 213−223. (47) Xu, J. Z.; Liao, L. X.; Tan, J. Y.; Shao, X. H. Ammonia volatilization in gemmiparous and early seedling stages from direct seeding rice fields with different nitrogen management strategies. A pots experiment. Soil Tillage Res. 2013, 126, 169−176. 7624

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625

Article

Journal of Agricultural and Food Chemistry (48) Shoji, S.; Delgado, J.; Mosier, A.; Miura, Y. Use of controlled release fertilizers and nitrification inhibitors to increase nitrogen use efficiency and to conserve air and water quality. Commun. Soil Sci. Plant Anal. 2001, 32, 1051−1070. (49) Tian, X. F.; Li, C. L.; Zhang, M.; Li, T.; Lu, Y. Y.; Liu, L. F. Controlled release urea improved crop yields and mitigated nitrate leaching under cotton-garlic intercropping system in a 4-year field trial. Soil Tillage Res. 2018, 175, 158−167. (50) Geng, J. B.; Ma, Q.; Zhang, M.; Li, C. L.; Liu, Z. G.; Lyu, X. X.; Zheng, W. K. Synchronized relationships between nitrogen release of controlled release nitrogen fertilizers and nitrogen requirements of cotton. Field Crop. Res. 2015, 184, 9−16. (51) Yang, Y. C.; Zhang, M.; Li, Y. C.; Fan, X. H.; Geng, Y. Q. Controlled release urea improved nitrogen use efficiency, activities of leaf enzymes, and rice yield. Soil Sci. Soc. Am. J. 2012, 76, 2307−2317. (52) Zheng, W. K.; Sui, C. L.; Liu, Z. G.; Geng, J. B.; Tian, X. F.; Yang, X. Y.; Li, C. L.; Zhang, M. Long-term effects of controlled-release urea on crop yields and soil fertility under wheat-corn double cropping systems. Agron. J. 2016, 108, 1703. (53) Gao, X.; Li, C. L.; Zhang, M.; Wang, R.; Chen, B. C. Controlled release urea improved the nitrogen use efficiency: yield and quality of potato (Solanum tuberosum L) on silt loamy soil. Field Crop. Res. 2015, 181, 60−68. (54) Fan, J. L.; Xiao, J.; Liu, D. Y.; Ye, G. P.; Luo, J. F.; Houlbrooke, D. J.; Laurenson, S.; Yan, J.; Chen, L. J.; Tian, J. P.; Ding, W. X. Effect of application of dairy manure, effluent and inorganic fertilizer on nitrogen leaching in clayey fluvo-aquic soil: a lysimeter study. Sci. Total Environ. 2017, 592, 206−214. (55) Yang, X.; Lu, Y.; Tong, Y. A.; Yin, X. A 5-year lysimeter monitoring of nitrate leaching from wheat-maize rotation system: comparison between optimum N fertilization and conventional farmer N fertilization. Agric., Ecosyst. Environ. 2015, 199, 34−42. (56) Sanderson, K. R.; Fillmore, S. A. E. Slow-release nitrogen fertilizer in carrot production on Prince Edward Island. Can. J. Plant Sci. 2012, 92, 1223−1228. (57) Perego, A.; Basile, A.; Bonfante, A.; De-Mascellis, R.; Terribile, F.; Brenna, S.; Acutis, M. Nitrate leaching under maize cropping systems in Po Valley (Italy). Agric., Ecosyst. Environ. 2012, 147, 57−65. (58) Grant, C. A.; Wu, R.; Selles, F.; Harker, K. N.; Clayton, G. W.; Bittman, S.; Zebarth, B. J.; Lupwayi, N. Z. Crop yield and nitrogen concentration with controlled release urea and split applications of nitrogen as compared to non-coated urea applied at seeding. Field Crop. Res. 2012, 127, 170−180. (59) McKenzie, R. H.; Bremer, E.; Middleton, A. B.; Pfiffner, P. G.; Dowbenko, R. E. Controlled-release urea for winter wheat in southern Alberta. Can. J. Soil Sci. 2007, 87, 85−91. (60) Madusanka, N.; Sandaruwan, C.; Kottegoda, N.; Sirisena, D.; Munaweera, I.; De Alwis, A.; Karunaratne, V.; Amaratunga, G. A. J. Urea−hydroxyapatite-montmorillonite nanohybrid composites as slow release nitrogen compositions. Appl. Clay Sci. 2017, 150, 303−308.

7625

DOI: 10.1021/acs.jafc.9b01432 J. Agric. Food Chem. 2019, 67, 7616−7625