Next Article in Journal
Effects of Marine Residue-Derived Fertilizers on Strawberry Growth, Nutrient Content, Fruit Yield and Quality
Previous Article in Journal
First Report of Resistance to Glyphosate in Several Species of the Genus Echinochloa in Argentina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Biodegradable Plastic Mulch Film on Cabbage Agronomic and Nutritional Quality Traits, Soil Physicochemical Properties and Microbial Communities

1
Institute of Vegetable Crops, Jiangsu Academy of Agricultural Sciences, Nanjing 210014, China
2
Jiangsu Provincial Agricultural Technology Extension Station, Nanjing 210036, China
3
College of Life Sciences, Nanjing Agricultural University, Nanjing 210095, China
4
Yixing Bureau of Agriculture and Rural Affairs, Yixing 214206, China
5
Jiangsu Yihe Agricultural Technology Co., Ltd., Yixing 214263, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2023, 13(5), 1220; https://doi.org/10.3390/agronomy13051220
Submission received: 7 March 2023 / Revised: 21 April 2023 / Accepted: 23 April 2023 / Published: 26 April 2023
(This article belongs to the Section Soil and Plant Nutrition)

Abstract

:
The long-term use of polyethylene mulch (PEM) films can cause plastic film residual pollution that has an adverse effect on soil health and crop quality. To address this issue, poly(butylene adipate-co-terephthalate) (PBAT), an aliphatic–aromatic copolyester, is widely used in the production of commercially biodegradable plastic mulch (BDM) films. The use of BDMs can alleviate soil plastic pollution and reduce the labor cost of retrieving plastic film residues from the field. The effects of BDM and PEM on the agronomic and nutritional quality traits of cabbage cultivar ‘Sugan No. 35’, as well as the physicochemical properties and microbial communities of the soil were analyzed during two consecutive years of the experiment. No significant difference was observed in the cabbage agronomic and nutritional quality traits among three mulching treatments. Nonetheless, the mulching and degradation of BDM reduced the pH value and increased the organic matter content of the soil samples compared with PEM mulching. In the soil bacterial and fungal communities, Proteobacteria and Ascomycota were the most abundant bacterial phylum and fungal phylum across all the soil samples, respectively; the use of BDM increased the relative abundance of soil Proteobacteria and Ascomycota compared with PEM mulching. The overall cost of BDM mulching was much lower than that of PEM mulching during the cabbage production.

1. Introduction

Plastic film mulching, which is a globally applied agricultural practice since 1960s, can produce huge productive and economic advantages [1,2]. Plastic film mulching can increase soil temperature, reduce the consumption of irrigation water [3,4], lower weed pressure, reduce the use of herbicides [5], and increase crop yields [6]. Polyethylene mulch (PEM) films have been commonly used in agricultural production because of their high durability, easy processability, flexibility, and low price. However, their long-term use has resulted in widespread plastic film residual pollution because the plastic film residues are not completely removed from fields at the end of the growing season and do not readily biodegrade [7,8,9]. China is currently the largest PEM consumer [10]. Nonetheless, the excessive use of PEMs has resulted in the accumulation of plastic film residues in soils >250 kg/ha [7,11]. To alleviate agricultural plastic pollution and reduce the labor arising from recovering the plastic film residues, biodegradable plastic mulch (BDM) films have been developed as a potential and sustainable alternative to PEMs [12,13,14]. Poly(butylene adipate-co-terephthalate) (PBAT) is an aliphatic–aromatic copolyester, that is widely used in the production of commercially available BDMs [10]. In general, the BDMs are only used once per crop-growing season. At the end of the crop cycle, PBAT-based BDMs can be tilled and incorporated into the soil directly, unlike PEMs, which require removal and disposal. Over time, they may be degraded and utilized by resident soil microorganisms, thereby simplifying the farmers’ operations [9,15].
The soil microorganisms play vital roles in the soil ecosystem. Studies have shown that the incorporation of mulch film results in altered abundance of microorganisms in the soil depending on the soil type, soil environment, and agricultural management [9,16,17,18,19,20]. The soil microorganisms secrete extracellular microbial enzymes that depolymerize the PBAT-based BDMs into low-molecular-weight compounds, which are further utilized and eventually converted into microbial biomass, CO2, and water, under ideal circumstances [21,22,23,24]. However, the complete biodegradation of BDMs, which is largely affected by soil microorganisms and soil parameters, takes time [10,25,26]. Using soil microcosms, the mineralization levels of the PBAT film in lou soil, fluvo-aquic soil, black soil, and red soil after 120 days were 16, 9, 0.3, and 0.9%, respectively [10]. Sěrá et al. (2016) found that the mineralization level of pure PBAT film reached the point of approximately 6% after 100 days of incubation [27]. The PBAT-degrading bacteria, especially for several members of Proteobacteria harbored in soil environments, were identified as promising degraders [10]. As for the PBAT-degrading fungi, the surface of PBAT films is mainly enriched by Ascomycota, which may degrade PBAT through cutinase and play a leading role in the biodegradation of the PBAT [16,28].
This study determines the agronomic and nutritional quality traits of the cabbage crop, the physicochemical properties and microbial communities of the soil under three mulching treatments during two consecutive years of the experiment (2021–2022). This study mainly aims to: (i) explore whether BDM mulching affects the cabbage agronomic and nutritional quality traits compared with PEM mulching; (ii) investigate whether the mulching and degradation of BDM alter the physicochemical properties and microbial communities of the soil compared with PEM mulching; and (iii) comparatively analyze the cost–benefit of BDM mulching and PEM mulching.

2. Material and Methods

2.1. Plastic Mulch Films and Sample Collection

The BDM and conventional PEM used in this study were purchased from Nantong Huasheng Plastic Products Co., Ltd., Nantong, China. The major constituents of the BDM and PEM were PBAT and polyethylene, respectively. These two mulch films were black in color, 0.01 mm thick, and 1.2 m wide. The experimental field (119°57′37.52″ E, 31°22′30.63″ N) was located in Yixing City, Jiangsu Province, China. The cabbage cultivar ‘Sugan No. 35’ was used as the test crop and was harvested approximately 60 days after the seedlings were transplanted into the field. The plant spacing and row spacing of this cultivar in the experimental field were set as 35 cm and 40 cm, respectively. The field area of each treatment covered with PEM or BDM was 1/15 ha with a border width of 80 cm. During the two consecutive years of the experiment (2021–2022), three mulching treatments were designed and named ‘PEM-PEM’, ‘PEM-BDM’, and ‘BDM-BDM’, respectively (Table 1). All the trials were arranged in a randomized complete block design, with each treatment comprising three plots. The procedures of the cabbage planting and sample collection were as follows. (a) First, the experimental field was prepared by mixing the base fertilizer (commercial organic fertilizer 6000 kg/ha, compound fertilizer 375 kg/ha) in the field and covering it with PEM and BDM before transplanting. (b) Then, the cabbage seedlings were transplanted into the field on 20 March 2021 and 20 March 2022. (c) The cabbage heads from the nine plots were harvested on 20 May 2021 and 20 May 2022; the cabbage heads harvested in 2022 were collected to determine their agronomic and nutritional quality traits. (d) After removing cabbage wastes removed manually, PEM residues were removed manually, and BDM residues were tilled into the soil on 1 June 2021 and 1 June 2022. The surface-layer soil samples (0–20 cm) were collected from each treatment of three plots on 20 October 2022. The experimental field was unused and uncovered after the cabbage planting period during the two consecutive years of the experiment. The collected soil samples were sieved through a 2 mm sieve and divided into two parts: one part for a soil physicochemical property analysis, and the second part for a soil microbial community analysis.

2.2. Determination of Agronomic and Nutritional Quality Traits and Physicochemical Properties

All the cabbage heads from each plot were collected, weighed, and converted into yield (ton/ha). Five cabbage heads were randomly selected from each plot and used for head weight (kg), head vertical diameter (cm), head transverse diameter (cm), core length (cm), and core width (cm) determination following standard methods [29]. After that, the five cabbage heads from the same plot were pooled together and used for the nutritional quality traits determination. The contents of total soluble solids, soluble sugar, soluble protein, and vitamin C were determined by refractometer, anthrone colorimetry, the Coomassie brilliant blue method, and 2, 6-dichlorophenol-indophenol titration method, respectively [30]. The soil samples (three soil samples per treatment) were subjected to a soil physicochemical property analysis following standard soil testing procedures [31]. The soil pH in a 1:2.5 w/v soil–water suspension was determined by a pH meter (ST3100-F, OHAUS Co., Parsippany, NJ, USA). The nitrate nitrogen content of the soil was measured by the ultraviolet spectrophotometry method [32]. The available phosphorus content of the soil was measured by the sodium bicarbonate–molybdenum antimony colorimetric method. The available potassium content of the soil was determined using a flame photometer with ammonium acetate. The organic matter content of the soil was determined by the potassium dichromate oxidation–external heat method. The total nitrogen content of the soil was determined by the Kjeldahl nitrogen method. The average values were calculated for each trait from three plots per treatment. Where necessary, the analysis of variance (ANOVA) and Duncan’s test (p < 0.05) were performed using SPSS 20.0 software.

2.3. Soil DNA Extraction, PCR, and Sequencing

The DNA extraction from nine soil samples was completed using the Omega Mag-Bind Soil DNA kit (M5635-02, Omega Bio-tek, Norcross, GA, USA) with inhibitor removal technology, following the manufacturer’s instructions. The V3-V4 region of bacterial 16S rRNA and the fungal internal transcribed spacers (ITS1) region were amplified using relevant primers with different barcodes [33,34]. The polymerase chain reaction (PCR) amplification was performed in a 25 μL reaction system containing 0.25 μL of Q5 DNA Polymerase, 2 μL of dNTP mixture (2.5 mM), 1 μL of forward primer (10 μM), 1 μL of reverse primer (10 μM), 2 μL of template DNA (10–30 ng), 5 μL of 5×reaction buffer, 5 μL of 5×GC buffer, and 8.75 μL of ddH2O. Subsequently, PCR was performed under the following conditions: initial denaturation at 98 °C for 2 min, 35 cycles of denaturation at 98 °C for 15 s, annealing at 55 °C for 30 s, and extension at 72 °C for 30 s, with a final extension at 72 °C for 5 min. The PCR products were purified by gel electrophoresis and sequenced on the Illumina NovaSeq platform.

2.4. Bioinformatics Analysis

After the initial trimming, FLASH v1.2.7 software (http://ccb.jhu.edu/software/FLASH/) (accessed on 10 February 2023) was used to assemble the paired reads into a sequence to obtain tags. The similarity of effective tags higher than 97% was defined as Operational Taxonomic Units (OTUs) by Usearch v8.0.1517 software (http://www.drive5.com/usearch/) (accessed on 10 February 2023) [35]. To get the corresponding species classification information of each OUT from the bacterial 16S rRNA sequences and fungal ITS sequences, the representative OTU sequences were aligned to the SILVA database (http://www.arb-silva.de) (accessed on 10 February 2023) and UNITE database (Release 8.0, https://unite.ut.ee/) (accessed on 10 February 2023), respectively. Then, the community composition and species richness of each sample was counted at the phylum and genus levels. The R language tool was used to draw a community structure graph of the samples at the phylum and genus levels, and only the species with the top ten most abundant levels are shown in the graph for clarity. The alpha diversity of the microbial community was assessed by calculating the ACE, Chao1, Shannon, and Simpson indices using Mothur v.1.30 software (http://www.mothur.org/) (accessed on 10 February 2023). A principal coordinates analysis (PCoA) was performed to compare the beta diversity of the microbial communities based on the Bray–Curtis distances using the labdsv package [36].

3. Results

3.1. Effects of BDM and PEM on Agronomic and Nutritional Quality Traits of Cabbage

To investigate the effect of BDM and PEM on the agronomic and nutritional quality traits of cole vegetables, the cabbage cultivar ‘Sugan No. 35’ was selected and measured under different mulching treatments. To determine the agronomic traits of cabbage, its head weight, head vertical diameter, head transverse diameter, core length, core width, and yield ranged from 2.19–2.32 kg, 17.50–19.07 cm, 18.90–20.50 cm, 6.77–7.20 cm, 3.37–3.67 cm, and 84.68–87.84 ton/ha, respectively. For the nutritional quality traits of cabbage, the total soluble solids, soluble sugar, soluble protein, and vitamin C ranged from 5.57–6.05%, 245.30–254.13 mg/g DW, 1.32–1.57 mg/g FW, and 74.66–81.07 mg/100g FW, respectively (Table 2). No significant difference was observed on the agronomic and nutritional quality traits of the cabbage under different mulching treatments.

3.2. Analysis of the Physicochemical Properties of Soil

The physicochemical properties of the soil under different mulching treatments were measured (Table 3). The study revealed that the pH value of all the soil samples was weakly acidic (5.27–5.93). The nitrate nitrogen, available phosphorus, available potassium, organic matter, and total nitrogen concentrations of the soil samples were 30.47–37.20 mg/kg, 24.00–33.57 mg/kg, 56.00–64.33 mg/kg, 24.03–32.23 g/kg, and 0.18–0.21%, respectively. The soil pH value of the PEM-PEM sample (5.93) was significantly (p < 0.05) higher than those in the PEM-BDM (5.27) and BDM-BDM (5.27) samples, whereas the content of soil organic matter in the PEM-PEM (24.03 g/kg) sample was significantly lower than those in the PEM-BDM (30.47 g/kg) and BDM-BDM (32.23 g/kg) samples. The mulching and degradation of BDM and lowered the pH value and increased the organic matter content of the soil. No significant difference was observed in the contents of nitrate nitrogen, available phosphorus, available potassium, and total nitrogen among the soil samples (Table 3).

3.3. Analysis of Soil Microbial Community Diversity

A total of 719,136 paired-end reads and 692,552 effective tags from nine soil samples for an average of 76,950 effective tags per sample were obtained for bacterial communities (Supplementary Materials: Table S1). A total of 590,036 paired-end reads and 533,783 effective tags from nine samples for an average of 59,309 effective tags per sample were obtained for fungal communities. There were 19,461 and 7543 OTUs obtained in the bacterial and fungal communities, respectively (Supplementary Materials: Table S1). The bacterial and fungal communities from the three samples (e.g., PEM-PEM, PEM-BDM, and BDM-BDM) shared 2221 and 902 OTUs, respectively (Figure 1).

3.4. Alpha Diversity Comparison among Different Soil Samples

The results showed that the indices of Chao1, Simpson, and Shannon indices of soil bacterial and fungal communities among the PEM-PEM, PEM-BDM, and BDM-BDM samples were not significantly different (p > 0.05) (Table 4; Supplementary Materials: Table S2). The abundance-based coverage estimator (ACE) index of fungi was lower in the PEM-PEM sample than those in the PEM-BDM sample. Therefore, the alpha diversity of soil bacteria and fungi was largely similar among the PEM-PEM, PEM-BDM, and BDM-BDM samples. The rarefaction curve of all the samples in the bacteria and fungi tended to be flat, indicating that the sample sequencing depth essentially covered all the species (Supplementary Materials: Figure S1).

3.5. Beta Diversity Comparison among Different Soil Samples

Principal coordinates analysis (PCoA) was used to evaluate the dissimilarity of the bacterial and fungal communities among three different mulching treatments (Figure 2). In the bacteria, the first principal coordinate (PCoA1) and the second principal coordinate (PCoA2) explained 22.74% and 20.66% of all the variances, respectively. The cumulative contribution rate of the variance of the two axes reached 43.40% (Supplementary Materials: Figure S2). In the fungi, the PCoA1 and PCoA2 explained 41.36% and 21.28% of the variance, respectively. The cumulative contribution rate of the variance of the two axes reached 62.64% (Supplementary Materials: Figure S2).

3.6. Effects of BDM and PEM on Soil Microbial Community Composition

In the bacterial community, the ten most abundant bacterial phyla, other phyla, and unassigned phyla were calculated at the phylum level under the three different treatments. Proteobacteria, Acidobacteria, Chloroflexi, Gemmatimonadetes, and Bacteroidetes were the major phyla (proportion more than 5.0%) in the three samples. Proteobacteria was the most abundant phylum, accounting for more than 34.8% of all the bacterial taxa identified in the three samples. Acidobacteria accounted for 21.8–22.3% of all the bacterial taxa in the three samples, followed by Chloroflexi, which accounted for 8.1–11.4% of all the bacterial taxa (Figure 2; Supplementary Materials: Figure S3 and Table S3). The relative abundance of the Proteobacteria and Acidobacteria phyla was higher in the BDM-BDM sample than those in the PEM-PEM and PEM-BDM samples, indicating that the use of BDM increased the relative abundance of the Proteobacteria and Acidobacteria phyla compared with the PEM mulching, whereas the use of BDM reduced the relative abundance of the Chloroflexi, Gemmatimonadetes, Bacteroidetes, Planctomycetes, Nitrospirae, Verrucomicrobia, and Latescibacteria phyla (Supplementary Materials: Table S3). In addition, the ten most abundant bacterial genera, other genera, unclassified genera, and unassigned genera in the three samples were calculated at the genus level. Uncultured_bacterium_c_Subgroup_6 was the most abundant genus, with more than 10.3% of the total bacterial taxa in the three samples. Uncultured_bacterium_f_Gemmatimonadaceae accounted for 4.4–4.6% of all the bacterial taxa in the three samples, followed by MND1, which accounted for 3.0–3.4% of all the bacterial taxa (Figure 2; Supplementary Materials: Figure S3 and Table S3). The use of BDM increased the relative abundance of MND1, RB41, uncultured_bacterium_f_TRA3-20, and uncultured_bacterium_c_Alphaproteobacteria compared with the PEM mulching and reduced the relative abundance of Nitrospira, uncultured_bacterium_f_A4b, and Sphingomonas (Supplementary Materials: Table S3).
In the fungal community, the ten most abundant fungal phyla, other phyla, and unclassified phyla in the three samples were calculated at the phylum level. Ascomycota was the most abundant phylum, which accounted for 66.6–81.1% of all the fungal taxa in the three samples, followed by Mortierellomycota, which accounted for 2.8–6.5% of all the bacterial taxa (Figure 3; Supplementary Materials: Figure S4 and Table S4). The use of BDM increased the relative abundance of Ascomycota, Basidiomycota, and Olpidiomycota compared with the PEM mulching, while it reduced the relative abundance of Mortierellomycota and Chytridiomycota (Supplementary Materials: Table S4). In addition, the ten most abundant fungal genera, other genera, and unclassified genera in the three samples were calculated at the genus level. Metarhizium was the most abundant genus, accounting for 12.8–17.6% of the total fungal taxa. Mortierella accounted for 2.7–6.0% of all the fungal taxa, followed by Ophiocordyceps, which accounted for 3.3–4.6% of all the fungal taxa (Figure 3; Supplementary Materials: Figure S4 and Table S4). The use of BDM increased the relative abundance of Ophiocordyceps, Gibberella, Plectosphaerella, Fusarium, Preussia, and Schizotheciumgenera compared with the PEM mulching, while it reduced the relative abundance of Mortierella (Supplementary Materials: Table S4).

3.7. The Cost–Benefit Analysis of Cabbage Production Covered with BDM and PEM

The cost–benefit of cabbage production covered with BDM and PEM was analyzed. The cost of seeds, vegetable seedling substrate, fertilizer, herbicides, insecticides, fungicides, equipment operation, and labor required for the sowing, transplanting, field management, and cabbage harvest was the same between the use of BDM and PEM. The cost of purchasing the BDM and PEM was 3150 and 1050 Chinese yuan (CNY, 1 US dollar ≈ 6.9 CNY) per hectare, respectively. The cost of labor arising from recovering the PEM residues was 3600 CNY per hectare. Nevertheless, no labor was need for BDM because the residues were left in the field. Although the price of purchasing BDM is more expensive than PEM, the overall cost of BDM mulching was much lower than that of PEM mulching, saving 1500 CNY per hectare during cabbage production.

4. Discussion

Plastic film mulching is a common practice in agriculture, and its benefits include pest limitation, weed suppression, and the conservation of soil temperature and moisture, among others [2,37,38,39]. Although PEMs are the most common mulches used in commercial agricultural systems [15], they are not largely prone to biodegradation and are typically recovered at the end of the crop cycle. However, plastic film residues in soil that resulted from the inadequate removal of PEMs adversely affect soil health and crop growth [40]. In this study, BDM was largely similar to PEM in many respects in terms of cabbage production. The agronomic and nutritional quality traits of the cabbage were not significantly altered under different mulching treatments (Table 2). A meta-analysis by Tofanelli and Wortman (2020) showed that crop yields (292 observations across 66 studies) under the BDM mulching were greater than those in bare soil but are comparable with those under the PEM mulching [41]. A previous study also revealed similar maize yield between BDM and PEM, and black BDM can improve maize quality [42].
A higher pH value was observed in the PEM-PEM sample than those in the PEM-BDM and BDM-BDM samples (Table 3). In the PEM-BDM and BDM-BDM samples, the premature rupture of BDM before cabbage harvest causes the soil more likely to undergo the leaching of alkaline salts from rainfall, thereby resulting in decreased pH in the soil. Thus, the content of soil organic matter in the PEM-PEM sample was significantly lower than those in the PEM-BDM and BDM-BDM samples. Previous findings showed that plastic mulch films decrease the content of soil organic matter [43,44]. The impact of BDMs on soil properties varies depending on multiple factors, including location, time, and planting systems [45,46,47]. A six-year study revealed that BDM coverage resulted in a 12.09–17.17% decrease in the soil bulk density of the 10–20 cm soil layer, a 14.75–28.37% increase in the soil’s total nitrogen, a 64.20% increase in the soil’s available phosphorus, and a 108.8% increase in the potassium content [48]. Previous studies showed the temperature, pH value, and organic matter content of the soil can affect soil enzyme activities [49,50,51]. Plastic mulch films can increase the soil temperature, and the high temperature can accelerate the microbial metabolic processes and the decomposition of organic matter in the soil [52,53]. Previous studies showed the PBAT-based BDM hydrolases have low or no activity at pH values < 5.0 [16,54,55]. In this study, the pH values of all the soil samples were >5.0 under different mulching treatments. Much of the organic matter may be able to regulate the production of BDM hydrolase [10]. The physicochemical properties of the soil, except the pH value and organic matter content, were rarely affected by the BDM mulching and degradation compared with the PEM mulching during the two consecutive years of the experiment, although it was tilled into the soil after the cabbage growing season.
In the bacterial community, Proteobacteria was the most abundant bacterial phylum, accounting for 37.1% of the abundance in the PEM-PEM sample. The use of BDM increased the relative abundance of Proteobacteria compared with the PEM mulching, which accounted for 41.5% of the abundance in the BDM-BDM sample. The potential PBAT-degrading bacteria Proteobacteria harbor in soil environments, possess a wide range of PBAT hydrolase catalogs, and play key roles in PBAT degradation [10]. In the BDM-BDM sample, the relative abundance of Proteobacteria increased, possibly because of their PBAT biodegradation activity. In the fungal community, Ascomycota was the most abundant fungal phylum, accounting for 66.6% of the abundance in the PEM-PEM sample. The use of BDM increased the relative abundance of Ascomycota compared with the PEM mulching, which accounted for 81.1% and 78.3% of the abundance in the PEM-BDM and BDM-BDM samples, respectively. A previous study showed that the surface of PBAT films was preferentially enriched by the fungi Ascomycota, which may degrade PBAT through cutinase [16]. In the PEM-BDM and BDM-BDM samples, the relative abundance of Ascomycota increased after the BDM mulching and may accelerate the degradation of the PBAT film. After the BDMs were incorporated into the soil, the soil environment was affected [22], and the abundance of soil fungal communities increased [16,20]. The research by Bandopadhyay et al. (2020) detected minimal differences in soil microbial communities between BDM and PEM mulching [56]. Liang et al. (2022) revealed that BDM mulching had minimal effect on the overall system of soil microorganisms [44]. Although the cost of purchasing BDM is currently more expensive than PEM, the overall cost of BDM mulching is much lower during cabbage production. The labor reduction from the no recovery of the plastic film residues incentives the growers to use BDMs. BDMs that are readily biodegradable and do not harm the soil environment could become a potential and sustainable alternative to PEMs [12,13,14]. However, previous research and this research have focused on the effects of short-term BDMs on the physicochemical and microbial properties of soil, generally one or two crop growing seasons; thus, the long-term impacts of BDM incorporation are unknown [9]. To address the current knowledge gaps, long-term BDM research is needed in future studies. Evaluating the effects of long-term BDMs incorporation on soil health and ecosystems is a critical part of the extended application of BDMs.

5. Conclusions

In conclusion, we focused on the effects of BDM and PEM on the agronomic and nutritional quality traits of cabbage, physicochemical properties and microbial communities of soil. No significant difference was observed on the cabbage agronomic and nutritional quality traits among the three mulching treatments. We found the mulching and degradation of BDM decreased the soil pH value and increased the soil organic matter content compared with the PEM mulching. Interestingly, in the soil bacterial and fungal communities, the bacteria Proteobacteria and fungi Ascomycota were the most abundant phyla across all the soil samples; the relative abundance of soil Proteobacteria and Ascomycota under the BDM mulching was higher than that under the PEM mulching. Notably, the overall cost of the BDM mulching was much lower than that of the PEM mulching, thereby saving 1500 CNY per hectare during cabbage production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agronomy13051220/s1. Figure S1. The rarefaction curves of bacterial and fungal communities under three mulching treatments; Figure S2. Principal coordinates analysis (PCoA) of the bacterial and fungal communities based on Bray–Curtis distances. Figure S3. Statistical analysis of the relative abundance at phylum and genus levels of the bacterial community under three mulching treatments; Figure S4. Statistical analysis of the relative abundance at phylum and genus levels of the fungal community under three mulching treatments; Table S1. Statistics of sequencing data processing results of nine soil samples in bacteria and fungi under three mulching treatments; Table S2. The alpha diversity indices of bacterial and fungal communities of nine soil samples under three mulching treatments; Table S3. The relative abundance of dominant bacterial phyla and genera under three mulching treatments; Table S4. The relative abundance of dominant fungal phyla and genera under three mulching treatments.

Author Contributions

Conceptualization, W.Z., J.M., S.W., and X.Z.; formal analysis, W.Z.; funding acquisition, S.W. and X.Z.; investigation, W.Z., J.M., L.X., and Q.L.; resources, J.L.; supervision, S.W. and X.Z.; writing—original draft, W.Z. and J.M.; writing—review and editing, Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was, in part, supported by the Key Research and Development Program of Jiangsu Province (BE2021376), the Science and Technology Planning Project of Nanjing City (202109021), the Agroecological Protection and Resource Utilization Project of Jiangsu Province (2021-SJ-102), and the China Agriculture Research System (CARS-23-G42).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lamont, W.J. Plastics: Modifying the microclimate for the production of vegetable crops. HortTechnology 2005, 15, 477–481. [Google Scholar] [CrossRef]
  2. Kasirajan, S.; Ngouajio, M. Polyethylene and biodegradable mulches for agricultural applications: A review. Agron. Sustain. Dev. 2012, 32, 501–529. [Google Scholar] [CrossRef]
  3. Kader, M.A.; Senge, M.; Mojid, M.A.; Ito, K. Recent advances in mulching materials and methods for modifying soil environment. Soil Tillage Res. 2017, 168, 155–166. [Google Scholar] [CrossRef]
  4. Zong, R.; Han, H.; Li, Q. Grain yield and water-use efficiency of summer maize in response to mulching with different plastic films in the North China Plain. Exp. Agric. 2021, 57, 33–44. [Google Scholar] [CrossRef]
  5. Ngouajio, M.; Auras, R.; Fernandez, R.T.; Rubino, M.; Counts, J.W.; Kijchavengkul, T. Field performance of aliphatic-aromatic copolyester biodegradable mulch films in a fresh market tomato production system. HortTechnology 2008, 18, 605–610. [Google Scholar] [CrossRef]
  6. Jiang, R.; Li, X.; Zhou, M.; Li, H.J.; Zhao, Y. Plastic film mulching on soil water and maize (Zea mays L.) yield in a ridge cultivation system on Loess Plateau of China. Soil Sci. Plant Nutr. 2016, 621, 1–15. [Google Scholar] [CrossRef]
  7. Liu, E.K.; He, W.Q.; Yan, C.R. ‘White revolution’ to ‘white pollution’—Agricultural plastic film mulch in China. Environ. Res. Lett. 2014, 99, 091001. [Google Scholar] [CrossRef]
  8. Steinmetz, Z.; Wollmann, C.; Schaefer, M.; Buchmann, C.; David, J.; Tröger, J.; Muñoz, K.; Frör, O.; Schaumann, G.E. Plastic mulching in agriculture. Trading short-term agronomic benefits for long-term soil degradation? Sci. Total Environ. 2016, 550, 690–705. [Google Scholar] [CrossRef]
  9. Bandopadhyay, S.; Martin-Closas, L.; Pelacho, A.M.; DeBruyn, J.M. Biodegradable plastic mulch films: Impacts on soil microbial communities and ecosystem functions. Front. Microbiol. 2018, 9, 819. [Google Scholar] [CrossRef]
  10. Han, Y.; Teng, Y.; Wang, X.; Ren, W.; Wang, X.; Luo, Y.; Zhang, H.; Christie, P. Soil type driven change in microbial community affects poly(butylene adipate-co-terephthalate) degradation potential. Environ. Sci. Technol. 2021, 55, 4648–4657. [Google Scholar] [CrossRef]
  11. Yan, C.; He, W.; Turner, N. Plastic-film mulch in Chinese agriculture: Importance and problems. World Agric. 2014, 4, 32–36. [Google Scholar]
  12. Albertsson, A.C.; Hakkarainen, M. Designed to degrade. Science 2017, 358, 872–873. [Google Scholar] [CrossRef] [PubMed]
  13. Gross, R.A.; Kalra, B. Biodegradable polymers for the environment. Science 2002, 297, 803–807. [Google Scholar] [CrossRef]
  14. Haider, T.P.; Völker, C.; Kramm, J.; Landfester, K.; Wurm, F.R. Plastics of the future? The impact of biodegradable polymers on the environment and on society. Angew. Chem. Int. Ed. Engl. 2019, 58, 50–62. [Google Scholar] [CrossRef]
  15. Sintim, H.Y.; Flury, M. Is biodegradable plastic mulch the solution to agriculture’s plastic problem? Environ. Sci. Technol. 2017, 51, 1068–1069. [Google Scholar] [CrossRef] [PubMed]
  16. Muroi, F.; Tachibana, Y.; Kobayashi, Y.; Sakurai, T.; Kasuya, K. Influences of poly(butylene adipate-co-terephthalate) on soil microbiota and plant growth. Polym. Degrad. Stab. 2016, 129, 338–346. [Google Scholar] [CrossRef]
  17. Somanathan, H.; Sathasivam, R.; Sivaram, S.; Mariappan Kumaresan, S.; Muthuraman, M.S.; Park, S.U. An update on polyethylene and biodegradable plastic mulch films and their impact on the environment. Chemosphere 2022, 307, 135839. [Google Scholar] [CrossRef]
  18. Xue, Y.; Jin, T.; Gao, C.; Li, C.; Zhou, T.; Wan, D.; Yang, M. Effects of biodegradable film mulching on bacterial diversity in soils. Arch. Microbiol. 2022, 204, 195. [Google Scholar] [CrossRef]
  19. Shan, X.; Zhang, W.; Dai, Z.; Li, J.; Mao, W.; Yu, F.; Ma, J.; Wang, S.; Zeng, X. Comparative analysis of the effects of plastic mulch films on soil nutrient, yields and soil microbiome in three vegetable fields. Agronomy 2022, 12, 506. [Google Scholar] [CrossRef]
  20. Li, C.; Moore-Kucera, J.; Miles, C.; Leonas, K.; Lee, J.; Corbin, A. Degradation of potentially biodegradable plastic mulch films at three diverse US locations. Agroecol. Sustain. Food Syst. 2014, 38, 861–889. [Google Scholar] [CrossRef]
  21. Sander, M. Biodegradation of polymeric mulch films in agricultural soils: Concepts, knowledge gaps, and future research directions. Environ. Sci. Technol. 2019, 53, 2304–2315. [Google Scholar] [CrossRef] [PubMed]
  22. Bandopadhyay, S.; Liquet, Y. González. J.E.; Henderson, K.B.; Anunciado, M.B.; Hayes, D.G.; DeBruyn, J.M. Soil microbial communities associated with biodegradable plastic mulch films. Front. Microbiol. 2020, 11, 587074. [Google Scholar] [CrossRef] [PubMed]
  23. Mueller, R.J. Biological degradation of synthetic polyesters-Enzymes as potential catalysts for polyester recycling. Process. Biochem. 2006, 41, 2124–2128. [Google Scholar] [CrossRef]
  24. Zumstein, M.T.; Schintlmeister, A.; Nelson, T.F.; Baumgartner, R.; Woebken, D.; Wagner, M.; Kohler, H.E.; McNeill, K.; Sander, M. Biodegradation of synthetic polymers in soils: Tracking carbon into CO2 and microbial biomass. Sci. Adv. 2018, 4, eaas9024. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Gao, W.; Mo, A.; Jiang, J.; He, D. Degradation of polylactic acid/polybutylene adipate films in different ratios and the response of bacterial community in soil environments. Environ. Pollut. 2022, 313, 120167. [Google Scholar] [CrossRef] [PubMed]
  26. Saadi, Z.; Cesar, G.; Bewa, H.; Benguigui, L. Fungal degradation of poly(butylene adipate-co-terephthalate) in soil and in compost. J. Polym. Environ. 2013, 21, 893–901. [Google Scholar] [CrossRef]
  27. Šerá, J.; Stloukal, P.; Jancova, P.; Verney, V.; Pekarova, S.; Koutny, M. Accelerated biodegradation of agriculture film based on aromatic-aliphatic copolyester in soil under mesophilic conditions. J. Agric. Food Chem. 2016, 64, 5653–5661. [Google Scholar] [CrossRef] [PubMed]
  28. Nikolić, M.A.; Gauthier, E.; Colwell, J.M.; Halley, P.; Bottle, S.E.; Laycock, B.; Truss, R. The challenges in lifetime prediction of oxodegradable polyolefin and biodegradable polymer films. Polym. Degrad. Stabil. 2017, 145, 102–119. [Google Scholar] [CrossRef]
  29. Li, X.; Fang, Z. Descriptors and Data Standards for Cabbage (Brassica oleracea L. var. capitata L. and Brassica oleracea L. var. gemmifera Zenk); China Agriculture Press: Beijing, China, 2007. [Google Scholar]
  30. Pearson, D. The Chemical Analysis of Foods; Longman Group Ltd.: London, UK, 1976. [Google Scholar]
  31. Lu, R. Analysis Method of Agricultural Chemistry in Soil; Agricultural Science and Technology Press: Beijing, China, 2000. [Google Scholar]
  32. Norman, R.J.; Edberg, J.C.; Stucki, J.W. Determination of nitrate in soil extracts by dual-wavelength ultraviolet spectrophotometry. Soil Sci. Soc. Am. J. 1985, 49, 1182–1185. [Google Scholar] [CrossRef]
  33. Sinclair, L.; Osman, O.A.; Bertilsson, S.; Eiler, A. Microbial community composition and diversity via 16S rRNA gene amplicons: Evaluating the illumina platform. PLoS ONE 2015, 10, e0116955. [Google Scholar] [CrossRef]
  34. Walters, W.; Hyde, E.R.; Berg-Lyons, D.; Ackermann, G.; Humphrey, G.; Parada, A.; Gilbert, J.A.; Jansson, J.K.; Caporaso, J.G.; Fuhrman, J.A.; et al. Improved bacterial 16S rRNA gene (V4 and V4–5) and fungal internal transcribed spacer marker gene primers for microbial community surveys. Msystems 2015, 1, e00009-15. [Google Scholar] [CrossRef]
  35. Edgar, R. UPARSE: Highly accurate OTU sequences from microbial amplicon reads. Nat. Methods 2013, 10, 996–998. [Google Scholar] [CrossRef] [PubMed]
  36. Roberts, D.W.; Roberts, M.D.W. Package ‘labdsv’. Ordination Multivar. 2016, 775, 1–68. [Google Scholar]
  37. Kader, M.A.; Singha, A.; Begum, M.A.; Jewel, A.; Khan, F.H.; Khan, N.I. Mulching as water-saving technique in dryland agriculture. Bull. Natl. Res. Cent. 2019, 43, 147. [Google Scholar] [CrossRef]
  38. Hanson, B.D.; Gerik, J.S.; Schneider, S.M. Effects of reduced-rate methyl bromide applications under conventional and virtually impermeable plastic film in perennial crop field nurseries. Pest Manag. Sci. 2010, 66, 892–899. [Google Scholar] [CrossRef] [PubMed]
  39. Yin, W.; Feng, F.; Zhao, C.; Yu, A.; Hu, F.; Chai, Q.; Gan, Y.; Guo, Y. Integrated double mulching practices optimizes soil temperature and improves soil water utilization in arid environments. Int. J. Biometeorol. 2016, 60, 1423–1437. [Google Scholar] [CrossRef]
  40. Qi, R.; Jones, D.L.; Li, Z.; Liu, Q.; Yan, C. Behavior of microplastics and plastic film residues in the soil environment: A critical review. Sci. Total Environ. 2020, 703, 134722. [Google Scholar] [CrossRef]
  41. Tofanelli, M.B.D.; Wortman, S.E. Benchmarking the agronomic performance of biodegradable mulches against polyethylene mulch film: A meta-analysis. Agronomy 2020, 10, 1618. [Google Scholar] [CrossRef]
  42. Wang, Z.; Li, M.; Flury, M.; Schaeffer, S.M.; Chang, Y.; Tao, Z.; Jia, Z.; Li, S.; Ding, F.; Wang, J. Agronomic performance of polyethylene and biodegradable plastic film mulches in a maize cropping system in a humid continental climate. Sci. Total Environ. 2021, 786, 147460. [Google Scholar] [CrossRef]
  43. Li, F.M.; Song, Q.H.; Jjemba, P.K.; Shi, Y.C. Dynamics of soil microbial biomass C and soil fertility in cropland mulched with plastic film in a semiarid agro-ecosystem. Soil Biol. Biochem. 2004, 3611, 1893–1902. [Google Scholar] [CrossRef]
  44. Liang, J.; Zhang, J.; Yao, Z.; Luo, S.; Tian, L.; Tian, C.; Sun, Y. Preliminary findings of polypropylene carbonate (PPC) plastic film mulching effects on the soil microbial community. Agriculture 2022, 12, 406. [Google Scholar] [CrossRef]
  45. Kapanen, A.; Schettini, E.; Giuliano, V.; Itävaara, M. Performance and environmental impact of biodegradable films in agriculture: A field study on protected cultivation. J. Polym. Environ. 2008, 16, 109–122. [Google Scholar] [CrossRef]
  46. Moreno, M.M.; Moreno, A. Effect of different biodegradable and polyethylene mulches on soil properties and production in a tomato crop. Sci. Hortic. 2008, 116, 256–263. [Google Scholar] [CrossRef]
  47. Sintim, H.Y.; Bandopadhyay, S.; English, M.E.; Bary, A.I.; DeBruyn, J.M.; Schaeffer, S.M.; Miles, C.A.; Reganold, J.P.; Flury, M. Impacts of biodegradable plastic mulches on soil health. Agric. Ecosyst. Environ. 2019, 273, 36–49. [Google Scholar] [CrossRef]
  48. Zhang, M.; Xue, Y.; Jin, T.; Zhang, K.; Li, Z.; Sun, C.; Mi, Q.; Li, Q. Effect of long-term biodegradable film mulch on soil physicochemical and microbial properties. Toxics 2022, 10, 129. [Google Scholar] [CrossRef] [PubMed]
  49. Yang, W.; Zhou, H.; Gu, J.; Liao, B.; Zhang, J.; Wu, P. Application of rapeseed residue increases soil organic matter, microbial biomass, and enzyme activity and mitigates cadmium pollution risk in paddy fields. Environ. Pollut. 2020, 264, 114681. [Google Scholar] [CrossRef] [PubMed]
  50. Qi, R.; Li, J.; Lin, Z.; Li, Z.; Li, Y.; Yang, X.; Zhang, J.; Zhao, B. Temperature effects on soil organic carbon, soil labile organic carbon fractions, and soil enzyme activities under long-term fertilization regimes. Appl. Soil Ecol. 2016, 102, 36–45. [Google Scholar] [CrossRef]
  51. Nannipieri, P.; Giagnoni, L.; Renella, G.; Puglisi, E.; Ceccanti, B.; Masciandaro, G.; Fornasier, F.; Moscatelli, M.C.; Marinari, S. Soil enzymology: Classical and molecular approaches. Biol. Fertil. Soils 2012, 48, 743–762. [Google Scholar] [CrossRef]
  52. Ramakrishna, A.; Tam, H.M.; Wani, S.P.; Long, T.D. Effect of mulch on soil temperature, moisture, weed infestation and yield of groundnut in northern Vietnam. Field Crops Res. 2006, 95, 115–125. [Google Scholar] [CrossRef]
  53. Wang, X.; Li, Z.; Xing, Y. Effects of mulching and nitrogen on soil temperature, water content, nitrate-N content and maize yield in the Loess Plateau of China. Agric. Water Manag. 2015, 161, 53–64. [Google Scholar]
  54. Müller, C.A.; Perz, V.; Provasnek, C.; Quartinello, F.; Guebitz, G.M.; Berg, G. Discovery of polyesterases from moss-associated microorganisms. Appl. Environ. Microbiol. 2017, 83, e02641-16. [Google Scholar] [CrossRef] [PubMed]
  55. Perz, V.; Hromic, A.; Baumschlager, A.; Steinkellner, G.; Pavkov-Keller, T.; Gruber, K.; Bleymaier, K.; Zitzenbacher, S.; Zankel, A.; Mayrhofer, C.; et al. An esterase from anaerobic Clostridium hathewayi can hydrolyze aliphatic-aromatic polyesters. Environ. Sci. Technol. 2016, 50, 2899–2907. [Google Scholar] [CrossRef] [PubMed]
  56. Bandopadhyay, S.; Sintim, H.Y.; Debruyn, J.M. Structural and functional responses of soil microbial communities to biodegradable plastic film mulching in two agroecosystems. PeerJ 2020, 8, e9015. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The Venn diagram of OTU numbers in the bacterial and fungal communities under three mulching treatments.
Figure 1. The Venn diagram of OTU numbers in the bacterial and fungal communities under three mulching treatments.
Agronomy 13 01220 g001
Figure 2. The composition and relative abundance of the bacterial community at phylum and genus levels under three mulching treatments.
Figure 2. The composition and relative abundance of the bacterial community at phylum and genus levels under three mulching treatments.
Agronomy 13 01220 g002
Figure 3. Composition and relative abundance of the fungal community at phylum and genus levels under three mulching treatments.
Figure 3. Composition and relative abundance of the fungal community at phylum and genus levels under three mulching treatments.
Agronomy 13 01220 g003
Table 1. Three mulching treatments during two consecutive years of the experiment (2021–2022).
Table 1. Three mulching treatments during two consecutive years of the experiment (2021–2022).
TreatmentFirst Year (2021)Second Year (2022)
PEM-PEMPEM covered (removed after cabbage harvest)PEM covered (removed after cabbage harvest)
PEM-BDMPEM covered (removed after cabbage harvest)BDM covered (tilled after cabbage harvest)
BDM-BDMBDM covered (tilled after cabbage harvest)BDM covered (tilled after cabbage harvest)
Table 2. The agronomic and nutritional quality traits of cabbage under three mulching treatments.
Table 2. The agronomic and nutritional quality traits of cabbage under three mulching treatments.
TraitPEM-PEMPEM-BDMBDM-BDM
Head weight (kg)2.32 ± 0.16 a2.21 ± 0.12 a2.19 ± 0.08 a
Head vertical diameter (cm)17.50 ± 1.06 a18.37 ± 0.85 a19.07 ± 1.26 a
Head transverse diameter (cm)20.50 ± 0.66 a19.50 ± 1.35 a18.90 ± 1.67 a
Core length (cm)6.77 ± 0.91 a7.20 ± 0.46 a6.97 ± 0.57 a
Core width (cm)3.37 ± 0.40 a3.53 ± 0.32 a3.67 ± 0.31 a
Yield (ton/ha)87.84 ± 1.31 a84.68 ± 1.76 a86.74 ± 1.55 a
Total soluble solids (%)5.86 ± 0.11 a5.57 ± 0.32 a6.05 ± 0.22 a
Soluble sugar (mg/g DW)251.70 ± 14.88 a245.30 ± 13.12 a254.13 ± 1.67 a
Soluble protein (mg/g FW)1.32 ± 0.12 a1.43 ± 0.14 a1.57 ± 0.10 a
Vitamin C (mg/100g FW)74.66 ± 4.21 a78.13 ± 5.87 a81.07 ± 3.98 a
Note: The data were expressed as means ± SD. Different lowercase letters indicate significant differences (p < 0.05).
Table 3. The physicochemical properties of the soil samples under three mulching treatments.
Table 3. The physicochemical properties of the soil samples under three mulching treatments.
TreatmentpH ValueNitrate Nitrogen (mg/kg)Available Phosphorus (mg/kg)Available Potassium (mg/kg)Organic Matter (g/kg)Total Nitrogen (%)
PEM-PEM5.93 ± 0.29 a37.20 ± 6.06 a24.00 ± 4.03 a64.33 ± 9.50 a24.03 ± 0.55 b0.18 ± 0.04 a
PEM-BDM5.27 ± 0.31 b30.47 ± 6.23 a33.57 ± 6.84 a59.33 ± 7.02 a30.47 ± 1.42 a0.20 ± 0.03 a
BDM-BDM5.27 ± 0.23 b33.37 ± 7.60 a30.93 ± 6.34 a56.00 ± 8.72 a32.23 ± 1.44 a0.21 ± 0.03 a
Note: The data were expressed as means ± SD. Different lowercase letters indicate significant differences (p < 0.05).
Table 4. The alpha diversity indices of bacterial and fungal communities under three mulching treatments.
Table 4. The alpha diversity indices of bacterial and fungal communities under three mulching treatments.
TypeSampleACEChao1SimpsonShannon
BacteriaPEM-PEM2192.48 ± 4.89 a2201.92 ± 7.68 a0.0022 ± 0.0001 a6.87 ± 0.01 a
PEM-BDM2191.97 ± 20.84 a2200.62 ± 18.80 a0.0022 ± 0.0003 a6.87 ± 0.07 a
BDM-BDM2178.28 ± 20.62 a2191.58 ± 15.57 a0.0026 ± 0.0002 a6.75 ± 0.06 a
FungiPEM-PEM843.36 ± 17.82 b849.02 ± 15.89 a0.0433 ± 0.0133 a4.58 ± 0.08 a
PEM-BDM877.66 ± 13.75 a884.97 ± 7.82 a0.0820 ± 0.0694 a4.26 ± 0.76 a
BDM-BDM863.75 ± 15.86 ab877.72 ± 29.57 a0.0553 ± 0.0051 a4.44 ± 0.14 a
Note: The data were expressed as means ± SD. Different lowercase letters indicate significant differences (p < 0.05).
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

Zhang, W.; Ma, J.; Cui, Z.; Xu, L.; Liu, Q.; Li, J.; Wang, S.; Zeng, X. Effects of Biodegradable Plastic Mulch Film on Cabbage Agronomic and Nutritional Quality Traits, Soil Physicochemical Properties and Microbial Communities. Agronomy 2023, 13, 1220. https://doi.org/10.3390/agronomy13051220

AMA Style

Zhang W, Ma J, Cui Z, Xu L, Liu Q, Li J, Wang S, Zeng X. Effects of Biodegradable Plastic Mulch Film on Cabbage Agronomic and Nutritional Quality Traits, Soil Physicochemical Properties and Microbial Communities. Agronomy. 2023; 13(5):1220. https://doi.org/10.3390/agronomy13051220

Chicago/Turabian Style

Zhang, Wei, Jinjun Ma, Zhongli Cui, Langtao Xu, Qian Liu, Jianbin Li, Shenyun Wang, and Xiaoping Zeng. 2023. "Effects of Biodegradable Plastic Mulch Film on Cabbage Agronomic and Nutritional Quality Traits, Soil Physicochemical Properties and Microbial Communities" Agronomy 13, no. 5: 1220. https://doi.org/10.3390/agronomy13051220

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop