Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Biochemical Engineering Journal 130 (2018) 104–112

Contents lists available at ScienceDirect

Biochemical Engineering Journal


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

Regular article

Effects of pH, dissolved humic acid and Cu2+ on the adsorption of


norfloxacin on montmorillonite-biochar composite derived from
wheat straw
Jinghuan Zhang ∗ , Mingyi Lu, Jun Wan, Yuhuan Sun, Huixia Lan, Xiaoyan Deng
College of Environment and Safety Engineering, Qingdao University of Science and Technology, Qingdao 266042, China

a r t i c l e i n f o a b s t r a c t

Article history: A montmorillonite-biochar (MT-BC) composite was prepared from wheat straw heating at 400 ◦ C. Adsorp-
Received 10 August 2017 tion isotherms and effects of pH, dissolved humic acid (DHA) and Cu2+ on norfloxacin (NOR) sorption
Received in revised form to biochar and MT-BC composite were carried out. Montmorillonite modified biochar showed higher
25 November 2017
adsorption affinity for NOR. The qmax value increased from 10.58 mg g−1 to 25.53 mg g−1 after surface
Accepted 27 November 2017
Available online 2 December 2017
modified by montmorillonite. MT-BC composite had rich pore structure, with specific surface area (SSA)
and pore volume values of 112.6 m2 g−1 and 0.604 cm3 g−1 , respectively. The optimum pH range for NOR
adsorption on MT-BC composite is 5–11. The presence of DHA (20–150 mg L−1 ) greatly suppressed the
Keywords:
Montmorillonite-biochar composite NOR adsorption. The decrease of NOR sorption is attributed to the “competitive adsorption” of DHA with
Adsorption mechanism NOR and “pore blockage” by sorbed DHA. Similarly, reduced NOR sorption with Cu2+ was probably due
Norfloxacin to the competition for sorption sites on biochar surface. Compared to DHA, Cu2+ resulted to a smaller
Dissolved humic acid decrease in NOR sorption. This is because the formation of NOR-Cu2+ complexes can slightly increase
Cu2+ NOR sorption capacity. The dominant adsorption mechanisms of NOR on MT-BC composite are electro-
static interaction, H-bond and pore-filling. This study provided valuable guidance and effective method
for the removal of NOR from aquatic environments.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction A prior study showed that MgO-impregnated biochar had


maximum adsorption capabilities of 398 mg g−1 for phosphate,
Norfloxacin (NOR), one of the fluoroquinolone antibiotics, was 22 mg g−1 for ammonium, and 247 mg g−1 for humate, respectively.
widely used in medical and veterinary practice to treat several Struvite crystallization, electrostatic attraction, and ␲-␲ interac-
infectious diseases in both human and animals [1]. About 30–90% tions contributed to the adsorption enhancement [10]. It is reported
of the antibiotics, acted as active compounds, are stable and can that coated minerals covered on the biochar surface modified
not be easily degraded, thus they can stay in the environment for a the pore diameter and surface functional groups, thus improv-
long time [2]. High concentrations of antibiotics in dissolved phase ing the adsorption of propiconazole on biochars [11,12]. Li et al.
had been detected in surface water [3]. Antibiotics were found to be also reported that attapulgite-biochar derived from potato stem
toxic to aquatic organisms such as algae, fish, and daphnia, contam- had great adsorption affinity for NOR, with the maximum qe of
inate drinking water and threaten human health [4]. Thus, removal 5.24 mg g−1 , which is about 1.68 times higher than primary biochar
of antibiotics from the water environment is necessary. [13]. Previous results showed that biochar coated with montmo-
Biochars can be used to remove antibiotics from aquatic rillonite was a highly effective adsorbent for methylene blue and
environment via adsorption technology. Biochars derived from dif- ammonium [14,15]. It is interesting and essential to explore the
ferent sources are reported to be strong adsorbents for organic adsorption affinity of clay-biochar composite for NOR and the influ-
pollutants in potable and reuse water [5–7]. Surface modified encing environmental factors. In our study, montmorillonite was
biochar composite with high surface area exhibited stronger used to modify the surface of wheat straw- derived biochar.
adsorption than original biochar for organic pollutants [8,9]. Dissolved organic matter (DOM) and heavy metal are widely
existing compounds in aquatic environment, which may interact
with water, solid matrix and antibiotics. Prior studies showed that
high pH and the presence of DOM (humic acid and sodium algi-
∗ Corresponding author.
nate) and Cu2+ greatly inhibited the sorption of sulfonamides onto
E-mail address: vickyhuan@163.com (J. Zhang).

https://doi.org/10.1016/j.bej.2017.11.018
1369-703X/© 2017 Elsevier B.V. All rights reserved.
J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112 105

Table 1
Selected physiochemical properties of NOR.

Chemical MW (g mol−1 ) Sw (mg L−1 ) logKow Tm (◦ C) Chemical structure

NOR 319.3 161000 (pH = 5) −1.7 (pH = 5) 220


400 (pH = 7) −1.0 (pH = 7)
910 (pH = 9) −1.63(pH = 9)

Notes: MW: molecular weight; Tm : melting point; Kow : octanol–water partition coefficient; Sw : aqueous solubility.

pine wood char and graphene oxides [16,17]. This can be explained using an ASAP 2010 system and interpreted with the Brunauer-
by the competition with antibiotic molecules for surface sorption Emmet-Teller (BET) model. The FTIR spectra were recorded in the
sites. Lian et al. reported that high levels of humic acid suppressed transmission mode by a NEXUS 670X spectrophotometer using
the adsorption of sulfonamides by crop straw-derived biochars, but KBr pellet. The 13 C NMR spectrum was acquired on a Bruker AU-
low levels of humic acid promoted the adsorption [18]. Moreover, 300 spectrometer using cross-polarization magic angle spinning
introduction of humic acid was able to decrease the adsorption (CP/MAS) techniques.
of NOR probably due to blockage of sorption sites by DOM [19].
Therefore, this is strongly necessary to investigate the impacts of 2.3. Sorption experiments
pH, coexisted dissolved humic acid and divalent heavy metal cation
Adsorption isotherms of NOR on BC, MT-BC composite and
Cu2+ on the adsorption of NOR by clay-biochar composite in aquatic
montmorillonite were conducted in replicates in 50 mL glass tubes
environment.
with Teflon-lined caps. 0.01 mol L−1 CaCl2 and 200 mg L−1 NaN3
The aims of this study are to (1) prepare montmorillonite-
were added to stock solutions to minimize biological activity [23].
biochar (MT-BC) composite from wheat straw heating at 400 ◦ C,
The ratios of water to solids (40 mL: 50 mg) were adjusted to
(2) determine the adsorption isotherms and the effects of pH, dis-
achieve 30–80% sorption of NOR. Our preliminary adsorption kinet-
solved humic acid (DHA) and Cu2+ on the adsorption of NOR on BC
ics experiments showed that apparent sorption equilibrium was
and MT-BC composite, and (3) explore the adsorption mechanisms
reached in less than 48 h (Fig. S1). The vials with adsorbents and
of NOR on MT-BC composite.
initial NOR solutions (0.4–15 mg L−1 ) were shaken at 150 rpm at
25 ± 0.5 ◦ C for 48 d and then centrifuged at 4500 rpm for 30 min
2. Materials and methods to separate solid and aqueous phases. NOR in the supernatant was
filtered through 0.45 ␮m nylon membrane and analyzed by high
2.1. Chemicals performance liquid chromatography (HPLC). The sorbed amounts
were computed from the difference of the initial and final solute
NOR (>98%; Aldrich Chemical Co., USA) was selected as flu- concentrations. Control reactors prepared similarly but with no sor-
oroquinolone antibiotics probe to determine the equilibrium bent were run simultaneously for assessing loss of solute to reactor
adsorption isotherms and impacts of pH, DHA and Cu2+ on sorption. during sorption. Results showed that average system losses were
DHA (>98%) was purchased from Aldrich Chemical Company in USA consistently less than 4% of the initial concentration, indicating that
and directly used in sorption experiments. The molecular weight of microbial degradation and volatilization during sorption and the
DHA was 1500–3000 g mol−1 . The C, H, N, O and ash contents of DHA uptake by the glass walls were negligible.
were 49.7%, 6.48%, 6.03%, 28.5% and 9.10%, respectively. Selected The effect of pH on the adsorption of NOR was determined by
physiochemical properties of NOR are given in Table 1 [20]. Stock varying the pH from 3.0 to 11.0 while containing 50 mg of adsorbent
solutions of NOR were prepared in HPLC-grade methanol. Methanol and the initial NOR concentration at 10 mg L−1 under 25 ◦ C. The ini-
concentrations in the aqueous solutions were always less than 0.2%, tial pH values of NOR solutions were adjusted using 0.1 mol L−1 HCl
a level at which methanol has no measurable effect on sorption [21]. or NaOH solutions. The impacts of DHA and Cu2+ on the adsorption
of NOR on BC and MT-BC composite were carried out at pH = 7.0
with adding concentrations of 20–150 mg L−1 . Sorption isotherms
2.2. Preparation and characterization of MT-BC composite from
of DHA on BC and MT-BC composite were also analyzed at 25 ◦ C.
wheat straw
Two replicates for each point were used in all sorption experiments.
Montmorillonite-biochar composites were prepared from 2.4. Analytical methods
wheat straw as described in a prior study [22]. First, a stable
montmorillonite suspension was prepared by adding 2.5 g of mont- Norfloxacin concentrations were analyzed by reversed-phase
morillonite powder to 100 mL deionized water and mixed in the HPLC, using a Waters 484 HPLC (Waters, America) with a UV detec-
ultrasonicator for 30 min. Then, 10 g of the wheat straw were tor [19]. A Diamon-sil 5U C18 column (4.6 mm × 250 mm, Dikma
dipped into the montmorillonite suspensions and stirred for 2 h. technologies) was used. Isocratic elution was performed at a flow
Finally, the wheat straw were separated from the mixture and oven rate of 1.0 mL min−1 using a mixed solution of 0.025 mol L−1 phos-
dried at 80 ◦ C. The mixture of montmorillonite and wheat straw phoric acid-acetonitrile (80:20, V/V) as the mobile phase. The UV
were heated in a muffle furnace at 400 ◦ C under a flow of N2 and detection wavelength was set at 278 nm. The retention time for
held there for 6 h. Then the biochars were washed with deionized NOR was 3.8 min. The concentration of humic acid was quantified
water, freeze-dried, passed though a 100-mesh sieve, and used for by UV absorption at 254 nm as described in the reference of Lin
sorption experiments. Wheat straw without montmorillonite was et al. [24].
observed through the same pyrolysis process. The biochars with-
out and with montmorillonite modification were denoted as BC and 2.5. Data analysis
MT-BC composite, respectively.
The elemental compositions (C, H, N, O) were analyzed with Both Langmuir and Freundlich models were used to quantify
a high-temperature combustion method (elementar Vario EL, the adsorption of NOR on BC and MT-BC composite. The Lang-
Germany). SSA was determined by N2 adsorption at −196 ◦ C muir model indicates a monolayer sorption of a target compound
106 J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112

Fig. 1. SEM images of BC (a) and MT-BC (b).

Table 2
Elemental composition and SSA of BC and MT-BC composite.

Sample elemental composition (%) SSA (m2 g−1 ) Pore volume Dp Clay content(%)
a a a a 3 −1 b
C H N O H/C O/C C/N (N + O)/C (cm g ) (nm)

BC 52.4 10.3 2.03 25.2 2.36 0.36 30.1 0.39 20.1 0.138 9.85 –
MT-BC 58.8 1.16 1.85 27.0 0.24 0.34 37.1 0.37 112.6 0.604 21.7 8.91
MT – – – – – – – – 148.3 0.202 48.2 –
a
Atomic ratio.
b
Dp is average pore diameter calculated from N2 adsorption isotherm by the Berret-Joyner-Halenda method.

on a homogenous surface, but the Freundlich model illustrates a ratio indicates a higher aromatic structure of MT-BC composite. The
multi-layer adsorption process [25]. The fitting curves and statisti- BC and MT-BC composite have polarity index (N + O)/C of 0.40 and
cal analysis were processed using SigmaPlot 12.5. 0.37. The SSA and pore volume values of BC were 20.1 m2 g−1 and
0.138 cm3 g−1 , respectively, which was lower than that of bamboo
Langmuirmodel :qe = (qmax ·K L ·C e )/(1 + K L ·C e ) (1)
char [27], but slightly higher than that of zein-derived biochar [28].
where Ce is the liquid phase concentration in mg L−1 , qe is the solid Surface treated by montmorillonite greatly increased the SSA and
phase concentration in mg kg−1 , pore volume values to 112.6 m2 g−1 and 0.604 cm3 g−1 , respectively.
KL is the Langmuir equilibrium constant (L mol−1 ), qmax is the The FTIR spectra of BC and MT-BC composite are presented
Langmuir maximum adsorption capacity in mg kg−1 . in Fig. 2. As shown in Fig. 2, the broad band observed at about
3400 cm−1 is assigned to the stretching vibration of hydroxyl
Freundlichmodel :qe = K f ·C e n (2) groups ( OH). Peaks at 2925 cm−1 and 2850 cm−1 are ascribed to
where Kf is the Freundlich sorption capacity-related parameter aliphatic −CH2 asymmetric and symmetric stretching and C H
[(mg kg−1 )/(mg L−1 ) n )], n is the isotherm nonlinearity index. band, respectively [29]. These peaks in MT-BC composite became
However, a precise comparison cannot be made according to the a decrease than those in the FTIR spectrum of BC. The sharp and
Kf values because of their different units as a result of nonlinear- strong band at 1640 cm−1 can be assigned to aromatic C C stretch-
ity. The concentration dependent sorption capacity coefficient Kd ing [30]. The sharp peak at 660 cm−1 is associated with mineral
(L g−1 ) at three selected concentration (Ce = 0.5, 1.5, and 5.0 mg L−1 ) compounds such as carbonates and phosphates. The absorbance
were employed to compare the sorption capacity. The Kd value is at 1190 cm−1 for MT-BC composite is the C H stretching and
calculated using the formula: OH deformation of COOH or C O stretching of aryl esters, which
was considered larger than that of BC. Introduction of montmoril-
K d = K f ·C e n−1 (3) lonite increased the surface oxygen-containing functional groups
of biochar.
In general, the Kd value decreased as a function of Ce because of
isotherm nonlinearity.
3.2. Effect of pH on the adsorption of NOR on BC and MT-BC
3. Results and discussion composite

3.1. Characteristics of BC and MT-BC composite NOR has two proton-binding sites (carboxyl and piperazinyl
group) with its acid dissociation constants pKa values of 6.22 and
The SEM images in Fig. 1 shows the different surface morphol- 8.51, respectively. Fig. 3 shows the molecular structure of NOR and
ogy of BC (a) and MT-BC composite (b). The BC sample had porous its ionic forms as a function of pH and pKa values [31]. NOR can
structure (Fig. 1a). Fig. 1b indicates that montmorillonite particles exist in cationic form, zwitterionic/neutral form, or anionic form
were deposited on the biochar surface. depending on the solution pH. The effects of pH on the adsorption
In Table 2, the C and O content of BC is 52.4% and 25.2%, which of NOR to BC, MT-BC composite and montmorillonite are given in
is close to the C content of rice straw-derived biochar, but lower Fig. 4. Montmorillonite modified biochar exhibited greater adsorp-
than that of pine wood-derived biochar [26]. The MT-BC composite tion affinity than original biochar for NOR at different pH (3–11).
had higher C and O content of 58.8% and 27.0% than original BC. This can be explained by the increase of SSA and pore volume shown
The H/C and O/C atomic ratios are 2.36 and 0.36 for BC, and 0.24 in Table 2. Modification by montmorillonite may lead to the forma-
and 0.34 for MT-BC composite, respectively. The lower H/C atomic tion of new pores and plenty of reactive sites on biochar surface.
J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112 107

Fig. 2. FTIR spectra analysis of BC (a) and MT-BC (b).

Fig. 3. Molecular structure of NOR and its ionic forms as a function of pH and pKa values.

density on a surface is zero. The pH of zero point charges (pHpzc) of


BC and MT-BC composite was determined by titration. The pHpzc of
BC and MT-BC composite is 8.86 and 6.92, respectively. The surface
of biochar samples was positively charged at pH lower than pHpzc
and negatively charged at pH higher than pHpzc. At pH<5.0, the
adsorption was depressed, because the NOR molecules and MT-BC
had the same charge and can repel each other. At pH>5.0, the NOR
molecules and MT-BC had opposite charges, which results in the
enhancement of adsorption. Thus, electrostatic attraction between
charged NOR and biochar surface was expected to be essential in
the adsorption process.
At pH = 6.22–8.51, NOR existed in zwitterioinic/neutral form.
The zwitterioinic/neutral form was much more hydrophobic than
the anionic or cationic form according to the Sw and logKow of NOR
presented in Table 1. The rich aromatic structure of biochar sam-
ples provided more hydrophobic sites, which can interact with NOR
molecules by hydrophobic effect. This suggested that hydrophobic
Fig. 4. Effects of pH on the adsorption of NOR on BC, MT-BC composite and mont- effect was an important factor determining the increase of NOR
morillonite. adsorption. Yang et al. also reported that the adsorption of NOR to
porous resins and carbon nanotubes enhanced with the increase of
The qe for MT-BC composite enhanced 2.29–2.60 times higher than zwitterionic form, with the highest sorption amount at pH = 7 [20].
BC with the initial NOR concentration of 10 mg L−1 . Compared to Our results showed that the optimum pH range for NOR adsorption
BC and MT-BC composite, montmorillonite had lower adsorption on MT-BC composite is 5–11. However, Wang et al. reported that
capacity for NOR with qe of 0.244–2.553 mg g−1 . the effect of pH on the adsorption of NOR on activated magnetic
As shown in Fig. 4, the maximum NOR sorption capacity on BC biochar derived from corn stalks, reed stalks and willow branches
was observed at pH = 7.0. The NOR adsorption on MT-BC composite was not obvious [1]. Good adsorption affinity of NOR was observed
enhanced with pH ranging from 3.0 to 5.0. When the pH increased in a wide pH range of 2.0–10.0. It was also reported that more than
to 5.0, the adsorption stayed high and steady with the maximum 80.1% of NOR was removed by a clay-biochar composite from aque-
amount at pH = 6.5. This is closely associated with the pH-regulated ous solution at 2.0–11.0, with the maximum amount at pH = 3.0
distribution of the deprotonated species of NOR. The point of zero [13].
charge (pzc) describes the condition when the electrical charge
108 J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112

Fig. 5. Effects of DHA on the adsorption of NOR on BC (a) and MT-BC composite (b).

3.3. Effects of DHA on the adsorption of NOR on BC and MT-BC


composites

Sorption isotherms and fitting parameters of NOR on BC and MT-


BC composite with and without DHA are given in Fig. 5 and Table 3.
Data points and fitting curves in Fig. 5 shows that sorption of NOR
was fitted better by the Langmuir model than Freundlich model.
The Langmuir model fitting results revealed that the monolayer
sorption was the main adsorption mechanism of NOR on BC and
MT-BC.
Montmorillonite modified biochar greatly increased the sorp-
tion of NOR with and without DHA. The Kd values of
NOR (0.5 mg L−1 < Ce < 5.0 mg L−1 ) on MT-BC composite were of
0.578–11.42 L g−1 , about 3.4–6.4 times higher than BC. More-
over, the qmax value for MT-BC composite with DHA enhanced
2.3–4.5 times higher than BC. As shown in Fig. 5 and Table 3,
the presence of DHA reduced the sorption of NOR on BC and
MT-BC composite. The qmax value reduced from 10.58 mg g−1 to Fig. 6. Sorption isotherms of DHA on BC and MT-BC composite.
1.586 mg g−1 for BC, and from 25.53 mg g−1 to 6.968 mg g−1 for MT-
BC composite, with DHA concentration increasing from 0 mg L−1
on BC and MT-BC composite depends on the extent and competi-
to 150 mg L−1 , respectively. Furthermore, when the concentration
tion of each interaction. To testify the hypothesis, the adsorption of
of DHA was 150 mg L−1 , the Kd value (at Ce = 0.5 mg L−1 ) decreased
DHA on BC and MT-BC composite was carried out. Notably, BC and
from 2.833 L g−1 to 0.213 L g−1 for BC, and from 11.42 L g−1 to
MT-BC composite showed high adsorption for DHA in Fig. 6. The
1.497 L g−1 for MT-BC composite, respectively. Moreover, negative
sorbed amounts of DHA are 5.822 mg g−1 and 9.445 mg g−1 on BC
correlations were found between qmax and Kd values of NOR and
and MT-BC composite, respectively, when the initial concentration
DHA concentration (R2 ≥ 0.936) (Fig. 8). Reduced NOR sorption with
of DHA is 200 mg L−1 . Therefore, the reduced NOR adsorption was
DHA was probably due to the fact that coexisted DHA may block the
mainly attributed to the competitive adsorption of DHA with NOR
adsorption sites on modified biochar surface. Peng et al. found that
and the pore blockage by DHA.
humic acid (50 mg L−1 ) showed a depressed effect on the adsorp-
tion of NOR onto titanium oxide [19]. Prior studies also showed
that the presence of low weight DOM was found to significantly 3.4. Effects of Cu2+ on the adsorption of NOR on BC and MT-BC
suppress the sorption of sulfonamides on wood-based biochar and composites
graphene oxides [17,32]. However, suspended carbon nanotubes
exhibited enhanced sorption of sulfamethoxazole in the presence Fig. 7 shows the impacts of Cu2+ with adding concentrations
of dissolved and coated humic acid [33]. Citric acid and malic acid of 20–150 mg L−1 on the adsorption of NOR on BC and MT-BC
(0–100 mmol L−1 ) also increased the adsorption of sulfamethoxa- composite. All sorption isotherms of NOR were nonlinear. The n
zole on crop straw biochar more than 5 times, which was attributed values of NOR with and without Cu2+ were of 0.446–0.648 for
to the elevated microporosity of biochar after treated by low weight BC, and of 0.500–0.611 for MT-BC composite, respectively. Com-
DOM [34]. pared to BC, MT-BC composite exhibited higher sorption for NOR
Hypothetically, the impact of DHA on NOR sorption may be with and without Cu2+ . As shown in Fig. 7, the adsorption of NOR
explained by the following mechanisms [35,36]: (1) competitive on BC and MT-BC composite greatly reduced in the presence of
sorption between NOR and DHA for sorptive sites on biochar sur- Cu2+ . With Cu2+ concentration of 150 mg L−1 , the qmax and Kd value
face, resulting in adsorption reduction; (2) coexisted DHA may (at Ce = 1.5 mg L−1 ) decreased from 10.58 mg g−1 and 1.703 L g−1 to
block the pores on biochar surface to decrease NOR adsorption; 1.779 mg g−1 and 0.350 L g−1 for BC, and from 25.53 mg g−1 and
and (3) partition of NOR to sorbed DHA on biochar surface, which 6.594 L g−1 to 8.076 mg g−1 and 1.878 L g−1 for MT-BC composite,
leads to adsorption enhancement. The apparent sorption of NOR respectively. Also, there were negative correlations between Cu2+
concentration with qmax and Kd values of NOR (R2 ≥ 0.968) (Fig. 8).
J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112 109

Table 3
Freundlich and Langmuir sorption parameters of NOR on BC and MT-BC composite in the presence of DHA and Cu2+ .

Coexisting Freundlich model Langmuir model


pollutants (mg/L)

Kd c (L g−1 )

n Kf R2
N a
Ce = 0.5 mg L−1 Ce = 1.5 mg L−1 Ce = 5.0 mg L−1 qmax (mg g−1 ) KL (L mol−1 ) R2 Na

BC
Control 0 0.637 ± 0.004b 2.055 ± 0.005 0.985 22 2.833 1.703 0.975 10.58 ± 0.02 0.194 ± 0.003 0.986 22
DHA 20 0.584 ± 0.005 1.213 ± 0.004 0.983 22 1.581 1.039 0.656 9.326 ± 0.03 0.120 ± 0.002 0.998 22
50 0.605 ± 0.006 0.959 ± 0.004 0.973 22 1.346 0.787 0.437 4.506 ± 0.007 0.219 ± 0.003 0.986 22
100 0.503 ± 0.003 0.744 ± 0.005 0.940 22 1.092 0.594 0.305 2.759 ± 0.004 0.300 ± 0.004 0.975 22
150 0.476 ± 0.004 0.167 ± 0.003 0.943 22 0.213 0.145 0.094 1.586 ± 0.005 0.092 ± 0.002 0.962 22
Cu2+ 20 0.618 ± 0.005 1.538 ± 0.004 0.990 22 2.052 1.299 0.787 9.536 ± 0.008 0.171 ± 0.002 0.992 22
50 0.511 ± 0.003 1.120 ± 0.005 0.990 22 1.472 0.953 0.593 7.586 ± 0.006 0.142 ± 0.001 0.995 22
100 0.446 ± 0.005 0.962 ± 0.004 0.969 22 1.408 0.769 0.397 3.369 ± 0.004 0.347 ± 0.004 0.980 22
150 0.648 ± 0.006 0.433 ± 0.003 0.920 22 0.623 0.350 0.186 1.779 ± 0.005 0.266 ± 0.004 0.980 22
MT-BC composite
Control 0 0.500 ± 0.003 8.076 ± 0.007 0.990 22 11.42 6.594 3.612 25.53 ± 0.11 0.474 ± 0.003 0.991 22
DHA 20 0.448 ± 0.004 4.995 ± 0.006 0.980 22 6.541 4.266 2.671 19.94 ± 0.07 0.252 ± 0.003 0.994 22
50 0.361 ± 0.003 3.674 ± 0.004 0.969 22 4.916 3.099 1.869 15.50 ± 0.04 0.204 ± 0.004 0.985 22
100 0.489 ± 0.005 2.599 ± 0.005 0.980 22 3.620 2.141 1.204 11.45 ± 0.01 0.252 ± 0.004 0.983 22
150 0.443 ± 0.003 1.124 ± 0.004 0.960 22 1.497 0.951 0.578 6.968 ± 0.005 0.158 ± 0.002 0.976 22
Cu2+ 20 0.611 ± 0.005 6.590 ± 0.005 0.970 22 9.662 5.268 2.711 22.64 ± 0.02 0.488 ± 0.005 0.975 22
50 0.580 ± 0.004 6.053 ± 0.003 0.960 22 9.426 4.671 2.164 17.23 ± 0.01 0.772 ± 0.007 0.970 22
100 0.522 ± 0.003 3.625 ± 0.005 0.975 22 5.166 2.947 1.593 13.95 ± 0.009 0.315 ± 0.004 0.987 22
150 0.587 ± 0.005 2.355 ± 0.004 0.955 22 3.465 1.878 0.961 8.076 ± 0.005 0.361 ± 0.005 0.987 22
a
Number of data.
b
Standard deviation.
c
the concentration dependent sorption capacity coefficient Kd = Kf Ce n−1 .

Fig. 7. Effects of Cu2+ on the adsorption of NOR on BC (a) and MT-BC composite (b).

The apparent decrease of NOR sorption with Cu2+ may be due to the In our study, complexes may also be formed in the BC-Cu2+ -NOR
“competitive adsorption” associated with sorbed Cu2+ on modified ternary system. Compared to DHA, the presence of Cu2+ resulted in
biochar surface. a smaller decrease in NOR sorption. With the adding concentra-
A prior study also showed that Cu2+ and DHA inhibited the tion of 100 mg L−1 , the qmax and Kd values (at Ce = 5.0 mg L−1 ) of
adsorption of sulfonamides to pine wood biochars attributed to NOR on MT-BC composite were 13.95 mg g−1 and 1.593 L g−1 for
the competition for adsorption sites on biochar surface [16]. Kong Cu2+ , and 11.45 mg g−1 and 1.204 L g−1 for DHA, respectively. This
et al. reported that the presence of divalent cations (Cu2+ , Zn2+ , can be explained by the formation of NOR-Cu2+ complexes, which
and Mg2+ , 0–20 mmol L−1 ) prominently suppressed the adsorption can slightly increase NOR sorption capacity [39].
of NOR by paddy and red soils [37]. Moreover, incorporation of
Sn4+ weakened the adsorption ability of NOR on Mg-Al layered 3.5. Adsorption mechanisms
double hydroxides due to the decrease of the SSA [32]. However,
Huang et al. (2017) reported that Cu2+ could significantly enhance According to the prior reported results, the adsorption mech-
the adsorption of tetracycline on Fe3 O4 @SiO2 -Chitosan/graphene anisms of antibiotics onto carbon materials mainly involved elec-
oxide (MSCG) nanocomposite [38]. The highest adsorption capac- trostatic interaction, hydrophobic effect, H-bond, ␲-␲ interaction,
ity of tetracycline increased about 2.72 times with Cu2+ . The surface precipitation, pore-filling and partition to non-carbonized
major mechanism is that the Cu2+ acts as a bridge between fractions [40,41]. In our study, the main mechanisms of NOR on
tetracycline and MSCG, and tetracycline-Cu2+ complexes were MT-BC composite may be explained in Fig. 9. FTIR spectra analy-
formed, which significantly improve the adsorption of tetracy- sis indicated that the −OH, C O, C H, −COOH bonds were covered
cline. on the surface of MT-BC composite. After modified by montmoril-
110 J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112

Fig. 8. Correlations of qmax and Kd values (Ce = 1.5 mg L−1 ) of NOR on BC and MT-BC composite with the concentration of Cu2+ and DHA.

Fig. 9. Adsorption mechanisms of NOR onto MT-BC composite with DHA and Cu2+.
J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112 111

Fig. 10. The relationship between the adsorption capacity of MT-BC composite for Fig. 11. The relationship between removal percentage of NOR from wastewater by
NOR and the number of regeneration cycles. MT-BC composite and the number of regeneration cycles.

lonite, the amount of surface oxygen-containing functional groups removed from MT-BC composite by methanol. The removal per-
greatly increased. It was reported that NOR molecules consisted of centage of NOR from MT-BC composite were 86.8%, 88.1%, 87.2%,
several moieties, which can be served as acceptors or donors for 87.9%, and 86.5% with the number of regeneration cycle from one
H bond. These moieties may interact with the functional groups to five, respectively. As a result, MT-BC composite prepared in this
on the surface of carbon materials to form H-bond, and greatly study had the potential to be recycled and reused for the removal
improved the adsorption affinity for NOR [20]. In our study, the of NOR from wastewater.
enhancement of the NOR adsorption was mainly due to the H-band
formation between NOR and oxygen-containing groups in MT-BC 3.7. Application of MT-BC composite in the removal of NOR from
composite. Thus, H-bond is one of the main mechanisms of NOR on wastewater
MT-BC composite as well as electrostatic attraction.
In addition to H-bond, the pore-filling effect also contributed to The application of MT-BC composite was evaluated for the
the adsorption of NOR on MT-BC composite. Prior studies showed removal of NOR from two wastewater samples obtained from Bohai
that pore-filling was the major factor for the great adsorption of Bay in China. It is reported that the concentrations of NOR in
antibiotics such as tetracycline and NOR onto biochars derived from polluted water of Bohai Bay were 0.003–6.8 ␮g L−1 [44]. In our
solid digestate and corn stalks [1,8]. Other carbon materials, such study, NOR concentrations in the two wastewater samples were
as porous resins, carbon nanotubes and animal hairs-based acti- 115 ng L−1 for NO.1 and 367 ng L−1 for NO.2. The relationship
vated carbon were also reported to have strong adsorption capacity between removal percentage of NOR and the number of regen-
for NOR mainly due to micropores [20,42]. Granular activated car- eration cycles is listed in Fig. 11. Fig. 11 reveals that the removal
bon with high surface area and pore volume of 852 m2 g−1 and percentage of NOR from wastewater by MT-BC composite was
0.671 cm3 g−1 , respectively, can improve the removal efficiency of 85.5–87.2% for NO.1 and 88.3–90.0% for NO.2. Therefore, the MT-
NOR and ciprofloxacin from water. It is believed that high SSA and BC composite had good stability and reusability for the removal of
favorable pore characters were the main reasons for the removal NOR from polluted water in Bohai Bay, China.
of antibiotics by adsorption technology [43]. The MT-BC composite
had a rich microporous structure with a large SSA and pore vol- 4. Conclusion
ume, and showed a strong adsorption capacity for NOR. Thus, like
This study reveals that montmorillonite modified biochar had
electrostatic interaction, H-bond and pore-filling effect were the
higher SSA and pore volume than original wheat straw biochar.
main mechanisms for NOR adsorption on MT-BC composite. Each
Compared to biochar, MT-BC composite exhibited greater adsorp-
interaction provides different contributions at different pH value.
tion capacity for NOR. The qmax and Kd values for MT-BC composite
At pH < 5, electrostatic attraction between negatively charged NOR
increased 2.41 times and 3.70–3.96 times, respectively. MT-BC
and biochar surface was the main adsorption mechanism. At pH > 5,
composite had a high adsorption capacity for NOR in a wide pH
H-bond between NOR molecules and oxygen-containing groups,
range of 5–11. The presence of DHA and Cu2+ (20–150 mg L−1 )
and pore-filling were the major reason for the high NOR adsorption.
reduced the adsorption of NOR on BC and MT-BC composite. Com-
petitive adsorption and pore blockage are the major reasons for
3.6. Regeneration of MT-BC composite
the reduction of NOR sorption. At the same concentration, DHA
leads to a higher decrease in NOR sorption than Cu2+ . Electrostatic
Methanol solution was selected as the washing agent for the
interaction, H-bond and pore-filling are the dominant adsorption
removal of NOR from the adsorbent with a three-stage washing
mechanisms of NOR on MT-BC composite surface. Our results
procedure reported in a prior study [25]. H-bond between NOR and
showed that both DHA and Cu2+ can inhibit the adsorption affinity
methanol was stronger, which facilitated the removal of NOR. The
for NOR on BC and MT-BC composite, therefore, effectively making
consecutive sorption processes were conducted for five times to
NOR more mobile in the environment especially for that with high
determine the regeneration and reuse value of MT-BC composite.
input of chars.
The adsorption capacity of MT-BC composite for NOR according to
the number of regeneration cycles is given in Fig. 10. Fig. 10 shows Acknowledgements
that the stability and reusability of MT-BC composite was good.
The adsorption capacity of MT-BC composite for NOR was steady The study was jointly supported by the Natural Science Foun-
along with the time for reuse. Futhermore, most of NOR could be dation of Shandong Province in China (No. ZR2016BB38; No.
112 J. Zhang et al. / Biochemical Engineering Journal 130 (2018) 104–112

ZR2013DQ002) and the Project of QingDao Minsheng Science and [21] R.D. Wauchope, W.C. Koskinen, Adsorption-desorption equilibria of
Technology (No. 14-2-3-70-nsh). herbicides in soil: a thermodynamic perspective, Weed Sci. 31 (1983)
504–512.
[22] Y. Yao, B. Gao, J. Fang, M. Zhang, H. Chen, Y. Zhou, A.E. Creamer, Y. Sun, L.
Appendix A. Supplementary data Yang, Characterization and environmental applications of clay-biochar
composites, Chem. Eng. J. 242 (2014) 136–143.
[23] J.H. Zhang, M.C. He, Effect of dissolved organic matter on sorption and
Supplementary data associated with this article can be found, in desorption of phenanthrene onto black carbon, J. Environ. Sci. 25 (12) (2013)
the online version, at https://doi.org/10.1016/j.bej.2017.11.018. 2378–2383.
[24] D.H. Lin, J. Ji, Z.F. Long, K. Yang, F.C. Wu, The influence of dissolved and
surface-bound humic acid on the toxicity of TiO2 nanoparticles to Chlorella
References sp, Water Res. 46 (2012) 4477–4487.
[25] A.A. Bazrafshan, S. Hajati, M. Ghaedi, Synthesis of regenerable Zn(OH)2
[1] B. Wang, Y.S. Jiang, F.Y. Li, D.Y. Yang, Preparation of biochar by simultaneous nanoparticle-loaded activated carbon for the ultrasound-assisted removal of
carbonization, magnetizationand activation for norfloxacin removal in water, malachite green: optimization, isotherm and kinetics, RSC Adv. 5 (2015)
Bioresour. Technol. 233 (2017) 159–165. 79119–79128.
[2] H.R. Pouretedal, N. Sadegh, Effective removal of amoxicillin, cephalexin. [26] L.F. Han, K. Sun, J. Jin, X. Wei, X.H. Xia, F.C. Wu, B. Gao, B.S. Xing, Role of
Tetracycline and penicillin G from aqueous solutions using activated carbon structure and microporosity in phenanthrene sorption by natural and
nanoparticles prepared from vine wood, J. Water Process Eng. 1 (2014) 64–73. engineered organic matter, Environ. Sci. Technol. 48 (2014) 11227–11234.
[3] D.M. Cheng, X.H. Liu, S.N. Zhao, B.S. Cui, J.H. Bai, Z.J. Li, Influence of the natural [27] C.P. Chen, W.J. Zhou, D.H. Lin, Sorption characteristics of
colloids on the multi-phase distributions of antibiotics in the surface water N-nitrosodimethylamine onto biochar from aqueous solution, Bioresour.
from the largest lake in North China, Sci. Total Environ. 578 (2017) 649–659. Technol. 179 (2015) 359–366.
[4] S. Zhou, S. Zhang, F. Liu, J. Liu, J. Xue, D. Yang, C. Chang, ZnO nanoflowers [28] M. Zhang, L. Shu, X.F. Shen, X.Y. Guo, S. Tao, B.S. Xing, X.L. Wang,
photocatalysis of norfloxacin: effect of triangular silver nanoplates and Characterization of nitrogen-rich biomaterial-derived biochars and their
watermatrix on degradation rates, J. Photochem. Photobiol. A: Chem. 328 sorption for aromatic compounds, Environ. Pollut. 195 (2014) 84–90.
(2016) 97–104. [29] M.C. Zhang J.H. and He, Effect of surfactants on sorption and desorption of
[5] M. Inyang, E. Dickenson, The potential role of biochar in the removal of phenanthrene onto black carbon, Water Environ. Res. 83 (2011) 15–22.
organic and microbial contaminants from potable and reuse water: a review, [30] T.J. Kinney, C.A. Masiello, B. Dugan, W.C. Hockaday, M.R. Dean, K. Zygourakis,
Chemosphere 134 (2015) 232–240. R.T. Barnes, Hydrologic properties of biochars produced at different
[6] M.B. Ahmed, J.L. Zhou, H.H. Ngo, W.S. Guo, M.A. Johir, K. Sornalingam, Single temperatures, Biomass Bioenergy 41 (2012) 34–43.
and competitive sorption properties and mechanism of functionalized [31] M.H. Sui, Y.F. Zhou, L. Sheng, B.B. Duan, Adsorption of norfloxacin in aqueous
biochar for removing sulfonamide antibiotics from water, Chem. Eng. J. 311 solution by Mg-Al layered double hydroxideswith variable metal composition
(2017) 348–358. and interlayer anions, Chem. Eng. J. 210 (2012) 451–460.
[7] P. Oleszczuk, M. Kołtowski, Effect of co-application of nano-zero valent iron [32] K.K. Shimabuku, J.P. Kearns, J.E. Martinez, R.B. Mahoney, L. Moreno-Vasquez,
and biochar on the total and freely dissolved polycyclic aromatic R.S. Summers, Biochar sorbents for sulfamethoxazole removal from surface
hydrocarbons removal and toxicity of contaminated soils, Chemosphere 168 water, stormwater, and wastewater effluent, Water Res. 96 (2016) 236–245.
(2017) 1467–1476. [33] B. Pan, D. Zhang, H. Li, M. Wu, Z.Y. Wang, B.S. Xing, Increased adsorption of
[8] D. Fu, Z. Chen, D. Xia, L. Shen, Y.P. Wang, Q.B. Li, A novel solid sulfamethoxazole on suspended carbon nanotubes by dissolved humic acid,
digestate-derived biochar-Cu NP composite activatingH2 O2 system for Environ. Sci. Technol. 47 (2013) 7722–7728.
simultaneous adsorption and degradation oftetracycline, Environ. Pollut. 221 [34] B. Sun, F. Lian, Q. Bao, Z. Liu, Z. Song, L. Zhu, Impact of low molecular weight
(2017) 301–310. organic acids (LMWOAs) on biochar micropores and sorption properties for
[9] F. Reguyala, A.K. Sarmaha, W. Gao, Synthesis of magnetic biochar from pine sulfamethoxazole, Environ. Pollut. 214 (2016) 142–148.
sawdust via oxidative hydrolysis of FeCl2 for the removal sulfamethoxazole [35] H. Haham, A. Oren, B. Chefetz, Insight into the role of dissolved organic matter
from aqueous solution, J. Hazard. Mater. 321 (2017) 868–878. in sorption of sulfapyridine by semiarid soils, Environ. Sci. Technol. 46 (2012)
[10] R.H. Li, J.J. Wang, B.Y. Zhou, Z.Q. Zhang, S. Liu, S. Lei, R. Xiao, Simultaneous 11870–11877.
capture removal of phosphate, ammonium and organic substances by MgO [36] Y. Gao, X. Yuan, X. Lin, B. Sun, Z. Zhao, Low-molecular-weight organic
impregnated biochar and its potential use in swine wastewater treatment, J. acidsenhance the release of bound PAH residues in soils, Soil Tillage Res. 145
Clean. Prod. 147 (2017) 96–107. (2015) 103–110.
[11] L. Zhao, X. Cao, W. Zheng, et al., Endogenous minerals have influences on [37] X.S. Kong, S.X. Feng, X. Zhang, Y. Li, Effects of bile salts and divalent cations on
surface electrochemistry and ion exchange properties of biochar, the adsorption of norfloxacinbyagricultural soils, J. Environ. Sci. 26 (2014)
Chemosphere 136 (2015) 133–139. 846–854.
[12] K. Sun, M. Kang, K.S. Ro, et al., Variation in sorption of propiconazole with [38] B.Y. Huang, Y.G. Liu, B. Li, et al., Effect of Cu(II) ions on the enhancement of
biochars: the effect of temperature, mineral, molecular structure, and tetracycline adsorption byFe3 O4 @SiO2 -chitosan/graphene oxide
nano-porosity, Chemosphere 142 (2016) 56–63. nanocomposite, Carbohydr. Polym. 157 (2017) 576–585.
[13] Y. Li, Z.W. Wang, X.Y. Xie, J.M. Zhu, R.N. Li, T.T. Qin, Removal of Norfloxacin [39] Z.G. Pei, X.Q. Shan, S.Z. Zhang, et al., Insight to ternary complexes of
from aqueous solution by clay-biochar composite prepared from potato stem co-adsorption of norfloxacin and Cu(II) onto montmorillonite at different pH
and natural attapulgite, Colloids Surf. A 514 (2017) 126–136. using EXAFS, J. Hazard. Mater. 186 (2011) 842–848.
[14] A. Benhouria, M.A. Islam, H. Zaghouane-Boudiaf, et al., Calcium [40] X. Tan, Y. Liu, G. Zeng, et al., Application of biochar for the removal of
alginate–bentonite–activated carbon composite beads as highly effective pollutants from aqueous solutions, Chemosphere 125 (2015) 70–85.
adsorbent for methylene blue, Chem. Eng. J. 270 (2015) 621–630. [41] M.B. Ahmed, J.L. Zhou, H.H. Ngo, W.S. Guo, Adsorptive removal of antibiotics
[15] S. Ismadji, D. Tong, F.E. Soetaredjo, et al., Bentonite hydrochar composite for from water and wastewater: progress and challenges, Sci. Total Environ. 532
removal of ammonium from Koi fish tank, Appl. Clay Sci. 119 (2016) 146–154. (2015) 112–126.
[16] M. Xie, W. Chen, Z. Xu, et al., Adsorption of sulfonamides to demineralized [42] H. Liu, W. Ning, P.F. Cheng, J. Zhang, Y. Wang, C.L. Zhang, Evaluation of animal
pine wood biochars prepared under different thermochemical conditions, hairs-based activated carbon for sorption of norfloxacin andacetaminophen
Environ. Pollut. 186 (2014) 187–194. by comparing with cattail fiber-based activated carbon, J. Anal. Appl. Pyrolysis
[17] F.F. Liu, J. Zhao, S.G. Wang, B.S. Xing, Adsorption of sulfonamides on reduced 101 (2013) 156–165.
graphene oxides as affected by pH and dissolved organic matter, Environ. [43] T.M. Darweesh, M.J. Ahmed, Adsorption of ciprofloxacin and norfloxacin from
Pollut. 210 (2016) 85–93. aqueous solution ontogranular activated carbon in fixed bed column,
[18] F. Lian, B. Sun, X. Chen, et al., Effect of humic acid (HA) on sulfonamide Ecotoxicol. Environ. Saf. 138 (2017) 139–145.
sorption by biochars, Environ. Pollut. 204 (2015) 306–312. [44] S.C. Zou, W.H. Xu, R.J. Zhang, et al., Occurrence and distribution of antibiotics
[19] H. Peng, S. Feng, X. Zhang, Y. Li, X. Zhang, Adsorption of norfloxacin onto in coastal water of the Bohai Bay China: impacts of river discharge and
titanium oxide: effect of drug carrier and dissolved humic acid, Sci. Total quaculture activities, Environ. Pollut. 159 (2011) 2913–2920.
Environ. 438 (2012) 66–71.
[20] W.B. Yang, Y.P. Lu, F.F. Zheng, X.X. Xue, N. Li, D.M. Liu, Adsorption behavior
and mechanisms of norfloxacin onto porous resins andcarbon nanotube,
Chem. Eng. J. 179 (2012) 112–118.

You might also like