Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Environmental Technology & Innovation 35 (2024) 103704

Contents lists available at ScienceDirect

Environmental Technology & Innovation


journal homepage: www.elsevier.com/locate/eti

Effect of electric field and humic acid on the mobility of biochar


particles in soil
Yifan Yang a, Min Yang a, Hongjia Bao a, Chen Chen a, Weimin Cao b, Xiaolei Zhang a,
Qiang Liu a, *
a
School of Environmental and Chemical Engineering, Shanghai University, Shanghai 200444, China
b
College of Sciences, Shanghai University, Shanghai 200444, China

A R T I C L E I N F O A B S T R A C T

Keywords: Biochar is a widely used material for the remediation of contaminated soils. However, the re­
Micro- and nano- biochar particles covery of biochar after its application in soil remains a significant challenge. To regulate the
Soil migration of particles and identify a possible strategy for the recovery of biochar from soil, this
Transport
study investigated the transport behavior of biochar microparticles (MPs) and nanoparticles (NPs)
Electric field
in real soil under different potential gradients (0, 0.5 and 1.0⋅V cm− 1) and humic acid (HA)
Humic acid
concentrations (0, 5 and 10 mg⋅L− 1). It was demonstrated that the electric field and the presence
of HA could enhance the mass recovery of biochar particles. Further improvement in the mobility
of biochar was observed upon increasing potential gradient and HA content, which resulted in an
increase in electrostatic repulsion between biochar and soil particles. In the optimal conditions,
the recovery of biochar MPs and NPs in the effluent could be significantly increased to 46.3
±1.2 % and 72±1.2 % at 1.0 V⋅cm− 1 in the presence of 5 mg⋅L− 1 HA. The transport of biochar
particles in soil was well explained by the convection-diffusion equation (CDE) model. The
enhanced migration was attributed to the increase of oxygen-containing functional groups on
biochar particles by the electric field, which reduced the zeta potential (ζ-potential) of biochar
and thus increased the electrostatic repulsion between them and the soil. Furthermore, the
application of direct current reduced the size of biochar particles, facilitating biochar migration
by reducing the size-blocking effect. The results of this study provide a basis for regulating the
migration of biochar in soil.

1. Introduction

Biochar is a carbon-rich material produced by pyrolysis of biomass in the absence or with limited oxygen supply (Bardi et al., 2023;
Nguyen et al., 2023). The potential of biochar for soil amendment and environmental remediation, particularly for the removal or
immobilization of contaminants such as heavy metals (Jiang et al., 2023; Liu et al., 2022b) and organic matters (Hung et al., 2023;
Zhang et al., 2023), has been widely reported in the literature. In some cases, engineering biochar incorporated with various metal
oxides has also been used for contaminant removal in environmental media (Tran et al., 2022). After application to soil, biochar
particles can be broken down and degraded by biotic and abiotic processes, which in turn can release micro- and nano-sized biochar
colloids into soil pore water or neighboring groundwater (Cross and Sohi, 2013). Previous studies have shown that biochar particles

* Corresponding author.
E-mail address: qliu@shu.edu.cn (Q. Liu).

https://doi.org/10.1016/j.eti.2024.103704
Received 11 April 2024; Received in revised form 3 June 2024; Accepted 3 June 2024
Available online 4 June 2024
2352-1864/© 2024 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

have strong mobility in aquatic and soil environments, and can act as viable carriers for transporting and diffusing adsorbed pollutants
(Ma et al., 2023). Therefore, in order to minimize the potential negative impacts of biochar applications in agricultural and envi­
ronmental practices, it is essential to understand the transport process of biochar particles in soil.
Using quartz sand as a representative porous medium, existing studies have shown that the migration of biochar is mainly influ­
enced by solution pH, ionic strength and species, natural organic matter and biochar properties (pyrolysis temperature and ageing)
(Yang et al., 2017; Yang et al., 2019). Wang et al. (2013) reported that nano-sized biochar with small particle size and low zeta po­
tential (ζ-potential) has a high potential to migrate in the sand column in contrast to micro-sized particles. Yang et al. (2019b) found
that lower ionic content in the background solution and lower pyrolysis temperature of biochar could enhance the mobility of biochar
in quartz sand. Furthermore, Wang et al. (2022) reported that the presence of humic acid could enhance the electrostatic repulsion
between the biochar particles and sand, thereby improving the mobility of biochar. These findings provide important insights into the
transport behavior of colloidal biochar. However, the manipulation and regulation of biochar migration to prevent the spread of
associated contaminants remains a great challenge.
Emerging evidence supports electrokinetics as a method to enhance colloidal particle transport in permeable soil matrices (Yuan
et al., 2012). The effectiveness of this strategy is attributed to its ability to generate electroosmotic flow in the low permeability soil
matrix by applying potential gradients, which enhances the mobility of nanoparticles (Reddy et al., 2011). Research has shown that a
low strength direct current (DC) electric field can dramatically enhance the mobility of synthetic particles in soil (Gomes et al., 2014).
Inspired by this, our previous study demonstrated that a DC electric field significantly enhanced the mass recovery of biochar particles
in saturated quartz sand with varying solution chemistries, highlighting the potential for regulating biochar transport in porous media
(Liu et al., 2023). Furthermore, in co-migration with heavy metals, the electric field could also significantly increase the mobility of
heavy metal-adsorbed biochar, especially nano-sized particles. By switching the electric field strength, the recoveries of NPs and Pb
were 90 % and 35 % higher, respectively, than in the absence of an electric field (Liu et al., 2024). Since natural soils are highly
heterogeneous and rich in humic acids, the transport behavior of biochar in quartz sand is significantly different from that in natural
soil (Sun et al., 2015). However, research on the effect of DC electric field on the movement of biochar particle in real soil is negligible.
Therefore, this study aims to investigate the influence of DC electric field on the transport of biochar microparticles (MPs) and
nanoparticles (NPs) in real soil under variable potential gradients and humic acid concentrations. The migration effect was evaluated
by theoretical calculations and model fitting. The results could be useful for the adoption of DC electric field as a regulatory mechanism
for biochar particle relocation in natural soil.

2. Materials and methods

2.1. Preparation of biochar particle suspension and porous soil media

In this study, corn straw was used as the feedstock for biochar production. This choice was based on its abundant availability and
the richness in surface functional groups of the derived biochar (Song et al., 2019). The corn straw was first washed with distilled water
to remove impurities. The water-washed biochar was dried for 24 h, then crushed and sieved through a 100-mesh screen. The biomass
pyrolysis process was carried out at 300 ℃ under a nitrogen atmosphere. During the pyrolysis, the furnace was continuously flushed
with N2 gas at a flow rate of 100 mL⋅min− 1. Pyrolysis was controlled by heating the biomass from room temperature to 100 ℃ at a rate
of 10 ◦ C⋅min− 1 and then stabilizing the temperature at 100 ℃ for 20 min to remove oxygen from the furnace. The sample was then
continuously heated to 300 ℃ at a rate of 10 ◦ C⋅min− 1 and held constant for a further 2 h. The resulting powder was crushed with a
mortar, and passed through a 400-mesh screen, and used as MPs for the experiment. Next, the ball mill was used to grind the MPs into
nano-sized powder and the resulting particles were called NPs.
Biochar suspensions (250 mg⋅L− 1) were prepared by suspending MPs and NPs in 10 mM NaCl solution with different background
levels of HA (0, 5 and 10 mg⋅L− 1). The low ionic strength (10 mM) was used to avoid the aggregation of biochar particles. Prior to the
column transport tests, the biochar suspension was sonicated for 10 min to allow complete dispersion of the biochar particles in the
solution.
Porous media for the biochar transport column tests were obtained from a real soil on the campus of Shanghai University, Shanghai.
The soil was air-dried and sieved to obtain particles with a size range of 600–710 μm (with a median grain size of 650 μm). Detail soil
properties are given in Section 3.1.

2.2. Characterization of the biochar particles and soils

The elemental component of MPs and NPs (C, H, O, N) was determined using an elemental analyzer (Euro Vector EA3000, Italy).
Surface chemical properties of biochar were characterized using a Fourier transform infrared spectroscopy (FTIR) spectrometer
(Nicolet 380, Thermo Scientific, USA) at 400–4000 cm− 1 with a resolution of 4 cm− 1. Pore volume and surface area of the biochar and
soil particles were determined using the Autosorb gas adsorption analyzer (SI, Quantachrome, USA). The hydrodynamic diameter and
ζ-potential of the biochar suspension and soil particles under different conditions were investigated using a zeta potential analyzer
(Nicomp 380 Z3000, Alpharmaca, USA). Soil particle crystallinity was investigated using an X ray diffractometer (Smartlab, Rigaku,
Japan). Sample surface morphologies were investigated using field electro scanning microscopy (IT800SHL, JEOL, Japan).

2
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

2.3. Biochar transport column tests

The experimental apparatus of biochar transport column test is shown in Fig. 1. Biochar migration in soil was evaluated in a Φ2 cm
(inner diameter) glass column with a height of 12 cm. A pair of graphite plates were installed at the top and bottom of the column as the
cathode and anode, and connected to the DC power supply by wires. During the tests, the potential gradient of the column was
controlled by the input voltage of the DC power supply. The glass column was filled with soil to a depth of 10 cm, with a porosity of
0.52 and a void volume of 16 mL.
Prior to the experiment, a 10 mM NaCl solution was circulated through the column for 3 days at a flow rate of 1.0 mL⋅min− 1 using a
peristaltic pump (BT100S-1-CE, Leadfluid, China) to ensure soil stabilization and washout of contaminants. To ensure negligible
impurity content, the column effluent was determined by UV spectrophotometry (Specord 210 plus, Jena, Germany). As a control
condition, biochar movement in the quartz sand column was investigated under identical experimental conditions. In this case, the
column was filled with ultrapure silica sand (99.8 % SiO2, 400–700 μm, Sinopharm Chemical Reagent Co., Ltd., China) with a porosity
of 0.49 and a void volume of 19 mL.
To determine the hydrodynamic parameters of the packed column, 3 pore volumes (PVs) of NaCl solution (10 mM) were used for
tracing at a rate of 1.0 mL⋅min− 1, which was then balanced with 5 PVs of deionized water. Tracer solutions were collected by an
automatic sample collector (BS-100A, Shanghai Huxi Analytical Instrument Factory Co., Ltd., China) at 3 min intervals. The tracer
(NaCl) content in the effluent was estimated from the electrical conductivity of the solution, which was measured using a conductivity
meter (FE38, Metler Toledo, Switzerland). The tracer breakthrough curve (Figure S1) was analyzed using the CXTFIT 2.1 plugin of the
STANMOD software to estimate the average pore water velocity V and the diffusion coefficient D within the soil and sand column.
To perform the column biochar transport experiments, 3 PVs of biochar suspension with different background HA concentrations
(0, 5, and 10 mg⋅L− 1) were first introduced into the column from bottom to top at a rate of 1.0 mL⋅min− 1 under different potential
gradients (0, 0.5 and 1.0 V⋅cm− 1). Subsequently, 5 PVs of NaCl solution (10 mM) were used to elute biochar particles entrapped in the
soil matrix. The effluent sample was collected every 3 min and the biochar content in the effluent was assessed using a UV spectro­
photometer at 282 nm (Specord 210 plus, Jena, Germany) according to an established calibration curve (Figure S2). In this study, 16
biochar transport soil column tests were conducted, and each test was performed in triplicate. The specific conditions of experiments
are shown in Table 1.
After each test, 5 soil segments were collected at equal intervals along the column to determine the amount of biochar retained in
the medium. The collected soil sample (0.2 g) was dried in an oven at 80 ℃ for 24 h and then oxidized with a 50 mL dichromate/
sulfuric acid mixture (0.1 M K2Cr2O7 + 2 M H2SO4) at 60 ℃ for 24 h (Song et al., 2002). The solution was then filtered and the
absorbance was measured at 585 nm using a UV spectrophotometer to quantify the total amount of oxidized carbon in the sample. The
baseline absorbance of the carbon content in the fresh soil was subtracted from each sample.

Fig. 1. Experimental setup of biochar transport column tests.

3
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

Table 1
Environmental conditions of biochar particles in soil column tests.
Trial Biochar Voltage (V⋅cm− 1) HA (mg⋅L− 1) pH dp/dsa

1 MPs 0 0 7.36 0.0056


2 MPs 0.5 0 7.36 0.0056
3 MPs 1.0 0 7.36 0.0056
4 MPs 0 5 7.07 0.0056
5 MPs 0.5 5 7.07 0.0056
6 MPs 1.0 5 7.07 0.0056
7 MPs 0 10 6.83 0.0056
8 MPs 0.5 10 6.83 0.0056
9 NPs 0 0 7.18 0.0012
10 NPs 0.5 0 7.18 0.0012
11 NPs 1.0 0 7.18 0.0012
12 NPs 0 5 6.87 0.0012
13 NPs 0.5 5 6.87 0.0012
14 NPs 1.0 5 6.87 0.0012
15 NPs 0 10 6.60 0.0012
16 NPs 0.5 10 6.60 0.0012
a
The ratio of the hydrodynamic diameter of biochar particles to the soil diameter

2.4. Theoretical calculations and model fitting

Derjaguin-Landau-Verwey-Overbeek (DLVO) theory was used to encapsulate the interaction energy between the soil surface and
biochar. The total DLVO interaction energy was derived as the sum of the electrostatic double layer repulsion and the van der Waals
attraction (Cao et al., 2023). The zeta potential of MPs and NPs and the hydrodynamic radius of biochar particles, were used for the
DLVO theory analysis (see Supporting Information Text S1 for details). All parameters relevant to the analysis are listed in Table S1.
Collision efficiency (α) denotes the propensity of biochar particles to adhere to the surface of soil when they collide with each other
(Tufenkji and Elimelech, 2004). Three mechanisms, i.e., interception, gravitational sedimentation, and Brownian diffusion are
generally responsible for the transport of colloidal particles from soil interstitial water to the surface of soil. The method used to
determine the collision efficiency of biochar particles with soil via these mechanisms are described in detail in Text S2 (Supporting
Information), and all the parameters used in the calculation are listed in Table S2.
The convection-diffusion equation (CDE) in CXTFIT 2.1 software was used to fit the breakthrough curve of biochar migration in the
soil column, which mainly includes the deterministic equilibrium CDE model and the deterministic nonequilibrium CDE model
(Supporting Information, Text S3) (Chung et al., 2016). The former was used to fit the breakthrough curve of NaCl to obtain the
average pore water flow velocity V and diffusion coefficient D, while the latter was used to fit the penetration curve of biochar in the
column. The model assumes that the liquid phase is divided into flow and non-flow regions, with solute migration governed by the
two-region formula, which allows exchange between mobile and immobile regions.
Multiple linear regression was used to analyze the significance of the changes in potential gradient and humic acid concentration on
the mass recovery of biochar particles (P < 0.05).

3. Results and discussion

3.1. Characterization of micro and nano biochar particles and soils

The physical and chemical properties of MPs and NPs are shown in Table 2. Despite similar pH and elemental profiles, the hy­
drodynamic diameter of NPs (~780 nm) is significantly lower than that of MPs (~3650 nm). Detailed morphological analysis
(Figures S3a and S3c) revealed that NPs are more uniform and slenderer in size compared to MPs. These results confirm that ball
milling is suitable for size reduction without altering the chemical composition of biochar (Xiao et al., 2020). The smaller size of NPs is
advantageous for their migration through the soil pores compared to MPs, as particles size is the main factor affecting their mobility in
porous media (Wang et al., 2013). After ball milling, the specific surface area and pore volume of NPs increased significantly compared
with MPs, possibly due to the exposure of more mesopores facilitated by the release of inherent minerals from biochar (Lyu et al.,

Table 2
Properties of biochar particles used in this study.
Biochar pH Elemental Content (%) SSAa BET (m2⋅g− 1) DLS Sizeb (nm) Pore Volume
(cm3⋅g− 1)
C H O N

MPs 7.4 54.8 4.2 38.4 1.91 4.9 3650 0.012


NPs 7.5 53.4 4.0 40.2 1.76 9.6 780 0.022
a
Specific surface area
b
Dynamic light scattering hydrodynamic diameter

4
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

2018).
The characteristics of the soil used in this study are detailed in Table 3. The soil is pH neutral and has organic carbon and organic
matter contents of 6.4 and 11.0 g⋅kg− 1, respectively. According to the X-ray diffraction pattern (Figure S4), the composition of the soil
was mainly quartz, clay minerals (kaolinite, bernalite, montmorillonite) and iron oxides (goethite, albite). Figure S5 clearly shows the
rougher surface texture of the soil particles, with fewer voids and more mineral components on their surface. Previous studies have
confirmed that the diversity of soil surface roughness plays a role in the migration of biochar (Chen et al., 2017). The SEM-EDS analysis
(Figure S5) also shows that iron, silicon, aluminum, magnesium, calcium, and manganese are present on the soil grains, indicating an
abundance of iron and aluminum oxides in the soil. These oxides are typically enriched with positive charges at neutral pH and could
potentially hinder the migration of negatively charged biochar particles within the soil pore.

3.2. Biochar migration under different potential gradients

Fig. 2 shows the breakthrough and retention curves of MPs and NPs in the soil column at different potential gradients (0, 0.5, and
1.0 V⋅cm− 1). From Figs. 2a and c, it can be seen that the breakthrough curves (BTCs) show an asymmetric model, particularly for MPs,
indicating a significant tailing of biochar transport through the soil matrix. This means that many biochar particles migrated out of the
soil column along with NaCl solution elution (5 PVs) under low ionic strength, highlighting a strong interaction between biochar
particles and soil. After application of DC electric field (0.5 and 1.0 V⋅cm− 1), the shape of the BTCs became steeper, especially for MPs
under higher potential gradient (1.0 V⋅cm− 1). This reflects that the site for the retention of biochar particles was progressively filled up
due to the soil blocking effect. In addition, the tailing phenomenon of the BTCs was more pronounced for both types of biochar
particles and increased with increasing potential gradient.
The BTC curves (Figs. 2a and c) also show the mobility of both MPs and NPs is significantly improved with increasing potential
gradient. For example, (Ci/C0)max, the dimensionless effluent concentration, of MPs (28.6 % at 0.5 V⋅cm− 1, 38.8 % at 1.0 V⋅cm− 1) and
NPs (50.9 % at 0.5 V⋅cm− 1, 57.7 % at 1.0 V⋅cm− 1) under two potential gradients was much higher than that of their counterpart
without electric field (20.7 % for MPs and 38.0 % for NPs). Meanwhile, the mass recovery of biochar particles in the effluent increased
significantly from 32.7±1.1 % to 44.1±1.8 % for MPs, and from 53.3±1.0 % to 68.9±2.3 % for NPs with the increase in potential
gradient from 0.5 to 1.0 V⋅cm− 1, respectively (Table 4). This trend is consistent with biochar transport in the silica sand column
(Figure S6a and S6c), indicating that a higher potential gradient indeed increased the mobility of biochar particles, regardless of the
type of porous media. The higher mobility of biochar is possibly caused by the electrostatic attraction from the DC electric field as well
as the washout force from the electroosmosis of electrolyte in the background solution (Liu et al., 2023), which potentially caused more
biochar particles to migrate out of the column with the eluate (10 mM NaCl).
The migration of biochar particles in soil is mainly influenced by the blocking effect, particle surface properties, and the secondary
potential well (Bradford et al., 2003). Typically, a ratio of biochar particle diameter (dp) to soil grain diameter (ds) greater than 0.002
indicates significant blocking of biochar migration in porous media. In this study, the dp/ds ratios of MPs and NPs are 0.0056 and
0.0012, respectively (Table 1), indicating that the blocking effect mainly affects MPs but not NPs. Table 5 summarizes the ζ-potential of
biochar and soil, collision efficiencies, and the parameters derived from DLVO theory analysis. As observed, DC electric field signif­
icantly decreased the collision coefficient (α) of MPs to soil from 0.54 (0 V⋅cm− 1) to 0.30 (1.0 V⋅cm− 1). This indicates that the
probability of adsorption of biochar particles to the soil substrate decreased significantly after each collision of MPs to soil grains,
thereby reducing the resistance of biochar migration in the soil. A similar decrease in the α value (from 0.48 to 0.19) was also found for
NPs, showing that although the blocking effect did not have a significant effect on the transport of NPs, the probability of NPs adhering
to soil particles was also reduced by the application of the DC electric field, which further improved the separation of biochar from soil
particles.
These experimental results were also supported by DLVO interaction energy calculations (Table 5 and Figure S7). After increasing
potential gradient from 0 to 1.0 V⋅cm− 1, the repulsive energy barrier (Φmax) between biochar and soil increased significantly from 268
to 975 kBT for MPs and from 105 to 221 kBT for NPs, and the secondary minimum value (Φmin2) increased from − 3.67 kBT to − 2.77 kBT
for MPs and from − 0.67 kBT to − 0.57 kBT for NPs, indicating that the electrostatic repulsion between the biochar particles and the soil
grains was enhanced while the attractive force became weak (Yuan et al., 2023). This was not conducive to the adhesion of MPs and
NPs to the soil grains, and thus biochar particles were not easily accumulated and deposited in the column. Therefore, some of the
trapped biochar particles may have migrated out due to the driving force provided by the electric field, and this attraction effect
became stronger as the potential gradient increased.
Figs. 2b and 2d illustrate the retention profiles of biochar along the depth of the soil column under different potential gradients.

Table 3
Properties of the soil used in this study.
Sample pH SOCa SOMb SSA BET Composition
(g⋅kg− 1) (g⋅kg− 1) (m2⋅g− 1)

Soil 7.7 6.4 11.0 9.6 quartz, kaolinite, bernalite,


montmorillonite, goethite, albite
Sand 7.3 - - 0.3 SiO2, Al2O3
a
Soil organic carbon
b
Soil organic matter

5
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

Fig. 2. The observed and fitted breakthrough curves (a, c) and the amounts (b, d) of MPs and NPs under different voltage gradients (0, 0.5, and
1.0 V⋅cm− 1) in the soil column tests.

Table 4
Mass recoveries of biochar particles in soil column tests.
Trial Mass recoverya (%)
Effluent Retained Total

1 22.5±0.8 80.8±1.8 103.3±2.6


2 32.7±1.1 70.4±2.0 103.1±3.1
3 44.1±1.8 52.5±0.9 96.6±2.7
4 28.9±0.7 67.6±3.3 96.5±4.0
5 39.4±1.5 58.3±3.0 97.7±4.5
6 46.3±1.2 51.4±2.6 97.7±3.8
7 36.1±2.3 61.9±2.9 98.0±5.2
8 44.2±2.0 55.7±3.5 99.9±5.5
9 38.9±1.8 56.0±3.5 94.9±5.3
10 53.3±1.0 46.5±2.0 99.8±3.0
11 68.9±2.3 32.3±2.1 101.2±4.4
12 43.7±2.0 52.1±1.5 95.8±3.5
13 60.3±2.5 41.6±2.5 101.9±5.0
14 72.2±1.2 26.4±3.4 98.6±4.6
15 51.9±1.1 46.0±2.9 97.9±4.0
16 63.3±2.7 36.8±1.7 100.1±4.4
a
Effluent, retained in soil column, and total percentage of biochar particles recovered from column tests

6
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

Table 5
Electrokinetic properties of biochar colloids and soil, the collision efficiency for packed-soil column experiments, and DLVO parameters calculated
under different environmental conditions.
Trial ζ-Potential (mV) αa Φmaxb (kBT) Φmin2c(kBT) hd (nm)

Influent Effluent Soil

1 -24.8±0.2 -22.6±0.5 -19.8±0.8 0.54 268 -3.67 65


2 -27.9±0.4 -24.8±0.1 0.41 552 -3.13 73
3 -33.7±1.2 -30.2±0.5 0.30 975 -2.77 79
4 -40.1±0.3 -25.1±0.1 -23.0±0.4 0.45 415 -3.34 70
5 -30.8±1.2 -28.5±0.4 0.34 790 -2.90 77
6 -37.5±0.5 -34.1±1.2 0.28 1338 -2.58 82
7 -43.3±0.5 -29.6±0.8 -25.8±0.6 0.37 640 -3.03 74
8 -32.8±1.0 -29.0±0.3 0.30 885 -2.82 78
9 -34.9±1.2 -28.7±1.5 -22.9±0.7 0.48 105 -0.67 75
10 -32.6±0.9 -26.9±0.3 0.32 163 -0.60 77
11 -35.2±1.2 -30.7±1.3 0.19 221 -0.57 80
12 -45.3±0.8 -31.1±0.5 -26.8±0.5 0.42 152 -0.61 76
13 -36.0±1.6 -31.4±1.0 0.26 235 -0.56 80
14 -42.5±0.7 -36.3±1.1 0.17 356 -0.51 85
15 -47.8±0.2 -34.7±0.4 -30.1±0.6 0.33 210 -0.57 79
16 -40.8±1.0 -35.4±0.2 0.23 327 -0.52 84
a
Collision efficiency (α) between biochar and soil resulting in attachment of the particles
b
Maximum energy barrier (Φmax)
c
Secondary minimum value (Φmin2)
d
Separation distance from biochar particles and soil surfaces to secondary minimum

Notably, the amount of biochar retained in the column decreased with increasing soil depth, regardless of whether the DC electric field
was on or off. This retention was pronounced at the entrance of the soil column (0–2 cm), as evidenced by higher biochar content (up to
0.6 mg⋅g− 1) in the collected soil samples. This result was consistent with previous studies (Kumar et al., 2023), and could be attributed
to the size exclusion effect resulting from biochar being trapped in smaller pores (Kretzschmar and Sticher, 1997). However, appli­
cation of a DC electric field significantly reduced biochar retention at the soil column entrance (0–2 cm). Specifically, approximately
0.53 mg⋅g− 1 of MPs and 0.33 mg⋅g− 1 of NPs were retained in the first segment of the column at 0 V⋅cm− 1, whereas only 0.44 mg⋅g− 1 of
MPs and 0.10 mg⋅g− 1 of NPs were retained in the same part at 1.0 V⋅cm− 1, respectively. This is partly due to the electrolysis of water in
the anodic region of the DC electric field, where the generated OH- can potentially increase the pH of the background solution
(Khodadoust et al., 2006). Following the increase in pH, the ζ-potential of biochar and soil particles further decreased (negative)
(Table 5), which increased the electrostatic repulsion between them and hindered biochar attachment to soil and aggregation. Under
different potential gradients, the amount of biochar particles retained in soil was apparently altered (P < 0.05).
Compared to quartz sand (Figure S6), biochar particles exhibited slightly lower mobility in the natural soil under equivalent
environmental conditions and potential gradient. This is illustrated by slightly lower (Ci/C0)max values and biochar mass recoveries of
MPs and NPs in the soil (Table 4) compared to silica grains (Table S5). Meanwhile, the excess biochar retained in the soil column
(Figs. 2b and 2d) compared to the sand column (Figures S6b and S6d) further confirmed the reduced mobility of biochar in the natural
soil. This could possibly be attributed to the larger specific surface area (SSA), greater surface charge heterogeneity, and rougher
surface of the soil than clean quartz. As shown in Table 3, the soil has a much higher SSA (9.6 m2⋅g− 1) than the quartz sand
(0.3 m2⋅g− 1), which may provide more retention sites conducive to biochar retention. In addition, the soil surface was coated with
layered Al or Fe oxides (Figure S5), which are more positively charged at neutral pH compared to negatively charged quartz sand
(Zhang et al., 2010). At the same time, the positively charged coating of clay minerals in the soil enhanced the adhesion of biochar at
their interface since biochar nano particles are typically negatively charged (Cornelis et al., 2013). In addition, the coarser surface
texture of soil particles (Figure S5) also limited the mobility of colloids in the soil column. This is because it has been confirmed that
surface rugosity alters the flow field near the media surface, which could subsequently affect the hydrodynamics of biochar during its
migration (Tian et al., 2023). This result is consistent with Sun et al. (2015), who reported that ZnO NPs were more readily to be
deposited in soil than in silica sand, and that retention was dominated by the surface geometry of the medium and the hydrodynamic
forces of the solution, which influenced mutual contact to overcome the boundary energy barrier (Sun et al., 2015).

3.3. HA-enhanced transport of biochar colloids

The effect of HA contents (0, 5, and 10 mg⋅L− 1) in the background solution on the transport of MPs and NPs under different po­
tential gradients is shown in Fig. 3. With increasing HA concentration, the mobility of both MPs and NPs in the soil increased slightly
compared to the condition without HA. Under 0 V⋅cm− 1, the (Ci/C0)max of MPs and NPs BTCs was increased from 20.7 % and 38.0 %
(HA free) to 24.6 % and 43.7 % at the HA concentration of 5 mg⋅L− 1, and additionally promoted to 34.6 % and 50.4 % at the HA
concentration of 10 mg⋅L− 1. In addition, further application of DC electric field significantly facilitated the transport of biochar
particles in soil in the presence of humic acid. For example, at the HA concentration of 5 mg⋅L− 1, the mass recoveries of MPs and NPs in
the effluent escalated from 28.9±0.7 % and 43.7±2.0 % at 0 V⋅cm− 1 to 39.4±1.5 % and 60.3±2.5 % at 0.5 V⋅cm− 1, and further to 46.3

7
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

1
Fig. 3. The observed and fitted breakthrough curves (a, c, e, g) and the amounts (b, d, f, h) of MPs and NPs in the presence of 5 mg⋅L− HA (a, b, c,
d) and 10 mg⋅L− 1 HA (e, f, g, h) under different voltage gradients (0, 0.5, and 1.0 V⋅cm− 1) in the soil column tests.

8
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

±1.2 % and 72.2±1.2 % at 1.0 V⋅cm− 1, respectively. A further increase in HA concentration to 10 mg⋅L− 1 increased the mass recovery
(Table 4) of MPs and NPs in the effluent to 44.2±2.0 % and 63.3±2.7 % at 0.5 V⋅cm− 1. Thus, HA was found to increase the mobility of
MPs and NPs for the NaCl electrolyte, and the increase was further enhanced by the introduction of a DC electric field. This
enhancement was associated with a more significant detachment of biochar particles from the soil surface by NaCl elution under the
DC electric field, as evidenced by an extended breakthrough curve tail. In addition, the enhancement of biochar migration could be
reflected by the amount of biochar particles entrapped in the soil column, which decreased significantly with increasing HA and
potential gradient levels (Figs. 3b, d, f and h). At the concentration of 5 mg⋅L− 1 HA, there was an apparent change in the amount of MPs
and NPs retained in the soil column (P < 0.05).
The increase in the mobility of biochar in the presence of HA is likely due to the zeta potentials of biochar and soil becoming more
negative as a result of HA adsorption on biochar colloids and soil particles, leading to greater electrostatic repulsion between these
particles. Following the addition of HA, the pH of the influent suspension of MPs and NPs decreased slightly from the initial 7.36 and
7.18–6.83 and 6.60 in the presence of 10 mg⋅L− 1 HA (Table 1). At lower pH, biochar colloids may be destabilised and deposited on the
surface of porous media due to an increase in negative charge, which could be reflected by a decrease in energy barrier (Cao et al.,
2023b). However, it can be confirmed that the presence of HA significantly increased the ζ-potential of both biochar and soil (negative
value) (Table 5). For example, as the HA content increased from 0 to 10 mg⋅L− 1, the zeta potentials of biochar increased from − 24.8
±0.2 to − 43.3±0.5 mV for MPs and from − 34.9±1.2 to − 47.8±0.2 mV for NPs. This was also be verified by DLVO theory analysis.
According to the calculation, in the absence of a DC electric field, Φmax for MPs increased from 268 to 415 kBT under 5 mg⋅L− 1 HA, and
further to 640 kBT under 10 mg⋅L− 1 HA. Φmax for NPs showed the same trends. Therefore, the addition of HA could increase the
interaction energy of repulsion between the particles, producing the opposite effect to the retention induced by chemical solutions. The
first reason was that the presence of HA reduced the ζ-potential of biochar and soil to mask the surface charge heterogeneity (Table 4),
thus improving the electrostatic repulsion effect of particles (Zhang et al., 2019b). Chatterjee et al. (2010) reported that surface charge
heterogeneity significantly affected the deposition of colloidal particles on the filter media. Similarly, the zeta potentials of biochar and
soil particles increased after application of DC electric field, and the Φmax value was further improved to a higher level with the
addition of HA. This result shows that electric field could also increase the electrostatic repulsion energy between biochar and soil
particles in the presence of humic acid, which was not conducive to mutual adhesion. Secondly, the steric effect prevented biochar
particles from adhering to the soil surface (Cao et al., 2022). In addition, humic acid has abundant redox-active groups, such as
quinone, phenol and carboxyl, which could increase the content of oxygen-containing functional groups (OCFGs) on the surface of
biochar upon contact (Wang et al., 2023). Following the application of a DC electric field, HA adsorbed on the surface of biochar could
potentially be oxidized, resulting in an increase in OCFGs. The dissociation of these OCFGs could further increase the density of
negative charges on the biochar, which in turn facilitate their migration in the soil (Yi et al., 2015).
As expected, the mobility of biochar particles in the soil in the presence of 5 and 10 mg⋅L− 1 HA was also slightly weaker than that in
quartz sand (Figure S8, Table S5), where the maximum biochar recovery can reach 81.4±2.2 % under optimal condition. Nevertheless,
the mass recovery of biochar particles in the effluent still reached a relatively high level under favorable conditions (72.2±1.2 % for
NPs at 1.0 V⋅cm− 1 and 5 mg⋅L− 1 HA, Table 4), demonstrating the effectiveness of applying DC electric field to enhance the mobility of
fine-textured biochar particles in real soil under specific environmental conditions.
Through multiple linear regression analysis, the significance values of two variables (potential gradient and HA) on the change of
mass recovery are listed in Table S3. It can be observed that the values of potential gradient and the concentration of HA on the mass
recovery of biochar particles in soil and quartz sand column tests are all less than 0.05 (P < 0.05), which emphasizes the important role

Fig. 4. Hydrodynamic diameter of MPs and NPs in effluents under different voltage gradients (0 and 0.5 V⋅cm− 1) in the presence of HA (0
and 5 mg⋅L− 1).

9
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

of the two factors in promoting the migration of biochar particles in the media.

3.4. Characterization of biochar colloids in the effluents

Following the transport tests, we measured the hydrodynamic estimated size of biochar particles in the effluent, as shown in Fig. 4.
In general, the diameter of biochar colloids in the effluent decreased dramatically (500–640 and 420–520 nm for MPs and NPs,
respectively) compared to their initial input levels (~3650 and 780 nm for MPs and NPs, respectively), especially for MPs. This was
confirmed by the results of the SEM analysis (Figure S3), where no large biochar aggregates were found in the effluent, whether for
MPs or NPs. Compared to NPs, the migration of larger sizes MPs was more difficult in real soil, so that only small fragments could
migrate out through the narrow soil pores. The main explanation for this is the size exclusion effect of the soil matrix on biochar
colloids, as larger particles and aggregates are more likely to be retained by the porous media (Wang et al., 2015). Fig. 4 also shows that
HA had little effect on the hydrodynamic diameters of biochar in the effluent. Since the addition of HA did not significantly change the
hydraulic diameter of biochar particles in the influent, the reason for the enhanced biochar migration in the soil in the presence of HA
was not due to the change in particle size. A possible reason for this is that the DC electric field promoted the disintegration or
dissociation of the carbonaceous and mineral fragments attached to the surface of the biochar particles (Liu et al., 2023), resulting in
some reduction in particle size.
FTIR spectra of MPs and NPs in the effluent at 0 and 0.5 V⋅cm− 1 in the presence of 0 and 5 mg⋅L− 1 HA are shown in Fig. 5. At
0 V⋅cm− 1, the spectral analysis revealed stronger stretching vibrations of hydroxyl -OH at 3420 cm− 1, -CH2- at 2925 cm− 1, C-H of
methylene at 2853 cm− 1 and 790 cm− 1, carboxyl -COOH at 1730 cm− 1, -C-O at 1100 cm− 1 and P-O at 470 cm− 1 in NPs compared to
MPs. Apparently, the characteristic peaks of these functional groups on the surface of NPs were more obvious after ball milling, which
in turn introduced a strong repulsive force that prevented the aggregation of biochar particles (Zhang et al., 2019a; Yang et al., 2019),
which can be consolidated by the enhanced ζ-potential of NPs. The decrease in the intensity of C– –C/C– –O may be due to the release of
aliphatic carbon, carboxyl carbon or tar particles in the pores of the biochar during ball milling (Wang et al., 2018; Jung et al., 2019;
Piccirillo et al., 2017). After the application of 0.5 V⋅cm− 1, the intensity of the OCFGs such as -OH, -COOH, C––O and P-O of MPs and
NPs was further increased, indicating that the DC electric field increased the content of OCFGs on the surface of biochar. The addition
of 5 mg⋅L− 1 HA, compared to the two biochar particles in the effluent at 0 V⋅cm− 1, further enhanced the stretching vibrations of -OH
and -C-O, indicating the increase in the content of OCFGs on the surface of biochar. The changes in the characteristic peaks were similar
to the trend of previous studies (Afzal et al., 2023). Similarly, the higher abundance of -OH, -CH2-, C––C/C– –O, -C-O, -CH and -P-O was
obtained for biochar particles under the stimulation of electric field at 0.5 V⋅cm− 1.

1
Fig. 5. FTIR spectra of MPs (a) and NPs (b) without HA, MPs (c) and NPs (d) in the presence of 5 mg⋅L− HA in the effluent under voltage gradient of
0 and 0.5 V⋅cm− 1.

10
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

The application of a DC electric field could increase the number of functional groups on the surface of biochar. This was mainly
because the dissociation of these acidic OCFGs (especially hydroxyl and carboxyl groups) under alkaline conditions resulting from the
application of the electric field increased the negative charges on the biochar (Liu et al., 2023). Electrodynamics were confirmed to
increase the number of oxygen-containing functional groups and the surface negative charge of biochar particles (Liu et al., 2022a). As
a result, the mobility of biochar in natural soil was enhanced, which was consistent with what we had observed in the column tests.

3.5. Fitting of transport modeling

It is clear from the BTC curves that biochar retention at the soil surface was reversible. Therefore, a non-equilibrium deterministic
model was used to simulate the breakthrough of biochar particles through the soil column under different conditions (Figs. 2 and 3).
The key model parameters R, β, ω, k1, and k2 are summarized in Table S4. The correlation coefficient (R2) of the breakthrough curve for
all experiments was greater than 0.95, indicating strong agreement between the data and the model. The parameter R indicates the
resistance of biochar migration within the soil column. A larger R indicates a more effective blockage from porous media to biochar
particles. β represents the distribution of adsorption sites; the lower β, the more abundant these sites. ω represents the dispersion
strength of biochar, with an increase in ω corresponding to a reduced aggregation potential. k1 and k2 represent the adsorption and
desorption coefficients of the biochar particle, respectively.
Derived from Table S4, our results show that the application of DC electric field decreased the retention factor R (for MPs from 1.89

Fig. 6. Correlation between DLVO maximum energy barrier (Φmax) and model parameters R (a), β (b), ω (c), k1 (d), and k2 (e) of soil column tests.

11
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

at 0 V⋅cm− 1 to 1.11 at 1.0 V⋅cm− 1, and for NPs from 1.31 at 0 V⋅cm− 1 to 0.68 at 1.0 V⋅cm− 1) and dispersion strength ω (for MPs from
12.80 at 0 V⋅cm− 1 to 6.32 at 1.0 V⋅cm− 1, and for NPs from 3.35 at 0 V⋅cm− 1 to 1.69 at 1.0 V⋅cm− 1), implying enhanced colloid
dispersion and suppressed column retention of biochar under the action of DC electric field. At the same time, a substantial increase in
β (from 0.05 and 0.50–0.53 and 0.80 for MPs and NPs, respectively) after application of DC electric field indicated a reduction in soil
sites for biochar sorption, resulting in decreased sorption of colloid particles by the soil, which was also favorable for biochar transport
in the soil matrix. In addition, we observed that the adsorption coefficient k1 decreased from 1.38 and 1.03–0.71 and 0.40, while
desorption coefficient k2 increased from 3.32×10− 8 and 1.00×10− 7 to 1.00×10− 7 and 5.21×10− 2 for MPs and NPs, respectively, after
increasing the potential gradient up to 1.0 V cm− 1, further supporting these observations.
With the addition of humic acid, the values of R, ω, and k1 decreased, while β and k2 increased, implying that the electric double
layer repulsion between biochar particles was enhanced and the relatively high ζ-Potential (negative) of biochar promoted the
electrostatic attraction in the electric field. At the same time, HA increased the electronegativity of the soil surface (Chen et al., 2017),
which further improved the migration of biochar colloids. Furthermore, the same trend occurred for the fitting parameters in the
presence of HA concentration at different potential gradients. For example, β, which represents the distribution of adsorption sites,
increased from 0.32 and 0.71 at 0 V⋅cm− 1 to 0.53 and 0.92 at 1.0 V⋅cm− 1 for MPs and NPs in the presence of 5 mg⋅L-1 HA, respectively.
And the adsorption coefficients k1 increased from 1.00 and 0.81 at 0 V⋅cm-1–0.53 and 0.13 at 1.0 V⋅cm-1 for MPs and NPs in the
presence of 5 mg⋅L-1 HA, respectively. This suggests that in the case of humic acid, the DC electric field similarly reduces the resistance
of the soil to biochar particles and improves the dispersion of biochar and the desorption of biochar into the medium, allowing more
biochar particles to migrate out of the soil. The fit of the parameters from the biochar migration experiments in the sand column
(Table S6) is also consistent with the trend.
To explore the correlation between DLVO theory and CDE model, the linear relationship between Φmax and five parameters (R, β, ω,
k1, and k2) in the soil (Fig. 6) and sand (Figure S10) column tests was analyzed by linear fitting. Clearly, most of the linear correlations
between Φmax from the DLVO theory and five parameters from the CDE model of biochar particles have reached a relatively high
degree (R2 > 0.9), indicating that the DLVO theory and the CDE model have a high correlation in jointly explaining the migration of
biochar in the column tests.

3.6. Environmental significance

In this study, we demonstrated that the introduction of a DC electric field significantly improved the mobility of colloidal biochar
particles in natural soil. Furthermore, it was shown that the migration of these particles could be controlled by potential gradient and
background humic acid concentration. This suggests that the application of a DC electric field has the potential to recover the biochar
particles being applied in the soil. Considering that biochar colloids can induce the adsorption of inorganic and organic pollutants on
them, we can also use DC electric field to deliver or relocate biochar particles in the soil for remediation purpose (Qin et al., 2020).
However, due to the heterogeneity of real soil, the actual transport behavior of biochar in different soil environmental conditions is
quite different because the physical and chemical properties (such as ionic strength, pH and natural organic matter) of different kind of
soil are highly diversified (Yang et al., 2017).
We must admit that our study only provided basic insights into the transport behavior of biochar particles under DC electric field in
the soil. The impact of DC electric field on the co-transport behavior of biochar and the adsorbed contaminants in different soil types
need to be discovered. The effect of DC electric field on the properties and long-term stability of biochar remains less known. These
aspects of the research should be further investigated in the future.

4. Conclusion

This study showed that the transport behavior of biochar micro- and nanoparticles in real soil was effectively promoted by the
application of a DC electric field. The increase in potential gradients further enhanced the mass recovery of biochar particles, especially
nano-sized particles, in soil by reducing the surface negative charge of biochar and soil. In addition, the enhanced mobility of NPs by
electric field in the presence of humic acid implied that biochar particles could be better recovered and regulated in the organic-rich
soil. Under the environmental conditions, the optimal mass recovery of biochar NPs could be significantly up to 72.2±1.2 % under
1.0 V⋅cm− 1 in the presence of 5 mg⋅L− 1 HA. The increasing content of oxygen-containing functional groups on the surface of biochar
improved the electrostatic repulsion between biochar and soil, and the reduced particle size mitigated the size-blocking effect between
biochar and soil matrix. The results of our study could provide theoretical and technical support for the application and recovery of
biochar in contaminated soils.

Funding

We greatly appreciate the financial support from National Natural Science Foundation of China (No. 41877123).

CRediT authorship contribution statement

Xiaolei Zhang: Validation. Weimin Cao: Writing – review & editing, Conceptualization. Chen Chen: Writing – review & editing,
Conceptualization. Hongjia Bao: Validation, Investigation. Qiang Liu: Funding acquisition, Data curation. Min Yang: Writing –
review & editing, Supervision, Conceptualization. Yifan Yang: Writing – original draft, Methodology, Conceptualization.

12
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements

We are grateful for the help from the Instrumental Analysis & Research Center of Shanghai University for sample characterization.

Appendix A. Supporting information

Supplementary data associated with this article can be found in the online version at doi:10.1016/j.eti.2024.103704.

References

Afzal, M.Z., Hameed, S., Mohiuddin, M., Abbasi, A., 2023. Simultaneous adsorptive removal of three fluoroquinolones using humic acid modified hydrogel beads.
Environ. Sci. Pollut. Res. 30, 24398–24407. https://doi.org/10.1007/s11356-022-23855-3.
Bardi, M.J., Mutunga, J.M., Ndiritu, H., Koch, K., 2023. Effect of pyrolysis temperature on the physiochemical properties of biochar and its potential use in anaerobic
digestion: a critical review. Environ. Technol. Innov. 32, 103349 https://doi.org/10.1016/j.eti.2023.103349.
Bradford, S.A., Simunek, J., Bettahar, M., Van Genuchten, M.T., Yates, S.R., 2003. Modeling colloid attachment, straining, and exclusion in saturated porous media.
Environ. Sci. Technol. 37, 2242–2250. https://doi.org/10.1021/es025899u.
Cao, Y., Ma, C., Yao, J., Chen, W., Gu, L., Liu, H., Liu, C., Xiong, J., Huangfu, X., 2023a. Impact of biochar colloids on thallium(I) transport in water-saturated porous
media: effects of pH and ionic strength. Chemosphere 311, 137152. https://doi.org/10.1016/j.chemosphere.2022.137152.
Cao, G., Qiao, J., Ai, J., Ning, S., Sun, H., Chen, M., Zhao, L., Zhang, G., Lian, F., 2022. Systematic research on the transport of ball-milled biochar in saturated porous
media: effect of humic acid, ionic strength, and cation types. Nanomaterials 12 (6), 988. https://doi.org/10.3390/nano12060988.
Chatterjee, J., Abdulkareem, S., Gupta, S.K., 2010. Estimation of colloidal deposition from heterogeneous populations. Water Res 44, 3365–3374. https://doi.org/
10.1016/j.watres.2010.03.025.
Chen, M., Wang, D., Yang, F., Xu, X., Xu, N., Cao, X., 2017. Transport and retention of biochar nanoparticles in a paddy soil under environmentally-relevant solution
chemistry conditions. Environ. Pollut. 230, 540–549. https://doi.org/10.1016/j.envpol.2017.06.101.
Chung, J., Kim, Y.J., Lee, G., Nam, K., 2016. Experimental determination of nonequilibrium transport parameters reflecting the competitive sorption between Cu and
Pb in slag-sand column. Chemosphere 154, 335–342. https://doi.org/10.1016/j.chemosphere.2016.03.108.
Cornelis, G., Pang, L., Doolette, C., Kirby, J.K., McLaughlin, M.J., 2013. Transport of silver nanoparticles in saturated columns of natural soils. Sci. Total Environ.
463–464, 120–130. https://doi.org/10.1016/j.scitotenv.2013.05.089.
Cross, A., Sohi, S.P., 2013. A method for screening the relative long-term stability of biochar. GCB Bioenergy 5, 215–220. https://doi.org/10.1111/gcbb.12035.
Gomes, H.I., Dias-Ferreira, C., Ribeiro, A.B., Pamukcu, S., 2014. Influence of electrolyte and voltage on the direct current enhanced transport of iron nanoparticles in
clay. Chemosphere 99, 171–179. https://doi.org/10.1016/j.chemosphere.2013.10.065.
Hung, C.M., Huang, C.P., Hsieh, S.L., Chen, Y.T., Chen, C.W., Dong, C.D., 2023. Water hyacinth derived biochar for polycyclic aromatic hydrocarbons removal and
oxidative stress study. Environ. Technol. Innov. 29, 103027 https://doi.org/10.1016/j.eti.2023.103027.
Jiang, X., Liu, Y., Yin, X., Deng, Z., Zhang, S., Ma, C., Wang, L., 2023. Efficient removal of chromium by a novel biochar-microalga complex: mechanism and
performance. Environ. Technol. Innov. 31, 103156 https://doi.org/10.1016/j.eti.2023.103156.
Jung, K.W., Lee, S.Y., Choi, J.W., Lee, Y.J., 2019. A facile one-pot hydrothermal synthesis of hydroxyapatite/biochar nanocomposites: adsorption behavior and
mechanisms for the removal of copper(II) from aqueous media. Chem. Eng. J. 369, 529–541. https://doi.org/10.1016/j.cej.2019.03.102.
Khodadoust, A.P., Reddy, K.R., Narla, O., 2006. Cyclodextrin-enhanced electrokinetic remediation of soils contaminated with 2,4-dinitrotoluene. J. Environ. Eng. 132,
1043–1050. https://doi.org/10.1061/(ASCE)0733-9372(2006)132:9(1043).
Kretzschmar, R., Sticher, H., 1997. Transport of humic-coated iron oxide colloids in a sandy soil: influence of Ca2+ and trace metals. Environ. Sci. Technol. 31,
3497–3504. https://doi.org/10.1021/es970244s.
Kumar, R., Sharma, P., Rose, P.K., Sahoo, P.K., Bhattacharya, P., Pandey, A., Kumar, M., 2023. Co-transport and deposition of fluoride using rice husk-derived biochar
in saturated porous media: effect of solution chemistry and surface properties. Environ. Technol. Innov. 30, 103056 https://doi.org/10.1016/j.eti.2023.103056.
Liu, Y., Bao, H., Chen, C., Cao, W., Zhang, X., Xu, Y., Ngo, H.H., Liu, Q., 2024. Recovery of biochar particles laden with lead in saturated porous media by DC electric
field. Chemosphere 355, 141890. https://doi.org/10.1016/j.chemosphere.2024.141890.
Liu, Q., Jiang, Y., Liu, Y., Zhang, X., Xu, Y., Chen, H., Cao, W., 2022a. Enhanced 2,4,6-trichlorophenol removal from soil by electrokinetic remediation coupled with
biochar in a permeable reactive barrier. Environ. Technol. Innov. 28, 102835 https://doi.org/10.1016/j.eti.2022.102835.
Liu, Q., Zhang, Q., Jiang, S., Du, Z., Zhang, X., Chen, H., Cao, W., Nghiem, L.D., Ngo, H.H., 2022b. Enhancement of lead removal from soil by in-situ release of
dissolved organic matters from biochar in electrokinetic remediation. J. Clean. Prod. 361, 132294 https://doi.org/10.1016/j.jclepro.2022.132294.
Liu, Y., Zhang, X., Xu, Y., Liu, Q., Ngo, H.H., Cao, W., 2023. Transport behaviors of biochar particles in saturated porous media under DC electric field. Sci. Total
Environ. 856, 159084 https://doi.org/10.1016/j.scitotenv.2022.159084.
Lyu, H., Gao, B., He, F., Zimmerman, A.R., Ding, C., Tang, J., Crittenden, J.C., 2018. Experimental and modeling investigations of ball-milled biochar for the removal
of aqueous methylene blue. Chem. Eng. J. 335, 110–119. https://doi.org/10.1016/j.cej.2017.10.130.
Ma, P., Qi, Z., Wu, X., Ji, R., Chen, W., 2023. Biochar nanoparticles-mediated transport of organic contaminants in porous media: Dependency on contaminant
properties and effects of biochar aging. Carbon Res 2, 4. https://doi.org/10.1007/s44246-023-00036-6.
Nguyen, T.B., Nguyen, V.T., Hoang, H.G., Cao, N.D.T., Nguyen, T.T., Vo, T.D.H., Nguyen, N.K.Q., Pham, M.D.T., Nghiem, D.L., Vo, T.K.Q., Dong, C.Di, Bui, X.T., 2023.
Recent development of algal biochar for contaminant remediation and energy application: a state-of-the art review. Curr. Pollut. Rep. 9, 73–89. https://doi.org/
10.1007/s40726-022-00243-6.
Piccirillo, C., Moreira, I.S., Novais, R.M., Fernandes, A.J.S., Pullar, R.C., Castro, P.M.L., 2017. Biphasic apatite-carbon materials derived from pyrolysed fish bones for
effective adsorption of persistent pollutants and heavy metals. J. Environ. Chem. Eng. 5, 4884–4894. https://doi.org/10.1016/j.jece.2017.09.010.
Qin, C., Wang, H., Yuan, X., Xiong, T., Zhang, J., Zhang, J., 2020. Understanding structure-performance correlation of biochar materials in environmental remediation
and electrochemical devices. Chem. Eng. J. 382, 122977 https://doi.org/10.1016/j.cej.2019.122977.

13
Y. Yang et al. Environmental Technology & Innovation 35 (2024) 103704

Reddy, K.R., Darko-Kagya, K., Cameselle, C., 2011. Electrokinetic-enhanced transport of lactate-modified nanoscale iron particles for degradation of dinitrotoluene in
clayey soils. Sep. Purif. Technol. 79, 230–237. https://doi.org/10.1016/j.seppur.2011.01.033.
Song, B., Chen, M., Zhao, L., Qiu, H., Cao, X., 2019. Physicochemical property and colloidal stability of micron- and nano-particle biochar derived from a variety of
feedstock sources. Sci. Total Environ. 661, 685–695. https://doi.org/10.1016/j.scitotenv.2019.01.193.
Song, J., Peng, P., Huang, W., 2002. Black carbon and kerogen in soils and sediments. 1. Quantification and characterization. Environ. Sci. Technol. 36, 3960–3967.
https://doi.org/10.1021/es025502m.
Sun, P., Shijirbaatar, A., Fang, J., Owens, G., Lin, D., Zhang, K., 2015a. Distinguishable transport behavior of zinc oxide nanoparticles in silica sand and soil columns.
Sci. Total Environ. 505, 189–198. https://doi.org/10.1016/j.scitotenv.2014.09.095.
Tian, Y., Chen, N., Yang, X., Li, C., He, W., Ren, N., Liu, G., Yang, W., 2023. Migration electric-field assisted electrocoagulation with sponge biochar capacitive
electrode for advanced wastewater phosphorus removal. Water Res 231, 119645. https://doi.org/10.1016/j.watres.2023.119645.
Tran, D.T., Pham, T.D., Dang, V.C., Pham, T.D., Nguyen, M.V., Dang, N.M., Ha, M.N., Nguyen, V.N., Nghiem, L.D., 2022. A facile technique to prepare MgO-biochar
nanocomposites for cationic and anionic nutrient removal. J. Water Process Eng. 47, 102702 https://doi.org/10.1016/j.jwpe.2022.102702.
Tufenkji, N., Elimelech, M., 2004. Correlation equation for predicting single-collector efficiency in physicochemical filtration in saturated porous media. Environ. Sci.
Technol. 38, 529–536. https://doi.org/10.1021/es034049r.
Wang, C., Cheng, T., Zhang, D., Pan, X., 2023. Electrochemical properties of humic acid and its novel applications: a tip of the iceberg. Sci. Total Environ. 863, 160755
https://doi.org/10.1016/j.scitotenv.2022.160755.
Wang, X., Dan, Y., Diao, Y., Liu, F., Wang, H., Sang, W., 2022. Transport and retention of microplastics in saturated porous media with peanut shell biochar (PSB) and
MgO-PSB amendment: Co-effects of cations and humic acid. Environ. Pollut. 305, 119307 https://doi.org/10.1016/j.envpol.2022.119307.
Wang, B., Gao, B., Wan, Y., 2018. Entrapment of ball-milled biochar in Ca-alginate beads for the removal of aqueous Cd(II). J. Ind. Eng. Chem. 61, 161–168. https://
doi.org/10.1016/j.jiec.2017.12.013.
Wang, D., Jin, Y., Jaisi, D.P., 2015. Effect of size-selective retention on the cotransport of hydroxyapatite and goethite nanoparticles in saturated porous media.
Environ. Sci. Technol. 49, 8461–8470. https://doi.org/10.1021/acs.est.5b01210.
Wang, D., Zhang, W., Hao, X., Zhou, D., 2013. Transport of biochar particles in saturated granular media: effects of pyrolysis temperature and particle size. Environ.
Sci. Technol. 47, 821–828. https://doi.org/10.1021/es303794d.
Xiao, J., Hu, R., Chen, G., 2020. Micro-nano-engineered nitrogenous bone biochar developed with a ball-milling technique for high-efficiency removal of aquatic Cd
(II), Cu(II) and Pb(II). J. Hazard. Mater. 387, 121980 https://doi.org/10.1016/j.jhazmat.2019.121980.
Yang, W., Shang, J., Sharma, P., Li, B., Liu, K., Flury, M., 2019. Colloidal stability and aggregation kinetics of biochar colloids: effects of pyrolysis temperature, cation
type, and humic acid concentrations. Sci. Total Environ. 658, 1306–1315. https://doi.org/10.1016/j.scitotenv.2018.12.269.
Yang, X., Wan, Y., Zheng, Y., He, F., Yu, Z., Huang, J., Wang, H., Ok, Y.S., Jiang, Y., Gao, B., 2019. Surface functional groups of carbon-based adsorbents and their
roles in the removal of heavy metals from aqueous solutions: a critical review. Chem. Eng. J. 366, 608–621. https://doi.org/10.1016/j.cej.2019.02.119.
Yang, W., Wang, Y., Shang, J., Liu, K., Sharma, P., Liu, J., Li, B., 2017. Antagonistic effect of humic acid and naphthalene on biochar colloid transport in saturated
porous media. Chemosphere 189, 556–564. https://doi.org/10.1016/j.chemosphere.2017.09.060.
Yi, P., Pignatello, J.J., Uchimiya, M., White, J.C., 2015. Heteroaggregation of cerium oxide nanoparticles and nanoparticles of pyrolyzed biomass. Environ. Sci.
Technol. 49, 13294–13303. https://doi.org/10.1021/acs.est.5b03541.
Yuan, S., Long, H., Xie, W., Liao, P., Tong, M., 2012. Electrokinetic transport of CMC-stabilized Pd/Fe nanoparticles for the remediation of PCP-contaminated soil.
Geoderma 185–186 18–25. https://doi.org/10.1016/j.geoderma.2012.03.028.
Yuan, R., Salam, M., Miao, X., Yang, Y., Li, H., Wei, Y., 2023. Potential disintegration and transport of biochar in the soil-water environment: A case study towards
purple soil. Environ. Res. 222, 115383 https://doi.org/10.1016/j.envres.2023.115383.
Zhang, W., Morales, V.L., Cakmak, M.E., Salvucci, A.E., Geohring, L.D., Hay, A.G., Parlange, J.Y., Steenhuis, T.S., 2010. Colloid transport and retention in unsaturated
porous media: effect of colloid input concentration. Environ. Sci. Technol. 44, 4965–4972. https://doi.org/10.1021/es100272f.
Zhang, X., Shu, S., Hou, D., Chen, H., Cao, W., Mameda, N., Nghiem, L.D., Liu, Q., 2023. Role of the surface characteristics of hyper-crosslinked polymers on the
transformation of adsorbed trichlorophenol: implications for understanding the surface reactivity of biochar derived from waste biomass. Sci. Total Environ. 886,
163864 https://doi.org/10.1016/j.scitotenv.2023.163864.
Zhang, Q., Wang, J., Lyu, H., Zhao, Q., Jiang, L., Liu, L., 2019a. Ball-milled biochar for galaxolide removal: sorption performance and governing mechanisms. Sci.
Total Environ. 659, 1537–1545. https://doi.org/10.1016/j.scitotenv.2019.01.005.
Zhang, R., Zhang, H., Tu, C., Luo, Y., 2019b. The limited facilitating effect of dissolved organic matter extracted from organic wastes on the transport of titanium
dioxide nanoparticles in acidic saturated porous media. Chemosphere 237, 124529. https://doi.org/10.1016/j.chemosphere.2019.124529.

14

You might also like