Next Article in Journal
Ecological Paradox in the Reuse of Internal Combustion Engines from Scrapped Vehicles for Electric Power Generation—Circular Economy Potential Versus Emission Certification Barriers
Previous Article in Journal
Towards an Impact Performance Measurement Approach for Impact Investing: Results from a Benchmarking Study for Credit Finance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Decoding the Sustainability Code: Enzyme Thermodynamic and Kinetic Parameters Reveal the Efficacy of Straw, Biochar, and Nanocarbon in Black Soil

College of Life Science and Technology, Harbin Normal University, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Sustainability 2025, 17(23), 10436; https://doi.org/10.3390/su172310436
Submission received: 9 October 2025 / Revised: 10 November 2025 / Accepted: 13 November 2025 / Published: 21 November 2025

Abstract

For sustainable soil management, the link between carbon amendment structure and soil health is paramount, yet how the particle size of carbon governs hydrolase activity through kinetic and thermodynamic mechanisms remains poorly understood. A three-year field experiment with four treatments, including Control, Straw, Biochar, and Nanocarbon, was conducted in black soil. After harvest, the activities of invertase (INV), urease (URE), and acid phosphatase (ACP) were assayed from 15 to 55 °C. Kinetic parameters—including half-saturation constant (Km), maximal reaction rate (Vmax) and catalytic efficiency (Ka)—and thermodynamic parameters—including Gibbs free energy (ΔG), enthalpy (ΔH) and entropy (ΔS)—were determined. INV and ACP activities increased with temperature, peaking at 55 °C, whereas URE peaked at 45 °C. The Vmax, Ka, and ΔG of the enzymes also increased with temperature. With straw, INV activity remained stable, whereas INV-Ka, INV-ΔH, and INV-ΔS increased with decreased INV-Km. URE activity declined with thermodynamic elevation. For ACP, ACP-Km and ACP-Vmax increased, whereas ACP-Ka and ACP-ΔG decreased. With biochar or nanocarbon, the enzyme activities, Vmax, and Ka decreased, whereas ∆G increased, with stronger inhibition by nanocarbon. Correlation analysis revealed ∆G as the dominant factor for activity after carbon addition, while redundancy analysis identified organic carbon (OC) and total phosphorus (TP) as the key regulators. Overall, straw, biochar, and nanocarbon had different sustainable values on hydrolase systems, with thermodynamic parameters, especially ∆G, better reflecting system shifts than kinetic traits.

1. Introduction

Long-term excessive application of nitrogen (N) fertilizer is one of the key drivers of soil degradation, such as soil acidification, reduced soil organic matter, and C/N ratio imbalance, this not only restricts the exertion of soil ecological functions but also poses a challenge to the sustainable development of agriculture [1,2,3]. Previous studies have demonstrated that the incorporation of exogenous organic carbon can effectively mitigate the decline in the soil C/N ratio and the disruption of elemental balance resulting from prolonged excessive application of nitrogen fertilizer [4]. Straw is one of the most important agricultural byproducts for soil improvement. Composed primarily of cellulose, hemicellulose, lignin, and other macromolecular carbon sources, straw plays a critical role in regulating the soil carbon cycle and sustaining microbial activity [5]. Lignin contributes to stable organic carbon pools through its aromatic structure, while cellulose and hemicellulose serving as degradable components stimulate short-term microbial proliferation [6]. The release of these compounds supplies energy to microorganisms and accelerates aggregate formation. For example, three consecutive years of straw return increased the proportion of large aggregates (>2 mm) by 27% and increased the organic carbon content by 15.6% [7]. Biochar is produced by pyrolyzing straw at 300–500 °C. Its surface area is 50–300 times greater than that of straw [8,9]. Owing to its porous microstructure, biochar exhibits high stability, adsorption, and retention capacity, thereby improving soil nutrient dynamics [10,11]. In soil remediation, biochar reduces atmospheric CO2, enhances fertility, promotes microbial activity, and strengthens plant resistance to diseases [12]. Once introduced into the environment, biochar may degrade into nanoparticles ≤ 100 nm [13,14,15]. These nanocarbon particles possess graphene-like sp2 skeletons, abundant oxygen-containing functional groups, and hierarchical porosity. Nanocarbon surface area is 5–8 fold higher than biochar’s, ranging from 800–1500 m2/g, and resulting in strong adsorption and electron transfer capacities [16,17]. These properties suggest significant potential for environmental remediation. However, current applications of nanocarbons focus mainly on energy storage (e.g., lithium-ion batteries) and catalysis (e.g., improving hydrogen production efficiency), while their transformation mechanisms and agronomic effects in soil systems remain unclear [18,19]. Therefore, further research is required to clarify its practical value in agricultural production.
Soil enzymes are essential drivers of soil biogeochemical cycles [20,21]. As “biosensors” of ecosystem functioning, enzyme activities respond sensitively to environmental changes including land use, pollution stress, and climate variation, while simultaneously influencing soil fertility and carbon storage capacity [22,23]. Compared with oxidoreductases and lyases, hydrolases exhibit unique traits by directly mediating immediate nutrient supply. They cleave glycosidic, peptide, and ester bonds in organic compounds, converting complex substrates into bioavailable small molecules for plants and microorganisms [24,25]. This property renders hydrolase activity as one of the most sensitive indicators of nutrient turnover efficiency [26]. In recent years, studies on exogenous carbon inputs and hydrolase activity have expanded significantly, driven by the demands for global carbon cycling and fertility management. Distinct effects of carbon sources with different particle sizes have been documented [27,28]. For instance, Wu et al. [29] observed that straw increased β-glucosidase, cellulase, urease, and alkaline phosphatase activities in bauxite residues by 7.2–9.1, 5.8–7.1, 11.1–12.5, and 1.1–2.2 folds, respectively. Chen et al. [30] reported that straw markedly enhanced α-glucosidase, β-glucosidase, cellulase, leucine aminopeptidase, and N-acetylglucosaminidase activities. Jaffar et al. [31,32] demonstrated that biochar from citrus peel and sugarcane bagasse activated soil urease and phosphatases. However, biochar effects are context-dependent and strongly influenced by pyrolysis temperature. One study suggested that most enzyme activities decreased as pyrolysis temperature increased, with biochar prepared at 200 °C being more favorable for fertility [33]. Nanocarbons applied with fertilizers have also been shown to increase soil urease, catalase, phosphatase, and sucrase activities [34]. Graphite nano-additives (GNA) significantly enhance carbon-related enzyme activities in rhizosphere soils [35]. Nonetheless, most investigations of novel carbon materials are limited to short-term, small-scale experiments, restricting the comprehensive evaluation of their practical potential.
Soil enzyme kinetics and thermodynamics provide quantitative methods for describing catalytic and energy changes [36]. Evaluating these parameters yields comprehensive insights into intrinsic enzyme properties and resistance to environmental stress, thereby advancing our understanding of enzyme stability [37,38]. Research on how exogenous carbon affects kinetic and thermodynamic characteristics of soil hydrolases remains limited. Available studies have shown that the addition of exogenous carbon can alter kinetic parameters (Km, Vmax, and Ka) and thermodynamic properties (ΔG, ΔH, and ΔS). For example, Guo et al. [39] found that straw, biochar, and nanocarbon all increased the Vmax and Ka of soil catalase, while decreasing its Km, ΔG, ΔH, and ΔS. In the study by Chintala et al. [40], biochar application led to a significant decline in the kinetic indices (Ka, Vmax) of soil denitrifying enzymes, whereas their ΔH was markedly increased. Additionally, Yan et al. [41] found that biochar addition led to a higher Km for soil invertase in grey desert soil, with the Vmax of alkaline phosphatase and catalase being notably enhanced; in contrast, in aeolian soil, there was a distinct reduction in the Km, Vmax, and Ka of urease, phosphatase, invertase, and catalase. However, systematic comparisons of macro-sized straw, micro-sized biochar, and nano-sized carbon materials remain scarce, leaving a knowledge gap in understanding how carbon source properties govern hydrolase responses.
Our previous work has reported that sustained warming and addition of straw, biochar, and nanocarbon can promote catalase activity and its kinetic and thermodynamic processes in black soil [39]. However, their effects on hydrolases remain unclear. Here, we selected invertase (INV), urease (URE), and acid phosphatase (ACP) to examine activity, kinetics, and thermodynamics under straw, biochar, and nanocarbon treatments in cold-region black soil in northeast China. Three hypotheses were proposed as follows. (1) Straw, biochar, and nanocarbon each demonstrated the ability to enhance soil hydrolase activity. (2) Enzyme responses to different particle sizes of carbon addition were all regulated by both kinetic and thermodynamic processes. (3) Changes in hydrolase activity were closely related to ∆G and soil physicochemical properties after different carbon sources addition. This study can further supplement and improve the theory of soil carbon cycling in low-temperature environments, and at the same time provide certain scientific guidance for black soil fertility maintenance and sustainable agricultural development.

2. Materials and Methods

2.1. Experimental Site

The experiment was conducted at the long-term fertilizer field experiment station at Harbin Normal University, Heilongjiang Province, China (45°51′ N, 126°33′ E). The area has a temperate continental monsoon climate, with mean annual temperature of −1 °C, a frost-free period of 110 d, and annual precipitation of 450 mm. The adopted soil type was Haplic Chernozem. The test field previously used for non-agricultural purposes was replanted with a 15-year rotation cycle, and tillage followed a maize–soybean rotation system.

2.2. Experimental Materials

The materials included corn straw (C content: 40%), corn straw biochar (C content: 79.68%), and nanocarbon (C content: 14.06%). Corn straw was pre-treated by the experimenters themselves, it from the previous season was cut into 1 cm × 1 cm fragments. Corn straw biochar was procured from Liyang Yuanfeng Activated Carbon Co., Ltd. (Liyang, China), prepared by carbonizing corn straw at 600 °C. Nanocarbon was procured from Liyang Yuanfeng Activated Carbon Co., Ltd. (Liyang, China), it consisted of granular C blended with corn straw, columnar activated C, coconut sheath activated C, and nano mineral crystals. Fertilizers included urea (N ≥ 46.4%) and Red Square compound fertilizer (N + P + K ≥ 45%; Zhongyang Red Square Co., Ltd., Hefei, China).

2.3. Experiment Preparation and Design

A randomized block design was applied, with 12 test plots of 24 m2 each. Four treatments were established: (1) control with no carbon source (CK), (2) straw addition only (Straw), (3) biochar addition only (Biochar), and (4) nanocarbon addition only (Nanocarbon). Each treatment was replicated three times. A soybean-corn rotation system was employed, in which the fields were preceded by corn from 2022 to 2023 and were planted with soybean in 2024. Based on the above-ground corn straw yield (6 t ha−1 dry weight) and its air-dried carbon concentration (40.3%), the total carbon input from full stover return was estimated at ≈2500 kg·hm−2 (6000 kg × 0.403). Using this value as the target carbon-equivalent input (2500 kg·hm−2). After the autumn harvest in 2022–2023, C sources were applied for two consecutive years at a rate equivalent to 2500 kg·hm−2 of carbon. Accordingly, 15 kg of straw, 7.53 kg of biochar, and 42.67 kg of nanocarbon were added to each plot. In May 2024, the basic physical and chemical properties of the topsoil (0–20 cm) were initially determined: pH 6.51, organic carbon 9.26 g·kg−1, available nitrogen 57.83 mg·kg−1, available phosphorus 72.65 mg·kg−1, and available potassium 101.20 mg·kg−1. Then compound fertilizer (375 kg·hm−1; N:P2O5:K2O = 15:15:15) was applied before sowing and incorporated into the top 20 cm of soil. The CK treatment received only the base fertilizer with no C addition or topdressing during crop growth. Other field practices followed standard local agricultural management.

2.4. Soil Collection and Chemical Analysis

On the 28 September 2024, following the soybean harvest, soil samples (0–20 cm) were collected from each field plot. Three randomly selected sites within each plot were sampled, and crop residues were removed before collection. The samples were placed in containers and transported to the laboratory, where they were divided into two parts. One portion was air-dried and sieved (2 mm) for chemical property analysis, whereas the other portion was sieved (1 mm) and stored at −20 °C for enzyme activity determination.
The soil pH was measured in a 1:2.5 soil-to-water suspension using a digital pH meter (STARTER 300, OHAUS, Shanghai, China). The bulk density (BD) was determined using the cutting ring method based on the weight of the air-dried soil per unit volume. Soil organic carbon (OC) was determined using the potassium dichromate volumetric method, and total nitrogen (TN) was determined using the Kjeldahl method. Available nitrogen (AN) was measured using alkaline hydrolysis diffusion. Total phosphorus (TP) was determined using the molybdenum-antimony colorimetric method after H2SO4-HClO4 digestion. Available phosphorus (AP) was extracted with 0.5 mol·L−1 sodium bicarbonate and quantified by the molybdenum blue-ascorbic acid method. All analyses were conducted in triplicate. These soil physicochemical properties were determined using a reported method [42].

2.5. Enzyme Activity Assay

The activities of INV, URE, and ACP were determined using a reported method [26]. Enzyme activities were measured at five temperatures (15, 25, 35, 45, and 55 °C), with 35 °C as the optimum. Each measurement was conducted with three replicates to ensure the robustness of the experimental findings. The INV activity was quantified using the 3,5-dinitrosalicylic acid colorimetric method. Briefly, 2 g of soil was incubated for 24 h at different temperatures with 1 mL toluene, 10 mL of sucrose solution at varying concentrations (0.01, 0.05, 0.1, 0.2 mol·L−1), and 10 mL acetate buffer (pH 5.0). After incubation, shaking, and filtration, 2 mL of the filtrate was mixed with 3 mL of 3,5-dinitrosalicylic acid and boiled for 10 min, and the optical density was measured by spectrophotometer (UV-2000, UNIC, Shanghai, China) at 508 nm.
The URE activity was determined using the indophenol blue method. The soil (2 g) was incubated with 1 mL toluene, 10 mL of urea solution at different concentrations (0.1, 0.2, 0.4, 0.8 mol·L−1), and 10 mL citrate buffer (pH 6.7) for 24 h at different temperatures. After incubation, shaking, and filtration, 10 mL of the filtrate was mixed with 4 mL of sodium phenolate and 3 mL of sodium hypochlorite solution, and the optical density was measured by spectrophotometer (UV-2000, UNIC, Shanghai, China) at 578 nm.
ACP activity was measured using the sodium phenyl phosphate colorimetric method. Specifically, 2 g of soil was incubated with 0.2 mL toluene, 10 mL of disodium phenyl phosphate hydrate solution at varying concentrations (5, 10, 20, and 40 mmol·L−1) and 10 mL of acetate buffer (pH 5.0) for 24 h at different temperatures. After incubation, shaking, and filtration, 2 mL of the filtrate was mixed with 5 mL borate buffer (pH 9.6) and 1 mL Gibb’s reagent, and optical density was measured by spectrophotometer (UV-2000, UNIC, Shanghai, China) at 578 nm.

2.6. Calculations and Statistical Analyses

The half-saturation constant (Km) indicates substrate affinity, with smaller values reflecting a higher affinity. Vmax represents the maximal reaction rate as enzyme-substrate complexes decompose, thus indicating total enzyme concentration and production. Catalytic efficiency (Ka), defined as Vmax/Km, reflects enzymatic performance. Thermodynamic parameters, such as ∆G, ∆H, and ∆S, represent the Gibbs free energy, enthalpy, and entropy, respectively, which indicate spontaneity, structural change, and transition-state disorder [36].
The enzyme kinetic parameters (Km and Vmax) were calculated by measuring the enzyme activity at different substrate concentrations using the Michaelis-Menten equation [36,43]:
V 0   =   V max · [ S ] K m   +   [ S ]
where V0 is the initial reaction velocity (mg product release g−1 soil 24 h−1), [S] is the substrate concentration (mol·L−1 or mmol·L−1), Km is the Michaelis-Menten constant (mol·L−1 or mmol·L−1), and Vmax is the maximum velocity of the enzymatic reaction [mg·(g·24 h)−1]. Catalytic efficiency (Ka) was calculated as the ratio of Vmax to Km.
The potential enzyme activity was measured at 15, 25, 35, 45, and 55 °C. The Arrhenius equation was used to determine ∆G, ∆H, and ∆S of INV, URE, and ACP [36]:
lnk   =   lnA     E a RT
G = RT ln RT Nhk
where k represents the enzymatic reaction constant at a given temperature, A represents the pre-exponential factor, Ea represents the activation energy (kJ·mol−1), R represents the gas constant (8.314 J·mol−1·K−1), T represents the absolute temperature (Kelvin) 273 + t (K), ∆G represents the activation free energy (kJ·mol−1), N represents the Avogadro constant (6.023·1023 mol−1), h represents the Planck constant (6.626·10−34 J·s).
H =   E a     RT
where ∆H represents the activation enthalpy (kJ·mol−1), Ea represents the activation energy (kJ·mol−1), R represents the gas constant (8.314 J·mol−1·K−1), T represents the absolute temperature (Kelvin) 273 + t (K).
S = H G T
where ∆S represents the activation entropy (J·mol−1), ∆H represents the activation enthalpy (kJ·mol−1), ∆G represents the activation free energy (kJ·mol−1), T represents the absolute temperature (Kelvin) 273 + t (K).
Normality was tested using the Kolmogorov–Smirnov test, variance homogeneity with Levene’s test, and correlation analysis with SPSS 27 (IBM, Armonk, NY, USA). The treatment differences were assessed by two-factor ANOVA, and the means were compared using the least significant difference (LSD) test at p < 0.05. The variance was expressed using Waller Duncan’s method. Graphs were generated using Origin 2021 (Origin Lab Company, Northampton, MA, USA), and redundancy analysis was performed using Canoco 5 (Microcomputer Power, Ithaca, NY, USA).

3. Results

3.1. Soil Physicochemical Properties

Physicochemical characteristics of soil samples are listed in Table 1. Straw, biochar, and nanocarbon did not alter soil pH but markedly reduced soil bulk density (BD) (p < 0.05). Total phosphorus (TP) significantly increased under all carbon treatments. After straw addition, soil organic carbon (OC), total nitrogen (TN), available nitrogen (AN) was notably elevated. Following biochar amendment, AN and available phosphorus (AP) increased substantially, while TN declined markedly. With nanocarbon addition, only AP rose significantly, whereas OC and TN all decreased (p < 0.05). No appreciable changes were observed in other parameters.

3.2. Soil Enzyme Activities

Temperature, carbon source, and their interaction had significant effects on the activities of the three hydrolases (p < 0.05, Figure 1). With increasing temperature, invertase (INV) and acid phosphatase (ACP) activities consistently increased under all treatments, reaching a maximum at 55 °C. Urease (URE) activity first increased and then declined, with a peak at 45 °C.
At each temperature tested, straw markedly reduced URE activity but increased ACP activity (p < 0.05), whereas INV activity exhibited no clear change. Biochar and nanocarbon both decreased the activities of all three hydrolases, with nanocarbon showing stronger inhibition than biochar (Figure 1).

3.3. Soil Enzyme Kinetic Parameters

Carbon sources and the interaction of temperature with carbon sources had no pronounced effects on INV-Km and URE-Km, whereas ACP-Ka remained unaffected. Marked effects were observed for other parameters (p < 0.05, Figure 2). Across treatments, as temperature increased, INV-Km decreased and then increased, reaching the minimum at 35 °C; URE-Km and ACP-Km increased and then decreased, peaking at 25 °C. INV-Vmax and ACP-Vmax increased with temperature, whereas URE-Vmax first increased and then declined, with the maximum at 45 °C. INV-Ka and URE-Ka demonstrated an increasing–decreasing trend, peaking at 45 °C, whereas ACP-Ka continued to increase.
Under straw addition, ACP-Km showed an appreciable increase compared with CK (p < 0.05). INV-Vmax and URE-Vmax exhibited no pronounced changes, whereas ACP-Vmax increased significantly (p < 0.05). INV-Ka increased at 15–45 °C, while no significant change was observed at 55 °C. URE-Ka decreased to a certain extent at 25–55 °C. For ACP-Ka, it decreased at 15–35 °C but showed no obvious change at 45–55 °C.
With biochar or nanocarbon, ACP-Km increased pronouncedly compared with CK, the Vmax values of all three hydrolases and the Ka values of INV and URE significantly decreased (p < 0.05), with a stronger reduction under nanocarbon. ACP-Ka decreased at 15–35 °C.

3.4. Soil Enzyme Thermodynamic Parameters

Figure 3 depicted the Eyring plots for three hydrolases after three carbon sources addition at 15–55 °C. Under different carbon source treatments, the ln(Ka/T) of the three hydrolases showed a linear decreasing trend with the increase of 1/T, the temperature dependence of hydrolases was consistent with the theoretical prediction of the Eyring equation when the temperature rises.
Further quantifying the thermodynamic parameters in depth, the temperature, carbon source, and their interaction showed a statistically meaningful effect on the ∆G of the three hydrolases (p < 0.05). Carbon sources markedly influenced ∆H and ∆S, whereas temperature and the interaction of temperature with carbon sources had no statistically significant effects (Figure 4). In all treatments, ∆G gradually increased with temperature, reaching a maximum at 55 °C.
After straw addition, INV-∆G showed no marked change, whereas ACP-∆G decreased notably at all tested temperatures compared with CK (p < 0.05). URE-∆G increased to a certain extent at 15–55 °C. Both ∆H and ∆S of INV and URE increased, whereas ACP-∆H and ACP-∆S decreased slightly.
With biochar or nanocarbon, the ∆G of all three hydrolases substantially increased at each tested temperature compared with CK (p < 0.05). For INV, both the ΔH and ΔS exhibited a moderate decrease, whereas for URE and ACP, both parameters notably exhibited a certain degree of elevation.

3.5. Correlation Analysis Between Hydrolase Activity and Catalytic Characteristics

After three carbon sources additions, the correlation analysis of three hydrolase activities with kinetic and thermodynamic parameters was shown in Figure 5. In CK treatment, the activities of three hydrolases all exhibited a distinct negative correlation with ΔG (p < 0.01). Additionally, URE activity showed a marked negative correlation with Ka, and ACP activity displayed a significant negative correlation with ΔS. After adding straw, biochar, and nanocarbon, the activities of three hydrolases only demonstrated a conspicuous negative correlation with ΔG (p < 0.01), showing a consistent correlation trend, and had no apparent relationship with other parameters.

3.6. Redundancy Analysis

Redundancy analysis indicated that Axis-1 and Axis-2 explained 53.93% and 23.71% of the total variation in the soil environmental factors and enzyme parameters, respectively (Figure 6). Compared with CK, distinct directional separations were observed in the hydrolase reaction system under different carbon source amendments. Soil physicochemical properties were significantly correlated with enzyme activities and with their kinetic and thermodynamic parameters. The relative contribution of soil physicochemical properties to the variations in enzyme characteristics followed the order OC > AP > TP > pH > AN > TN > BD. Among these, OC and AP exerted highly notable influences (p < 0.05), explaining 41.1% and 24.4% of the variation in carbon-treated soils, respectively (Figure 7).

4. Discussion

4.1. Response of Hydrolase Activities to Straw, Biochar, and Nanocarbon in Black Soil

Previous studies have reported that incorporating straw into saline-alkali soil increases INV, URE, and ACP activities at the optimal temperature (37 °C) [44,45,46]. Conversely, Ortega et al. [47] found that straw mulching in post-fire soil did not affect URE but reduced ACP. However, the effect of straw on hydrolase activity under warm conditions remains unclear. In this study, straw addition caused the INV and ACP activities to increase progressively with temperature (15–55 °C), peaking at 55 °C, while the URE activity increased and then decreased, peaking at 45 °C. This was consistent with observations without straw and followed the general rule that enzyme activity could increase with temperature within a certain range [48], indicating that straw could not alter the temperature response trend. Compared with CK, straw addition resulted in no significant change in INV activity, a decrease in URE, or an increase in ACP across all temperatures. These results are not completely consistent with previous findings [44,45]. This newly observed phenomenon suggests that the enhanced enzyme activity after straw incorporation is not absolute and may depend on straw properties, application rates, and the specific conditions of cold-region black soil [49,50].
Biochar and nanocarbon, characterized by high surface area and porosity, have been shown to improve soil structure and nutrient efficiency [51,52]. However, due to the significant variations in soil properties across different regions, the applicability of existing research findings in other areas remains to be verified. Research on the underlying internal mechanisms associated with these differences is also relatively scarce. Considering severe degradation and organic matter loss at the study site, a three-year field experiment was undertaken to assess how continuous biochar and nanocarbon application could influence hydrolase activity under warming. The results showed that as temperature increased (15–55 °C), the hydrolase activities under biochar and nanocarbon followed trends similar to straw, with no change in the overall temperature response. At all tested temperatures, biochar and nanocarbon reduced the enzyme activities, with nanocarbon exhibiting stronger inhibition. These results are not entirely consistent with our initial hypothesis that “carbon sources can enhance soil hydrolase activity,” suggesting that the influence of carbon sources on enzyme activity may be more complex. Our findings contrast with most previous reports that biochar and nanocarbon can stimulate enzyme activity [31,32,34]. The discrepancy may be due to their stable, inert, aromatic-rich structures [53,54]. In black soils of northeast China, such inert carbons may fail to provide energy to microorganisms, instead suppressing enzyme-substrate interactions through the adsorption of enzymes or substrates, leading to lower apparent activity [55]. Notably, nanocarbon caused greater inhibition than biochar, suggesting stronger sensitivity of hydrolases to nanocarbon. This may be explained by its larger surface area and smaller particle size, which enhances adsorption and coupling with soil nutrients and microelements. By preferentially adsorbing low-molecular-weight organics, nanocarbons likely reduce the availability of substrates to microbes, thereby indirectly lowering enzyme activity [56,57]. These findings offered a scientific basis for making informed decisions regarding the balanced management of straw incorporation alongside biochar and nanocarbon amendments.

4.2. Response of Hydrolase Kinetic and Thermodynamic Characteristics to Straw, Biochar, and Nanocarbon in Black Soil

The kinetic and thermodynamic properties of soil enzymes provide effective indicators of how exogenous carbon inputs can influence apparent activities. After straw addition, the changes in INV and ACP were opposite: Ka, ΔH, and ΔS of INV increased, while INV-Km decreased. The Km and Vmax of ACP increased, whereas Ka and ΔG decreased at all temperatures. These results suggested that INV stabilization was driven by accelerated kinetic reactions combined with slower thermodynamic processes, while ACP enhancement reflected increased enzyme production and concentration and reduced activation free energy. In contrast, the URE kinetic parameters remained unchanged, whereas the thermodynamic parameters increased. Thus, the reduced URE activity under straw addition was mainly associated with higher energy demand, enthalpy, and entropy. This indicated distinct response mechanisms among hydrolases, which could be consistent with the resource allocation theory of microbial metabolism [58,59]. Previous studies have confirmed that straw incorporation can modify enzyme kinetics and thermodynamics through multiple pathways, including altering substrate adsorption efficiency, modulating catalytic rates, affecting activation energy, and shifting enthalpy and entropy [39,60]. Although it remains difficult to determine whether straw universally enhances or suppresses enzyme activities, the observed measurable changes in kinetic and thermodynamic traits confirmed that hydrolase responses were shaped by both processes, supporting our second hypothesis.
With the addition of biochar or nanocarbon, the three hydrolases exhibited consistent responses, where ΔG increased, whereas Vmax and Ka decreased. Hence, the reduction in hydrolase activity reflected slower kinetic and thermodynamic processes, in agreement with our second hypothesis. Compared with straw, biochar and nanocarbon mainly decreased enzyme secretion and catalytic efficiency, increased activation energy barriers, and exerted similar effects on soil C, N, and P metabolism. Few studies have examined enzyme kinetics and thermodynamics under biochar and nanocarbon, and the reported trends remain inconsistent. For example, Raiesi found that biochar in sandy soil increased the Vmax and Ka of β-glucosidase while decreasing Km and thermodynamic parameters [61], whereas Khadem suggested that biochar increased the thermodynamic parameters of arylsulfatase in sandy soil [62]. In our recent study, both biochar and nanocarbon increased the Vmax and Ka of catalase but reduced Km, ΔG, ΔH, and ΔS in northeastern black soil [39]. Therefore, the responses of enzyme kinetic and thermodynamic parameters to biochar and nanocarbon appeared to vary depending on enzyme type and soil properties, which could explain the differences between our results and those reported previously.

4.3. Factors Influencing Changes in Soil Hydrolase Activity

Correlation analysis revealed a significant negative correlation between soil hydrolase activity and ΔG after the addition of straw, biochar, and nanocarbon, ΔG consistently served as a key regulatory factor. This suggests that the energy barrier of enzymatic reactions is the fundamental mechanism by which carbon sources alter soil enzyme functionality. This ΔG-dominated mechanism of carbon-mediated regulation of hydrolase activity supports our hypothesis. Particularly notable was the consistent increase in ΔG following carbon amendments, indicating a trend towards reduced spontaneous enzymatic reactions, in contrast with most previous studies. Our three-year experiment on continuously cultivated boreal black soil in northeast China suggested that the ΔG elevation may result from (1) the long-term cultivation-induced depletion of organic matter and microenvironments protecting enzymes, coupled with microbial imbalance, which may delay electron transformation and elevate transition-state energy barriers [63,64,65]; and (2) low temperatures restricting microbial metabolism, disrupting functional balance, and leading to insufficient enzyme synthesis and increased activation energy [66,67]. Collectively, these mechanisms contribute to the thermodynamic constraints observed in this unique agroecosystem. This discovery refines our traditional perspective on the regulation of soil enzyme activity by establishing a critical thermodynamic constraint for biogeochemical processes within global carbon cycle models. It thereby significantly improves the predictive accuracy of these models for carbon turnover dynamics in high-latitude agricultural ecosystems.
The activity of soil hydrolases is regulated by soil physicochemical properties, a conclusion that is further supported by redundancy analysis. The analysis indicated that in the unique boreal black soil ecosystem of northeast China, both OC and AP played critical roles in modulating the hydrolase system. Prolonged low temperatures and tillage practices substantially slowed OC mineralization rates, establishing OC as the primary limiting factor for hydrolase systems. OC functioned not only as an enzyme substrate and stabilizer but also reflected the carbon preservation mechanism typical of cold-region soils, where microbial activity was suppressed by low temperatures, resulting in the accumulation of partially decomposed organic matter [68]. In contrast, AP emphasized the unique phosphorus cycling dynamics of cold soils, where seasonal freeze-thaw cycles induced asynchrony between P fixation-release patterns and crop growth seasons. This positioned AP as the second major limiting factor after OC [69], directly constraining the energy metabolism required for microbial hydrolase synthesis.
Therefore, the activity of soil hydrolases is jointly determined by soil physicochemical properties, kinetic parameters, and thermodynamic parameters. Among these, thermodynamic factors, particularly ΔG of enzyme reactions, are critical determinants of enzyme activity changes. Although this study has achieved some progress, it still has several limitations. Moving forward, continuous monitoring of the seasonal fluctuation rules of soil enzyme reactions is required, along with further expansion of the research scope and systematic conduct of correlation studies between soil enzymes, crop yield, and nutrient use efficiency, so as to construct a more comprehensive soil science analysis system.

5. Conclusions

Based on investigating soil hydrolase activities, their kinetic and thermodynamic parameters after adding straw, biochar, and nanocarbon, the study conclusions are summarized as follows:
(1)
Although all carbon inputs did not alter the temperature response trends of INV, URE, and ACP, their effects on enzyme activities were distinct. Straw exerted different influences on the enzymes (no significant change in INV, a decrease in URE, and an increase in ACP), whereas both biochar and nanocarbon consistently inhibited all three enzymes, with nanocarbon exhibiting a stronger inhibitory effect.
(2)
The changes in hydrolase activity following carbon addition were regulated through the coupling of kinetic and thermodynamic processes.
(3)
Thermodynamic properties, particularly ∆G, were identified as the key limiting factors driving enzyme activity changes, whereas soil OC and TP acted as important physicochemical regulators.
These findings elucidate the mechanisms by which exogenous carbon inputs could influence enzymatic processes in cold-region black soils, underscoring the critical role of thermodynamic constraints and soil specificity. In addition, this study provides the scientific basis for designing carbon amendment strategies that maximize soil sustainability. By linking carbon particle size to enzymatic energy pathways, our work offers a framework for balancing short-term productivity with long-term soil health, guiding the development of resilient agricultural systems in cold-region black soil.

Author Contributions

Writing—original draft, data curation, formal analysis and validation, J.X.; writing—review and editing, X.W.; methodology, P.W.; sample collection and arrangement, J.Z.; project administration and funding acquisition, Z.Y. and X.B.; proofreading, J.L., Y.Y. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Provincial Universities of Heilongjiang, China in 2024, grant number 1305123257 and supported by Heilongjiang Provincial Natural Science Foundation of China, grant number JJ2024LH1618.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
INVInvertase
UREUrease
ACPPhosphatase
BDBulk density
OCOrganic carbon
TNTotal nitrogen
ANAvailable nitrogen
TPTotal phosphorus
APAvailable phosphorus

References

  1. Smerald, A.; Kraus, D.; Rahimi, J.; Fuchs, K.; Kiese, R.; Butterbach, B.K.; Scheer, C. A redistribution of nitrogen fertiliser across global croplands can help achieve food security within environmental boundaries. Commun. Earth Environ. 2023, 4, 315. [Google Scholar] [CrossRef]
  2. Zhao, Q.; Hao, X.Y.; Ali, I.; Iqbal, A.; Ullah, S.; Huang, M.; Kong, F.Y.; Li, T.Y.; Xuan, Y.; Li, F.Q.; et al. Characterization and grouping of all primary branches at various positions on a rice panicle based on grain growth dynamics. Agronomy 2020, 10, 223. [Google Scholar] [CrossRef]
  3. Shen, W.Z.; He, J.; Li, S.S.; Zhuang, Y.H.; Wang, H.Y.; Liu, H.B.; Zhang, L.; Kappler, A. Opportunity and shift of nitrogen use in China. Geogr. Sustain. 2024, 5, 33–40. [Google Scholar] [CrossRef]
  4. Li, Q.S.; Wang, D.Q.; Du, C.Y.; Li, X.D.; Zhang, G.C.; Wang, S.K.; Guan, E.S.; Wang, S.Q. Effects of exogenous carbon sources on nitrogen transformation and N2O emissions in tobacco-planting soil. Chin. Tob. Sci. 2020, 41, 13–19. [Google Scholar] [CrossRef]
  5. Jin, Z.Q.; Shah, T.; Zhang, L.; Liu, H.Y.; Peng, S.B.; Nie, L.X. Effect of straw returning on soil organic carbon in rice–wheat rotation system: A review. Food Energy Secur. 2020, 9, e200. [Google Scholar] [CrossRef]
  6. Rivka, B.F.; David, A.L.; Timothy, B.P. Impact of six lignocellulosic biochars on C and N dynamics of two contrasting soils. GCB Bioenergy 2017, 9, 1279–1291. [Google Scholar] [CrossRef]
  7. Zhang, P.; Chen, P.; Nie, T.Z.; Zhang, Z.X.; Li, T.C.; Dai, C.L.; Jiang, L.L.; Wu, Y.; Sun, Z.Y.; Yin, S. Long-term straw incorporation under controlled irrigation improves soil quality of paddy field and rice yield in northeast China. Plants 2024, 13, 1357. [Google Scholar] [CrossRef]
  8. Arshad, U.; Altaf, M.T.; Liaqat, W.; Ali, M.; Shah, M.N.; Jabran, M.; Ali, M.A. Biochar: Black gold for sustainable agriculture and fortification against plant pathogens—A review. J. Crop Health 2023, 76, 385–396. [Google Scholar] [CrossRef]
  9. Silvana, A.; Omar, G.P. Bamboo-based biochar: A still too little-studied black gold and its current applications. J. Xenobiotics 2024, 14, 416–451. [Google Scholar] [CrossRef]
  10. Schmidt, M.W.; Skjemstad, J.O.; Czimczik, C.I.; Glaser, B.; Prentice, K.M.; Gelinas, Y.; Thomas, A.J.K. Comparative analysis of black carbon in soils. Glob. Biogeochem. Cycles 2001, 15, 163–167. [Google Scholar] [CrossRef]
  11. Lehmann, J.; Rillig, M.C.; Thies, J.; Masiello, C.A.; Hockaday, W.C.; Crowley, D. Biochar effects on soil biota—A review. Soil Biol. Biochem. 2011, 43, 1812–1836. [Google Scholar] [CrossRef]
  12. Ahmed, I.O.; Mohamed, F.; Ahmed, K.R. Life cycle assessment of biochar as a green sorbent for soil remediation. Curr. Opin. Green Sustain. Chem. 2024, 46, 100882. [Google Scholar] [CrossRef]
  13. Wang, D.J.; Zhang, W.; Hao, X.Z.; Zhou, D.M. Transport of biochar particles in saturated granular media: Effects of pyrolysis temperature and particle size. Environ. Sci. Technol. 2013, 47, 821–828. [Google Scholar] [CrossRef]
  14. Spokas, K.A.; Novak, J.M.; Masiello, C.A. Physical disintegration of biochar: An overlooked process. Environ. Sci. Technol. Lett. 2014, 1, 326–332. [Google Scholar] [CrossRef]
  15. Shen, Y.; Tang, H.Y.; Wu, W.H.; Shang, H.P.; Zhang, D.; Zhan, X.H.; Xing, B.S. Role of nano-biochar in attenuating the allelopathic effect from Imperata cylindrica on rice seedlings. Environ. Sci. Nano 2019, 7, 116–126. [Google Scholar] [CrossRef]
  16. Xin, R.J.; Nam, H.N.; Phung, Q.M.; Tang, J.; Ma, S.C.; Markus, J.; Dai, Y.C.; Alowasheeir, A.; Khaorapapong, N.; Wang, J.; et al. Trimodal Hierarchical Porous Carbon Nanoplates with Edge Curvature for Faster Mass Transfer and Enhanced Oxygen Reduction. ACS Nano 2025, 19, 11648–11663. [Google Scholar] [CrossRef] [PubMed]
  17. Santhi, M.B.; Abhishek, H.; Akhil, R.; Mousumi, B.; Binoy, K.S. Hierarchical porous carbon derived from petroleum coke via one-step chemical activation for the fabrication of a supercapacitor and real time clock application. RSC Adv. 2024, 14, 21411–21424. [Google Scholar] [CrossRef]
  18. Wang, Y.P.; Liu, Y.; Zhang, H.Y.; Duan, X.G.; Ma, J.; Sun, H.Q.; Tian, W.J.; Wang, S.B. Carbonaceous materials in structural dimensions for advanced oxidation processes. Chem. Soc. Rev. 2025, 54, 2436–2482. [Google Scholar] [CrossRef]
  19. Kumar, D.; Abraham, J.E.; Varghese, M.; George, J.; Balachandran, M.; Cherusseri, J. Nanocarbon assisted green hydrogen production: Development and recent trends. Int. J. Hydrogen Energy 2024, 50, 118–141. [Google Scholar] [CrossRef]
  20. Sainju, U.M.; Liptzin, D.; Dangi, S.M. Enzyme activities as soil health indicators in relation to soil characteristics and crop production. Agrosyst. Geosci. Environ. 2022, 5, e20297. [Google Scholar] [CrossRef]
  21. Li, J.; Wu, J.Z.; Yu, J.Y.; Wang, K.B.; Li, J.P.; Cui, Y.X.; Shangguan, Z.P.; Deng, L. Soil enzyme activity and stoichiometry in response to precipitation changes in terrestrial ecosystems. Soil Biol. Biochem. 2024, 191, 109321. [Google Scholar] [CrossRef]
  22. Zhang, Y.L.; Wang, Y.S. Soil enzyme activities with greenhouse subsurface irrigation. Pedosphere 2006, 16, 512–518. [Google Scholar] [CrossRef]
  23. Jia, R.; Zhou, J.; Yang, L.; Blagodatskaya, E.; Jones, D.L.; Razavi, B.S.; Yang, Y.D.; Kuzyakov, Y.; Zeng, Z.H.; Zang, H.D. Trade-off between soil enzyme activities and hotspots area depends on long-term fertilization: In situ field zymography. Sci. Total Environ. 2024, 954, 176386. [Google Scholar] [CrossRef]
  24. Marx, M.C.; Kandeler, E.; Wood, M.; Wermbter, N.; Jarvis, S.C. Exploring the enzymatic landscape: Distribution and kinetics of hydrolytic enzymes in soil particle-size fractions. Soil Biol. Biochem. 2004, 37, 35–48. [Google Scholar] [CrossRef]
  25. Zhang, Y.Y.; Wang, X.; Sun, Y.X.; Wu, J.H.; Deng, T.; Yuan, M.H.; Duan, W.H.; Zhao, Y.F. Hydrolases control soil carbon sequestration in alpine grasslands in the Tibetan plateau. Sustainability 2024, 16, 3508. [Google Scholar] [CrossRef]
  26. Guan, S.Y. Soil Enzymes and Their Research Methods, 1st ed.; Chinese Agricultural Science and Technology Press: Beijing, China, 1986; pp. 19–21. [Google Scholar]
  27. Bai, M.X.; Situ, G.M.; Li, S.H.; Wu, Q.F.; Liang, C.F.; Qin, H.; Chen, J.H. Effects of combined application of biochar with organic amendments on enzyme activity and microbial metabolic function of carbon sources in infertile red soil. Chin. J. Appl. Ecol. 2022, 33, 1283–1290. [Google Scholar] [CrossRef]
  28. Shaon, K.D.; Goutam, K.G.; Ravikant, A.; Burhan, U.C.; Vinay, K.M.; Manik, C.K.; Aniruddha, R.; Tilak, M.; Achal, L.; Dhakre, D.S. Organic nutrient sources and biochar technology on microbial biomass carbon and soil enzyme activity in maize-black gram cropping system. Biomass Convers. Biorefin. 2023, 13, 9277–9287. [Google Scholar] [CrossRef]
  29. Wu, H.; Sun, W.; Zhu, F.; Jiang, Y.F.; Huang, S.W.; Goloran, J.; Xue, S.G. Straw addition increases enzyme activities and microbial carbon metabolism activities in bauxite residue. J. Environ. Sci. 2024, 135, 332–344. [Google Scholar] [CrossRef] [PubMed]
  30. Chen, X.; Tang, Z.W.; Weng, Y.; Cai, H.M.; Wu, Y.; Zheng, B.Q.; Li, J.C. Effects of 15-year straw incorporation on soil carbon composition and microbial community under wheat-maize rotation system in the Huang-Huai-Hai Plain. BMC Plant Biol. 2025, 25, 522. [Google Scholar] [CrossRef] [PubMed]
  31. Jaffar, M.T.; Zhen, S.A.; Han, J.L.; Zhang, J.G.; Dar, A.; Mushtaq, Z.; Hussain, Q.; Zahir, A.; Siddique, K.H.M. Orange peel biochar: An effective amendment to improve the maize resilience by regulating the soil enzymatic activities, nutrient uptake, and ionic homeostasis under salinity stress. Ind. Crops Prod. 2024, 222, 120081. [Google Scholar] [CrossRef]
  32. Jaffar, M.T.; Chang, W.Q.; Zhang, J.G.; Mukhtar, A.; Mushtaq, Z.; Ahmed, M.; Zahir, A.; Siddique, K.H.M. Sugarcane bagasse biochar boosts maize growth and yield in salt-affected soil by improving soil enzymatic activities. J. Environ. Manag. 2024, 363, 121418. [Google Scholar] [CrossRef]
  33. Rahmanian, M.; Khadem, A. The effects of biochar on soil extra and intracellular enzymes activity. Biomass Convers. Biorefin. 2023, 14, 21993–22005. [Google Scholar] [CrossRef]
  34. Chen, S.; Dong, C.W.; Gao, Y.; Li, Y.J.; Shi, Y. Effects of nanocarbon and nano-calcium carbonate on soil enzyme activities and soil microbial community in wheat (Triticum aestivum L.) rhizosphere soil. J. Plant Nutr. Soil Sci. 2023, 187, 639–652. [Google Scholar] [CrossRef]
  35. Das, P.; Barker, C.; Park, Y.; Perreault, F.; Westerhoff, P.; Penton, C.R. Impact of graphite nano amendments on soil enzyme activities, functional genes and microbiome composition in a soil-plant system. Soil Biol. Biochem. 2025, 203, 109714. [Google Scholar] [CrossRef]
  36. Zhu, M.E. Soil Enzyme Kinetics and Thermodynamics, 1st ed.; Science Press: Beijing, China, 2011; pp. 1–41. [Google Scholar]
  37. Stone, M.M.; Plante, A.F. Changes in phosphatase kinetics with soil depth across a variable tropical landscape. Soil Biol. Biochem. 2014, 71, 61–67. [Google Scholar] [CrossRef]
  38. Allison, S.D.; Romero-Olivares, A.L.; Lu, Y.; Taylor, J.W.; Treseder, K.K. Temperature sensitivities of extracellular enzyme Vmax and Km across thermal environments. Glob. Change Biol. 2018, 24, 2884–2897. [Google Scholar] [CrossRef]
  39. Guo, Y.X.; Zhao, J.Y.; Li, J.W.; Dong, Y.L.; Yue, Z.H.; Yin, Y.; Li, W.; Ren, Q.N.; Wu, X.Y. Straw, biochar, and nanocarbon altered the enzymatic reaction kinetics and thermodynamic process of catalase in the black soil under continuous warming. Arch. Agron. Soil Sci. 2023, 69, 3637–3650. [Google Scholar] [CrossRef]
  40. Chintala, R.; Owen, R.K.; Schumacher, T.E.; Spokas, K.A.; McDonald, L.M.; Kumar, S.; Clay, D.E.; Malo, D.D.; Bleakley, B. Denitrification kinetics in biomass- and biochar-amended soils of different landscape positions. Environ. Sci. Pollut. Res. Int. 2015, 22, 5152–5163. [Google Scholar] [CrossRef]
  41. Yan, Q.T.; Tian, H.X.; Huang, Y.; Mu, X.L.; Tang, G.M.; Ma, H.G.; Megharaj, M.; Xu, W.L.; He, W.X. Recycled wheat straw biochar enhances nutrient-poor soil: Enzymatic kinetics of carbon, nitrogen, and phosphorus cycling. J. Environ. Manag. 2025, 380, 124950. [Google Scholar] [CrossRef] [PubMed]
  42. Bao, S.D. Soil and Agricultural Chemistry Analysis, 3rd ed.; China Agriculture Press: Beijing, China, 2008; pp. 30–34, 42–58, 74–81. [Google Scholar]
  43. Tabatabai, M.A.; Bremner, J.M. Michaelis constants of soil enzymes. Soil Biol. Biochem. 1971, 3, 317–323. [Google Scholar] [CrossRef]
  44. Shaaban, M.; Wu, Y.P.; Delgado, A.; Kuzyakov, Y.; Peng, Q.A.; Lin, S.; Hu, R.G. Enzyme activities and organic matter mineralization in response to application of gypsum, manure and rice straw in saline and sodic soils. Environ. Res. 2023, 224, 115393. [Google Scholar] [CrossRef]
  45. Xu, H.D.; Sun, J.N.; Zhao, Z.Q.; Gao, Y.; Tian, L.J.; Wei, X.M. Long-term straw return promotes soil phosphorus cycling by enhancing soil microbial functional genes responsible for phosphorus mobilization in the rice rhizosphere. Agric. Ecosyst. Environ. 2025, 381, 109422. [Google Scholar] [CrossRef]
  46. Xie, X.F.; Xu, Z.Q.; Tian, Z.Y.; Bu, X.G.; Xu, F.; Liang, J.; Pu, L.J. Effects of Supplementation of Different Amendments on Soil Heavy Metals and Enzyme Activities in Coastal Saline Land. Huanjing Kexue 2023, 44, 5649–5656. [Google Scholar] [CrossRef]
  47. Ortega, R.; Miralles, I.; Soria, R.; RodríguezBerbel, N.; Villafuerte, A.B.; Zema, D.A.; Borja, M.E.L. Short-term effects of post-fire soil mulching with wheat straw and wood chips on the enzymatic activities in a Mediterranean pine forest. Sci. Total Environ. 2022, 857, 159489. [Google Scholar] [CrossRef] [PubMed]
  48. Li, H.Y.; Tian, H.X.; Wang, Z.Q.; Liu, C.Y.; Nurzhan, A.; Megharaj, M.; He, W.X. Potential effect of warming on soil microbial nutrient limitations as determined by enzymatic stoichiometry in the farmland from different climate zones. Sci. Total Environ. 2022, 802, 149657. [Google Scholar] [CrossRef] [PubMed]
  49. Jing, Y.L.; Zhang, Y.H.; Han, I.; Wang, P.; Mei, Q.W.; Huang, Y.J. Effects of different straw biochars on soil organic carbon, nitrogen, available phosphorus, and enzyme activity in paddy soil. Sci. Rep. 2020, 10, 8837. [Google Scholar] [CrossRef]
  50. Liu, L.; Cheng, M.; Yang, L.; Gu, X.Y.; Jin, J.Y.; Fu, M.J. Regulation of straw decomposition and its effect on soil function by the amount of returned straw in a cool zone rice crop system. Sci. Rep. 2023, 13, 15673. [Google Scholar] [CrossRef]
  51. Bekchanova, M.; Campion, L.; Bruns, S.; Kuppens, T.; Lehmann, J.; Jozefczak, M.; Cuypers, A.; Malina, R. Biochar improves the nutrient cycle in sandy-textured soils and increases crop yield: A systematic review. Environ. Evid. 2024, 13, 3. [Google Scholar] [CrossRef]
  52. Bhandari, G.; Gangola, S.; Dhasmana, A.; Rajput, V.; Gupta, S.; Malik, S.; Slama, P. Nano-biochar: Recent progress, challenges, and opportunities for sustainable environmental remediation. Front. Microbiol. 2023, 14, 1214870. [Google Scholar] [CrossRef]
  53. Song, Y.B.; Tan, J.W.; Jin, M.Y.; Liu, Z.; Zhu, J.F.; Mohamed, E.A.; Abdelhafeez, I.A.; Zhang, Y.L.; Shen, Z. Effect of pyrolysis temperature and heating rate on the physicochemical properties of alkali lignin-derived biochar: A comparative study of fast and slow pyrolysis. J. Anal. Appl. Pyrolysis 2025, 191, 107236. [Google Scholar] [CrossRef]
  54. Yan, J.F.; Xu, J.D.; Zhang, R.Y.; Zhou, P.; Yuan, Y.F.; Li, Y.M. Nanocarbon molecules—The fascination of synthetic chemistry. Prog. Chem. 2023, 35, 699–708. [Google Scholar] [CrossRef]
  55. Zeng, L.Q.; Zimmerman, A.R.; Huang, R.X. Adsorption of extracellular enzymes by biochar: Impacts of enzyme and biochar properties. Geoderma 2024, 451, 117082. [Google Scholar] [CrossRef]
  56. Raliya, R.; Saharan, V.; Dimkpa, C.; Biswas, P. Nanofertilizer for precision and sustainable agriculture: Current state and future perspectives. J. Agric. Food Chem. 2017, 66, 6487–6503. [Google Scholar] [CrossRef] [PubMed]
  57. Hu, Y.; Cao, Y.N.; Ma, C.X.; Yan, W.D. Nano-biochar as a potential amendment for metal(loid) remediation: Implications for soil quality improvement and stress alleviation. J. Environ. Manag. 2024, 351, 119658. [Google Scholar] [CrossRef]
  58. Malik, A.A.; Martiny, J.B.H.; Brodie, E.L.; Martiny, A.C.; Treseder, K.K.; Allison, S.D. Defining trait-based microbial strategies with consequences for soil carbon cycling under climate change. ISME J. 2020, 14, 1–9. [Google Scholar] [CrossRef] [PubMed]
  59. Weverka, J.R.; Moeller, H.V.; Schimel, J.P. Chemodiversity controls microbial assimilation of soil organic carbon: A theoretical model. Soil Biol. Biochem. 2023, 187, 109161. [Google Scholar] [CrossRef]
  60. Shang, W.H.; Razavi, B.S.; Yao, S.H.; Hao, C.K.; Kuzyakov, Y.; Blagodatskaya, E.; Tian, J. Contrasting mechanisms of nutrient mobilization in rhizosphere hotspots driven by straw and biochar amendment. Soil Biol. Biochem. 2023, 187, 109212. [Google Scholar] [CrossRef]
  61. Raiesi, F.; Khadem, A. Short-term effects of maize residue biochar on kinetic and thermodynamic parameters of soil β-glucosidase. Biochar 2019, 1, 213–227. [Google Scholar] [CrossRef]
  62. Khadem, A.; Besharati, H.; Khalaj, M.A. Biochar application changed arylsulfatase activity, kinetic and thermodynamic aspects. Eur. J. Soil Biol. 2019, 95, 103134. [Google Scholar] [CrossRef]
  63. Mbuthia, L.W.; Acosta-Martínez, V.; DeBruyn, J.; Schaeffer, S.; Tyler, D.; Odoi, E.; Mpheshea, M.; Walker, F.; Eash, N. Long term tillage, cover crop, and fertilization effects on microbial community structure, activity: Implications for soil quality. Soil Biol. Biochem. 2015, 89, 24–34. [Google Scholar] [CrossRef]
  64. Qiu, L.P.; Wang, Y.Q.; Liu, J.; Zhang, X.C. The dynamic and thermodynamic characteristics of soil reactions catalyzed by soil enzymes under long-term fertilization in Loess Plateau. Plant Nutr. Fertil. Sci. 2007, 13, 1028–1034. [Google Scholar]
  65. Panettieri, M.; Sosa, D.L.L.; Domínguez, T.M.; Madejón, E. Long-term impacts of conservation tillage on Mediterranean agricultural soils: Shifts in microbial communities despite limited effects on chemical properties. Agric. Ecosyst. Environ. 2020, 304, 107144. [Google Scholar] [CrossRef]
  66. Zhao, Z.W.; Wu, Y.; Chen, W.J.; Sun, W.; Wang, Z.H.; Liu, G.B.; Xue, S. Soil enzyme kinetics and thermodynamics in response to long-term vegetation succession. Sci. Total Environ. 2023, 882, 163542. [Google Scholar] [CrossRef] [PubMed]
  67. Misiak, M.; Goodall Copestake, W.P.; Sparks, T.H.; Worland, M.R.; Boddy, L.; Magan, N.; Convey, P.; Hopkins, D.W.; Newsham, K.K. Inhibitory effects of climate change on the growth and extracellular enzyme activities of a widespread Antarctic soil fungus. Glob. Change Biol. 2021, 27, 1111–1125. [Google Scholar] [CrossRef] [PubMed]
  68. Guo, Y.X.; Yu, G.H.; Hu, S.J.; Liang, C.; Kappler, A.; Jorgenson, M.T.; Guo, L.D.; Guggenberger, G. Deciphering the intricate control of minerals on deep soil carbon stability and persistence in Alaskan permafrost. Glob. Change Biol. 2024, 30, e17552. [Google Scholar] [CrossRef]
  69. Mooshammer, M.; Hofhansl, F.; Frank, A.H.; Wanek, W.; Hämmerle, I.; Leitner, S.; Schnecker, J.; Wild, B.; Watzka, M.; Keiblinger, K.M.; et al. Decoupling of microbial carbon, nitrogen, and phosphorus cycling in response to extreme temperature events. Sci. Adv. 2017, 3, e1602781. [Google Scholar] [CrossRef]
Figure 1. Enzyme activity values of (a) invertase (INV), (b) urease (URE), and (c) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Figure 1. Enzyme activity values of (a) invertase (INV), (b) urease (URE), and (c) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Sustainability 17 10436 g001
Figure 2. Enzyme kinetic parameters of (a,d,g) invertase (INV), (b,e,h) urease (URE), and (c,f,i) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Figure 2. Enzyme kinetic parameters of (a,d,g) invertase (INV), (b,e,h) urease (URE), and (c,f,i) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Sustainability 17 10436 g002
Figure 3. Eyring plots for (a) invertase (INV), (b) urease (URE), and (c) acid phosphatase (ACP) activities after three carbon sources addition at 15–55 °C. The horizontal axis represents 1/T, the vertical axis represents ln(Ka/T), R2 represents the degree of fit of the linear equation. T: temperature; Ka: catalytic efficiency.
Figure 3. Eyring plots for (a) invertase (INV), (b) urease (URE), and (c) acid phosphatase (ACP) activities after three carbon sources addition at 15–55 °C. The horizontal axis represents 1/T, the vertical axis represents ln(Ka/T), R2 represents the degree of fit of the linear equation. T: temperature; Ka: catalytic efficiency.
Sustainability 17 10436 g003
Figure 4. Enzyme thermodynamic parameters of (a,d,g) invertase (INV), (b,e,h) urease (URE), and (c,f,i) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Figure 4. Enzyme thermodynamic parameters of (a,d,g) invertase (INV), (b,e,h) urease (URE), and (c,f,i) acid phosphatase (ACP) after three carbon sources addition at 15–55 °C. The results of the two-way ANOVA are attached at the top of the figures. C, the main effects of carbon source; T, the main effects of temperature; C × T, the interaction effects of carbon source and temperature. Different uppercase letters indicate significant differences between treatments with different carbon sources at the same temperature, and different lowercase letters indicate significant differences between temperature at the same treatment (p < 0.05). Values are mean ± standard deviation (n = 3).
Sustainability 17 10436 g004
Figure 5. Correlation analysis between three hydrolase activities and catalytic properties after three carbon sources addition. **, * are significant at the 1% and 5%, respectively. INV: invertase; URE: urease; ACP: acid phosphatase.
Figure 5. Correlation analysis between three hydrolase activities and catalytic properties after three carbon sources addition. **, * are significant at the 1% and 5%, respectively. INV: invertase; URE: urease; ACP: acid phosphatase.
Sustainability 17 10436 g005
Figure 6. Redundancy analysis of hydrolase catalytic characteristics and physical chemical properties of soil. The geometric figure represents the different carbon source treatments. Red lines represent soil physicochemical properties, green lines represent soil hydrolase activities, blue lines represent kinetic and thermodynamic parameters of soil hydrolases. BD: bulk density; OC: organic carbon; TN: total nitrogen; AN: available nitrogen; TP: total phosphorus; AP: available phosphorus; INV: invertase; URE: urease; ACP: acid phosphatase.
Figure 6. Redundancy analysis of hydrolase catalytic characteristics and physical chemical properties of soil. The geometric figure represents the different carbon source treatments. Red lines represent soil physicochemical properties, green lines represent soil hydrolase activities, blue lines represent kinetic and thermodynamic parameters of soil hydrolases. BD: bulk density; OC: organic carbon; TN: total nitrogen; AN: available nitrogen; TP: total phosphorus; AP: available phosphorus; INV: invertase; URE: urease; ACP: acid phosphatase.
Sustainability 17 10436 g006
Figure 7. The proportion of total variability in soil hydrolase activities and their kinetic and thermodynamic properties explained by different soil physicochemical characteristics. OC: organic carbon; AP: available phosphorus; TP: total phosphorus; AN: available nitrogen; TN: total nitrogen; BD: bulk density.
Figure 7. The proportion of total variability in soil hydrolase activities and their kinetic and thermodynamic properties explained by different soil physicochemical characteristics. OC: organic carbon; AP: available phosphorus; TP: total phosphorus; AN: available nitrogen; TN: total nitrogen; BD: bulk density.
Sustainability 17 10436 g007
Table 1. Variations in soil physicochemical properties after three carbon sources addition.
Table 1. Variations in soil physicochemical properties after three carbon sources addition.
TreatmentpHBulk Density
(BD)
(g·cm−3)
Organic Carbon
(OC)
(g·kg−1)
Total Nitrogen
(TN)
(g·kg−1)
Available Nitrogen
(AN)
(mg·kg−1)
Total Phosphorus
(TP)
(g·kg−1)
Available Phosphorus
(AP)
(mg·kg−1)
CK6.66 ± 0.02 a1.52 ± 0.01 a9.28 ± 0.10 b0.84 ± 0.03 ab66.22 ± 0.88 b0.92 ± 0.02 b74.72 ± 1.26 b
Straw6.65 ± 0.00 a 1.47 ± 0.03 ab11.11 ± 0.10 a0.97 ± 0.06 a85.17 ± 6.15 a1.03 ± 0.06 a72.93 ± 3.79 b
Biochar6.69 ± 0.08 a1.45 ± 0.04 b9.88 ± 0.10 b0.76 ± 0.03 b83.71 ± 3.15 a1.02 ± 0.02 a100.71 ± 6.94 a
Nanocarbon6.66 ± 0.00 a1.46 ± 0.02 b8.31 ± 0.60 c0.73 ± 0.08 b61.83 ± 8.75 b1.01 ± 0.02 a98.49 ± 4.73 a
All values are given as mean ± standard deviation (n = 3). The different lowercase letters corresponding to the treatment bars are significantly different at a 0.05 probability level.
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

Xu, J.; Wu, X.; Wang, P.; Zhao, J.; Yue, Z.; Bai, X.; Li, J.; Yin, Y.; Huang, J. Decoding the Sustainability Code: Enzyme Thermodynamic and Kinetic Parameters Reveal the Efficacy of Straw, Biochar, and Nanocarbon in Black Soil. Sustainability 2025, 17, 10436. https://doi.org/10.3390/su172310436

AMA Style

Xu J, Wu X, Wang P, Zhao J, Yue Z, Bai X, Li J, Yin Y, Huang J. Decoding the Sustainability Code: Enzyme Thermodynamic and Kinetic Parameters Reveal the Efficacy of Straw, Biochar, and Nanocarbon in Black Soil. Sustainability. 2025; 17(23):10436. https://doi.org/10.3390/su172310436

Chicago/Turabian Style

Xu, Jia, Xiangyu Wu, Pengwei Wang, Jingyi Zhao, Zhonghui Yue, Xin Bai, Jiawang Li, Yuan Yin, and Jianhao Huang. 2025. "Decoding the Sustainability Code: Enzyme Thermodynamic and Kinetic Parameters Reveal the Efficacy of Straw, Biochar, and Nanocarbon in Black Soil" Sustainability 17, no. 23: 10436. https://doi.org/10.3390/su172310436

APA Style

Xu, J., Wu, X., Wang, P., Zhao, J., Yue, Z., Bai, X., Li, J., Yin, Y., & Huang, J. (2025). Decoding the Sustainability Code: Enzyme Thermodynamic and Kinetic Parameters Reveal the Efficacy of Straw, Biochar, and Nanocarbon in Black Soil. Sustainability, 17(23), 10436. https://doi.org/10.3390/su172310436

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop