Development of A New Hydrophobic Magnetic Biochar For Removing Oil Spills On The Water Surface
Development of A New Hydrophobic Magnetic Biochar For Removing Oil Spills On The Water Surface
Abstract
More technologies are urgently needed for combined use to effectively eliminate the effect of oil spills, an envi-
ronmental problem of widespread concern. Among these technologies, sorption methods are available to remove
residual oil and prevent the further spread on the water surface. In this study, biochars, prepared from different feed-
stock materials and pyrolysis temperatures, were screened and further modified to improve their application in the
water environment. Among cornstalk biochar (CSBC), corncob biochar (CCBC), Sophora sawdust biochar (SSBC), and
rice husk biochar (RHBC), the CSBC had excellent oil sorption capacity, especially prepared at 350℃ (CSBC350), which
has a complete and full pore structure. Furthermore, magnetic and silane agent modifications of CSBC350 (OMBC)
were performed to enhance the properties of the magnetic field controllability and hydrophobicity to increase oil
sorption. The OMBC exhibited satisfactory oil sorption capacities to crude oil, diesel oil, and engine oil in the water-oil
system of 8.77 g g−1, 4.01 g g−1, and 4.44 g g−1, respectively. The sorption process of CSBC350 and OMBC complied
with the pseudo-second-order kinetics (R2 > 0.97) and the Langmuir isotherm models (R2 > 0.80) based on the highest
regression coefficients. The sorption mechanisms are dominated by hydrophobic forces, pore intercepts, and hydro-
gen-bond interactions. The biochar adsorbent can availably cooperate with other physical methods to eliminate oil
contaminants, which can be an outstanding fuel source for producing heat.
Highlights
• Low temperature (350℃) pyrolyzed cornstalk biochar has the highest oil sorption ability among the various bio-
chars.
• Modified cornstalk biochar has excellent hydrophobicity and floatage with a higher oil adsorptivity.
• The oil sorption process of the modified biochar conforms to pseudo-first-order kinetics and the Langmuir iso-
therm models.
Keywords: Oil spill, Cornstalk biochar, Hydrophobic magnetic biochar, Oil sorption
*Correspondence: [email protected]
1
Frontiers Science Center for Deep Ocean Multispheres and Earth System,
and Key Laboratory of Marine Chemistry Theory and Technology, Ministry
of Education, Ocean University of China, Qingdao 266100, China
Full list of author information is available at the end of the article
© The Author(s) 2022. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which
permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the
original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or
other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line
to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this
licence, visit http://creativecommons.org/licenses/by/4.0/.
Sun et al. Biochar (2022) 4:60 Page 2 of 17
Graphical Abstract
contaminants containing antibiotics (Liang et al. 2022), n-octadecyl trimethoxy silane (OTMS) was obtained
dyes (Wathukarage et al. 2019), persistent organic pol- from Rhawn (China). Ferrous sulfate heptahydrate
lutants (POPs) (Shi et al. 2016; Zhang et al. 2018), heavy (FeSO4·7H2O), ferric chloride hexahydrate (FeCl3·6H2O),
metals (Agrafioti et al. 2014), nitrogen and phosphorus sodium hydroxide, and ethanol were purchased from Sin-
(Dai et al. 2020), based on its porosity, high surface area, opharm Chemical Reagent Limited Corporation (Shang-
and abundant functional groups, etc. hai, China). Diesel and engine oil were provided by a local
In recent years, biochar has gradually attracted much petrol station (Qingdao, China). Crude oil was obtained
attention on adsorbing oil. In previous studies, various from an oilfield in the Bohai Sea (China).
biochars have been studied as adsorbents for the puri-
fication of oil pollution, such as rice husk BC (Angelova
et al. 2011), peat-derived BC (AlAmeri et al. 2019), and 2.2 Preparation of biochar
sawdust BC (Silvani et al. 2017). The oil sorption effi- The process of biochar preparation is shown in Addi-
ciency of pristine BC to a great extent relies on its hydro- tional file 1: Fig. S1. The feedstock material, with impu-
phobicity, which guarantees that biochar does not sink rities removed, was chopped into small chunks, washed
into the aqueous phase (Madhubashani et al. 2021). The with deionized water 3 times, and dried at 105℃ over-
enhancement of hydrophobicity by modifying the bio- night. A crucible was filled with feedstock material and
char surface can facilitate the sorption of the oil. Hence, was wrapped in aluminum foil. Then pyrolysis was per-
research focused on biochar modification had been con- formed at 350℃, 500℃, and 650℃ in a muffle furnace
ducted by researchers. Fir BC, red pine BC, and lettuce (Zhonghuan, SX-G07103) under oxygen-limited condi-
BC decorated with lauric acid, coconut oil, and polydi- tions achieved by nitrogen gas (N2) purging in advance.
methylsiloxane (PDMS) have been studied to improve The heating rate of the reactor was 10℃ min−1 and the
sorption (Chen et al. 2020a, b; Gurav et al. 2021a, b; target pyrolysis temperature was maintained for 4 h.
Navarathna et al. 2020). However, there are few research After the reactor was cooled to room temperature, the
concerns about the oil sorption variation resulting from biochars were removed and milled, passing through a
the difference in precursor materials and pyrolysis tem- 0.55 mm mesh screen for later use. According to the
perature. Accordingly, we investigated the characteris- pyrolysis temperature and feedstock, the prepared pris-
tics of the inherent properties and sorption capacity of tine biochars were referred to as CSBC350, CSBC500,
biochars produced from different representative feed- CSBC650, CCBC500, SSBC500, and RHBC500. Fur-
stock materials and pyrolysis temperatures. The water thermore, the specific biochars were screened into three
repellency of biochar has been unsatisfactory for a long different sizes, small (0.55–0.25 mm), medium (0.25–
time. n-Octadecyl trimethoxy silane (OTMS) is a hydro- 0.15 mm), and large (< 0.15 mm).
phobic modification reagent with excellent performance
and eco-friendliness. Hydrolysis and condensation reac- 2.3 Modification of biochar
tions easily occur to form a hydrophobic film layer on the The magnetic biochar was prepared as described by
material surface, which makes the material have better Mohan (2014). The main process is presented in Fig. 1a.
water resistance. Then, biochar modification was per- FeCl3·6H2O (1.80 g) and F eSO4·7H2O (1.85 g) were dis-
formed by employing n-Fe3O4 and OTMS for control- solved in 150 mL distilled water for 5 min at 60℃ to
lability and hydrophobicity. The surface morphology, prepare the F e2+/Fe3+ solution. Then, 2.0 g of CSBC350
surface element distribution, BET surface area, functional was slowly added to the Fe2+/Fe3+ solution and stirred
groups, crystal structure, magnetic properties, thermal for 30 min, forming a mixed suspension. After that, 1 M
analysis, flotation, and wettability of the prepared adsor- NaOH was carefully added to the mixed suspension
bents were investigated. The performances of the sorp- until the pH reached ~ 11, and the suspension succes-
tion capacity, availability, sorption kinetics, and sorption sively became brown and black at ~ pH 6 and ~ pH 10,
isotherm were tested. The target adsorbent is expected to respectively. Subsequently, the suspension was heated
float on the water to adsorb spilled oil individually or as to 80℃ for 1 h and aged at room temperature for 2 h.
a complement to other techniques, which is beneficial to Then, a magnet was used to separate the magnetic bio-
the thoroughness of oil spill clean-up work. char (MBC), and the product was washed several times
with distilled water and ethanol. Finally, the prepared
2 Materials and methods MBC was filtered by a vacuum suction filter and dried
2.1 Feedstock and chemicals at 50℃ in a vacuum drying chamber. The fabrication of
Cornstalk (CS), corncob (CC), Sophora sawdust (SS), hydrophobic magnetic biochar was as follows: the mag-
and rice husk (RH) were collected from a local farm- netic biochar and OTMS were mixed and heated for 3 h
land and wood-working factory (Qingdao, China). The in 100 mL of 25% aqueous ethanol solution, centrifuged
Sun et al. Biochar (2022) 4:60 Page 4 of 17
Fig. 1 General scheme for the preparation of the hydrophobic magnetic biochar (a). The schematic diagram of the silylation reaction mechanism
for the magnetic biochar modification with OTMS (b)
for 10 min at 6000 rpm and washed several times with The sorption capacity of these adsorbents was calcu-
ethanol. The obtained biochar was named OMBC. lated as follows (Eq. 1):
mc −mb −ma
2.4 Sorption experiment Q= (1)
ma
The adsorptive performance of several pristine biochars
was determined in a pure oil system. The detailed steps where Q is the mass sorption capacity of the adsorbent (g
were performed according to the standard test method g−1), ma is the weight of the adsorbent (g), mb is the mass
for the sorption performance of sorbents as given by sorption capacity of the control grid (g), and mc is the
ASTM F726 (Bazargan et al. 2015). A certain quantity final weight of the grid and adsorbent after sorption (g).
of biochar was placed on a preweighed stainless steel The experiments to determine the effect of salinity, pH,
grid and immersed in the oil for 1 h to equilibrate at and temperature on the sorption of crude oil by OMBC
room temperature. Afterward, before being weighed, were carried out by changing a single condition based
the excess oil was drained from the grid for 30 min on the above description. After reaching adsorption
until no drops fell. The sorption capacity was obtained equilibrium, a magnet was used to collect the OMBC oil
by subtracting the sorption capacity of the grid. BC and mixture. Then, 20 mL of n-hexane was used to elute the
modified biochars (MBC, OMBC) were employed for adsorbed oil for three times, and the sorbent was washed
the removal of petroleum oil from the water surface. again with deionized water. Finally, the cleaned sorbent
Measurement of the adsorptive capacity in the oil- was dried and weighed.
water system was similar to that in the oil system. In
brief, 200 mL of freshwater was poured into a beaker
(250 mL) and 2.0 g of oil was added to simulate spilled 2.5 Sorption kinetics and isotherm models
oil. The stainless steel grid carrying 0.2 g of adsorbent The kinetics and isotherm models for oil sorption on bio-
was placed in the oil and was allowed to float on the char and modified biochar were studied in a water–oil
surface of water for 2 h at room temperature. Next, system at 25℃. The adsorbent kinetics were determined
the excess oil was allowed to drain off the grid for by adding 200 mL of distilled water and 2 g of oil to a
30 min, and the grid was dried in a drying oven at 60℃ 250 mL glass beaker and varying the contact time from
to remove adsorbed water. In the same way, the blank 0 to 240 min. For the isotherm study, adsorbent uptake
net was used as a control to measure the real sorption capacities were determined by changing the quantity
capacity. For the modified biochars (MBC, OMBC), a of each added oil from 0.30 to 7 g with 0.20 g of mate-
magnet was used to separate and collect the biochars rial while holding other factors constant. The detailed
after sorption equilibrium.
Sun et al. Biochar (2022) 4:60 Page 5 of 17
contents of the models are presented in the Supplemen- full holes became crevices, which is attributed to the
tary Information. formation of fine stomates that connect and destroy
the original structure (Xiong et al. 2021). The prop-
2.6 Characterization methods erties and structure of biochar rely on the type and
The surface morphology and elemental compositions of nature of the lignocellulosic feedstock used. Agricul-
the adsorbent material were observed through scanning tural residues, such as corn stalks, contain more cel-
electron microscopy (SEM, TESCAN, TESCAN MIRA4, lulose and hemicellulose and less lignin. During the
Czech Republic) equipped with an energy dispersive preparation process, the cellulose and hemicellulose,
spectrometer (EDS, Xplore). The N2 adsorption–desorp- which needed a low pyrolysis temperature (200–400℃)
tion experiments were conducted using a surface area first decomposed to form pores. Subsequently, lignin
analyzer (Micromeritics, APSP 2460, USA). The surface gradually decomposed with increasing pyrolysis temper-
area and pore size of the biochars were obtained via the ature (> 500℃), which left a portion of air holes (Yaashi-
Brunauer–Emmett–Teller (BET) and Barrett-Joyner- kaa et al. 2019). Under high temperature conditions,
Halenda (BJH) analysis. The composition of CSBC350, cellulose and hemicellulose were completely gasified, and
MBC, and OMBC was analyzed using X-ray diffractom- lignin began to change, which caused the catastrophic
etry (XRD, Rigaku, SmartLab-SE, Japan) from 5° to 90°. collapse of the porous structure.
The changes in the functional groups of the samples were The SEM images of MBC and OMBC are shown in
determined by Fourier transform infrared spectroscopy Fig. 3. As shown in Fig. 3a, abundant magnetic parti-
(FT-IR, MAGNA-560, Nicolet, USA) within the range cles and their aggregates were deposited on the surface
of 4000 to 400 cm−1 in transmittance mode. The hydro- of the biochar. Likewise, there was a depositional plane
phobicity of the CSBCs and modified BC was measured at the opening of the channel, as seen in Fig. 3b. Nota-
by a contact angle measuring device (CA, OCA 20, Data bly, there was no blocking phenomenon in the porous
Physics ES, Germany) at room temperature by releasing structure, and the original porous construction was
2 μL of water on the sample surface. The magnetic prop- maintained to a large extent. As seen in Fig. 3c, the
erties of the samples were recorded by vibrating sample magnetic nanoparticles had a lamellate structure, with
magnetometer (VSM, Lake Shore, 7404, USA) in a field a nano magnitude size of less than 100 nm (Karuna-
of 15,000 to 15,000 Oe. The thermal stability of adsor- nayake et al. 2019; Navarathna et al. 2019). After fur-
bent was characterized by thermogravimetric analysis ther modification using OTMS, the MBC appeared to
(TGA and DTG, METTLER TOLEDO, TGA2/DSC3, have a smoother surface that coated the rough magnetic
Switzerland) under nitrogen conditions from 30 to 800℃. layer (Fig. 3d). At the same time, the macropores left by
The iron content of the water sample was measured by pyrolysis had well-defined borders as before, as shown in
inductively coupled plasma-mass spectrometry (ICP‒ Fig. 3d and e. The surface element compositions of the
MS, Agilent, 7700, USA). OMBC were analyzed by EDS spectroscopy. The distri-
bution of elemental C, O, Fe, and Si was observed from
3 Results and discussion the corresponding SEM‒EDS mapping images (Fig. 3 g),
3.1 Characterization of the adsorbents with their peaks appearing at approximately 0.267, 0.521,
3.1.1 Surface morphology and surface area analysis 0.710, and 1.74 keV, respectively (Additional file 1: Fig.
The surface appearances of the CSBCs produced at S2). Elemental C and O were the main component ele-
three pyrolysis temperatures (350℃, 500℃, and 650℃) ments, mainly from the biochar substrate. Elemental Fe
were observed by SEM, as shown in Fig. 2. The images and Si were found in trace amounts in the biochar and
show that all CSBCs had an obvious three-dimensional were mainly a result of the magnetic nano iron oxide and
macropore structure, regardless of the preparation the OTMS coating, respectively. In summary, according
temperature. However, there were still some evident to their elemental mapping images (Fig. 3g), iron oxide
structural distinctions that can be found among them. nanoparticles were evenly distributed on the surface of
Figure 2 shows that with the increase in pyrolysis tem- the biochar, and homogeneously covered by the OTMS.
perature from 350℃ to 650℃, the porous structure of The modification reaction process is presented in Fig. 1b.
CSBCs appeared to gradually shrink, and even collapse. The -SiOCH3 functional group of the OTMS hydrolysis
Specifically, CSBC350 had a tidy, full, and sharply mar- became -SiOH. Then, the dehydration reaction occurred
ginated cellular pore structure (Fig. 2a and b). CSBC500 between the -SiOH and the hydrophilic -OH functional
appeared as a disordered, fragmented surface under groups on the surface of the MBC.
continuous pyrolysis (Fig. 2c and d). Furthermore, the The specific surface area and pore size of the biochars
situation of CSBC650 was more prominent (Fig. 2e were obtained via the Brunauer–Emmett–Teller (BET)
and f ): numerous pores existed on the surface, and
Sun et al. Biochar (2022) 4:60 Page 6 of 17
Fig. 2 The SEM images of the CSBC350 (a, b), CSBC500 (c, d), and CSBC650 (e, f)
Sun et al. Biochar (2022) 4:60 Page 7 of 17
Fig. 3 SEM micrographs of MBC (a–c) and OMBC (d–f). A selected area on the OMBC for EDS measurement (f). Elemental mapping images of
OMBC for C, O, Fe, and Si (g)
Table 1 BET surface area and pore size of biochars and Barrett-Joyner-Halenda (BJH) analyses. The detailed
data are shown in Table 1.
Biochars BET surface Pore Average pore BJH pore
area (m2 volume diameter (nm) diameter The biochar obtained at low pyrolysis temperature-
g−1) (cm3 g−1) (nm) shas more macropores, relatively few mesopores and
micropores, and a smaller BET surface area. With the
CSBC350 3.1982 0.005847 5.7095 12.3097
increase of the pyrolysis temperature, the hemicellulose
CSBC500 7.8631 0.012912 5.6284 14.7617
and lignin in the cornstalk gradually volatilized, which
CSBC650 208.3195 0.096402 1.8256 3.2387
resulted in the shrinkage or fragmentation of macropores
MBC 48.2159 0.209931 14.9065 17.1995
and the increase in the ratio of mesopore and micropore,
OMBC 1.8881 0.002258 4.2508 5.9620
thus BET surface area and pore volume increased. This
Sun et al. Biochar (2022) 4:60 Page 8 of 17
is the reason why the CSBC650 had a higher BET sur- compound-rich carbon material with lower polarity (Kei-
face area and pore volume, reaching 208.3194 m2 g−1 and luweit et al. 2010; Sun et al. 2011). For the MBC spectra,
0.096402 cm3 g−1, respectively. However, at higher pyrol- the most noteworthy was the Fe–O characteristic peak
ysis temperatures, the macropores and mesopores were (575 cm−1), which indicates the success of magnetic
reduced, and the micropores became the main pores. loading on CSBC350 (Tapia et al. 2019). In the OMBC
Hence, the average pore diameter and mesopore diame- spectrum, the peaks at approximately 2917 and 2846,
ter were smaller for CSBC650, 1.8256 nm and 3.2387 nm, 1466 cm−1 were attributed to aliphatic CH and C H2
respectively. The above results were also verified in Fig. 2 stretching vibrations. Furthermore, the peaks at 1027,
SEM images. 790 cm−1, and 880 cm−1 were associated with Si–O–Si
In addition, the BET surface area, pore volume, and (Villegas et al. 1988) and Si–C (Zhou et al. 2018) stretch-
BJH pore diameter of MBC were increased to 48.2159 ing vibrations, respectively. Simultaneously, the peak at
m2 g−1, 0.209931 nm, and 17.1995 nm, due to the depo- around 575 cm−1 was also from the Fe–O band. The pri-
sition of magnetic nanoparticles on the surface and the mary peaks in the OMBC spectrographic curve strongly
macropore’s inner wall of CSBC350 (Fig. 3a–c). How- suggested that the magnetic nanoparticles and the long-
ever, the modification of OTMS caused the rough sur- chain hydrophobic alkyl groups of OTMS had been
face of MBC to become smooth, destroying the existing grafted onto the surface of biochar as expected.
mesoporous and microporous structures. Therefore, XRD analysis was used to determine the crystalline
the BET surface area, pore volume, and pore diameter structures of the samples. As shown in Fig. 4b, the peak
of OMBC were decreased compared with before the at around 22.7° in the XRD pattern of CSBC350 was the
modification. diffraction from the transformation of the lignocellu-
losic crystal structure to a more carbonaceous structure
3.1.2 FT‑IR and XRD analysis after pyrolysis (Regmi et al. 2012). This broad peak dis-
Figure 4a demonstrates the FT-IR spectra of CSBC350, appeared after magnetite precipitated due to magnetite’s
CSBC500, CSBC650, MBC and OMBC. From the spec- highly intense crystalline peaks (Navarathna et al. 2020).
trogram of the CSBCs, it can be easily concluded that It can be seen from the spectra of CSBCs that the peaks
the organic functional groups either faded away or were at 28.41°, 40.51°, 50.25°, and 58.66°, etc. indicate the pres-
eliminated, and the curves became smooth as the pyrol- ence of potassium chloride, especially in CSBC500. This
ysis temperature increased. These organic and inor- is due to the release of potassium from biomass in the
ganic functional groups were predominated by –OH form of potassium chloride during the pyrolysis process
(3406 cm−1), aliphatic CH stretching (2927 cm−1), ester (Fatehi et al. 2015). It is more pronounced at 500℃ and
C=O (1698 and 1562 cm−1), aromatic C=C (1580 cm−1), alleviated due to volatilization at higher pyrolysis tem-
and C–O (1200 to 1060 cm−1) (Qian et al. 2016; Xie peratures (650℃) (Jensen et al. 2000). For both modified
et al. 2021). The biochar produced at lower tempera- adsorbents, it was completely obvious that there were
tures had lower aromaticity, compared with the biochar multiple typical iron oxide diffraction peaks. The peak
prepared at higher temperatures, which was an aromatic shapes of MBC and OMBC indicates that a precipitated
Fig. 4 The FT-IR characterization of the CSBCs and OMBC (a). The XRD spectra of CSBCs, MBC, and OMBC (b)
Sun et al. Biochar (2022) 4:60 Page 9 of 17
layer of magnetite particles had formed. The major peak separation and the saturation magnetization was 14.19
at 35.62° was attributed to the crystalline plane of Fe3O4 emu g−1, indicating outstanding magnetic properties (Yin
(Karunanayake et al. 2017). The other peaks at 30.16° et al. 2021). Furthermore, the OMBC iron loading rate
(220), 43.24° (400), 53.79° (422), 57.15° (511), 62.83° was 16.86%, which is slightly lower than the theoretical
(440), and 74.26° (620) were also observed. Remarkably, value (18.26%).
the peak at 21.5° (002) in OMBC was attributed to S iO2 The thermal gravimetric analyses (TGA) results of
(Alchouron et al. 2020). CSBC, MBC, and OMBC at 30–800℃ are shown in
The crystallite sizes of the F
e3O4 deposited on MBC Fig. 5b. In the initial stage, CSBC, MBC, and OMBC had
and OMBC were analyzed through the Scherrer equation a small amount of weight loss below 150℃ to varying
(Eq. 2) (Navarathna et al. 2020). The magnetite crystallite degrees, which was caused by the evaporation of physi-
sizes were deduced from the peak of D311 for MBC, and cally adsorbed water. During this period, it is noticeable
OMBC, which were 8.65 nm and 8.85 nm, respectively. that the OMBC lost relatively little weight, which may be
a result of its hydrophobic nature. A weight loss occurred
k
Dhkl = (2) in the temperature range of 170–260℃ was mainly due
βcosθhkl to the dihydroxylation and the loss of unreacted oxygen-
where k is the shape factor, with k = 0.89 for cubic crys- containing species on the OMBC (Yang et al. 2021). In
tals, is the XRD analysis wavelength (nm), β is the full the temperature range of 350–550℃, there was great
width at half-maximum value (rad) for XRD lines, and θ weight loss for CSBC (15.85%) and MBC (9.87%), which
is half the diffraction angle 2 θ. was attributed to the pyrolysis of the biochar produced at
350℃. Over the same temperature range, the OMBC dra-
3.1.3 Magnetic moment and thermal analysis
matically lost almost 31% of its weight, which was mainly
The magnetic properties of OMBC were investigated by attributed to the thermal decomposition of the OTMS
vibrating sample magnetometry (VSM). The loading of chain grafted on the MBC surface (Liu et al. 2015). This
magnetic nanoparticles offered OMBC paramagnetic phenomenon exactly indicates that the alkyl chains (–
behavior with no hysteresis while removing the magnetic CH2(CH2)16CH3) of OTMS were efficaciously grown
field (Fig. 5a). The magnetic moment illustrates that the on the surface of the magnetic biochar. The weight loss
saturation magnetization of OMBC was 24.21 emu g−1, at temperatures above 550℃ resulted from the thermal
which was less than value of iron oxide (66.75 emu g −1). decomposition of the biochar substrate, whereby almost
The reason for this phenomenon is the coating effect of 74%, 69%, and 54% of the weight was lost for MBC,
OTMS on Fe3O4 and the low proportion of magnetic CSBC, and OMBC, respectively, at temperatures up to
particles. However, the saturation magnetization of 800℃.
OMBC was sufficiently high to be separated by a magnet. As seen in the CSBC curve, three stages typically
Similarly, Raj et al. had prepared a magnetic adsorbent existed in the process. In the first stage (below 200℃), the
nanocomposite material, in which the magnetization decrease in mass was owning to the evaporation of mois-
was 15.00 emu g −1 (Raj et al. 2015). Yin et al. had pro- ture and light volatiles (Cárdenas-Aguiar et al. 2017). The
duced carbonized Fe3O4@cotton for use in the oil-water next stage was from approximately 200 to 500℃, during
Fig. 5 Magnetic hysteresis loop of magnetic Fe3O4 and OMBC (a). Thermogravimetric analysis curves (10℃ min−1 heating rate under N2
atmosphere) of CSBC (350℃), MBC, and OMBC (b)
Sun et al. Biochar (2022) 4:60 Page 10 of 17
which CSBC had a significant mass loss. As the primary the hydrophobic characteristic that allows the material
process, polymerized organic compounds, such as hemi- to float on the surface of the water. Additional file 1: Fig.
cellulose and cellulose, were volatilized and decomposed S3b shows the situations of CSBC350, MBC, and OMBC
at a fast rate. In the third stage above 500℃, the slow in water after stirring to test the floatability. CSBC350
weight loss was caused by the decomposition of lignin and MBC were mixed into the water phase, and OMBC
and other organic matter with stronger chemical bonds, was still floating on the surface yet. The silane reagent’s
and residual organic matter was further transferred into hydrocarbon chains on the surface of OMBC prevent
biochar gradually (Yuan et al. 2019). Overall, although water from adhering to the surface and entering the pore
the weight loss can be caused by the modifying agent, the network, whereas the OMBC remains buoyant and does
TG curve for the OMBC corresponded to the change in not sink. The excellent hydrophobicity and floatabil-
the weight of the biochar during the pyrolysis process. ity help to guarantee efficient oil sorption on the water
surface. In addition, even when entering the water, since
3.1.4 Wettability measurement hydrocarbons are entrapped inside the pore network,
As depicted in Additional file 1: Fig. S3a, the wettabil- biochar particles also provide a favorable environment
ity of the CSBCs and OMBC was studied by water con- for the colonization of oil-degrading microorganisms
tact angle (WCA). Hydrophilicity is defined as a water (Nguyen et al. 2013).
contact angle less than 90° and vice versa. The WCA
of CSBC350 was 86.23° (< 90°), appearing hydrophilic. 3.2 Oil sorption capability
However, it was significant that the WCA of the CSBCs The oil sorption ability of different biochar was
gradually increased with increasing pyrolysis tempera- researched. We selected three representative oils with
ture (Additional file 1: Fig. S3a). The hydrophilicity and different densities and viscosities to study the adsorption
hydrophobicity of biochar are determined by the organic properties of biochar on typical oils. Figure 6a shows that
groups on the surface (Yuan et al. 2019). However, a there were obvious differences in the oil sorption abil-
higher pyrolysis temperature leads to a decline in polar ity among the biochars produced from different biomass
functional groups on the biochar surface, which enhances materials. At the same pyrolysis temperature (500℃), the
the hydrophobicity of the biochar (Ahmad et al. 2014). In CSBC had the maximum oil sorption ability for crude
detail, the biochar produced at low temperatures had an oil (8.45 g g−1), being almost 2 times that of CCBC and
abundance of hydrophilic groups, such as oxhydryl and SSBC, while RHBC was the most inefficient. Based on
carboxyl groups. In contrast, the biochar produced at the results, the CSBCs produced at different tempera-
higher temperatures was aromatic compound rich with tures were employed to adsorb the crude oil in the pure
high hydrophobicity. In Additional file 1: Fig. S3a, the oil system. CSBC350 had the strongest ability to adsorb
water contact angle of OMBC decorated with OTMS oil, as seen in Fig. 6a. The main reason for this result was
reached 138.00°, which indicates a stronger hydrophobic- that the CSBC350 had a more complete macropore struc-
ity compared with the substrate material (CSBC350). A ture to hold the oil, as mentioned before. The oil sorption
large number of long alkyl chains have been grafted on ability of biochar is closely associated with the pyrolysis
the surface of OMBC and in the crevices, which result in temperature and feedstock material (Lian et al. 2017).
Fig. 6 The oil capacity of biochar and modified biochar. The crude oil sorption capacity of biochar that pyrolyzed from various feedstock materials
prepared at 500℃ and the corn stock biochar that was pyrolyzed at different temperatures (a). The sorption capacity of different sizes of CSBC350
for three oils. (Small: 0.55–0.25 mm, Medium: 0.25–0.15 mm, and Large: < 0.15 mm) (b). The sorption capacities of CSBC (350℃), MBC, and OMBC for
three types of oil in the oil–water system (c)
Sun et al. Biochar (2022) 4:60 Page 11 of 17
Generally, feedstock materials, such as agricultural stalks oil molecules on the surface of the material. In this case,
that are composed of cellulose and hemicelluloses, only the sorption effect is greater than the absorption effect.
need low preparation temperatures to form a fully porous Therefore, increasing the viscosity of the oil can increase
structure. Conversely, wood-like materials, contain- the oil absorption of the biochar (Ceylan et al. 2009; Zhu
ing a high proportion of lignin and celluloses, are often et al. 2011). The viscosity of crude oil is relatively high,
pyrolyzed at high temperatures to achieve volatilization and it easily attaches and accumulates inside the biochar.
and form pores (Yaashikaa et al. 2019). In addition, the Engine oil and diesel oil have lower densities and viscosi-
biochar particle size is a significant and neglected factor ties and, of course, lower storage ability in biochar.
in the sorption of organic pollutants, so the correlation
between removal efficiency and biochar particle size has 3.3 Sorption kinetics and sorption isotherm experiments
aroused controversy (Jin et al. 2022). The CSBC350 was As seen in Fig. 7, the sorption of crude oil by CSBC and
sieved into three different sizes: large (0.55–0.25 mm), OMBC reached equilibrium at 60 min, while the sorp-
medium (0.25–0.15 mm), and small (< 0.15 mm). Dif- tion equilibrium of diesel oil and engine oil was com-
ferent types of oil were used to measure the oil sorption pleted within 10 min. During the sorption of crude oil,
ability of the three biochar sizes in the pure oil environ- the sorption was rapid at the beginning, and after 30 min
ment. As shown in Fig. 6b, the oil sorption ability of bio- the sorption was sluggish. The reason for this process is
char in all sizes was the greatest for crude oil, with the that initially the surface of biochar contained many blank
results all being greater than over 10 g g−1. The medium- active sites, which contribute to the fast progress of sorp-
sized biochar has the best sorption to crude oil and diesel tion. With continuous sorption, the sites were occupied
oil. The small-sized biochar had a higher ability to adsorb by crude oil, and the crude oil had a high viscosity, hence,
engine oil. Overall, the medium-sized biochar had a bet- the sorption gradually slowed down. The increased sorp-
ter sorption effect on the three kinds of oil. The differ- tion after 30 min was probably attributed to the breakage
ences in the sorption of the three oils by different sizes of crude oil droplets, which reduces the diameter of the
were mainly due to the pore size. The undersized biochar crude oil droplets causing more interfacial pore area for
broke the initial macroporous structure, which made it the sorption to occur (Stang et al. 1994). However, in the
inefficient for oil sorption (Jin et al. 2022). presence of low viscosity oils, such as diesel and engine
The experimental results of CSBC350, MBC, and oil, the sorption was relatively fast. The pseudo-first-order
OMBC oil sorption in the oil-water system are shown in and pseudo-second-order models were used to simulate
Fig. 6c. It was obvious that the oil sorption ability of MBC oil sorption onto CSBC and OMBC, as shown in Fig. 7,
presented an obvious downward trend, due to the hydro- and the results and detailed parameters are summarized
philicity of the nano iron oxide and magnetite deposition in Additional file 1: Table S1. From the fitting results, the
enhancing the adsorbent weight. The MBC decorated by sorption kinetics of the CSBC and OMBC both coincided
OTMS had an oil sorption ability that was almost equal with the pseudo-second-order model, and the correlation
to that of CSBC350, reaching 8.54 g crude oil g −1, 5.03 g coefficients (R2) were high (> 0.99). When comparing the
engine oil g−1, and 4.30 g diesel oil g−1. Although the oil theoretical equilibrium sorption capacity produced from
sorption capacity of CSBC350 was slightly greater than the two kinetic models, the theoretical capacities fitted
that of OMBC, the capacity sharply declined when the from the pseudo-second-order model were closer to the
CSBC faced the turbulence of the water wave, which can experimental values. These fitting results illustrate that
provide a greater probability of contact with water. How- the sorption process was a chemical sorption process,
ever, OMBC can selectively adsorb oil on the water sur- not just simple physical sorption.
face in a floating posture due to its hydrophobic surface. The sorption isotherms of CSBC and OMBC are
Regarding oil types, crude oil always has the highest level shown in Fig. 8. The sorption isotherms are significant
of sorption capacity followed by engine oil and diesel oil, for describing the interaction between adsorbate and
regardless of the size of the original biochar or modifi- adsorbent (Sidik et al. 2012). Nonlinear Langmuir and
cation. In general, the difference in oil sorption capac- Freundlich isotherm models were employed to analyze
ity is related to the density and viscosity of the oil. The the experimental isotherm data. The detailed contents
greater the density of the oil is, the greater the quality of are listed in Additional file 1: Table S2. The Langmuir
the oil stored in the material space and the greater the isotherm model was most suitable for the sorption of
calculated oil sorption capacity. The viscosity is another three oils, with regression coefficient values (R2) that
important influencing factor. Although a higher viscos- were both greater than 0.95 for CSBC and greater than
ity of oil hinders the sorption of oil molecules into the 0.80 for OMBC, which indicates that the sorption of oils
internal structure of the material, for macroporous mate- was monolayer coverage onto the homogenous distribu-
rials, a higher viscosity is conducive to the sorption of tion of active sites on the CSBC or OMBC surface, where
Sun et al. Biochar (2022) 4:60 Page 12 of 17
Fig. 7 The pseudo-first-order kinetics model of oils sorption on CSBC (350℃) (a) and OMBC (c). The pseudo-second-order kinetics model of oil
sorption on CSBC (350℃) (b) and OMBC (d)
all active sites are energetically equivalent (Gurav et al. easily adsorbed. In addition, the influences of pH and
2021a, b). OMBC’s maximum oil sorption capacity (qm, temperature on crude oil sorption were surveyed. As
9.25 g crude oil g−1, 4.05 g diesel oil g−1, 4.41 g engine Fig. 9c shows, the sorption ability of OMBC in the low
oil g−1) calculated from the Langmuir model was close to pH range was greater than that in the high pH range,
the experimental value (qe, Exp, 8.77 crude oil g g−1, 4.01 and this result may be related to the charge change.
diesel oil g g −1, 4.44 engine oil g g−1). Certainly, the qm When the temperature was gradually increased, the vis-
values of CSBC were also nearly the same as the experi- cosity of crude oil decreased, and the sorption capacity
mental values. increased slightly (Fig. 9d).
The recovery rate of the sorbent is an important
3.4 Availability of sorbent parameter for magnetic oil sorbents. Hence, the recov-
The oil sorption capacity of OMBC in the seawater- ery rate of OMBC after oil sorption was tested, and the
oil system was measured, as shown in Fig. 9a. The results show that the recovery rate was almost stable
sorption ability of OMBC for the three oils showed a at approximately 75% regardless of the oil type (Addi-
slight decrease compared with the performance in the tional file 1: Fig. S4a). Furthermore, considering that
deionized water-oil environment (Fig. 6c). A similar the adsorbent will be used in the water environment,
result was also reported by other researchers (Navar- there may leach supported metals. Hence, after the
athna et al. 2020). Furthermore, the effect of salinity sorption of crude oil by OMBC, the iron contents of
on OMBC sorption of crude oil was investigated. Fig- deionized water and seawater were measured by ICP‒
ure 9b shows that the effect of salinity change on the MS (Additional file 1: Fig. S4b). Compared with the
sorption capacity of crude oil was small, but gener- blank water sample, it can be seen that iron was leached
ally showed a trend of decreasing and then increasing out neither in deionized water or seawater environment
in the range of 30–120‰. The reason for this may be after OMBC adsorbed crude oil.
that the hydrophobicity of the oil is enhanced under the Table 2 summarizes the capacities of oil sorbents
action of static electricity, which makes the oil more reported in the literature. It can be observed that our
Sun et al. Biochar (2022) 4:60 Page 13 of 17
Fig. 8 The Langmuir isotherm model of oil sorption on CSBC (350℃) (a), OMBC (c). The Freundlich isotherm model of oil sorption on CSBC (350℃)
(b), OMBC (d)
sorbent exhibited a great sorption performance, in on the surface and prevent the attachment and entry of
comparison with previous sorbent capacities. water. Moreover, the three-dimensional multiporous
structure of the biochar is one of the essential points
3.5 Sorption mechanism that dominates the oil removal efficacy. Specifically, this
Biochar is an ideal sorption material due to its high poros- three-dimensional structure increases the pore volume
ity, high specific surface area, abundant surface organic and provides more effective sites for oil sorption (Ha
functional groups, and outstanding cation exchange et al. 2016). For instance, higher pyrolysis temperatures
capacity and has extensive application in the remediation can lead to the destruction of the partial pore structure
of soil pollution, water treatment, and carbon sequestra- of rice husk biochar, which results in a further decrease
tion. The differences in the biochar characteristics were in removal (Zhao et al. 2013), similar to CSBC. The SEM
important reasons for the different sorption mechanisms. figure of OMBC shows that the porous structure was
As shown in Fig. 10, the oil sorption mechanisms of bio- unbroken, even though modification occurred. With the
char were mainly dominated by hydrophobic forces, pore sorption of the surface hydrophobic layer, the intact pore
intercepts, π–π electron attraction, hydrogen bonds, and structure associated with the presence of macropores can
electrostatic attraction. As another critical factor affect- hold more oil. Moreover, hydrogen bonds are formed
ing oil sorption capability, the hydrophobicity of the between polar components of the oil (especially resins
adsorbent is determined by the heterogeneity of the mor- and asphaltenes) and the biochar’s oxygen-containing
phology and surface chemistry and can be represented functional groups, including –COOH and –OH (Jiang
by water contact angle (Doshi et al. 2018). Through the et al. 2016). The graphene content of biochar is electron
aforementioned results of WCA analysis, the OMBC rich and interacts with electron-deficient components of
was identified as a hydrophobic material and can specifi- the oil, enhancing the π–π electron effect (Ahmad et al.
cally adsorb oil due to the fatty acid functional groups 2014). In addition, electrostatic interactions may have
Sun et al. Biochar (2022) 4:60 Page 14 of 17
Engine oil
Diesel oil
120 (b)
Salinity (‰)
90
60
30
11 (c)
8
pH
5
2
Temperature (℃)
55 (d)
40
25
10
0 2 4 6 8 10 12
Capacity (g g-1)
Fig. 9 Sorption of oils by OMBC in the seawater-oil system (a). Sorption of OMBC on crude oil under different conditions, salinity (b), pH (c), and
temperature (d)
Coconut oil-modified wood biochar Crude oil Freshwater 25 5.32 Gurav et al. (2021a, b)
Coconut shell biochar-iron oxide composite Used motor oil Deionized water 25 7.65 Raj et al. (2015)
Lauric acid-coated magnetic biochar Crude oil Deionized water 25 6.30 Navarathna et al. (2020)
Rice husks biochar Crude oil Seawater 25 6.00 Angelova et al. (2011)
Rice husk biochar Crude oil Deionized water 25 3.23 Kandanelli et al. (2018)
Wood chip biochar Toluene Deionized water 25 0.18 Silvani et al. (2017)
Synthetic seawater 0.09
Spent coffee grounds biochar Diesel oil Deionized water 25 1.50 Lee et al. (2022)
OMBC Diesel oil Deionized water 25 4.30 This study
Engine oil 25 5.03
Crude oil 25 8.54
Sun et al. Biochar (2022) 4:60 Page 15 of 17
Navarathna CM, Karunanayake AG, Gunatilake SR, Pittman CU, Perez F, Mohan Xiong Z, Huanhuan Z, Jing W, Wei C, Yingquan C, Gao X, Haiping Y, Hanping
D, Mlsna T (2019) Removal of Arsenic (III) from water using magnetite pre- C (2021) Physicochemical and adsorption properties of biochar from
cipitated onto Douglas fir biochar. J Environ Manage 250:109429. https:// biomass-based pyrolytic polygeneration: effects of biomass spe-
doi.org/10.1016/j.jenvman.2019.109429 cies and temperature. Biochar 3(4):657–670. https://doi.org/10.1007/
Navarathna CM, Bombuwala Dewage N, Keeton C, Pennisson J, Henderson s42773-021-00102-5
R, Lashley B, Zhang X, Hassan EB, Perez F, Mohan D, Pittman CU, Mlsna Xu R, Xiao S, Yuan J, Zhao A (2011) Adsorption of methyl violet from aqueous
T (2020) Biochar adsorbents with enhanced hydrophobicity for oil spill solutions by the biochars derived from crop residues. Bioresource Tech-
removal. Acs Appl Mater Inter 12(8):9248–9260. https://doi.org/10.1021/ nol 102(22):10293–10298. https://doi.org/10.1016/j.biortech.2011.08.089
acsami.9b20924 Yaashikaa PR, Senthil Kumar P, Varjani SJ, Saravanan A (2019) Advances in
Nguyen HN, Pignatello JJ (2013) Laboratory tests of biochars as absorbents for production and application of biochar from lignocellulosic feedstocks
use in recovery or containment of marine crude oil spills. Environ Eng Sci for remediation of environmental pollutants. Bioresource Technol
30(7):374–380. https://doi.org/10.1089/ees.2012.0411 292:122030. https://doi.org/10.1016/j.biortech.2019.122030
Qian L, Zhang W, Yan J, Han L, Gao W, Liu R, Chen M (2016) Effective removal of Yang Y, Chen X, Li Y, Yin Z, Bao M (2021) Construction of a superhydrophobic
heavy metal by biochar colloids under different pyrolysis temperatures. sodium alginate aerogel for efficient oil absorption and emulsion separa-
Bioresource Technol 206:217–224. https://doi.org/10.1016/j.biortech. tion. Langmuir 37(2):882–893. https://doi.org/10.1021/acs.langmuir.0c032
2016.01.065 29
Raj KG, Joy PA (2015) Coconut shell based activated carbon–iron oxide mag- Yim UH, Ha SY, An JG, Won JH, Han GM, Hong SH, Kim M, Jung J, Shim WJ
netic nanocomposite for fast and efficient removal of oil spills. J Environ (2011) Fingerprint and weathering characteristics of stranded oils after
Chem Eng 3(3):2068–2075. https://doi.org/10.1016/j.jece.2015.04.028 the Hebei Spirit oil spill. J Hazard Mater 197:60–69. https://doi.org/10.
Regmi P, Garcia Moscoso JL, Kumar S, Cao X, Mao J, Schafran G (2012) Removal 1016/j.jhazmat.2011.09.055
of copper and cadmium from aqueous solution using switchgrass Yin Z, Pan Y, Bao M, Li Y (2021) Superhydrophobic magnetic cotton fabricated
biochar produced via hydrothermal carbonization process. J Environ under low carbonization temperature for effective oil/water separation.
Manage 109:61–69. https://doi.org/10.1016/j.jenvman.2012.04.047 Sep Purif Technol 266:118535. https://doi.org/10.1016/j.seppur.2021.
Shi K, Qiu Y, Li B, Stenstromb MK (2016) Effectiveness and potential of straw- 118535
and wood-based biochars for adsorption of imidazolium-type ionic Yuan P, Wang J, Pan Y, Shen B, Wu C (2019) Review of biochar for the manage-
liquids. Ecotox Environ Safe 130:155–162. https://doi.org/10.1016/j. ment of contaminated soil: preparation, application and prospect. Sci
ecoenv.2016.04.017 Total Environ 659:473–490. https://doi.org/10.1016/j.scitotenv.2018.12.
Sidik SM, Jalil AA, Triwahyono S, Adam SH, Satar MAH, Hameed BH (2012) 400
Modified oil palm leaves adsorbent with enhanced hydrophobicity for Zhang P, Sun H, Min L, Ren C (2018) Biochars change the sorption and
crude oil removal. Chem Eng J 203:9–18. https://doi.org/10.1016/j.cej. degradation of thiacloprid in soil: insights into chemical and biologi-
2012.06.132 cal mechanisms. Environ Pollut 236:158–167. https://doi.org/10.1016/j.
Silvani L, Vrchotova B, Kastanek P, Demnerova K, Pettiti I, Papini MP (2017) Char- envpol.2018.01.030
acterizing biochar as alternative sorbent for oil spill remediation. Sci Rep. Zhao MY, Cui YB, Li X, Liu YH, Li HX (2013) Study on purification of oil-polluted
https://doi.org/10.1038/srep43912 sea water by rice hull adsorbent. Appl Mech Mater 295–298:1245–1248.
Stang M, Karbstein H, Schubert H (1994) Adsorption kinetics of emulsifiers at https://doi.org/10.4028/www.scientific.net/AMM.295-298.1245
oil-water interfaces and their effect on mechanical emulsification. Chem Zhou S, You T, Zhang X, Xu F (2018) Superhydrophobic cellulose nanofiber-
Eng Process 33:307–311. https://doi.org/10.1016/0255-2701(94)02000-0 assembled aerogels for highly efficient water-in-oil emulsions separation.
Sun K, Ro K, Guo M, Novak J, Mashayekhi H, Xing B (2011) Sorption of bisphe- ACS Appl Nano Mater 1(5):2095–2103. https://doi.org/10.1021/acsanm.
nol A, 17α-ethinyl estradiol and phenanthrene on thermally and hydro- 8b00079
thermally produced biochars. Bioresource Technol 102(10):5757–5763. Zhu H, Qiu S, Jiang W, Wu D, Zhang C (2011) Evaluation of Electrospun Polyvi-
https://doi.org/10.1016/j.biortech.2011.03.038 nyl Chloride/Polystyrene Fibers As Sorbent Materials for Oil Spill Cleanup.
Tapia JI, Alvarado-Gómez E, Encinas A (2019) Non-expensive hydrophobic Environ Sci Technol 45(10):4527–4531. https://doi.org/10.1021/es2002343
and magnetic melamine sponges for the removal of hydrocarbons and
oils from water. Sep Purif Technol 222:221–229. https://doi.org/10.1016/j.
seppur.2019.04.008
Tran H, Lin C, Bui X, Ngo H, Cheruiyot NK, Hoang H, Vu C (2021) Aerobic
composting remediation of petroleum hydrocarbon-contaminated soil.
Current and future perspectives. Sci Total Environ 753:142250. https://doi.
org/10.1016/j.scitotenv.2020.142250
Veldkornet D, Rajkaran A, Paul S, Naidoo G (2020) Oil induces chlorophyll
deficient propagules in mangroves. Mar Pollut Bull 150:110667. https://
doi.org/10.1016/j.marpolbul.2019.110667
Villegas MA, Navarro JMF (1988) Characterization of B2O3-SiO2 glasses pre-
pared via sol-gel. J Mater Sci 23:2464–2478
Wang F, Harindintwali JD, Yuan Z, Wang M, Wang F, Li S, Yin Z, Huang L, Fu Y,
Li L, Chang SX, Zhang L, Rinklebe J, Yuan Z, Zhu Q, Xiang L, Tsang D, Xu
L, Jiang X, Liu J, Wei N, Kastner M, Zou Y, Ok YS, Shen J, Peng D, Zhang
W, Barcelo D, Zhou Y, Bai Z, Li B, Zhang B, Wei K, Cao H, Tan Z, Zhao LB,
He X, Zheng J, Bolan N, Liu X, Huang C, Dietmann S, Luo M, Sun N, Gong
J, Gong Y, Brahushi F, Zhang T, Xiao C, Li X, Chen W, Jiao N, Lehmann J,
Zhu YG, Jin H, Schaffer A, Tiedje JM, Chen JM (2021) Technologies and
perspectives for achieving carbon neutrality. Innovation (n Y) 2(4):100180.
https://doi.org/10.1016/j.xinn.2021.100180
Wathukarage A, Herath I, Iqbal MCM, Vithanage M (2019) Mechanistic under-
standing of crystal violet dye sorption by woody biochar: implications for
wastewater treatment. Environ Geochem Hlth 41(4):1647–1661. https://
doi.org/10.1007/s10653-017-0013-8
Xie J, Lin R, Liang Z, Zhao Z, Yang C, Cui F (2021) Effect of cations on the
enhanced adsorption of cationic dye in Fe3O4-loaded biochar and
mechanism. J Environ Chem Eng 9(4):105744. https://doi.org/10.1016/j.
jece.2021.105744