You are on page 1of 9

Journal of Colloid and Interface Science 551 (2019) 155–163

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Regular Article

Influence of fulvic acid on Pb(II) removal from water using a


post-synthetically modified MIL-100(Fe)
Haibo Zhang a,b,1, Jia Wen a,b,⇑,1, Ying Fang a,b, Siyu Zhang a,b, Guangming Zeng a,b
a
College of Environmental Science and Engineering, Hunan University, Changsha 410082, PR China
b
Key Laboratory of Environmental Biology and Pollution Control (Hunan University), Ministry of Education, Changsha 410082, PR China

g r a p h i c a l a b s t r a c t

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

Article history: Hypothesis: Modification of MIL-100(Fe) with ethanediamine (ED) is expected to improve the removal
Received 20 March 2019 percentage of Pb(II) ions from aqueous solution. Nevertheless, the adsorption performance of ED-MIL-
Revised 4 May 2019 100(Fe) in natural waters will be affected by natural organic matter such as fulvic acids (FA). Hence, it
Accepted 4 May 2019
is necessary to study the influence of FA on the physicochemical property, adsorption behavior and mech-
Available online 6 May 2019
anisms of ED-MIL-100(Fe) towards the removal of Pb(II) from water.
Experiments: Batch adsorption experiments were conducted to compare adsorption capacity of MIL-100
Keywords:
(Fe) and ED-MIL-100(Fe) for Pb(II). In addition, XRD, FTIR, SEM and XPS techniques were applied to inves-
Pb(II)
Adsorption
tigate the characteristics, adsorption isotherms, and kinetics in the presence and absence of FA.
Post-synthetic modification Findings: ED-MIL-100(Fe) enabled 99% removal of Pb(II) from water at or below 60 mg L1, whereas the
MIL-100(Fe) MIL-100(Fe) hardly adsorbed any Pb(II). Additionally, the adsorption percentage of the Pb(II) ions (76%)
Fulvic acid outcompeted that of other metal ions (i.e. Co, Cu, Ni, Zn, and Cd) (each  25%) in the mixed solution. The
negative surface charge of ED-MIL-100(Fe) was not affected by FA. Increasing FA concentration steadily
increased the functional groups surrounding the ED-MIL-100(Fe) via electrical interaction, thus facilitat-
ing the Pb(II) ions adsorption.
Ó 2019 Elsevier Inc. All rights reserved.

⇑ Corresponding author at: College of Environmental Science and Engineering,


Hunan University, Changsha 410082, PR China.
E-mail address: jwen@hnu.edu.cn (J. Wen).
1
Equal first author.

https://doi.org/10.1016/j.jcis.2019.05.016
0021-9797/Ó 2019 Elsevier Inc. All rights reserved.
156 H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163

1. Introduction lating the removal kinetics and adsorption capacity for Pb(II) ions.
This study provides critical information for optimal design of ED-
Lead (Pb) in water environment has posed a significant threat to MIL-100(Fe) that is capable of removing Pb(II) ions in environmen-
human and other animals due to the fact that they can be tal water samples.
accumulated in living organisms [1,2]. Compared with other heavy
metals, the acute and chronic toxicity of Pb(II) ions can cause 2. Experimental
anemia, renal disorders, and can damage central nervous system
[3]. Therefore, the World Health Organization (WHO) has set the 2.1. Materials
limitation for Pb(II) ions in drinking water to be as low as
0.01 mg L1 [4]. There are many reported methods concerning All solvents and reactants obtained were of analytical grade and
the removal of Pb(II) ions from the water environment such as used without further purification. Ferric nitrate ninehydrate (Fe
chemical precipitation, ion exchange, membrane filtration, and (NO3)39H2O, Hushi, Shanghai, China, 99%), trimesic acid (H3BTC,
flocculation [5]. Compared to these methods, adsorption technol- Aladdin Industrial Corporation, Shanghai, China, 98%), ethylenedi-
ogy is the most widely used in the purification processes owing amine (ED, Hushi, Shanghai, China, 99%) were used to synthesize
to its easy application and superior efficiency [6–9]. samples. Lead nitrate (Pb(NO3)3, Hushi, Shanghai, China, AR) was
In recent years, metal organic frameworks (MOFs) have used as a target heavy metal pollutant. FA was purchased from
emerged as a group of promising materials for adsorptive removal Cool Seoul Bio [90%, C14H12O8 (MW = 308)]. The ultrapure water
of hazardous materials from the water environment due to their (18.25 MX cm) was produced by Millipore Milli-Q water purifica-
high porosity, large surface areas and adjustable properties [10– tion system (UPT-11-40).
14]. Meanwhile, post-synthetic modification (PSM) is a foremost
approach to enlarge the potential application of MOFs. During 2.2. Preparation of the adsorbents
the PSM process, MOFs species undergo chemical transformations
while retaining the crystallinity and original framework structure 2.2.1. Synthesis of MIL-100(Fe)
[15]. The physico-chemical properties of MOFs can be altered by Considering that handling of hydrofluoric acid (HF) is harmful
PSM approach either by developing metal binding or functional to nature and environment, herein, we chose a new strategy to
organic sites or both sites on the surface of MOFs [16]. Previously, synthesize MIL-100(Fe) under HF-free conditions according to
ethanediamine (ED) modified MIL-101(Cr) successfully improved the study of Fang et al.[30]. The procedure was shown in Support-
Pb(II) adsorption with high selectivity [17]. The introduction of ing Information (SI) Text 1.
the thiol (ASH) to the surface of MOF-5 also improved the stability
of MOF-5 material in water [18]. MIL-100(Fe) is a more 2.2.2. Procedure for post-synthetic modification of ED-MIL-100(Fe)
environmental-friendly MOF compared to the above-mentioned, For the synthesis of various amine-functionalized MIL-100(Fe),
and has been regarded as a great candidate for heavy metal 1 g MIL-100(Fe) was dehydrated at 150 °C for 12 h, and then sus-
removal due to its high water stability, non-toxicity and good pended in absolute ethanol (100 mL). Following that, various con-
adsorption capacity [19]. However, its ability for Pb(II) removal centrations of ED solution (v/v) (0.4% for sample A; 0.6% for sample
from wastewater is still a subject that hasn’t been investigated. B; 0.8% for sample C; and 1% for sample D) was individually intro-
Humic substances (HS), further operationally divided into fulvic duced to each suspension, and the mixture was transferred into
acids (FA), humic acids (HA) and humin according to their solubil- autoclave and heated at 90 °C for 12 h. The final product, named
ity [20], are a group of natural organic matter (NOM) found ubiq- ED-MIL-100(Fe) (A, B, C, and D), was washed with ethanol several
uitous in the natural environment. The concentrations of HS in times, and then dried at 90 °C prior to characterization.
natural waters vary from a few mg/L of dissolved organic C
(DOC) to more than a few hundreds mg/L DOC [21]. In waters, con- 2.3. Characterization
taminants will inevitably interact with NOM thus changing their
properties (such as electrophoretic mobility, transport and interac- X-ray powder diffraction (XRD) patterns of the as-synthesized
tions with colloids) [22]. Therefore, the influence of HS on the four samples were obtained on a Rigaku D/Max 2500 diffractometer
removal of contaminants has been reported with controversial using CuKa radiation in a scanning range of 5–40° at a scanning rate
results. Depending on the condition, either positive or negative of 1° min1. Scanning electron microscope (SEM) (FEI QuANTA 200)
effect of HS on the adsorbent was found in aqueous solutions was used to characterize the particle morphology and crystal size
[23–26]. For example, Pignatello et al. [27] reported that HA combined with energy dispersive spectroscopy (EDS) for elemental
exerted an inhibition effect on adsorbents of strong sorption affini- analysis. Fourier-transform infrared (FTIR) spectra were recorded on
ties. In contrary, Dong et al. [28] reported that HA promoted the a Nicolet Nexus 670 FTIR spectrometer using KBr pellets containing
removal of Cr(VI) by zero valent iron. Li et al. [29] showed that 1% weight sample. Zeta potential of the MIL-100(Fe) and ED-MIL-
the presence of FA led to a strong increase in Cu(II) sorption on 100(Fe) samples were determined using dynamic light scattering
Fe3O4 nanoparticles (GO/Fe3O4) at pH < 6 but a decrease at (DLS) (Zetasizer nano-ZS90, Malvern) at 25 °C. The surface electronic
pH > 8. However, there is lack of studies regarding the influence status was investigated by X-ray photoelectron spectroscopy (XPS,
of FA as a model NOM on MOFs. MOFs are currently explored as ES CALAB 250Xi, USA) and the XPS data were internally calibrated
an advanced adsorption material applied in aqueous system, thus by fixing the binding energy of C1s at 284.8 eV.
it is imperative to understand the influence of HS on MOFs with
respect to adsorption patterns and the mechanisms involved. 2.4. Adsorption experiments
The iron trimmers in the cages of MIL-100(Fe) can generate
open metal sites or coordinatively unsaturated sites (CUSs) as 2.4.1. Batch adsorption of Pb(II) ions
Lewis acids in the structure upon vacuum heating, which are Pb(II) ions stock solution (1000 mg g1) was prepared by dis-
usable for surface modification. Therefore, ED used as the common solving analytical reagent 1.5985 g Pb (NO3)2 into 1000 mL ultra-
functional groups can be easily coordinated on the CUSs of the MIL- pure water, and diluted to 5–750 mg g1 for subsequent use.
100(Fe). The adsorption capacity, kinetics, and selectivity of the Adsorption capacity of each sample material was examined at
modified MIL-100(Fe) towards the removal of Pb(II) from aqueous the following experimental condition (pH = 5; solution volume:
solutions were studied. FA was investigated for its effect on regu- 20 mL; shaker speed: 180 rpm; time: 24 h; and temperature:
H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163 157

25 °C). Our preliminary experiment indicated that the FA used in 2.4.3. Adsorption selectivity of the ED-MIL-100(Fe) for heavy metals
the study could easily pass through 1000 Da ultrafiltration mem- The capacity of the various ED-MIL-100(Fe) to selectively
brane due to its small molecular weight, and Pb(II) contents in remove heavy metal ions from aqueous solution was determined
the samples containing FA with or without microfiltration by using a batch process with nitrate salts of Co(II), Cu(II), Pb(II),
(0.22 lm) had no significant difference (P > 0.05). Therefore, after Ni(II), Zn(II), and Cd(II) ions. Ten miligrams of each ED-MIL-100
adsorption, the adsorbent was filtered by 0.22 lm membrane filter, (Fe) sample was added to 20 mL solution of mixed metal ions (con-
and concentrations of Pb(II) ions were analyzed using inductively centration of each metal: 40 mg L1). The suspension was shaken
coupled plasma-optical emission spectrometry (ICP-OES) at room temperature for 6 h, and then filtered through 0.22 lm
(PE5300D, Pelkin Elmer). The amounts of Pb(II) ions adsorbed on membrane filter. Concentrations of each metal ion in the remain-
the original MIL-100(Fe) and ED-MIL-100(Fe) (sample A, B, C, and ing samples were then detected by using ICP-OES.
D) were calculated according to Eq. (1):

ðC0  Ce ÞV 2.4.4. Effects of FA


Qe ¼ ð1Þ Kinetics experiments studying the effects of FA were conducted
m
using 300 mL of 250 mg L1 Pb(II) solution with an increasing con-
where Qe (mg g1) represents the adsorption capacity; C0 (mg L1) centration of FA from 1 to 10 mg L1 at pH 5, following the above-
is the initial concentration of Pb(II) ions; Ce(mg L1) is the equilib- mentioned adsorption procedure. Moreover, the adsorption behav-
rium concentration of Pb(II) ions; m(g) is the weight of adsorbents; ior of the sample C was investigated in the absence or the presence
and V(L) is the volume of solution. of FA with initial concentrations of 1, 5 and 10 mg L1, and data
The adsorption isotherms of the materials were fitted with were fit with Langmuir and Freundlich isotherms. Initial concen-
Langmuir, Freundlich and Dubinin-Radushkevich models to ana- trations of Pb(II) solution were set from 25 to 600 mg L1 and
lyze the adsorption behavior (see SI Text 2). the pH was adjusted to 5 for the isotherm experiment. At predeter-
To study the adsorption kinetics of Pb(II) ions, 150 mg of adsor- mined time interval in the kinetics experiment and 24 h after iso-
bent was added to 300 mL of 250 mg L1 Pb(II) ions solution at therm experiment, samples were filtered using 0.22 lm membrane
25 °C with constant stirring. Analytical samples were taken from filters and the solution concentrations were detected. The precipi-
the mixture solution at given time intervals and immediately fil- tates were collected, washed with Milli-Q water several times and
tered to remove the adsorbent by 0.22 lm membrane filter. The then dried at 90 °C before being subjected to FTIR analysis.
concentrations of Pb(II) ions were measured by ICP-OES. To accu-
rately evaluate the adsorption mechanism, the pseudo-first-order 2.5. Data analysis
and the pseudo-second-order models were applied to fit the
kinetic data (see SI Text 3). All experiments were run in triplicate. Data were statistically
analyzed using Statistica 12.0. One-way analysis of variance
2.4.2. Effect of pH on Pb(II) adsorption (ANOVA) was performed using Origin version 8.0. Least significant
To investigate the effect of pH on Pb(II) adsorption onto the ED- difference (LSD) was used to separate means where ANOVA
MIL-100(Fe), batch adsorption as a function of pH from 3 to 7 was showed significance at P < 0.05. Similar letters on top of the his-
conducted. The solution pH was firstly adjusted by adding 1 M togram mean values of no significant difference (P > 0.05).
NaOH or 1 M HCl dropwise. Following that, 10 mg of the adsorbent
was added to each 20 mL Pb(II) solution (200 mg g1) of different
3. Results and discussion
pH. The shaking regime followed the above-mentioned procedure.
The removal rate (R) of the Pb(II) ions was calculated using the
3.1. Characterization
Eq. (2):

C0  Ce 3.1.1. X-ray powder diffraction


R¼  100% ð2Þ The XRD patterns of the MIL-100(Fe) and ED-MIL-100(Fe)
C0
(sample A, B, C and D) were shown in Fig. 1a. The diffraction peaks
where C0 (mg g1) is the initial concentration of Pb(II) ions; and Ct of MIL-100(Fe) before and after PSM with ED matched well with
(mg g1) is the concentration of the Pb(II) ions at time t (min). the simulation peaks, indicating the MIL-100(Fe) was synthesized

Fig. 1. The (a) XRD patterns and (b) FTIR spectra of the pristine MIL-100(Fe) and ED-MIL-100(Fe)s (Sample A, B, C and D).
158 H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163

successfully and the structure was maintained after PSM. As evi- confirmed that the adsorbent had taken up Pb(II) ions directly.
denced by the relative intensities, the patterns of the modified As for the Fig. 3f, the N 1s peak around 400 eV was comprised of
MIL-100(Fe) still obtained a well crystal structure even if the NAH (399.6 eV) and N-metal (401.8 eV). In particular, the peak at
increasing dose of ED was added during PSM. In comparison, the 401.8 eV could be attributed to the formation of NH-Pb complexes,
same MOF-series MIL-101 was not resistant to increasing ED doses in which a lone pair of electrons in the nitrogen atom was donated
as shown in the study of Luo et al. [17], indicating the better stabil- to the shared bond between the N atom and metals, leading to a
ity of MIL-100(Fe) than MIL-101 when subject to the ED solution. higher BE peak observed [40]. Hence, the XPS spectra of N1s sug-
gested that one possible adsorption mechanism could be through
3.1.2. FTIR analysis the formation of metal complexes by the Pb(II) ions with the nitro-
Post-synthetic modification by ED was used to implant amino gen atoms, which was also supported by the result of FTIR where
functional groups on the structure of MIL-100(Fe). As shown in the ANH2 functional group was identified for element N. Therefore,
Fig. 1b, the appearance of new peaks located at 1562 and the electron-donor ED-MIL-100(Fe) could have an excellent
1042 cm1 correspond to the NAH plane stretching and CAN bond adsorption capacity toward Pb(II) ions in water.
stretching frequency in the amino functional group, respectively.
The extended stretching regions of v(ANH) at 2950 cm1 and v 3.2. Adsorption experiments
(ACH) at 2890 cm1 from sample A to D also suggested the ED
grafting on MIL-100(Fe) [31]. The implanted amino groups were 3.2.1. Adsorption isotherms
in the form of neutral ANH2 rather than protonated ANH+3, evi- Adsorption of Pb(II) ions on various ED-MIL-100(Fe) were com-
denced by a missing peak at around 2100 cm1 that belongs to pared using 1 mg of ED-MIL-100(Fe) (sample A, B, C, and D) added
the protonated ammonia [32–34]. to 20 mL of an aqueous solution containing Pb(II) ions of various
concentration. The original MIL-100(Fe) almost had no adsorption
3.1.3. SEM and EDS analysis for the cationic Pb(II) ions in the solution (with concentration
Surface morphologies of the five samples were investigated by below detection limit), whereas the functionalized ED-MIL-100
SEM technique. There was not much difference in surface morphol- (Fe) remarkably increased the adsorption of Pb(II) ions, owing to
ogy between the pristine MIL-100(Fe) (Fig. 2a) and the functional- the covalent interaction occurring between the Pb(II) ions and
ized sample C (Fig. 2b; others were identical so figures are not the neutral ANH2 (Fig. 4a). There were significant differences of
shown). Additionally, the percentage of N increased from 6.24% adsorption between sample A, B and C (P < 0.05) but no difference
to 7.33% with the increasing ED added during the modification of between sample C and D (P > 0.05), suggesting that the grafted
MIL-100(Fe) (Fig. 2c). Combined with the analysis of SEM and FTIR, capacity may have reached saturation. Therefore, from an eco-
it is confirmed that ED had been grafted on the surface of MIL-100 nomic point of view, sample C was elected for all subsequent
(Fe). experiments.
Langmuir and Freundlich isotherm models were applied to fur-
3.1.4. XPS analysis ther understand the adsorption process of Pb(II) on ED-MIL-100
The XPS survey of the chemical composition of C, O, N, and Fe in (Fe) (sample C). The fitting results from the Langmuir and Fre-
the ED-MIL-100(Fe) (sample C) before and after adsorption of Pb(II) undlich models were shown in Fig. 4b and Table 1. The calculated
was shown in Fig. 3a. These peaks shifted slightly to higher binding maximum Pb(II) adsorption capacity of ED-MIL-(100) sample C
energy compared with the initial MOFs before adsorption, which was 378.8 mg g1, much higher than that of many other original
could be attributed to the interaction between the N and Pb(II) or functionalized MOFs (Table 2). The values of correlation coeffi-
[35]. As shown in Fig. 3b, the Fe 2p energy peaks can be deconvo- cients (R2) indicated that the experimental data fitted better with
luted into three peaks centered at the band energy (BE) of 712.3 eV, the Langmuir model than the Freundlich model (Table 1), suggest-
725.8 eV and 718.7 eV, corresponding to the peaks of Fe 2p3/2, Fe ing a homogeneous monolayer adsorption process of Pb(II) ions on
2p1/2 and the satellite peaks of Fe 2p3/2 [36]. In the O 1s spectrum the ED-MIL-100(Fe). Based on the value of E from the linear fitting
(Fig. 3c), the energy peaks at 531.2 eV and 531.5 eV before and result of Dubinin-Radushkevich model, we can justify that the
after Pb(II) adsorption suggested the presence of AOH, and H2O mechanism of adsorption was a chemical process due to the com-
adsorbed on the surface of ED-MIL-100(Fe) [37]. The appearance plexation of ANH2 and Pb(II) on the surface of MIL-100(Fe) [41].
of C 1s peaks at 284.8 eV and 288.5 eV represented the phenyl
and carboxyl signals, respectively (Fig. 3d) [38]. After adsorption, 3.2.2. Adsorption kinetics
the two well-defined and symmetrically centered peaks at The adsorption of Pb(II) ions on the ED-MIL-100(Fe) was
143.9 eV and 138.8 eV (Fig. 3e), corresponding to the Pb 4f5/2 and very fast in the first 10 min and reached equilibrium within
Pb 4f7/2 signals [39], were detected in the wide-scan spectrum. It 40 min (Fig. 4c). The high adsorption rate can be attributed to

Fig. 2. SEM images of (a) the pristine MIL-100(Fe) and (b) ED-MIL-100(Fe) (Sample C); (c) weight percentage of C, O, Fe and N in the samples.
H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163 159

Fig. 3. XPS spectra for sample C (a) before and after adsorption; (b) Fe 2p; (c) O 1s; (d) C 1s; (e) Pb 4f; (f) N 1s.

the abundant adsorption sites on the sample C and the strong surface, and the reaction is dominantly controlled by chemical
complexation between Pb(II) ions and the amino groups. The reaction [47].
adsorption kinetics of Pb(II) on the ED-MIL-100(Fe) was found to
fit better with the pseudo-second-order than the pseudo-first- 3.2.3. Effect of pH on the adsorption of Pb(II)
order kinetic model (SI Table 1). Again, it confirms that the adsorp- The effect of initial pH on Pb(II) removal from ED-MIL100(Fe)
tion process is proportional to the number of active sites on its was shown in Fig. 5a. Generally, the removal percentage of the
160 H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163

Fig. 4. (a) Pb(II) adsorption isotherms for samples A–D; (b) Langmuir and Freundlich model fitting for the adsorption of Pb(II) on sample C; (c) adsorption kinetic of Pb(II) on
sample C.

Table 1 3.2.4. Selective adsorption of Pb(II)


Adsorption isotherm parameters from Langmuir, Freundlich and Dubinin-Radushke- The ED-MIL-100(Fe) showed selective adsorption toward Pb(II)
vich models.
as compared to other cationic metal ions (Fig. 6). The adsorption
Adsorption isotherms Isotherm constants Values percentage of Pb(II) ions on the ED-MIL-100(Fe) reached over
Langmuir Qm (mg g1) 378.8 76%, while equal or less than 25% of each of the competing metal
KL (L mg1) 0.098 ion was adsorbed. This phenomenon can be attributed to fact that
R2 0.9787 the Pb(II) ions possess the biggest absolute electronegativity and
Freundlich n 4.7267 the largest radius amongst these metal ions, indicating that Pb(II)
KF (mg g1) 111.78 ions have the strongest attraction to the lone electron pairs on
R2 0.9168
the N doner, and consequently form the most stable complexes
Dubinin-Radushkevich Qs (mg g1) 391.5 [50]. For this reason, it is hopefully that the application of ED-
E (kJmol1) 9.407
K (mol2kJ2) 0.00565
MIL-100(Fe) in wastewater of mixed metals may be only selective
R2 0.9578 for the removal of Pb(II).

3.2.5. Effects of FA
Table 2 The influence of FA on the Pb(II) ions sorption by ED-MIL-100
Comparison of MOFs and other adsorbents for Pb(II) removal. (Fe) was studied at pH 5, at which condition the ANH2 groups on
Adsorbents Adsorption capacities (mg g1) References the MOF might become protonated. Because FA is negatively
UiO-66 8.4 [42]
charged over the whole range of pH [51,52], it is likely that the pro-
UiO-66-NH2 31.2 [42] tonated ANH2 groups could attract the FA through electrostatic
UiO-66-NHC(S)NHMe 90.8 [42] attraction. Indeed, it was found that the adsorption capacity was
Zr-MOFs 135.0 [42] slightly increased with the increasing concentration of FA from 0
NH2 functionalized Zr-MOFs 166.7 [43]
to 10 mg L1 (Fig. 7a), likely due to the increasing functional groups
MIL-101 15.8 [17]
ED-MIL-101 (5 mmol) 81.1 [17] provided by FA which are able to bind with Pb(II) ions. Fig. 7b
Fe3O4 @SiO2-EDTA 114.9 [44] clearly shows the influence of FA on the adsorption isotherms of
(Fe3O4-ED)/MIL-101(Fe) 198.0 [45] Pb(II) ions for the Freundlich and Langmuir models at pH 5, and
Melamine-MOFs 122.0 [46] the fitting constants are summarized in Table 3. Increasing FA con-
MIL-100(Fe) <detection limit This work
ED-MIL-100(Fe) 378.8 This work
centrations in the solution improved the adsorption capacity of ED-
MIL-100(Fe) from 405.3 mg g1 to 427.6 mg g1 towards Pb(II). For
all conditions, the Langmuir model fits better than the Freundlich
ED-MIL-100(Fe) for Pb(II) ions gradually increased with the model, suggesting that the adsorption of Pb(II) ions still followed
increase of pH from 3 to 7. We believe that the large amount of monolayer molecular adsorption under the influence of FA. We
hydrated hydrogen ions present at low pH (i.e. pH = 3) could be also compared the FTIR spectrum of the ED-MIL-100(Fe) in the
competitive with the Pb(II) ions for the same adsorption sites. Once absence and presence of FA (Fig. 7c) and found a slight decrease
the pH increased to above 7, Pb(II) can be gradually present in of the NAH plane stretching at 1562 cm1 and the disappearance

forms of Pb4(OH)4+ 3+
4 , Pb3(OH)4 , Pb(OH)2 and Pb(OH)3 in aqueous of CAN bond stretching at 1042 cm1. The variation indicates the
solution, then the removal of Pb(II) was mostly due to precipitation CAN bonds of ED might have been broken in the presence of FA,
but not adsorption [48]. Meanwhile, the zeta potential analysis with only one amino group in every grafted ED preserved intact
(Fig. 5b) showed that the surface charge of MIL-100(Fe) was posi- and attached to the MOF (Fig. 8). Therefore, the phenomena could
tive from pH 3 to 10, causing electrostatic repulsion against the be considered as three steps. Firstly, the ED groups grafted on the
cationic Pb(II) ions, therefore there was an absence of Pb(II) surface of MOFs were altered with one CAN bond broken upon the
adsorption on the MIL-100(Fe). In comparison, the surface of ED- arrival of FA; secondly, parts of the remaining ANH2 functional
MIL-100(Fe) was negatively charged throughout pH 3 to 11, which groups in ED-MIL-100(Fe) were protonated under acidic condi-
facilitated Pb(II) ions adsorption due to electrostatic attraction tions; thirdly, a small fraction of the FA molecules arrived at the
[49]. Therefore, the ultrahigh adsorption capacity of ED-MIL-100 surface of adsorbent may attach instantly to the protonated amino
(Fe) (Table 2) was derived from the strong electrostatic attraction groups, thereby increasing adsorption sites and causing the
(physically) and coordination interaction (chemically) between observed increase in Pb(II) ions adsorption [53,54]. However, it is
the MOFs and Pb(II). not known why the CAN bond would be broken (or CAN signal
H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163 161

Fig. 5. (a) Effect of pH on the removal rate of Pb(II) and (b) the zeta potential of MIL-100(Fe) and sample C (in the absence and presence of FA).

whatever the concentration of FA (Fig. 5b). Therefore, there was


no chance of FA attaching to the surface of ED-MIL-100(Fe).

4. Conclusion

In summary, the amino-functionalized MIL-100(Fe) was pre-


pared by a simple PSM method to efficiently remove Pb(II) from
water. Amino groups were successfully grafted on the MIL-100
(Fe), and thus changing the surface charge and chemistry of MIL-
100(Fe) by introducing the functional groups. Both physical (elec-
trostatic attraction) and chemical (coordination interaction) forces
existed between the ED-MIL-100(Fe) and Pb(II). Nevertheless, the
isotherm and kinetics fitting suggested that chemical adsorption
played the dominating role during the adsorption. The surface
charge of the material was not affected by FA at pH 5; meanwhile,
FA provided more adsorption sites and then promoted the removal
Fig. 6. Adsorption selectivity of sample C toward Co(II), Cu(II), Pb(II), Ni(II), Zn(II) of Pb(II). Taken together, the results showed that ED-MIL-100(Fe)
and Cd(II) ions. is an effective adsorbent for Pb(II) ions at natural water environ-
ment where FA is an nonnegligible factor regulating many chemi-
cal reactions. The influence of NOM on other MOF materials that
missing) in the presence of FA, requiring further research to study have been tested effective in contaminant removal, including the
the mechanism in depth. It is also worth mentioning that the inorganic and organic contaminant, should be further studied to
surface charge of ED-MIL-100(Fe) was not obviously changed at provide useful information for real application.

Fig. 7. Adsorption kinetics (a) and isotherms (b) for Pb(II) adsorption by sample C in the absence and presence of FA; (c) FT-IR spectra of sample C in the absence and presence
of FA after adsorption.
162 H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163

Table 3
The isotherm parameters obtained using Langmuir and Freundlich models in the absence and presence of FA.

FA (mg L1) Langmuir Freundlich


1 1
Qm (mg g ) KL (L mg ) R2
n KF (mg g1) R2
0 405.3 0.168 0.9006 5.9031 162.91 0.6777
1 403.8 0.180 0.9041 5.9343 163.98 0.6877
5 423.3 0.188 0.9428 6.0890 175.01 0.6836
10 427.9 0.229 0.9527 6.0882 179.68 0.7143

[5] F. Fu, Q. Wang, Removal of heavy metal ions from wastewaters: a review, J.
Environ. Manage. 92 (2011) 407–418.
[6] F.A. López, M.I. Martı́N, C. Pérez, A. López-Delgado, F.J. Alguacil, Removal of
copper ions from aqueous solutions by a steel-making by-product, Water Res.
37 (2003) 3883–3890.
[7] G. Li, Z. Zhao, J. Liu, G. Jiang, Effective heavy metal removal from aqueous
systems by thiol functionalized magnetic mesoporous silica, J. Hazard. Mater.
192 (2011) 277–283.
[8] J. Deng, Y. Liu, S. Liu, G. Zeng, X. Tan, B. Huang, X. Tang, S. Wang, Q. Hua, Z. Yan,
Competitive adsorption of Pb(II), Cd(II) and Cu(II) onto chitosan-pyromellitic
dianhydride modified biochar, J. Colloid. Interface Sci. 506 (2017) 355–364.
[9] W. Wang, P. Xu, M. Chen, G. Zeng, C. Zhang, C. Zhou, Y. Yang, D. Huang, C. Lai,
M. Cheng, L. Hu, W. Xiong, H. Guo, M. Zhou, Alkali metal-assisted synthesis of
graphite carbon nitride with tunable band-gap for enhanced visible-light-
driven photocatalytic performance, ACS Sustain. Chem. Eng. 6 (2018) 15503–
15516.
[10] N.A. Khan, Z. Hasan, S.H. Jhung, Adsorptive removal of hazardous materials
using metal-organic frameworks (MOFs): a review, J. Hazard. Mater. 244–245
(2013) 444–456.
[11] X. Qi, B. Huang, X. Yuan, W. Hui, Z. Wu, L. Jiang, T. Xiong, Z. Jin, G. Zeng, W. Hou,
Modified stannous sulfide nanoparticles with metal-organic framework:
toward efficient and enhanced photocatalytic reduction of chromium (VI)
under visible light, J. Colloid Interf. Sci. 530 (2018) 481–492.
[12] W. Hou, X. Yuan, W. Yan, G. Zeng, X. Chen, L. Leng, L. Hui, Synthesis and
applications of novel graphitic carbon nitride/metal-organic frameworks
mesoporous photocatalyst for dyes removal, Appl. Catal. B – Environ. 174–
175 (2015) 445–454.
[13] H. Wang, X. Yuan, Y. Wu, G. Zeng, H. Dong, X. Chen, L. Leng, Z. Wu, L. Peng, In
situ synthesis of In2S3@MIL-125(Ti) core–shell microparticle for the removal of
tetracycline from wastewater by integrated adsorption and visible-light-
driven photocatalysis, Appl. Catal. B – Environ. 186 (2016) 19–29.
Fig. 8. Possible mechanism associated with Pb(II) adsorption on ED-MIL-100(Fe) [14] Q. Xia, H. Wang, B. Huang, X. Yuan, J. Zhang, J. Zhang, L. Jiang, T. Xiong, G. Zeng,
under the influence of FA. State-of-the-art advances and challenges of iron-based metal organic
frameworks from attractive features, synthesis to multifunctional
applications, Small 15 (2019) 1803088.
[15] A.H. Chughtai, N. Ahmad, H.A. Younus, A. Laypkov, F. Verpoort, ChemInform
Funding abstract: metal—organic frameworks: versatile heterogeneous catalysts for
efficient catalytic organic transformations, Chem. Soc. Rev. 46 (2015) 6804–
This work was supported by Changsha Science and Technology 6849.
[16] J. Gascon, U. Aktay, M.D. Hernandez-Alonso, G.P.M.V. Klink, F. Kapteijn, Amino-
Program (grant number: kq1801006) and Natural Science Founda- based metal-organic frameworks as stable, highly active basic catalysts, J.
tion of Hunan Province, China (Grant No. 2019JJ50043). Catal. 261 (2009) 75–87.
[17] X. Luo, D. Lin, J. Luo, Adsorptive removal of Pb(II) Ions from aqueous samples
with amino-functionalization of metal-organic frameworks MIL-101(Cr), J.
Declaration of Competing Interest Chem. Eng. Data 60 (2015) 1732–1743.
[18] J. Zhang, Z. Xiong, C. Li, C. Wu, Exploring a thiol-functionalized MOF for
elimination of lead and cadmium from aqueous solution, J. Mol. Liq. 221
None. (2016) 43–50.
[19] J. Li, X. Wang, G. Zhao, C. Chen, Z. Chai, A. Alsaedi, T. Hayat, X. Wang, Metal-
organic framework-based materials: superior adsorbents for the capture of
Appendix A. Supplementary material toxic and radioactive metal ions, Chem. Soc. Rev. 47 (2018) 2322–2356.
[20] S. Amir, M.H. Lemee, G. Merlina, M. Guiresse, E. Pinelli, J.C. Revel, J.R. Bailly, A.
Supplementary data to this article can be found online at Ambles, Structural characterization of humic acids, extracted from sewage
sludge during composting, by thermochemolysis-gas chromatography-mass
https://doi.org/10.1016/j.jcis.2019.05.016. spectrometry, Process Biochem. 77 (2006) 149–158.
[21] N.A. Wall, G.R. Choppin, Humic acids coagulation: influence of divalent
cations, Appl. Geochem. 18 (2003) 1573–1582.
References [22] M.S. Rahman, M. Whalen, G.A. Gagnon, Adsorption of dissolved organic matter
(DOM) onto the synthetic iron pipe corrosion scales (goethite and magnetite):
[1] J. Wan, C. Zhang, G. Zeng, D. Huang, L. Hu, C. Huang, H. Wu, L. Wang, Synthesis effect of pH, C. Eng. J. 234 (2013) 149–157.
and evaluation of a new class of stabilized nano-chlorapatite for Pb [23] N.F. Adegboyega, V.K. Sharma, K. Siskova, R. Zbořil, M. Sohn, B.J. Schultz, S.
immobilization in sediment, J. Hazard. Mater. 320 (2016) 278–288. Banerjee, Interactions of aqueous Ag+ with fulvic acids: mechanisms of silver
[2] F. Ke, L. Qiu, Y. Yuan, F. Peng, X. Jiang, A. Xie, Y. Shen, J. Zhu, Thiol- nanoparticle formation and investigation of stability, Environ. Sci. Technol. 47
functionalization of metal-organic framework by a facile coordination-based (2013) 757–764.
postsynthetic strategy and enhanced removal of Hg2+ from water, J. Hazard. [24] J. Chen, Z. Xiu, G.V. Lowry, P.J.J. Alvarez, Effect of natural organic matter on
Mater. 196 (2011) 36–43. toxicity and reactivity of nano-scale zero-valent iron, Water Res. 45 (2011)
[3] R.A.K. Rao, M.A. Khan, F. Rehman, Batch and column studies for the removal of 1995–2001.
lead(II) ions from aqueous solution onto lignite, Adsorpt. Sci. Technol. 29 [25] I. Christl, R. Kretzschmar, Interaction of copper and fulvic acid at the hematite-
(2011) 83–98. water interface, Geochim. Cosmochim. Acta 65 (2001) 3435–3442.
[4] Ihsanullah, A. Abbas, A.M. Al-Amer, T. Laoui, M.J. Al-Marri, M.S. Nasser, M. [26] F. He, D. Zhao, J. Liu, C.B. Roberts, Stabilization of FePd nanoparticles with
Khraisheh, M.A. Atieh, Heavy metal removal from aqueous solution by sodium carboxymethyl cellulose for enhanced transport and dechlorination of
advanced carbon nanotubes: critical review of adsorption applications, Sep. trichloroethylene in soil and groundwater, Ind. Eng. Chem. Res. 46 (2007) 29–
Purif. Technol. 157 (2016) 141–161. 34.
H. Zhang et al. / Journal of Colloid and Interface Science 551 (2019) 155–163 163

[27] J.J. Pignatello, K. Seokjoon, L. Yufeung, Effect of natural organic substances on [41] E. Wibowo, M. Rokhmat, Sutisna, Khairurrijal, M. Abdullah, Reduction of
the surface and adsorptive properties of environmental black carbon (char): seawater salinity by natural zeolite (Clinoptilolite): adsorption isotherms,
attenuation of surface activity by humic and fulvic acids, Environ. Sci. Technol. thermodynamics and kinetics, Desalimation 409 (2017) 146–156.
40 (2006) 7757–7763. [42] H. Saleem, U. Rafique, R.P. Davies, Investigations on post-synthetically
[28] H. Dong, K. Ahmad, G. Zeng, Z. Li, G. Chen, Q. He, Y. Xie, Y. Wu, F. Zhao, Y. Zeng, modified UiO-66-NH2 for the adsorptive removal of heavy metal ions from
Influence of fulvic acid on the colloidal stability and reactivity of nanoscale aqueous solution, Micropor. Mesopor. Mat. 221 (2016) 238–244.
zero-valent iron, Environ. Pollut. 211 (2016) 363–369. [43] W. Ke, J. Gu, Y. Na, Efficient removal of Pb(II) and Cd(II) using NH2-
[29] J. Li, S. Zhang, C. Chen, G. Zhao, X. Yang, J. Li, X. Wang, Removal of Cu(II) and functionalized Zr-MOFs via rapid microwave-promoted synthesis, Ind. Eng.
fulvic acid by graphene oxide nanosheets decorated with Fe3O4 nanoparticles, Chem. Res. 56 (2017) 1880–1887.
ACS Appl. Mater. Inter. 4 (2012) 4991–5000. [44] L. Yu, R. Fu, S. Yue, X. Zhou, S.A. Baig, X. Xu, Multifunctional nanocomposites
[30] Y. Fang, J. Wen, G. Zeng, F. Jia, S. Zhang, Z. Peng, H. Zhang, Effect of mineralizing Fe3O4@SiO2-EDTA for Pb(II) and Cu(II) removal from aqueous solutions, Appl.
agents on the adsorption performance of metal-organic framework MIL-100 Surf. Sci. 369 (2016) 267–276.
(Fe) towards chromium(VI), Chem. Eng. J. 337 (2017) 532–540. [45] M. Babazadeh, R. Hosseinzadeh-Khanmiri, J. Abolhasani, E. Ghorbani-Kalhor, A.
[31] H. Young Kyu, H. Do-Young, C. Jong-San, J. Sung Hwa, S. You-Kyong, K. Hassanpour, Solid phase extraction of heavy metal ions from agricultural
Jinheung, V. Alexandre, D. Marco, S. Christian, F. Gérard, Amine grafting on samples with the aid of a novel functionalized magnetic metal–organic
coordinatively unsaturated metal centers of MOFs: consequences for catalysis framework, RSC Adv. 5 (2015) 19884–19892.
and metal encapsulation, Angew. Chem. 47 (2010) 4144–4148. [46] Y. Na, W. Ke, Y. Xia, Z. Li, Novel melamine modified metal-organic frameworks
[32] T. Hu, Q. Jia, S. He, S. Shan, H. Su, Y. Zhi, L. He, Novel functionalized metal- for remarkably high removal of heavy metal Pb(II), Desalination 430 (2018)
organic framework MIL-101 adsorbent for capturing oxytetracycline, J Alloy. 120–127.
Compd. 727 (2017) 114–122. [47] Z. Bai, L. Yuan, L. Zhu, Z. Liu, W. Shi, Introduction of amino groups into acid-
[33] L. Hajiaghababaei, A. Badiei, M.R. Ganjali, S. Heydari, Y. Khaniani, G.M. Ziarani, resistant MOFs for enhanced U(VI) sorption, J. Mater. Chem. A 3 (2014) 525–
Highly efficient removal and preconcentration of lead and cadmium cations 534.
from water and wastewater samples using ethylenediamine functionalized [48] M. Feng, P. Zhang, H.C. Zhou, V.K. Sharma, Water-stable metal-organic
SBA-15, Desalination 266 (2011) 182–187. frameworks for aqueous removal of heavy metals and radionuclides: a
[34] T. Ramanathan, F.T. Fisher, R.S.R. And, L.C. Brinson, Amino-functionalized review, Chemosphere 209 (2018) 783–800.
carbon nanotubes for binding to polymers and biological systems, Chem. [49] J.R. Rangelmendez, M. Streat, Adsorption of cadmium by activated carbon
Mater. 17 (2011) 1290–1295. cloth: influence of surface oxidation and solution pH, Water Res. 36 (2002)
[35] Y. Na, W. Ke, L. Wang, Z. Li, Amino-functionalized MOFs combining ceramic 1244–1252.
membrane ultrafiltration for Pb (II) removal, Chem. Eng. J. 306 (2016) 619– [50] X. Li, Z. Wang, Q. Li, J. Ma, M. Zhu, Preparation, characterization, and
628. application of mesoporous silica-grafted graphene oxide for highly selective
[36] F. Zhang, S. Jing, J. Yan, Y. Fu, Y. Zhong, W. Zhu, Facile synthesis of MIL-100(Fe) lead adsorption, Chem. Eng. J. 273 (2015) 630–637.
under HF-free conditions and its application in the acetalization of aldehydes [51] S.G. Wang, X.F. Sun, X.W. Liu, W.X. Gong, Chitosan hydrogel beads for fulvic
with diols, Chem. Eng. J. 259 (2015) 183–190. acid adsorption: behaviors and mechanisms, Chem. Eng. J. 142 (2008) 239–
[37] Z. Ai, Z. Gao, L. Zhang, W. He, J. Yin, Core-shell structure dependent reactivity of 247.
Fe@Fe2O3 nanowires on aerobic degradation of 4-chlorophenol, Environ. Sci. [52] W.S.W. Ngah, A. Musa, Adsorption of humic acid onto chitin and chitosan, J.
Technol. 47 (2013) 5344–5352. Applied. Polym. Sci. 69 (1998) 2305–2310.
[38] B.J. Zhu, X.Y. Yu, Y. Jia, F.M. Peng, B. Sun, M.Y. Zhang, T. Luo, J.H. Liu, X.J. Huang, [53] M. Hasan, A.L. Ahmad, B.H. Hameed, Adsorption of reactive dye onto cross-
Iron and 1,3,5-benzenetricarboxylic metal-organic coordination polymers linked chitosan/oil palm ash composite beads, Chem. Eng. J. 136 (2008) 164–
prepared by solvothermal method and their application in efficient As(V) 172.
removal from aqueous solutions, J. Phys. Chem. C. 116 (2012) 8601–8607. [54] M.N.V.R. Kumar, A review of chitin and chitosan applications, React. Funct.
[39] N. Gedam, N.R. Neti, M. Kormunda, J. Subrt, S. Bakardjieva, Novel lead dioxide- Polym. 46 (2000) 1–27.
graphite-polymer composite anode for electrochemical chlorine generation,
Electrochim. Acta 169 (2015) 109–116.
[40] H. Jing, Y. Lu, G. Luo, Ca(II) imprinted chitosan microspheres: an effective and
green adsorbent for the removal of Cu(II), Cd(II) and Pb(II) from aqueous
solutions, Chem. Eng. J. 244 (2014) 202–208.

You might also like